24.10.2013 Views

FORESTRY - Institute of Foresters of Australia

FORESTRY - Institute of Foresters of Australia

FORESTRY - Institute of Foresters of Australia

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

The IFA thanks the following sponsors:<br />

Gold Sponsors<br />

Silver Sponsors<br />

Conference Dinner Sponsor Field Trip Sponsors Exhibitors<br />

• Queensland Commodity Exports • Healthcare Insurance Limited<br />

• Griffi th University - Centre for • Narromine Transplants<br />

Forestry and Horticultural Research • Private Forestry Southern Queensland<br />

• National Forestry Masters Program<br />

Advertising Sponsor • Aerial Acquisitions<br />

• SciWest Consulting • LREye Imagery<br />

Poster Prize Student Support<br />

• Gary Bacon • Alan Irving<br />

Supported through DERM Forest Products, a<br />

commercial business group <strong>of</strong> the Department<br />

<strong>of</strong> Environment and Resource Management<br />

PROCEEDINGS OF THE INSTITUTE OF FORESTERS OF AUSTRALIA 2009 CONFERENCE<br />

IFA 2009<br />

<strong>FORESTRY</strong>: A CLIMATE OF CHANGE<br />

THE INSTITUTE OF FORESTERS OF AUSTRALIA 2009 CONFERENCE<br />

Proceedings<br />

6 - 10 September 2009 Caloundra, Queensland, <strong>Australia</strong>


Forestry: a climate <strong>of</strong> change<br />

Pre-Conference Proceedings <strong>of</strong> the<br />

Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong><br />

Caloundra, Queensland, <strong>Australia</strong><br />

6 – 10 September 2009<br />

Edited by:<br />

Robert Thistlethwaite, David Lamb, Russell Haines<br />

August 2009


Published by:<br />

The University <strong>of</strong> Queensland<br />

With the assistance <strong>of</strong>:<br />

Dr John Herbohn and Annerine Bosch, School <strong>of</strong> Integrative Systems, The University <strong>of</strong><br />

Queensland, Brisbane, Qld<br />

Edited by:<br />

Robert Thistlethwaite, David Lamb, Russell Haines<br />

National Library <strong>of</strong> <strong>Australia</strong> Cataloguing-in-Publication entry:<br />

Forestry: a climate <strong>of</strong> change. Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the <strong>Institute</strong> <strong>of</strong><br />

<strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, Queensland, <strong>Australia</strong>, 6 – 10 September 2009<br />

ISBN: 978-0-646-52024-7<br />

1. Trees-Forest Governance-Congresses.<br />

2. Sustainable Forestry-Climate Change-Congresses.<br />

I Thistlethwaite, R. II. <strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong><br />

© <strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong> 2009<br />

This publication and its individual papers should be cited as:<br />

1. General:<br />

Thistlethwaite, R., Lamb, D and Haines, R. (eds) 2009. Forestry: a climate <strong>of</strong> change. Proc.<br />

IFA Conference, Caloundra, Queensland, <strong>Australia</strong>, 6 – 10 September 2009.<br />

2. Specific:<br />

Feary, S., Kanowski, P., Baker, R. and Altman, J. 2009. Managing forest country: Aboriginal<br />

<strong>Australia</strong>ns and forests. In: Forestry: a climate <strong>of</strong> change, Thistlethwaite, R., Lamb, D. and<br />

Haines, R. (eds). pp. 422–432. Proc. IFA Conf., Caloundra, Queensland, <strong>Australia</strong>, 6 – 10<br />

September 2009.<br />

Cover design:<br />

Cameron Coward<br />

Electronic layout:<br />

Annerine Bosch, Alison Williams, Robert Thistlethwaite


Forestry: a climate <strong>of</strong> change<br />

PRE‐CONFERENCE PROCEEDINGS OF THE BIENNIAL CONFERENCE OF THE<br />

INSTITUTE OF FORESTERS OF AUSTRALIA<br />

CALOUNDRA, QUEENSLAND, 6–10 SEPTEMBER 2009<br />

TABLE OF CONTENTS<br />

KEYNOTE ADDRESS<br />

John Innes The role <strong>of</strong> sustainable forestry in climate change 1‐8<br />

David Brand The changing role <strong>of</strong> forests: state <strong>of</strong><br />

development <strong>of</strong> markets for ecosystem services<br />

9‐18<br />

Kathryn Adams Capturing the value <strong>of</strong> innovation: R&D for<br />

forest and wood products industries<br />

19‐27<br />

SESSION 1 Climate change panel<br />

SESSION 2 Mitigation opportunities and risks<br />

Rodney Keenan Native forest management options for climate<br />

mitigation in <strong>Australia</strong> and PNG<br />

28‐35<br />

Jody Bruce, Michael Battaglia, <strong>Australia</strong>’s plantation vulnerability to climate<br />

36<br />

Libby Pinkard<br />

change<br />

Paul Dargusch, Steve Harrison, Environmental markets and <strong>Australia</strong>n forestry: 37‐43<br />

John Herbohn<br />

the emergence <strong>of</strong> a new period <strong>of</strong> industry<br />

development<br />

Scott Arnold Lessons learned from building estate‐scale<br />

carbon accounting systems<br />

44‐53<br />

SESSION 3A Delivery <strong>of</strong> environmental and ecosystem services<br />

John Tredinnick and Jodie Carbon stores and bioenergy supplies:<br />

54‐63<br />

Mason<br />

opportunities to develop sustainable plantation<br />

industries in low rainfall agricultural landscapes<br />

Himlal Baral, Sabine Kasel, GIS‐based classification, mapping and valuation 64‐71<br />

Rodney Keenan, Julian Fox, <strong>of</strong> ecosystem services in production landscapes:<br />

Nigel Stork<br />

a case study <strong>of</strong> the Green Triangle region <strong>of</strong><br />

south‐eastern <strong>Australia</strong><br />

David Lee, John Huth, David Selecting species for carbon sequestration under 72‐82<br />

Osborne, Bruce Hogg<br />

climate change scenarios in sub‐tropical<br />

Queensland<br />

Frank Batini Responses to a drying climate in the northern<br />

Jarrah forest<br />

83‐90<br />

SESSION 3B Forestry Education<br />

Lyndall Bull, Peter Kanowski The National Forestry Masters Program: a new<br />

era <strong>of</strong> collaboration in pr<strong>of</strong>essional forestry<br />

education<br />

91‐96<br />

Antoinette Hewitt: Education and training pathways 97<br />

Chris Weston and Katherine Forestry is a handy word but not the right one for<br />

Whittaker<br />

attracting students<br />

Matthew Doig Training operations staff in the field: challenges<br />

and opportunities<br />

98‐102<br />

103‐110


SESSION 4 Forestry education panel<br />

SESSION 5A Resource assessment<br />

Phil West: Practical use <strong>of</strong> 3P sampling in forest inventory 111‐119<br />

Andrew Haywood and Michael Estimating forest characteristics in young<br />

120‐130<br />

Sutton<br />

Victorian ash regrowth forests using field plots<br />

and airborne laser scanning data<br />

Andrew Haywood and Christine Object‐based analysis <strong>of</strong> forest stand delineation 131‐141<br />

Stone<br />

on high spatial resolution imagery using open<br />

source s<strong>of</strong>tware<br />

Adam Gerrand, Erik Lindquist, The Global Forest Resource Assessment Remote 142‐150<br />

Mette Wilkie and Rod Keenan Sensing Survey<br />

Hugh Stone and David Wood Sustainable pole supply project 151‐160<br />

SESSION 5B Innovation in utilisation<br />

Russell Washusen, Andrew Processing performance and sawn product<br />

161‐169<br />

Morrow, Dung Ngo, Graeme recovery from thinned native forest regrowth logs<br />

Siemon, Tim Wardlaw, Mike<br />

Ryan, Martin Linehan and Daniel<br />

Tuan<br />

from southern <strong>Australia</strong><br />

Mark Brown Diesel‐electric hybrid drive technology for reduced<br />

fuel consumption and carbon emissions in forest<br />

operations<br />

170‐174<br />

Russell Wa+shusen Modelling mill door sawlog prices <strong>of</strong> <strong>Australia</strong>n<br />

plantation‐grown eucalypts with CSIROMILL<br />

175‐184<br />

Mark Brown Reduced fuel use in forestry transportation<br />

through the use <strong>of</strong> higher productivity vehicles<br />

(HPV)<br />

185‐190<br />

Martin Strandgard Onboard systems for <strong>Australia</strong>n forest operations 191‐197<br />

SESSION 5C Wood<br />

Gregory Nolan Timber from native forest and plantation<br />

eucalypts ‐ Users will quickly find that they are not<br />

the same thing<br />

198‐207<br />

Kevin Harding, Terry Copley, Wood property variation in mature Queensland 208‐216<br />

Marks Nester, Kerrie Catchpoole,<br />

Anton Zbonak<br />

exotic pine silviculture experiments<br />

Robert Evans, Josh Bowden, Modelling the influence <strong>of</strong> climate change on 217‐224<br />

Kevin Harding, Terry Copley,<br />

Kerrie Catchpoole, Marks Nester<br />

plantation wood properties<br />

Sasha Alexander Competition and collaboration in the <strong>Australia</strong>n<br />

timber furniture industry ‐ a value chain approach<br />

toward higher value creation for local and global<br />

markets and pathways for sustainable forest<br />

resource use<br />

225‐233<br />

Kevin Harding, Terry Copley, Caribbean Pine graded sawn recovery variation in 234‐243<br />

Marks Nester, Kerrie Catchpoole, a mature spacing by thinning experiment in<br />

Anton Zbonak<br />

Queensland<br />

SESSION 5D Biological performance<br />

Peter St. Clair Rehabilitation <strong>of</strong> forests in decline: Mt. Lindsay<br />

State Forest<br />

244‐254<br />

Simon Lawson Using a process‐based model, ParopSys, to predict<br />

the impact <strong>of</strong> climate change on the eucalypt<br />

plantation pest Paropsis atomaria<br />

255‐262<br />

Tim Smith and Ge<strong>of</strong>f Pegg Impact <strong>of</strong> tree boron status on tree growth and<br />

susceptibility to Quambalaria shoot blight<br />

(Quambalaria pitereka) in Corymbia sp.<br />

263‐270


Vivien de Remy de Courcelles,<br />

Bhupinderpal‐Singh, Mark<br />

The influence <strong>of</strong> climate change on soil respiration<br />

with Eucalyptus saligna<br />

271‐279<br />

Adams<br />

SESSION 5E Fire protection<br />

Roger Underwood Bushfire governance in <strong>Australia</strong> in 2009: a note to<br />

future historians<br />

280‐284<br />

Miguel Cruz and Jim Gould National Fire Behaviour Prediction System 285‐291<br />

Jennifer Hollis and Lachlan Woody fuel consumption and carbon in the<br />

292‐302<br />

McCaw<br />

changing climate <strong>of</strong> <strong>Australia</strong><br />

Mal Tonkin Victorian Bushfires 2009 ‐ A plantation company’s<br />

experience<br />

303‐308<br />

SESSION 5F Silvicultural aspects<br />

David Doley Silviculture for a climate <strong>of</strong> change 309‐320<br />

Paul Warburton, Paul Macdonell, Improving grey gums to sequester carbon on 321‐329<br />

Jeremy Brawner, John Huth,<br />

David Lee<br />

marginal sites in sub‐tropical <strong>Australia</strong><br />

Jeremy Brawner, David Bush, The utilisation <strong>of</strong> red mahogany for high value 330‐339<br />

Paul Macdonell, David Boden,<br />

Simon Potter, Paul Warburton,<br />

Paul Clegg<br />

plantation forestry in the tropics ‐<br />

Don Willis Growing teak for the roots – the jiffy solution 340<br />

SESSION 6A Reduced emission from deforestation and degradation<br />

Zoe Harkin Lessons learned from implementation <strong>of</strong> REDD – a<br />

collection <strong>of</strong> analogies, metaphors and clichés<br />

341‐346<br />

Grahame Applegate Developing a REDD scheme for post 2012: the<br />

Kalimantan forests and climate partnerships<br />

347‐354<br />

Majella Clarke Reducing emissions from deforestation and<br />

degradation (REDD) in Lao Peoples’ Democratic<br />

Republic<br />

355‐364<br />

Andrew Morton and Blair<br />

Freeman<br />

REDD in Asia Pacific: Challenges and implications 365‐373<br />

Majella Clarke A characterisation <strong>of</strong> the current market for<br />

reducing emissions from deforestation and<br />

degradation (REDD) and forest carbon <strong>of</strong>fsets<br />

374‐384<br />

SESSION 6B Changing needs in forest governance and management<br />

Aidan Flanagan Tasmania’s New Forest Industry Plan: Reshaping<br />

the forestry agenda<br />

385‐394<br />

Sean Ryan Reshaping the forestry agenda in Queensland 395‐402<br />

David Williams Improving business performance in s<strong>of</strong>twood<br />

plantations<br />

403‐412<br />

Dick McCarthy and Gabriel Samol Addressing climate change impacts on the<br />

business <strong>of</strong> PNG forestry: Implications for<br />

<strong>Australia</strong> and PNG<br />

413‐421<br />

SESSION 6C Indigenous interests and aspirations<br />

Sue Feary, Peter Kanowski, Managing forest country: engaging indigenous 422‐432<br />

Richard Baker and Jon Altman communities and the forests sector<br />

Mark Annandale Community consultation and indigenous forestry<br />

opportunities in Cape York Peninsula<br />

433‐440<br />

Peter Shepherd A Forester’s Guide to working with Indigenous<br />

<strong>Australia</strong>ns: An introduction<br />

441‐446<br />

Braden Jenkin Michael Blyth, An integrated approach to species regime<br />

447‐457<br />

Peter Kanowski and Don Yakuma selection in PNG: beginning with local people's<br />

needs


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 1<br />

THE ROLE OF SUSTAINABLE <strong>FORESTRY</strong> IN CLIMATE CHANGE<br />

John Innes 1<br />

ABSTRACT<br />

Forests will only help mitigate climate change if they can survive the changes in<br />

climate. Consequently, any attempts to consider mitigation must also consider<br />

adaptation. A recent report from the Collaborative Partnership on Forests has examined<br />

adaptation in detail. It is very difficult, if not impossible, to take a prescriptive approach<br />

to forest management when considering climate change. Instead, there are a wide range<br />

<strong>of</strong> possible actions that could be taken, and the precise combination will depend on the<br />

local conditions, the anticipated changes in climate, and the objectives for the forest<br />

(which may themselves have to be changed because <strong>of</strong> climate change). A key issue is<br />

that forestry regulations and codes <strong>of</strong> practice must be sufficiently flexible to allow<br />

foresters to try different strategies appropriate to their particular situations.<br />

INTRODUCTION<br />

For those living at higher latitudes, particularly in the northern hemisphere, there is a great deal <strong>of</strong><br />

evidence that the climate is changing. Meteorological records indicate that temperatures have<br />

increased significantly, with the most obvious change being a reduction in the coldest extremes during<br />

winter. Precipitation records are more variable and no generalization is possible. The changes <strong>of</strong><br />

temperature have been accompanied by changes in both the physical and biological phenomena<br />

present in the region. Glaciers are melting in most parts <strong>of</strong> the world, permafrost is melting, slope and<br />

stability has decreased. Tree-lines are moving upwards and northwards (there appears to be no<br />

documented evidence <strong>of</strong> southward extensions <strong>of</strong> tree-lines in the southern hemisphere), and<br />

aboriginal peoples are reporting that species never previously seen are now occurring. In some cases,<br />

these are invasive species that are spreading, such as the European starling (Sturnus vulgaris), in other<br />

cases a northward shift in the distributions <strong>of</strong> some species is clearly occurring.<br />

It is natural to assume that these changes are all linked to the phenomenon known as ‘global warming’,<br />

and the recent reports <strong>of</strong> the Intergovernmental Panel on Climate Change frequently cite such evidence<br />

(Intergovernmental Panel on Climate Change 2007), but some argue that they could equally well be<br />

natural changes following on from the ‘Little Ice Age’ that peaked in the 17 th and 18 th centuries.<br />

During this period, the Thames River in the UK froze over regularly, something that would be<br />

considered impossible today. Many glaciers reached their maximum Holocene extent, and many<br />

marginal agricultural areas had to be abandoned (Grove, 1988). In this paper, I do not wish to address<br />

the relative merits <strong>of</strong> the arguments surrounding the causes <strong>of</strong> today’s changing climate. That is dealt<br />

with by others at this conference. Instead, I focus on the implications for today’s foresters <strong>of</strong> a<br />

changing climate, regardless <strong>of</strong> cause.<br />

SOME IMPACTS<br />

We <strong>of</strong>ten speak about climate change as if it is a new phenomenon that will somehow affect us only in<br />

the future. It is not new: climate change has always occurred, and the implications <strong>of</strong> the possibility <strong>of</strong><br />

accelerated warming have been discussed in forestry for over 20 years (e.g., Booth and McMurtrie<br />

1988). However, in the northwest <strong>of</strong> the North American continent, there is already evidence <strong>of</strong><br />

substantial changes. These include:<br />

• A major outbreak <strong>of</strong> the Mountain Pine Beetle (Dendroctonus ponderosae) in central British<br />

Columbia that has resulted in the death <strong>of</strong> ca. 14.5 million ha <strong>of</strong> lodgepole pine (Pinus<br />

contorta), and the sudden availability <strong>of</strong> over 620 million m 3 <strong>of</strong> logs<br />

1 Pr<strong>of</strong>essor John L.Innes, Faculty <strong>of</strong> Forestry, University <strong>of</strong> British Columbia, 2045-2424 Main Mall, Vancouver,<br />

Canada V6T 1Z4. Tel: (+1) 604 822 6761. Email: john.innes@ubc.ca.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 2<br />

• Major outbreaks <strong>of</strong> the spruce beetle (Dendroctonus rufipennis) that has devastated large areas<br />

<strong>of</strong> white spruce (Picea glauca) in the Kenai Peninsula <strong>of</strong> Alaska (600,000 ha) and southwest<br />

Yukon (390,000 ha)<br />

• A series <strong>of</strong> droughts in coastal British Columbia that appears to be causing increased mortality<br />

in western red cedar (Thuja plicata)<br />

• Reductions in the length <strong>of</strong> the winter operations season, when snow and frozen ground makes<br />

forestry operations feasible and roads are passable.<br />

The bark beetle outbreaks remain controversial – they can be related to both past management<br />

activities (or lack there<strong>of</strong>) that resulted in the widespread occurrence <strong>of</strong> mature and old stands <strong>of</strong> pine<br />

and spruce. However, it is likely that changes in climatic conditions enabled the epidemics to become<br />

as severe as they have (Berg et al. 2006). In the case <strong>of</strong> the mortality, which is evident in a number <strong>of</strong><br />

species in western North America, the data are quite limited, and the precise role <strong>of</strong> climate change is<br />

difficult to determine (van Mantgem et al. 2009).<br />

FUTURE CHANGES<br />

If model predictions are correct, it is very likely that many parts <strong>of</strong> the Earth’s surface will experience<br />

significant changes in climate over the next 100 years. It is currently not possible to specify what these<br />

changes will be, as they will depend to a large extent on the manner in which the world develops and<br />

responds to climate change over the same period. As a result, modellers have developed a number <strong>of</strong><br />

scenarios based on different development and mitigation options. However, there is evidence that<br />

already carbon dioxide emissions are exceeding even the most pessimistic <strong>of</strong> these options. The choice<br />

<strong>of</strong> scenario is therefore a cause <strong>of</strong> considerable uncertainty.<br />

The predicted emissions <strong>of</strong> carbon dioxide and other greenhouse gases are fed into Global Change<br />

Models. There are a number <strong>of</strong> these, and the results they produce differ. Early models were very<br />

crude, but they have gradually become more and more sophisticated and better at reproducing the<br />

climate change that has occurred over the past 100 years. The predicted future climate for a given<br />

scenario does however vary depending on the choice <strong>of</strong> model.<br />

Global Change Models, by their very nature, work at coarse global scales, and cannot take into<br />

account fine-scale topography. Consequently, they are likely to be very poor in predicting climate<br />

change in mountainous areas. Also, if there are any local effects, such as land and sea breezes or<br />

anabatic and katabatic winds, these are unlikely to be accounted for. Consequently, downscaling is<br />

necessary, whereby the output from the global models is adjusted for local conditions. This is a further<br />

source <strong>of</strong> uncertainty.<br />

The net effect <strong>of</strong> these three major sources <strong>of</strong> uncertainty is that it is not possible to predict with any<br />

degree <strong>of</strong> confidence the future climate at a specific place. Yet such predictions are precisely what the<br />

forester, thinking <strong>of</strong> the next 100 years or so, needs. Much the same problem is faced by the engineer:<br />

structures being designed today need to be able to cope with climate change while remaining<br />

economically feasible (Hallegatte 2009).<br />

MITIGATION VS. ADAPTATION<br />

Given such uncertainty, what should a forester do? There are two major strategies that have been<br />

discussed in detail: mitigation and adaptation. Mitigation presupposes a causal link between<br />

greenhouse gas emissions and climate change and involves efforts by foresters to reduce the amount <strong>of</strong><br />

greenhouse gases, particularly carbon dioxide, in the atmosphere through selected forestry activities. It<br />

includes increased sequestration, through an increase in global forest cover and the increased use <strong>of</strong><br />

wood products in long-lived products such as houses and furniture, reduced emissions from forests<br />

through reductions in deforestation and forest degradation and reductions in the carbon dioxide<br />

released by forest fires, and substitution <strong>of</strong> substances associated with greenhouse gas emissions, such<br />

as coal, concrete and steel, by wood and wood composites.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 3<br />

The other major strategy is adaptation. This presumes that climate change is inevitable (regardless <strong>of</strong><br />

cause), and involves the adaptation <strong>of</strong> forest practices to enable forests and forestry to cope with these<br />

changes. Clearly, this is potentially problematic, as we do not know the nature or extent <strong>of</strong> future<br />

changes in climate. Consequently, we need to identify changes in management strategy that will be<br />

beneficial for forests and forestry whatever the future changes in climate are likely to be, or even in<br />

the unlikely event that there is no change.<br />

Is there such a “win-win” strategy? Fortunately there is, in the form <strong>of</strong> sustainable forest management.<br />

Many <strong>of</strong> the criteria associated with sustainable forest management, together with their associated<br />

indicators, could reduce the vulnerability <strong>of</strong> forests to climate change. There are a number <strong>of</strong> different<br />

definitions <strong>of</strong> sustainable forest management and also a number <strong>of</strong> criteria and indicator schemes.<br />

Here, I focus on the Montreal Process, as this is the scheme used in <strong>Australia</strong>, and is also used for<br />

more than 85% <strong>of</strong> the temperate and boreal forest area.<br />

THE MONTREAL PROCESS CRITERIA<br />

The Montreal Process contains seven criteria, all <strong>of</strong> which are relevant to climate change. However,<br />

climate change is creating some questions about the continued relevance <strong>of</strong> current interpretations <strong>of</strong><br />

the criteria. For example, ‘Maintain biodiversity’ is an admirable concept in a static environment, but<br />

may not be so applicable in the context <strong>of</strong> climate change. Within the context <strong>of</strong> climate change, this<br />

criterion needs to be interpreted as maintaining biodiversity appropriate for the site conditions. The<br />

emphasis then is on enabling the forest to adapt as climate changes, rather than preventing any<br />

changes that may represent a part <strong>of</strong> that adaptation. This will require some flexibility on the part <strong>of</strong><br />

managers, as highly desirable tree species may be replaced by less desirable but better adapted tree<br />

species.<br />

Having the flexibility to make such decisions is critical. Forestry institutions, broadly defined, include<br />

the governance structures, formal and informal education systems, pr<strong>of</strong>essional bodies and other actors<br />

involved in forests. Many <strong>of</strong> these are traditionally very conservative, with the result that adaptation to<br />

the rapid changes in forests induced by climate may be difficult. For example, in some jurisdictions<br />

(particularly in developed countries), governance systems are highly prescriptive, with detailed<br />

regulations on exactly how all aspects <strong>of</strong> forestry are to be undertaken. These may be amongst the<br />

most inflexible to adaptation. Conversely, where a results-based approach to forestry has been<br />

adopted, the focus is on particular management objectives, with the management options being taken<br />

to achieve these objectives being largely left to the manager to decide. Such systems will be much<br />

more flexible.<br />

Table 1 (appended) reveals some <strong>of</strong> the issues that may be involved, including both direct and indirect<br />

effects. For example, the inadequate communication between forest scientists and policy makers is<br />

partly a result <strong>of</strong> the rapidity <strong>of</strong> the changes in knowledge, but also a function <strong>of</strong> the lack <strong>of</strong><br />

appropriate institutions and the early failures <strong>of</strong> the forestry community to be actively involved with<br />

policy makers in this area. In addition, it should be noted that in many cases the institutional<br />

framework in other sectors may have a greater impact on the ability <strong>of</strong> the forest sector to adapt to<br />

climate change than that <strong>of</strong> the forest sector itself. Promotion <strong>of</strong> bi<strong>of</strong>uel production from agricultural<br />

products, for example, already has increased forest conversion into oil-palm plantations in Asia (Koh<br />

and Wilcove 2007, 2008). Free trade agreements may also affect directly or indirectly the local<br />

capacity to deal with climate change, in some cases with positive effects, in others not.<br />

STRATEGIES FOR FORESTERS<br />

Every forest manager (or forest policy maker) faces a unique set <strong>of</strong> problems, before climate change is<br />

even considered. As a result, the actions taken to adapt to climate change will vary from place to<br />

place, and no generalizations are possible. In fact, prescriptive approaches to climate change would be<br />

highly misleading, and most likely erroneous. While it is tempting to provide broad recommendations<br />

applicable to, for example, boreal forests, this would be disingenuous. The managers <strong>of</strong> forests in<br />

Alaska, Canada, Fennoscandia and Russia have very different objectives. Even within these units,<br />

there may be major differences, such as between the manager <strong>of</strong> a forest biological reserve and that <strong>of</strong><br />

a forest biomass plantation. Managers must be given the freedom to choose which management


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 4<br />

options are the most appropriate for their particular piece <strong>of</strong> land. Policy makers must ensure that<br />

appropriate mechanisms exist to enable this flexibility. This suggests strongly that forest policy must<br />

move away from prescriptive approaches to results-based approaches, allowing managers the freedom<br />

to adopt the best way to achieve those results, given local circumstances.<br />

The majority <strong>of</strong> tropical production forests are not being managed (sustainably), almost 20 years after<br />

ITTO defined the needs <strong>of</strong> sustainable forest management (e.g., International Tropical Timber<br />

Organization 1992, 1993). From the perspective <strong>of</strong> mitigation through reductions in deforestation and<br />

forest degradation, this represents a major challenge for the global forest policy community.<br />

Despite years <strong>of</strong> negotiation, the poor uptake <strong>of</strong> sustainable forest management has not been<br />

satisfactorily addressed by groups such as the United Nations Forum on Forests (Persson 2005,<br />

Humphreys 2006, Capistrano et al. 2007). Similarly, illegal logging remains rife, contributing to<br />

emissions from deforestation and forest degradation (Tacconi 2007). There is a need for effective<br />

policy initiatives to address these issues.<br />

While recommendations for strategies such as improved fire-fighting capabilities or the ability to<br />

combat insect infestations are possible, such actions would be appropriate in some circumstances, but<br />

not others. Some areas might indeed face such problems, but others will not. Consequently, general<br />

prescriptions run the risk <strong>of</strong> encouraging scarce resources to be allocated inappropriately.<br />

The key message is that although every situation is different, there are both procedures and<br />

management options that will help forest managers adapt to climate change. To date, few actions have<br />

been taken, and there is a major need to educate foresters <strong>of</strong> both the need to adapt to climate change<br />

and the options that are available to them. They need to be given the freedom to choose the options<br />

that are most applicable to their particular situations. However, there is a need to recognize that the<br />

definition <strong>of</strong> sustainable forest management is continuously evolving, and managers must be prepared<br />

to address these “shifting goal posts”.<br />

There is an urgent need to ensure that forests utilized for timber exploitation are managed sustainably.<br />

This is a task for the international forest policy community, but their actions to date suggest that other<br />

mechanisms may need to be involved. To a certain extent, this is already occurring. The current<br />

international interest in illegal logging and REDD occurred despite, rather than because <strong>of</strong>, the forest<br />

policy community.<br />

The Intergovernmental Panel on Climate Change was largely independent <strong>of</strong> the forest policy<br />

community, as was the Stern report (Stern 2006), but their reports contain many references to forests<br />

and their adaptation to climate change. There is a real danger <strong>of</strong> the forest policy community<br />

becoming increasingly distanced from the emerging mainstream forest policy issues, but climate<br />

change has provided an impetus that is now driving the forestry agenda.<br />

The feasibility and/or effectiveness <strong>of</strong> an adaptation practice will depend on the characteristics <strong>of</strong> the<br />

forested system, such as its soil, existing biodiversity, pre-existing natural and anthropogenic stresses,<br />

the status <strong>of</strong> the local ecosystem (e.g., early to late successional), and on the characteristics <strong>of</strong> the<br />

social, economic, and political system interacting with, or dependent on, the ecosystem.<br />

Increasing the adaptive capacity <strong>of</strong> stakeholders may involve improving access to appropriate<br />

information and technologies to address the impacts and opportunities <strong>of</strong> the changing climate,<br />

strengthening social networks and traditional forest management and governance institutions,<br />

increasing food security, alleviating poverty, and policy changes to expand or enhance current<br />

management opportunities. For example, existing management practices may need to be applied in<br />

new and different situations (such as prescribed fire outside <strong>of</strong> the ‘traditional’ fire management<br />

seasons).<br />

Understanding the opportunities and barriers that affect successful implementation <strong>of</strong> management<br />

practices to enhance adaptive capacity is a critical need. Enhancements or barriers to adaptation may<br />

include biological, physical, economic, social, cultural, institutional, or technological conditions. As<br />

such, it will be necessary to adopt cross-sectoral approaches to the issues.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 5<br />

CONCLUSIONS<br />

Climate change poses another set <strong>of</strong> uncertainties for foresters trying to plan over the medium- to<br />

long-term. These uncertainties should not be taken an as excuse for taking no action. Instead, foresters<br />

should choose those actions that are most likely to benefit sustainable forest management while at the<br />

same time maintaining the ability <strong>of</strong> the forest to adapt to climate change. No specific prescriptions are<br />

possible, partly because <strong>of</strong> the uniqueness <strong>of</strong> every situation, and partly because the future changes in<br />

climate remain unknown.<br />

Forestry is increasingly regulated, especially in a country such as <strong>Australia</strong>. It is important that these<br />

regulations do not hinder foresters in managing their forests to adapt to climate change. If the<br />

regulations are too prescriptive, then the ability <strong>of</strong> foresters to allow their forests and forestry practices<br />

to adapt to climate change may be compromised. At the same time, many groups see highly<br />

prescriptive regulations as the only way to control forestry. An alternative however is a results-based<br />

approach, where there is general agreement on the endpoint being sought and foresters are allowed the<br />

flexibility to try different paths to that point. This, however, requires a well-educated forestry<br />

workforce that is able to recognize the potential implications not only <strong>of</strong> their practices, but <strong>of</strong> the<br />

changes in the environment (specifically the climate) that are occurring around them.<br />

REFERENCES<br />

Berg, E.E, Henry, J.D., Christopher, L., Fastie, C.L, Volder, A.D. and Matsuoka, S.M. 2006. Spruce beetle<br />

outbreaks on the Kenai Peninsula, Alaska, and Kluane National Park and Reserve, Yukon Territory:<br />

Relationship to summer temperatures and regional differences in disturbance regimes. Forest Ecology<br />

and Management 227(3) 219–232<br />

Booth, T.H. and McMurtrie, R.E. 1988. Climate change and Pinus radiata plantations in <strong>Australia</strong>. In: Pearman,<br />

G.I. (ed.) Greenhouse: Planning for climate change. CSIRO, Melbourne, pp. 534-545.<br />

British Columbia Ministry <strong>of</strong> Forests and Range (BCMOF) 2006. Preparing for climate change: Adapting to<br />

impacts on British Columbia’s forest and range resources. Province British Columbia, Victoria, British<br />

Columbia, Canada. 80pp.<br />

Capistrano, D., Kanninen, M., Guariguata, M.R., Barr, C., Sunderland, T.C.H., Raitzer, D.A. 2007. Revitalizing<br />

the United Nations Forum on Forests: critical issues and ways forward. Centre for International Forestry<br />

Research, Bogor, Indonesia.<br />

Chapin, F.S. III, Peterson, G., Berkes, F., Callaghan, T.V., Angelstam, P., Apps, M., Beier, C., Bergeron, Y.,<br />

Crépin, A.-S., Danell, K., Elmqvist, T., Folke, C., Forbes, B., Fresco, N., Juday, G., Niemelä, J.,<br />

Shvidenko, A. & Whiteman, G. 2004. Resilience and vulnerability <strong>of</strong> northern regions to social and<br />

environmental change. Ambio 33(6): 344–349.<br />

FAO 2008. Conference Report for “Adaptation <strong>of</strong> forests and forest management to changing climate with<br />

emphasis on forest health: A review <strong>of</strong> science, policies and practices. Umeå, Sweden, 25–28 August<br />

2008. UN Food and Agriculture Organization, Rome, Italy.<br />

Grove, J.M. 1988. The Little Ice Age. Routledge. London and New York, 498 p.<br />

Hallegatte, S. 2009. Strategies to adapt to an uncertain climate change. Global Environmental Change – Human<br />

and Policy Dimensions 19 (2): 240 – 247.<br />

Humphreys, D. 2006. Logjam. Deforestation and the crisis in global governance. Earthscan Publications,<br />

London, UK.<br />

Intergovernmental Panel on Climate Change. 2007. Impacts, adaptation and vulnerability: Contribution <strong>of</strong><br />

Working Group II to the Fourth Assessment Report <strong>of</strong> the Intergovernmental Panel on Climate Change.<br />

Parry, M.L., Canziani, O.F., Palutik<strong>of</strong>, J.P., van der Linden, P.J. and Hanson, C.E. (eds.). Cambridge<br />

University Press, Cambridge, UK. p. 976.<br />

International Tropical Timber Organization 1992. ITTO guidelines for the sustainable management <strong>of</strong> natural<br />

tropical forests. International Tropical Timber Organization, Yokohama, Japan. p. 18.<br />

International Tropical Timber Organization 1993. ITTO guidelines for the establishment and sustainable<br />

management <strong>of</strong> planted tropical forests. International Tropical Timber Organization, Yokohama, Japan.<br />

p. 39.<br />

International Tropical Timber Organization 2006. Status <strong>of</strong> tropical forest management 2005. ITTO Technical<br />

Series No. 24. International Tropical Timber Organization, Yokohama, Japan. p. 302.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 6<br />

Johnston, M., Williamson, T., Price, D., Wellstead, A., Gray, P., Scott, D., Askew, S. & Webber, S. 2006.<br />

Adapting forest management to the impacts <strong>of</strong> climate change in Canada. BIOCAP Research<br />

Integration Program Synthesis Paper. 96 pp.<br />

Kellomaki, S., Peltola, H., Bauwens, B., Dekker, M., Mohreh, F., Badeck, F.W., Gracia, C., Sanchez, A., Pla, E.,<br />

Sabate, S., Lindner, M. & Pussinen, A. 2005. European Mitigation and Adaptation Potentials:<br />

Conclusions and Recommendations. In: Kellomaki, S. & Leinonen, S. (eds). SilviStrat: Management <strong>of</strong><br />

European Forests under changing climatic conditions. Tiedonantoja Research Notes No. 163. Faculty <strong>of</strong><br />

Forestry, University <strong>of</strong> Joensuu, Joensuu, Finland. 427 pp.<br />

Koh, L.P., Wilcove, D.S. 2007. Cashing in palm oil for conservation. Nature 448, 993-994.<br />

Koh, L.P., Wilcove, D.S. 2008. Is oil palm agriculture really destroying tropical biodiversity? Conservation<br />

Letters 1 (2), 60-64.<br />

Lemmen, D. & Warren, F. (eds.). 2004. Climate change impacts and adaptation: A Canadian perspective.<br />

Climate Change Impacts and Adaptation Directorate, Natural Resources Canada. Ottawa, Ontario.<br />

Ogden, A.E. & Innes, J.L. 2007. Incorporating climate change adaptation considerations into forest management<br />

planning in the boreal forest. International Forestry Review 9: 713–733.<br />

Persson, R. 2005. Where is the United Nations Forum on Forests going? International Forestry review 7 (4), 348-<br />

357.<br />

Spittlehouse, D.L. 2005. Integrating climate change adaptation into forest management. Forestry Chronicle<br />

81(5): 691–695.<br />

Spittlehouse, D.L. & Stewart, R.B. 2003. Adaptation to climate change in forest management. BC Journal <strong>of</strong><br />

Ecosystems and Management 4(1): 1–11.<br />

Stern, N. 2006. The economics <strong>of</strong> climate change. The Stern Review. Cambridge University Press, Cambridge.<br />

p. 692.<br />

Tacconi, L. (ed.) 2007. Illegal logging. Law enforcement, livelihoods and the timber trade. Eathscan<br />

Publications, London, UK.<br />

van Mantgem, P.J., Stephenson, N.L., Byrne, J.C., Daniels, L.D., Franklin, J.F., Fule, P.Z., Harmon, M.E.,<br />

Larson, A.J., Smith, J.M., Taylor, A. H. and Veblen, T.T. 2009. Widespread increase <strong>of</strong> tree mortality<br />

rates in the western United States. Science 323 (5913): 521-524.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 7<br />

Table 1. Strategic (S) and operational (O) climate change adaptation options that may be considered<br />

to achieve the management objective <strong>of</strong> ensuring the appropriate legal, institutional and economic<br />

framework is in place for forest conservation and sustainable management. Table adapted from Ogden<br />

and Innes (2007). Sources: BCMOF (2006), Chapin et al. (2004), FAO (2008), Johnston et al. (2006),<br />

Kellomaki et al. (2005), Lemmen and Warren (2004), Spittlehouse (2005), Spittlehouse and Stewart<br />

(2003).<br />

Impact S/O Adaptation Options<br />

Forest<br />

S Development <strong>of</strong> flexible forest management plans and policies that<br />

management plans incorporate uncertainty and are capable <strong>of</strong> responding to climate<br />

and policies lack<br />

change.<br />

the flexibility that Maintain an overall management process where targets and results<br />

is required to<br />

and the effectiveness <strong>of</strong> chosen options are revisited and assessed<br />

respond to climate<br />

change<br />

periodically against emerging issues and changing environmental<br />

conditions<br />

Provide long term tenures to encourage long term considerations<br />

within short term decisions<br />

Relax rules governing the movement <strong>of</strong> seed stocks from one area to<br />

another<br />

Measure, monitor and report on indicators <strong>of</strong> climate change and<br />

sustainable forest management to determine the state <strong>of</strong> the forest<br />

and identify when critical thresholds are reached<br />

Evaluate the adequacy <strong>of</strong> existing environmental and biological<br />

monitoring networks for tracking the impacts <strong>of</strong> climate change on<br />

forest ecosystems<br />

Practice adaptive management.<br />

O Conduct a priori vulnerability assessments<br />

Conduct advanced planning <strong>of</strong> salvage logging procedures<br />

Train forest managers to deal with uncertainty and to manage risk<br />

effectively<br />

Implement practices that reduce vulnerability and increase resilience<br />

even at the expense <strong>of</strong> growth<br />

Forest<br />

S Development <strong>of</strong> forest management plans that reduce vulnerability<br />

management plans <strong>of</strong> forest and forest dependent communities to climate change<br />

and policies<br />

enhance the<br />

vulnerability <strong>of</strong><br />

forests and forest<br />

dependent<br />

communities to<br />

Assess vulnerability <strong>of</strong> forests, forest dependent communities and<br />

society in general<br />

Support research on climate change, climate impacts, and climate<br />

change adaptations and increase resources for basis climate change<br />

impacts and adaptation science<br />

climate change<br />

Support knowledge exchange, technology transfer, capacity building<br />

and information sharing on climate change; maintain or improve<br />

capacity for communications and networking<br />

Incorporate new knowledge about the future climate and forest<br />

vulnerability into forest management plans and policies<br />

Involve the public in an assessment <strong>of</strong> forest management adaptation<br />

options<br />

O Diversify forest-based and non-forest based income sources<br />

Increase local governance <strong>of</strong> local forest resources<br />

Increase general capacity for the detection and management <strong>of</strong><br />

climate change impacts


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 8<br />

Forest<br />

management<br />

policies and<br />

incentives do not<br />

encourage<br />

adaptation to<br />

climate change<br />

Forest institutions<br />

are unable to cope<br />

with the needs <strong>of</strong><br />

climate change<br />

Inadequate<br />

communication<br />

between forest<br />

scientists and<br />

S<br />

S<br />

Remove barriers and develop incentives to adapt to climate change.<br />

Provide incentives and remove barriers to enhancing carbon sinks<br />

and reducing greenhouse gas emissions<br />

Provide opportunities for forest management activities to be<br />

included in carbon trading systems (as outlined in Article 3.4 <strong>of</strong> the<br />

Kyoto Protocol)<br />

Increase awareness <strong>of</strong> the implications <strong>of</strong> climate change to<br />

encourage behavioural change that will facilitate planned adaptation<br />

Encouragement local, regional and national forest management<br />

policies and plans to consider their role globally (think global, act<br />

local)<br />

Establish new or enhance existing institutions and infrastructure for<br />

proactive or proactive adaptation<br />

Develop multiple partnerships and break down institutional barriers<br />

to better enable the recovery from climate-mediated crises<br />

Develop mechanisms to encourage multilateral solutions to the<br />

problems brought about by climate change<br />

S Develop mechanisms to enable research results to be shared and<br />

communicated to decision- and policy-makers<br />

Elevate the importance <strong>of</strong> the social sciences in the adaptation <strong>of</strong><br />

communities, land-uses and forest management to climate change<br />

policy makers Raise the awareness <strong>of</strong> the research community to the type <strong>of</strong><br />

information needed by decision- and policy makers


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 9<br />

THE CHANGING ROLE OF FORESTS: STATE OF DEVELOPMENT<br />

OF MARKETS FOR ECOSYSTEM SERVICES<br />

David Brand 1<br />

INTRODUCTION<br />

For over twenty years there have been concerted efforts to reduce the loss and degradation <strong>of</strong> forests.<br />

While some success has been seen in countries like <strong>Australia</strong> where regulation <strong>of</strong> land clearing has<br />

been effective in reducing conversion rates (WWF, 2009), in other regions, particularly tropical<br />

forests, reduction in forest cover has continued (FAO, 2005). Much <strong>of</strong> the systematic degradation <strong>of</strong><br />

ecosystems, including forests, has been driven by the expansion <strong>of</strong> agriculture, and to a lesser extent<br />

by human settlement, mining, oil and gas and infrastructure (Millennium Ecosystem Assessment,<br />

2005). Ultimately the process <strong>of</strong> ecosystem degradation is driven by the size and expansion <strong>of</strong> the<br />

global economy, which is driven by human population and per capita economic growth.<br />

We now have a population <strong>of</strong> 6.7 billion, likely to rise to 10.5 billion before peaking (UN DESA,<br />

2009). The global economy is approximately $62.2 trillion (CIA, 2009), and growing on average<br />

between 2% and 3% per annum. This means that the global economy is likely to double every 24 to<br />

36 years. Many forms <strong>of</strong> consumption have been outstripped economic growth. For example meat<br />

production, energy consumption, and paper usage have all grown at or faster than economic growth<br />

(FAOSTAT, 2009). Growing consumption impacts ecosystems both directly and indirectly.<br />

Conversion <strong>of</strong> Brazilian rainforest to cattle grazing, or <strong>of</strong> Indonesian rainforest to palm oil, has direct<br />

and long-lasting consequences. Excessive pollution and climate change have more chronic impacts on<br />

ecosystems via potential changes to disturbance factors like wildfire and insect epidemic, and shifts in<br />

species distributions.<br />

In fact since the Earth Summit in 1992 we have been able to foresee that the global system would be<br />

systematically changed by consumption and that the three great challenges would be related to soil<br />

conservation and access to freshwater, biodiversity conservation and global climate change. Forest<br />

conservation fundamentally intersects these three challenges—forests are one <strong>of</strong> the key elements <strong>of</strong><br />

the global carbon cycle (Malhi, Y et al, 2002), regulate soil conservation and water quality, and<br />

provide the basis for something like 50% <strong>of</strong> the diversity <strong>of</strong> life on earth (Millennium Ecosystem<br />

Assessment, 2005). As our population grows and the economy becomes larger, there is a conundrum.<br />

On the one hand we know that these ecosystems provide services such as water purification, carbon<br />

storage and soil conservation, but on the other we continue to convert these ecosystems to production<br />

systems to support ever-increasing demand for food, fiber, human settlement and infrastructure.<br />

No one would realistically expect that we might end up with a planet that has no natural ecosystems<br />

left. But how do we diverge from the pattern <strong>of</strong> economic growth linked to conversion <strong>of</strong> ecosystems<br />

from conservation functions to production functions? The problem is that production systems<br />

generate pr<strong>of</strong>it and attract capital for development, while conservation functions do not. Effectively,<br />

nature is unpriced, and therefore is used wastefully as a low cost input to production. In our present<br />

economic system it is more cost effective to convert more land to production than it is to intensify<br />

production on the existing landbase. It is easy to argue that this represents a market failure.<br />

There have been many attempts to use planning, policies, aid programs and philanthropy to halt the<br />

degradation and loss <strong>of</strong> forests (e.g. the Forest Principles developed at the 1992 Earth Summit and the<br />

FAO Tropical Forest Action Plan which was developed in 1985). They have largely failed, because<br />

they don’t change the underlying price signals and economic drivers. As a result there is now a<br />

rapidly escalating effort to create price signals for ecosystems and for ecosystem services as a way <strong>of</strong><br />

actually making conservation functions valuable (Ecosystem Marketplace, 2009). Once something<br />

1<br />

Dr David Brand, Managing Director, New Forests Pty Limited, The Zenith Centre, Level 19 Suite 1905, 821 Pacific Hwy,<br />

Chatswood NSW 2067, <strong>Australia</strong>.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 10<br />

becomes valuable, it can attract capital and will become more costly to use as an input to production.<br />

This paper aims to discuss how these markets for ecosystem services are emerging, and how they<br />

might contribute to a kind <strong>of</strong> global end-game where production and conservation functions are<br />

stabilized, and human society effectively makes an accommodation with nature.<br />

THE CONCEPT OF MARKETS FOR ECOSYSTEM SERVICES<br />

The question is how to introduce these price signals into an economic system that has benefited from<br />

nature being unpriced. In some ways it is not as difficult as it may seem. Markets have long<br />

understood the concept <strong>of</strong> auctioning fishing quotas, timber rights, grazing rights and irrigation water.<br />

It is only a step more to commodify carbon storage, watershed or catchment management and species<br />

conservation. The challenge in any emerging market, however, is to create both supply and demand.<br />

The demand side <strong>of</strong> ecosystem services has been lacking.<br />

There are a growing range <strong>of</strong> examples <strong>of</strong> markets for ecosystem services, and while these remain<br />

small they are expanding rapidly. In these markets demand is generally created by either voluntary<br />

action or by government regulation. Voluntary action examples might include the sale <strong>of</strong> carbon<br />

credits from <strong>Australia</strong>n re-vegetation projects to automobile owners by Greenfleet (Greenfleet, 2009),<br />

payment to landowners by Coca-Cola for clean water (Clinton Global Initiative, 2008), and<br />

sponsorship <strong>of</strong> duck habitat conservation by duck hunters (Ducks Unlimited, 2009).<br />

Yet voluntary funding is invariably a marginal act, and does not transform price signals. That only<br />

occurs through regulation. During the 1970’s the US Government passed the Clean Water Act (US<br />

Environmental Protection Agency, 2008) and the Endangered Species Act (United States Fish and<br />

Wildlife Service, 2009). These two acts provide an interesting case study <strong>of</strong> the establishment <strong>of</strong> price<br />

signals for ecosystems. In each case the Act created a no net loss philosophy for wetlands and<br />

endangered species respectively. Over time regulatory instruments emerged to allow for wetlands<br />

mitigation banking and endangered species banking. These were market-based instruments. For<br />

example if I am a property developer wanting to convert habitat for an endangered species, I will need<br />

to pay someone else to enhance the conservation or expand the habitat for the same species elsewhere.<br />

These markets <strong>of</strong>ten worked by creating large areas <strong>of</strong> functional habitat, or by rehabilitating degraded<br />

areas into productive wetlands. In many cases the ‘banks’ created to sell wetland or endangered<br />

species conservation became highly valuable. Many developers found that their properties were more<br />

valuable if managed for conservation rather than property development.<br />

The mitigation banking industry in the US is now a well-regulated sector with hundreds <strong>of</strong> millions <strong>of</strong><br />

dollars in sales annually (speciesbanking.com, 2008). The sector has small and medium sized<br />

businesses, increasingly attracts private equity funding, and there is a growing sophistication in the<br />

industry. In fact it could be seen as a model for how an ecosystem market can unfold.<br />

The other evolving market that came from the US was the Clean Air Act (US Environmental<br />

Protection Agency, 2009) and its creation <strong>of</strong> a market for reductions in sulfur dioxide emissions<br />

causing acid rain. The regulations set a cap on emissions and allowed polluters to trade the allowances<br />

necessary to comply with the cap. This meant that a company who could reduce their emissions more<br />

inexpensively would reduce more than required and sell the excess reductions to companies with a<br />

higher cost <strong>of</strong> compliance. This market based approach proved highly successful and the cost <strong>of</strong><br />

meeting the targeted reductions was far lower than had been predicted by economic models (Gutt, E.<br />

et.al, 2000).<br />

The US government also promoted the use <strong>of</strong> markets to address greenhouse gas reductions in the<br />

1990’s. While the US ultimately pulled out <strong>of</strong> the Kyoto Protocol, the use <strong>of</strong> flexibility mechanisms<br />

like carbon trading was a legacy <strong>of</strong> US negotiating strategies. The US and other countries also<br />

promoted the incorporation <strong>of</strong> forest conservation and reforestation into the global carbon market as an<br />

<strong>of</strong>fset (or carbon credit against emissions). While this ultimately was curtailed in the Kyoto Protocol<br />

and rejected for incorporation to the European Union Emissions Trading Scheme (European Union


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 11<br />

Emissions Trading Scheme, 2009) forestry has now begun to see a resurgence <strong>of</strong> interest in non-<br />

European carbon markets such as the <strong>Australia</strong>n Carbon Pollution Reduction Scheme (Department <strong>of</strong><br />

Climate Change, 2009) , the NZ Emissions Trading Scheme (Ministry for the Environment, 2009), the<br />

California Forestry Protocols (Climate Action Reserve, 2009) and most recently and significant the US<br />

Federal Waxman-Markey Bill (Committee on Energy and Commerce, 2009). Significant innovation is<br />

occurring related to new concepts like Reductions in Emissions from Deforestation and Forest<br />

Degradation (“REDD”) (VCS Association, 2008 and Avoided Deforestation Partners, 2008) that could<br />

create whole new markets for forest conservation.<br />

So there are functional examples <strong>of</strong> regulatory systems across carbon, water and biodiversity<br />

attributes. The challenge is to make these instruments ubiquitous in the global economy, and subject<br />

to a comprehensive no net loss policy at national and international levels. This will not be easy, but<br />

the ultimate solution may well be found in the problem itself—effectively embedding ecosystem<br />

services into the supply chains to consumers.<br />

CONSERVATION AND PRODUCTION – LANDSCAPES AND HUMAN SOCIETY<br />

Several recent studies confirm the degree to which human society and its demand for goods and<br />

services now dominates nature (see for example the Millennium Ecosystem Assessment or the WWF<br />

Living Planet Report). Economic growth is inexorable, but ecosystems are finite. Therefore as the<br />

economy grows we have seen a steady erosion <strong>of</strong> natural systems resulting in a series <strong>of</strong> negative<br />

consequences. Loss <strong>of</strong> ecosystem function is linked with soil erosion, loss <strong>of</strong> productivity, the spread<br />

<strong>of</strong> weeds, feral animals and disease, increased flooding, nutrient leaching, hypoxic zones, and a host <strong>of</strong><br />

other outcomes. It is as if the demand for goods is now seriously impeding the capacity to maintain<br />

ecosystem services, and some form <strong>of</strong> balance must be sought. The lack <strong>of</strong> pricing for ecosystem<br />

services means that there is no optimal outcome where the marginal benefit <strong>of</strong> increased production is<br />

balanced against the marginal cost <strong>of</strong> lost ecosystem services. The market based pricing <strong>of</strong> goods will<br />

dominate the unpriced public services <strong>of</strong> ecosystems even if there is a huge negative net cost or<br />

‘externality’ to society. So the answer has to be to price ecosystems and their services.<br />

The approach that has the most merit is to regulate those industries that either impact or benefit from<br />

water quality or biodiversity, or which have greenhouse gas emissions, such that they have to avoid,<br />

reduce and ultimately mitigate their environmental impacts. In parallel we need to commoditize the<br />

ecosystem services so that they can be conserved and enhanced by private capital flows the same way<br />

that oil palm or beef is produced. The production functions then will need to acquire ecosystem<br />

services credits in order to continue to operate or expand operations. This embeds the conservation<br />

functions in the production functions and consumption ultimately must pay for the costs <strong>of</strong><br />

maintaining ecosystem services. As the global economy grows and demand rises, ecosystems should<br />

become steadily more valuable. This will lead to greater emphasis on resource use efficiency,<br />

production efficiency, recycling, etc. Effectively there should be some kind <strong>of</strong> economic balance<br />

where the cost <strong>of</strong> further impacts to ecosystems becomes too expensive to contemplate.<br />

The supply side instruments are starting to become clear. The first is the emerging forest carbon<br />

market. While the Kyoto protocol has provided limited innovation around forest carbon products,<br />

there has been substantial new thinking coming from the United States and the voluntary carbon<br />

market (see AD Partners website cited above).<br />

Standards for creating carbon conservation value related to reforestation, improved forest management<br />

and forest conservation are all emerging into a workable framework. The US Waxman Markey bill<br />

was a key step forward and comprehensively integrated both domestic and international forest carbon<br />

management into a legislative framework. The work undertaken by US legislators is likely to support<br />

a strong US position on the incorporation <strong>of</strong> forest carbon instruments into the negotiations at COP15<br />

on a successor international agreement to the Kyoto Protocol in 2013.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 12<br />

The second key instrument is a biodiversity product. Unlike carbon, biodiversity is not easily<br />

measurable or commodifiable. Biodiversity is also a surrogate itself for a suite <strong>of</strong> ecosystem services<br />

related to purification <strong>of</strong> water, pollination, disease control, and general maintenance <strong>of</strong> productivity<br />

and natural function. It is hard to comprehend a completely homogenized world with complex natural<br />

systems largely gone, and human managed exotic production systems completely dominant. It is<br />

probably hard to think through what such a world might look like, but there are certainly regions <strong>of</strong> the<br />

earth already well down the path <strong>of</strong> complete homogenization.<br />

The mitigation banking industry in the US is probably the most likely model for a global biodiversity<br />

conservation product. The approach is based on institutionalizing a philosophy <strong>of</strong> no net loss and<br />

then, based on assessments, allowing development activities to mitigate their impact by buying<br />

‘credits’ from mitigation bankers. The ultimate expression <strong>of</strong> this approach could be large scale<br />

protected areas funded by mitigation banking fees. New Forests has recently established the Malua<br />

Biobank (MWHCB Inc, 2009) in Sabah, Malaysia, that is trying to experiment with exporting the<br />

mitigation banking approach to conserving tropical rainforests with a globally significant suite <strong>of</strong> plant<br />

and animal species. In this case serialized Biodiversity Conservation Certificates are created and<br />

listed on the TZ1 exchange (TZ1 Market, 2009). Each certificate represents the rehabilitation and<br />

conservation management <strong>of</strong> 100 square metres <strong>of</strong> dipterocarp forest. Oil palm plantations can buy<br />

these certificates and attach them to their exports <strong>of</strong> crude palm oil. In fact the product works very<br />

well with the palm oil supply chain. Each hectare <strong>of</strong> palm oil plantation produces about 100 tonnes <strong>of</strong><br />

crude palm oil during its 25 year life. If a biodiversity certificate is attached to each tonne produced,<br />

the hectare <strong>of</strong> palm oil plantation effectively sponsors the rehabilitation and conservation management<br />

<strong>of</strong> one hectare <strong>of</strong> the biobank. This makes the palm oil producer the sponsor <strong>of</strong> biodiversity<br />

conservation rather than the cause <strong>of</strong> its depletion.<br />

The third key instrument is a water quality related product. The impacts <strong>of</strong> production activities on<br />

water quality include soil erosion and resultant turbidity, increases in dissolved nutrients, particularly<br />

nitrates and phosphates, and run<strong>of</strong>f <strong>of</strong> pesticides. These impacts affect water quality directly for<br />

downstream human use, but also affect aquatic system health via impacts on benthic fauna, fish, bioaccumulation<br />

<strong>of</strong> pesticides, algal blooms, hypoxia, and coral bleaching. There are also specific<br />

regional impacts on water quality from temperature changes, salinisation, acid sulphate soils and shifts<br />

in pH.<br />

Like the biodiversity related instruments, water quality problems are <strong>of</strong>ten localized and need to be<br />

regulated at the level <strong>of</strong> catchments or watersheds. In the US, again, there have been a number <strong>of</strong><br />

experiments and pilot programs related to water quality trading (Environmental Trading Network,<br />

2009). In general the nutrient trading regimes have been designed so that point source polluters like<br />

sewage treatment plants could buy <strong>of</strong>fsets after achieving a certain level <strong>of</strong> in-house pollution control.<br />

This reflects the reality that the last 2 or 3 percent reductions may have extremely high costs, and<br />

paying farmers to fence riparian zones, or use precision fertilization practices may deliver additional<br />

reductions in nutrient loss at a far lower cost.<br />

A major water quality initiative is currently being undertaken in the <strong>Australia</strong>n state <strong>of</strong> Queensland,<br />

where catchments draining into the Great Barrier Reef have been put under a stringent continuous<br />

improvement and monitoring system. All land management practices have been benchmarked for<br />

their water quality impacts, and payment schemes have been instituted to support landowners in<br />

making a transition to lower impact land management practices. The catchments have been modeled<br />

and the level <strong>of</strong> nutrients and other pollutants can be forecast under different levels <strong>of</strong> program takeup.<br />

The specific goal <strong>of</strong> the <strong>Australia</strong>n Government is to reduce nutrient loads by 50% in the Great<br />

Barrier Reef Lagoon. The Government is paying approximately $AU200 million over four years to<br />

landowners under the Reef Rescue Program (Caring for our Country, 2009). The question is what will<br />

happen as Government funding comes to an end. The Government funding was a by-product <strong>of</strong><br />

selling the national telephone company. Further initiatives would need to come from tax revenues,<br />

which is unlikely given the current budget deficit.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 13<br />

New Forests has put forward a concept for a water quality bank for the Great Barrier Reef Catchments<br />

that will perpetuate the price signals for sustainable land management via a Nutrient Bank. The<br />

concept is based on setting a baseline <strong>of</strong> current loads <strong>of</strong> nutrients, sediment and chemicals and<br />

creating a kind <strong>of</strong> water pollution unit. Monthly water quality monitoring at key points can be used to<br />

build an annual water quality rating. If the water quality rating is below the baseline, then there are<br />

credits created. These credits would be listed and saleable. The revenue gained from selling these<br />

credits would be used for three purposes—to pay landowners an upfront price for improved land<br />

management, to generate a return to private investors in the bank and to pay dividends to landowners<br />

who maintain their properties under higher levels <strong>of</strong> land management practice. Downstream<br />

beneficiaries <strong>of</strong> the improved water quality like tourism operators could be buyers or the credits could<br />

be embedded in the agricultural commodities, principally sugar. The government could also be a<br />

buyer, at least initially to act on behalf <strong>of</strong> the general public good.<br />

WHAT IS THE END-GAME?<br />

There have been continuing attempts to set targets and policy mechanisms for conservation <strong>of</strong><br />

ecosystems. The global goal <strong>of</strong> 10% protected areas in each bioregion is a major example, supported<br />

by the United Nations Convention on Biological Diversity, the IUCN and others. Conservation<br />

International has embraced the concept <strong>of</strong> protecting key ‘hotspots’ which are areas <strong>of</strong> unique<br />

endemism and biodiversity. The concept <strong>of</strong> protected areas works reasonably well in developed<br />

countries with substantial resources, but even there the protected areas tend to be over-represented in<br />

non-productive ecosystems (alpine, mountainous terrain, arctic regions, deserts, etc.) and underrepresented<br />

in productive systems used for agriculture, grazing and human settlement.<br />

Given that most productive areas are under private ownership or management, it proves expensive and<br />

controversial for governments to use tax-payers dollars to acquire these areas for rehabilitation or<br />

conservation. Some NGO’s have sought to augment formally protected areas in Government<br />

ownership with informal conservation areas <strong>of</strong>ten established by the purchase <strong>of</strong> conservation<br />

easements or development rights from property owners. These philanthropic conservation funds have<br />

made substantial contribution to protected areas networks, but the general view is that they are not<br />

capable <strong>of</strong> out-competing private investment in development activities. A recent study by the Union<br />

<strong>of</strong> Concerned Scientists and the UK government suggested that funding <strong>of</strong> $US20 to $US33 billion<br />

per annum would be needed to reduce deforestation by 50% over the next ten to twenty years. That is<br />

250% to 400% <strong>of</strong> the current combined funding <strong>of</strong> the World Bank, Overseas Development Assistance<br />

and Philanthropy. If we wish to completely stop all further rainforest conversion we would potentially<br />

be looking at funding <strong>of</strong> $50 billion per annum or more.<br />

Another way to look at the problem <strong>of</strong> forest conservation finance is to explore the goals for emissions<br />

reduction in 2030 and work backwards. It has been suggested that global agreements should aim for<br />

80% reductions in greenhouse gas emissions by 2050 and that this will require approximately 30<br />

billion tonnes per annum <strong>of</strong> greenhouse gas reductions relative to business as usual by 2030. Analysis<br />

<strong>of</strong> that target suggests that at least 20% <strong>of</strong> the net emission reduction will need to come from rainforest<br />

conservation by 2030. That is 5 billion tonnes <strong>of</strong> carbon dioxide equivalent reductions per annum. At<br />

a global carbon price <strong>of</strong> $US20 that represents a revenue stream <strong>of</strong> $US100 billion per annum. If we<br />

securitized such a cash flow at a 10% real cash yield, it would represent an asset value <strong>of</strong> the world’s<br />

rainforests <strong>of</strong> $US1 trillion. That is approximately equal to 5% <strong>of</strong> the institutional financial assets in<br />

the United States, or approximately equal to the entire assets <strong>of</strong> the <strong>Australia</strong>n Superannuation<br />

industry.<br />

On the other hand if we compare this with the value <strong>of</strong> agricultural commodities it is less daunting.<br />

For example, Malaysia exports around 15 million tonnes <strong>of</strong> crude palm oil for $US8 billion per<br />

annum. Globally oil from oils seeds is approximately a $US50 billion per annum industry (FAO).<br />

Meat, sugar, timber, and other commodities are even greater in value. The Union <strong>of</strong> Concerned<br />

Scientists recently suggested that global timber imports are $US160 billion per annum – a 5% tax


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 14<br />

would generate $US8 billion per annum for REDD or other forest conservation initiatives (Union <strong>of</strong><br />

Concerned Scientists USA, 2009).<br />

A tax is generally a blunt instrument as it affects both sustainable and unsustainable producers equally.<br />

A more effective approach would be to create certification or accreditation processes that allow<br />

sustainably produced goods to either trade at a premium, or to have lower costs associated with their<br />

lower environmental impacts. There are a number <strong>of</strong> these certification processes now in operations.<br />

The first and longest running is the Forest Stewardship Council (Forest Stewardship Council, 2009)<br />

which certifies the sustainable management <strong>of</strong> forests and provides for chain <strong>of</strong> custody certification<br />

so that end users can differentiate between wood products coming from Forest Stewardship Council<br />

Certified forests and those that are not. There are now a series <strong>of</strong> commodity Round-Tables (eg Soy,<br />

Palm Oil, Sugar) that are proposing certification standards and mechanisms to differentiate sustainable<br />

production from unsustainable production (RSPO, 2009).<br />

This is a critical first step, but not quite enough to achieve a no net loss end-game. The sustainable<br />

production systems are accredited to best practice in terms <strong>of</strong> not clearing high conservation value<br />

forests, minimizing pesticide use, implementing fair employment practices, etc. They do not however<br />

require zero environmental impact. This means that as the global economy doubles and redoubles in<br />

size, the consumption continuously chips away at natural systems, albeit less rapidly than under<br />

unsustainable production systems. That is why the idea <strong>of</strong> actually embedding the ecosystem services<br />

products described above can more realistically move us towards an ultimate solution.<br />

The way this might work is that each bio-region would have an inventory <strong>of</strong> its ecosystems and state<br />

<strong>of</strong> the ecosystem services. In some cases the goal might be to stabilize the system, in other cases it<br />

could be to rehabilitate and recover some ecosystem services to a higher level (eg in our Great Barrier<br />

Reef example). In some rare cases it may even be acceptable to draw down further on the ecosystems<br />

before determining a stabilization point. Whatever that point is, then the ecosystems providing those<br />

services would be commodified into REDD projects, bio-banks and water quality banks. The credits<br />

could then be attached to the commodities produced in those bio-regions to create a systematic<br />

sponsorship <strong>of</strong> the conservation functions by the production functions. As an example <strong>of</strong> this, the<br />

Malua bio-bank can generate a commercially acceptable return from selling its Biodiversity<br />

Conservation Certificates (BCC) at $US10 per 100 square metres <strong>of</strong> forest rehabilitation and<br />

conservation management. Therefore, attaching one BCC to one tonne <strong>of</strong> crude palm oil only adds<br />

1.5% to the current price (Figure 1).<br />

The bio-regions may be supporting not only local commodity agribusiness, and point source industries<br />

like mining, oil and gas, but could be servicing global industries like electricity generation via REDD<br />

markets. In this case the bio-banks could be established to sell multiple ecosystem commodities<br />

including biodiversity certificates to palm oil companies, REDD credits to overseas energy sector<br />

firms, and water quality credits to downstream water users. Over time as the global economy grows<br />

the eco-commodities could become hugely valuable in line with their growing importance in<br />

supporting an ever larger global economy. The result would be landscapes that integrate production<br />

and conservation functions on a commercial basis (Figure 2).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 15<br />

Figure 1. An example <strong>of</strong> how biodiversity conservation credits can be integrated into the crude<br />

palm oil supply chain.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 16<br />

Figure 2. Landscapes with more functional ecosystems can be made more valuable<br />

CONCLUSION<br />

We appear to be on the verge <strong>of</strong> some breakthrough ideas that could shift the balance in the economics<br />

<strong>of</strong> deforestation and forest degradation. Not only is there a renewed emphasis on incorporating forests<br />

into the global carbon market, but the global agri-business industry is under pressure to introduce third<br />

party certification <strong>of</strong> the sustainability <strong>of</strong> production systems. These trends need to be linked with<br />

mechanisms to standardize the ecosystem services products like REDD, bio-banks and water quality<br />

or watershed conservation banks. We can already see the potential instruments emerging in voluntary<br />

markets and in some national level regulatory experience. The challenge now is to move ahead with<br />

implementation fast enough that the tide can be turned before there is little left to save.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 17<br />

REFERENCES<br />

Avoided Deforestation Partners (2008). AD Partners, Washington D.C., viewed 13/07/2009,<br />

www.adpartners.org.<br />

Caring for our Country (2009). Caring for our Country, Canberra, viewed 13/07/2009, www.nrm.gov.au.<br />

Central Intelligence Agency (CIA) (2009). CIA, Washington D.C., viewed 13/07/2009,<br />

www.cia.gov/library/publications/the-world-factbook/geos/XX.html<br />

Climate Action Reserve (2009). Climate Action Reserve, Los Angeles, viewed 13/07/2009,<br />

www.climateactionreserve.org.<br />

Clinton Global Initiative (2008). Clinton Global Initiative, New York, viewed 13/07/2009,<br />

http://www.clintonglobalinitiative.org//Page.aspx?pid=2646&q=270340&n=x<br />

Committee on Energy and Commerce (2009). Committee on Energy and Commerce, Washington D.C., viewed<br />

13/07/2009, http://energycommerce.house.gov/index.php?option=com_content&task=view&id=1560.<br />

Department <strong>of</strong> Climate Change (2009). Carbon Pollution Reduction Scheme, Canberra, viewed 13/07/2009,<br />

www.climatechange.gov.au/emissionstrading/index.html.<br />

Ducks Unlimited (2009). Ducks Unlimited, Memphis, viewed 13/07/2009, www.ducksunlimited.org.<br />

Ecosystem Marketplace (2009). Ecosystem Marketplace, Washington D.C., viewed 13/07/2009,<br />

www.ecosystemmarketplace.com.<br />

Environmental Trading Network (2009). Environmental Trading Network, Michigan, viewed 13/07/09,<br />

http://www.envtn.org.<br />

European Union Emissions Trading Scheme (2009), European Commission, Brussels, viewed 13/07/2009,<br />

http://ec.europa.eu/environment/index_en.htm.<br />

FAO Statistics Division (2009). Data available online at http://faostat.fao.org<br />

Food and Agriculture Organization <strong>of</strong> the United Nations (FAO) (2005). Global Forest Resources Assessment<br />

2005. Rome: Food and Agriculture Organization <strong>of</strong> the United Nations. Available online at<br />

www.fao.org/forestry/fra2005/en/<br />

Food and Agriculture Organization <strong>of</strong> the United Nations (FAO) (1985). Tropical forestry action plan. Rome:<br />

Food and Agriculture Organization <strong>of</strong> the United Nations.<br />

Forest Stewardship Council (2009). FSC, Bonn, viewed 13/07/2009, www.fsc.org.<br />

Greenfleet (2009) Greenfleet, Melbourne, viewed 13/07/2009, www.greenfleet.com.au.<br />

Gutt, E., Patton, V. and Spencer, N. (2000), Building on 30 Years <strong>of</strong> Clean Air Act Success: The Case for<br />

Reducing NOx Air Pollution, viewed 13/07/2009, www.edf.org/documents/398_CAAReport.PDF.<br />

Malhi, Y., Meir, P. & Brown, S. (2002) Forests, carbon and global climate. Philosophical Transactions <strong>of</strong> the<br />

Royal Society. A 360, 1567–1591.<br />

Millennium Ecosystem Assessment (2005). Ecosystems and human well-being : current state and trends :<br />

findings <strong>of</strong> the Condition and Trends Working Group / edited by Rashid Hassan, Robert Scholes, Neville<br />

Ash. Island Press, Washington, DC<br />

Millennium Ecosystem Assessment (2005). Ecosystems and Human Well-being: Synthesis. Island Press,<br />

Washington, DC<br />

Ministry for the Environment, (2009). The New Zealand Emissions Trading Scheme, Wellington, viewed<br />

13/07/2009, www.climatechange.govt.nz/emissions-trading-scheme/index.html.<br />

MWHCB Inc (2009). Malua BioBank, viewed 13/07/2009, www.maluabiobank.com.<br />

RSPO (2009). RSPO, Kuala Lumpur, viewed 13/07/2009, www.rspo.org.<br />

SPECIESBANKING.COM (2008). The Katoomba Group, Washington D.C., viewed 13/07/09,<br />

www.speciesbanking.com.<br />

TZ1 Market (2009). TZ1 Market, viewed 13/07/2009, www.tz1market.com.<br />

Union <strong>of</strong> Concerned Scientists USA (2009). Union <strong>of</strong> Concerned Scientists, Cambridge, MA, viewed<br />

13/07/2009, www.ucsusa.org.<br />

United Nations (2000). Forest Principles, viewed 13/07/2009,<br />

http://www.un.org/documents/ga/conf151/aconf15126-3annex3.htm.<br />

United Nations, Department <strong>of</strong> Economic and Social Affairs, Population Division (2009). World Population<br />

Prospects: The 2008 Revision, Highlights, Working Paper No. ESA/P/WP.210


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 18<br />

United States Fish and Wildlife Service, (2009). USFWS, Washington D.C., viewed 13/07/2009,<br />

www.fws.gov/laws/lawsdigest/ESACT.html.<br />

US Environmental Protection Agency (2008). Clean Water Act, Washington D.C., viewed 13/07/2009,<br />

www.epa.gov/watertrain/cwa.<br />

US Environmental Protection Agency (2009). Clean Air Act, Washington D.C., viewed 13/07/2009,<br />

www.epa.gov/air/caa.<br />

VCS Association (2008). VCS, Washington D.C., viewed 13/07/2009, www.v-c-s.org.<br />

WWF-<strong>Australia</strong> (2009) WWF-<strong>Australia</strong>, <strong>Australia</strong>, viewed 10/07/2009, www.wwf.org.au/articles/feature08<br />

ABOUT NEW FORESTS<br />

New Forests (see www.newforests.com.au) is an investment management and advisory services firm<br />

specialising in forestry and land-based environmental markets, such as timber, carbon, biodiversity<br />

and water. The company’s investment philosophy seeks to deliver traditional timber returns as well as<br />

returns from eco products, such as certified timber, renewable energy, carbon credits, biodiversity<br />

benefits and water-quality improvements. The company is headquartered in Sydney, <strong>Australia</strong>, with<br />

<strong>of</strong>fices in Washington, D.C., San Francisco and Kota Kinabalu, Malaysia. New Forests holds an<br />

<strong>Australia</strong>n Financial Services Licence.<br />

New Forests manages over $250 million in sustainable forestry and eco product (carbon, biodiversity,<br />

and water) assets in the United States, <strong>Australia</strong>, New Zealand, Southeast Asia and the Pacific Islands.<br />

This includes the full value chain <strong>of</strong> services, from the development <strong>of</strong> investment theses and portfolio<br />

management to operational execution and asset management. New Forests Advisory Services<br />

business provides policy and market analysis, regulatory advice, technical modelling capabilities and<br />

investment strategy to external clients and internal investment management teams. The business line is<br />

focused on developing commercial solutions at the intersection <strong>of</strong> land-based investment, conservation<br />

and climate change mitigation by identifying and quantifying environmental asset opportunities.<br />

New Forests' staff includes experienced pr<strong>of</strong>essionals across timberland investment, operational<br />

forestry, environmental management and finance, and the company provides its clients with a unique<br />

combination <strong>of</strong> forestry, carbon and financial management skills. Staff members have been involved<br />

in forest carbon transactions since the 1990s and have contributed to the development <strong>of</strong> forestry<br />

<strong>of</strong>fset rules in <strong>Australia</strong>, New Zealand, California, the United States, Canada and through the<br />

UNFCCC, as well as the drafting <strong>of</strong> the Agriculture, Forestry and Other Land Use guidelines for the<br />

Voluntary Carbon Standard.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 19<br />

CAPTURING THE VALUE OF INNOVATION –<br />

R&D FOR FOREST AND WOOD PRODUCTS INDUSTRIES<br />

Kathryn Adams 1<br />

ABSTRACT<br />

<strong>Australia</strong> has invested significantly in the development <strong>of</strong> a well managed forest and<br />

wood products industry, backed by internationally recognised certification for sustainable<br />

forest management and chain <strong>of</strong> custody. To maintain and improve its competitiveness<br />

the industry should not lose sight <strong>of</strong> the need to continue to invest in its innovation capital<br />

through focussed scientific research.<br />

In <strong>Australia</strong> the government provides mechanisms through Forest and Wood Products<br />

<strong>Australia</strong> (FWPA) and the Cooperative Research Centre (CRC) program. However, this<br />

joint investment can dilute direct engagement <strong>of</strong> commercial end users and they may miss<br />

some <strong>of</strong> the value that this opportunity provides unless they actively participate.<br />

This paper provides a reminder <strong>of</strong> the need for a clear industry strategic direction on<br />

which to base research investment priorities, followed by answers to the questions ‘why<br />

are we doing it, who will use it, how will they use it, are the answers there already, how<br />

should we manage the intellectual property to properly capture the value’? If this is done<br />

well, project design will focus the research on impact and outcomes rather than around<br />

convenient inputs.<br />

ADDRESS<br />

Ladies and Gentlemen,<br />

It is a privilege for me to be here today as a Keynote Speaker for this session on ‘Promoting<br />

Innovation in Forest Management and Processing’. But before I start I would like to emphasise that<br />

the views that I express here are my own and are not those <strong>of</strong> any organisation with which I might be<br />

associated, including Forest and Wood Products <strong>Australia</strong> Ltd and <strong>Australia</strong>n Forestry Standards Ltd.<br />

This theme <strong>of</strong> ‘promoting innovation’ is critical for the future <strong>of</strong> the forestry and wood products sector<br />

if it is to realise its full potential in <strong>Australia</strong> and overseas. <strong>Australia</strong> has invested significantly in the<br />

industry to ensure that this renewable resource is adding value to the environment, carbon reduction,<br />

carbon storage and the economy. This is supported by internationally recognised certification schemes<br />

for sustainable forest management and chain <strong>of</strong> custody. To maintain the benefit from this investment,<br />

the sector needs to be able to continue to substantiate its position through ongoing investment in<br />

science based research and development (R&D). It needs to take care that this investment keeps pace<br />

with market needs and that it don't use up its current store <strong>of</strong> innovation assets faster than it is creating<br />

new ones.<br />

The Macquarie Dictionary defines ‘innovate as:<br />

‘to bring in something new; make changes to anything established’.<br />

This can happen at a very simple level by changing the way information is filed or answering the<br />

phone in a more consistent and informative way. At a higher level it can include introducing new<br />

technology or production systems. More complex innovation is generally based on scientific R&D<br />

from within a company, commissioned by a company or arising from joint investments such as<br />

through FWPA or a CRC. If a company undertakes or commissions the R&D it is more likely to<br />

adopt the outcomes as it is intimately involved in the project and has commercial reasons for<br />

1<br />

Director, Forests & Wood Products <strong>Australia</strong>, Level 4, 10-16 Queen Street, Melbourne VIC 3000; and Director, <strong>Australia</strong>n<br />

Forestry Standard Limited.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 20<br />

investing. Where the R&D comes from an external source, even though companies might be<br />

contributing, it is more difficult to ensure that the R&D is meeting a specific need. I intend to focus<br />

on this ‘shared investment’ R&D and to emphasise the need for industry engagement at the company<br />

level if the full value <strong>of</strong> the investment is to be captured.<br />

I don't propose to tell you anything new but rather to repeat some <strong>of</strong> the key attributes that the industry<br />

could address to help maximise the return through the opportunities provided in <strong>Australia</strong> for joint<br />

investment in R&D. I will look at the issues under four headings:<br />

1. The forest and wood products sector in <strong>Australia</strong> today<br />

2. Some key principles for capturing the value from R&D<br />

3. Promoting and marketing the benefits<br />

4. Challenges for industry<br />

THE FOREST AND WOOD PRODUCTS SECTOR IN AUSTRALIA TODAY<br />

Industry Overview<br />

The Forest and Wood Products <strong>Australia</strong> (FWPA) Strategic Plan 2009-2013 (2008, p. 7), has a table<br />

showing the contribution <strong>of</strong> the forest and wood products sector to the <strong>Australia</strong>n Economy:<br />

Table 1: Some industry statistics<br />

Industry turnover $19 billion/year<br />

Direct employment 83,000 people<br />

Proportion <strong>of</strong> Gross Domestic Product 1%<br />

Proportion <strong>of</strong> manufacturing industry 7%<br />

The sector (including pulp) contributes about 1% <strong>of</strong> GDP. In relation to the non-pulp sector, figure 1<br />

shows that the estimated gross value <strong>of</strong> log production in <strong>Australia</strong> has increased (not adjusted for<br />

inflation) from $1,337 m in 2001-02 to $1,872 m in 2007-8.<br />

(from numbers in ABARE, 2009, p.13).<br />

In 2007-08 the apparent consumption <strong>of</strong> sawn wood increased by almost 9 per cent, mainly from<br />

increased use <strong>of</strong> s<strong>of</strong>twood both from <strong>Australia</strong> and overseas (ABARE, 2009 p2). Although growth<br />

indicators for the current year are not as positive due to the economic downturn, the sector contributes


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 21<br />

significantly to the <strong>Australia</strong>n economy and the gross value <strong>of</strong> log production has been increasing (in<br />

unadjusted terms) at least since June 2002.<br />

Investment in R&D<br />

For the sector to continue to contribute in this way to the <strong>Australia</strong>n economy it will need to maintain<br />

its competitiveness and credibility both in <strong>Australia</strong> and overseas. Much depends on the industry’s<br />

ability to innovate and substantiate its practices through scientific analysis. There is extensive<br />

literature on the role <strong>of</strong> scientific R&D in assisting with industry growth and commercial viability. Of<br />

itself R&D is not enough and needs to be accompanied by innovative management, marketing and<br />

sales – it is an integral part <strong>of</strong> a well managed business.<br />

Historically, State governments in <strong>Australia</strong> have produced the wood, while processing has been with<br />

private sector companies, many <strong>of</strong> them small sawmillers. Research relating to production <strong>of</strong> timber<br />

was undertaken by government research organisations and some universities whereas processing<br />

research was undertaken predominantly by CSIRO, universities and individual companies. As<br />

plantation timber has increased, more companies and government corporations have become involved<br />

in growing timber.<br />

To try and increase the level <strong>of</strong> private sector investment in R&D, the Forest and Wood Products<br />

Research and Development Corporation (FWPRDC) was established in 1994 as a Commonwealth<br />

Statutory Authority to invest in R&D for the industry. It was based on the model <strong>of</strong> other Primary<br />

Industry R&D Corporations (RDCs), with a levy on production <strong>of</strong> logs for processing, which was<br />

matched by the <strong>Australia</strong>n Government. In addition there was also a levy on imported logs to<br />

minimise any perceived advantage or ‘free rider’ effect that might accrue to importers <strong>of</strong> logs. The<br />

pulp and paper sector opted not to contribute (and take advantage <strong>of</strong> the Commonwealth contribution)<br />

by setting its levy at zero. Although producers <strong>of</strong> wood (predominantly the States) did not contribute<br />

to the levy (for constitutional reasons), they benefited through support for their research programs<br />

where there was clear value to the processing sector.<br />

In recent years a number <strong>of</strong> RDCs have been replaced by an industry owned company limited by<br />

guarantee to integrate R&D with marketing and industry promotion. In September 2007 Forest and<br />

Wood Products <strong>Australia</strong> Ltd (FWPA) was established in this way to replace FWPRDC and expand its<br />

role into other industry services, including generic promotion and marketing. As part <strong>of</strong> the agreement<br />

with the <strong>Australia</strong>n Government, the investment in R&D was not to be lower in dollar terms than the<br />

average <strong>of</strong> the last 3 years <strong>of</strong> FWPRDC ie $6.73 m per annum. Expenditure on R&D is matched by the<br />

<strong>Australia</strong>n Government up to 0.5% <strong>of</strong> industry GVP (the levy is currently only about 0.2% GVP and<br />

the pulp sector still has a zero levy). Expenditure on promotions and marketing for the industry is not<br />

matched.<br />

In addition, for the first time the forest growers are contributing to FWPA (the States on a voluntary<br />

levy basis) so the whole industry, from ground to wood product, has an interest in the value add and<br />

increased return on investment that can be obtained through FWPA's investment.<br />

The other avenue for joint investment in R&D in <strong>Australia</strong> is through the CRC Program, where<br />

research providers (including universities), companies and other R&D investors come together to<br />

invest in a specific, focussed R&D program. The Commonwealth runs a competitive bid process and<br />

contributes to the successful CRCs for 7 years initially, provided that the CRC meets its agreed<br />

research objectives. There have been at least 3 CRCs focussed on the forest and wood products sector.<br />

In a recent draft report for FWPA, Turner and Lambert (2009 – unpublished p 24) found that R&D<br />

investment by the sector in <strong>Australia</strong> has increased in unadjusted dollars, since 1982, from about $44.5<br />

million per annum to $104.6 million per annum in 2007-8 (6.2% <strong>of</strong> value <strong>of</strong> logs harvested). However,<br />

in real terms this is a decrease in investment to about $37.5 million in 1982 terms (figure 2). In<br />

addition, the R&D effort is fragmented, with over 50 organisations and 500 scientists and technicians,<br />

all with different strategic agendas. In 2007-8, 52% <strong>of</strong> research expenditure was provided directly or<br />

indirectly by the Commonwealth, 28% by State Governments and 20% by private companies (Turner<br />

and Lambert 2009 unpublished p14).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 22<br />

Total expenditure ($ million)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

1982 1986 1990 1995 2002 2008<br />

Assessment Year<br />

Total Actual<br />

Total 1982 Dollars<br />

Figure 2. Total research costs estimated over the study periods from 1981/82 to 2007/08 as both<br />

actual dollars and adjusted dollars (1982 base). (from Turner and Lambert 2009<br />

unpublished, p15).<br />

CAPTURING THE VALUE<br />

Capturing the value from investment in R&D is not always easy. There is a general rule <strong>of</strong> thumb that<br />

says only about 1 in 100 projects will yield results that get to commercial pilot stage and that maybe 1<br />

in 1000 will actually get to market and make money. These figures differ depending on the proximity<br />

to market, the complexity <strong>of</strong> the innovation and the level <strong>of</strong> change that is required in an operating<br />

business to implement the innovation. For example, if a whole new processing line is required and the<br />

existing one still has 20 years <strong>of</strong> life, then the value add <strong>of</strong> the new line will need to be considerable.<br />

However if the innovation can be readily integrated into the existing system, then the likelihood <strong>of</strong><br />

more immediate change is much greater.<br />

In 2007 Agtrans completed a cost benefit study for FWPA which showed that for the 36 completed<br />

projects in the sample there were a range <strong>of</strong> quantitative and qualitative benefits. The weighted<br />

average benefit at 5% discount rate, using 2005-6 dollars, was 11:1. Similarly the RDCs have just<br />

completed a joint cost-benefit analysis <strong>of</strong> a range <strong>of</strong> programs. They too found a return <strong>of</strong> about 11:1<br />

from 32 randomly selected projects rated across their combined portfolio. In addition a cost benefit<br />

analysis was done on 36 projects rated as ‘highly successful’ across the 15 RDCs. For a total $465<br />

million investment ($265 million from RDCs plus $200 m from partners), the value added was $10.5<br />

billion (22:1) being $5.5 billion in direct industry benefits and $5.0 billion in other benefits (eg<br />

environmental and social benefits). The project was undertaken by seven independent consultant<br />

groups and the methodology was reviewed by a number <strong>of</strong> key economic development agencies<br />

including:<br />

Treasury<br />

Department <strong>of</strong> Finance and Deregulation<br />

Department <strong>of</strong> Agriculture Fisheries and Forestry<br />

Productivity Commission<br />

<strong>Australia</strong>n Bureau <strong>of</strong> Agricultural and Resource Economics.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 23<br />

In the project Summary Report (Rural R&D Corporations (2008), p3) it was noted that:<br />

The returns attributable to the RDCs’ $265 million investment – $5.9 billion – will more than pay<br />

for the entire $4.5 billion invested by RDCs across 600 projects over the past 10 years'.<br />

These types <strong>of</strong> analyses are undertaken by most rural RDCs, CRCs and others who invest in R&D and<br />

innovation both in the private and public sector. For both FWPA and the RDC model in general, this<br />

is an excellent result and these analyses are a valuable tool in assessing the potential return on R&D<br />

investment.<br />

However, for the forest and wood products sector, where Turner and Lambert have noted the<br />

fragmented nature <strong>of</strong> the R&D effort, the question has to asked as to whether there may be ways <strong>of</strong><br />

improving this return even further by placing more emphasis on simple principles that are included in<br />

the current investment processes but where companies, investors and researchers could place more<br />

emphasis.<br />

R&D investment principles<br />

I would like to look at whether there are some more basic steps that can be taken to improve the<br />

usability and value add <strong>of</strong> R&D outputs. I will focus particularly on research programs that are for the<br />

benefit <strong>of</strong> an industry and the <strong>Australia</strong>n community and are not company specific; for example where<br />

a group <strong>of</strong> interested partners are investing jointly eg through FWPA, CRCs, Joint Ventures. I would<br />

also like to concentrate on the industry benefits in particular, while recognising that joint funding with<br />

government also requires community benefits. My purpose here is to highlight the need for more<br />

commercial engagement, on the basis that the government is engaging to ensure that the community<br />

benefits are being achieved in a more direct way than the company investors.<br />

Where a company is investing directly in R&D for its own benefit, it will take a close interest in the<br />

direction and progress <strong>of</strong> the project. Where there is more than one investor the specific benefit to each<br />

one is a little further removed and responsibility for the success <strong>of</strong> the investment can become diluted<br />

the greater the number <strong>of</strong> investors. It is then left to the researcher, an industry association or the<br />

coordinating organisation to do their best to interpret what they think the end users want.<br />

To help bring the focus back to end user driven R&D, I have built a simple framework based on eight<br />

fundamental questions to be asked by those identifying investment priorities and preparing and<br />

assessing investment proposals:<br />

• What R&D should we be investing in, or from another perspective, why are we investing in<br />

this research?<br />

• What outcomes/impact do we want to achieve as a result?<br />

• Who will use the outputs to achieve these outcomes?<br />

• How will the outputs be used?<br />

• How is industry ‘usability’ monitored during the project?<br />

• How is intellectual property identified and managed to maximise benefits to industry?<br />

• Are we making the best use <strong>of</strong> the information already available in this area?<br />

• How to we design a project to achieve the outcomes and address each <strong>of</strong> the above issues?<br />

The effectiveness <strong>of</strong> this approach depends on how seriously those assessing the proposed investment<br />

are in requiring complete and substantiated answers. Focus on outcomes rather than inputs can result<br />

in much more directed investment and more usable, value-adding outputs for the end users.<br />

The approach I am taking can be applied to all levels <strong>of</strong> research, whether it be short-term problem<br />

solving or longer term ‘good ideas’ research. It also applies whether the outcomes are industry or<br />

community specific. The answers may not be as clear cut for the longer term issues and may change<br />

over time, but the important thing for an industry is to have a focus for R&D as an integrated part <strong>of</strong><br />

an industry strategic plan. It can then be monitored and modified as circumstances change.<br />

Experience with a number <strong>of</strong> R&D investing organisations suggests that there is a general acceptance<br />

that this approach is already being taken, but while at the Board and top management levels this may<br />

be the case, time and other pressures mean that there can be considerable deviation on the ground. For<br />

example, most application forms from RDCs or CRCs have some questions around commercialisation


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 24<br />

and path to market, but very few <strong>of</strong> the project milestones actually reflect an integrated approach to<br />

market need, focussed research and industry usability. Similarly, The project summary is written<br />

before the questions about commercialisation and adoption so it is clear that the project is designed<br />

and then paths to market and industry use are made to ‘fit’ the project.<br />

What research should we invest in and why are we investing in this research?<br />

These are fundamental questions but are <strong>of</strong>ten glossed over with greater focus being given to inputs<br />

rather than addressing real needs. How an investor decides where the priority investment areas are and<br />

then which particular projects to invest in should be identified by 'market need'. ‘Who wants this<br />

research and how will the outputs value-add’ are key questions? To help this process it is important<br />

for industry to be able to identify its strategic directions and then the market drivers.<br />

• Industry Strategic Directions<br />

The first questions to ask is ‘does the forest and wood products sector in <strong>Australia</strong> have an<br />

industry strategic plan to identify where it wants to be in 5-50 years and what R&D,<br />

promotion and marketing does it need to get there? There have been some sector based<br />

approaches to this, including the ‘Plantations 2020 Vision’ developed in 2002, and the 2008<br />

Forest and Forest Products Committee ‘Forest Research Strategic Directions 2008-2011’.<br />

Neither <strong>of</strong> these really provides an industry driven, integrated approach to a strategic future<br />

which would provide a firmer base for decision making. If the <strong>Australia</strong>n forest and wood<br />

products sector could develop this sort <strong>of</strong> strategic framework, it could be used for a range <strong>of</strong><br />

purposes, but would be particularly useful for organisations like FWPA when developing<br />

investment priorities.<br />

To be successful, such a strategy needs the CEOs <strong>of</strong> the companies to engage and drive the<br />

strategic direction – it cannot be driven by researchers or industry association staff; they may<br />

have the ideas for innovation but the companies, at a senior level, must provide the strategic<br />

commercial direction and market drivers for R&D focus. This is easier said than done as<br />

market forces change rapidly and response time, particularly in the tree growing area, is long.<br />

CEOs <strong>of</strong>ten do not have the luxury <strong>of</strong> time to focus on industry wide strategies and where they<br />

do they need to be cognisant <strong>of</strong> Trade Practices Act restrictions.<br />

• Is there a market?<br />

Once a strategic direction has been developed and the broad areas for research investment<br />

have been outlined, researchers and investors can take a market based approach to project<br />

development by asking 'what research is needed to meet industry market expectations'.<br />

What outcomes/impact do we want to achieve as a result?<br />

The next question is to clearly articulate the benefits (quantitative and qualitative) that the investment<br />

is meant to achieve. This may not always be directly translated into dollars but may include practice<br />

change within a community or contribution to the science behind policy development. The main thing<br />

is that the desired impact can be identified and measured so that investment in R&D has clear goals.<br />

Who is going to use the outputs from the research?<br />

While at the higher strategic level, an area <strong>of</strong> research might be identified as being valuable, it is also<br />

important to identify the potential end users at the project level and ask them if they would use the<br />

results from this research in their business? If not, why not? Could the outputs be changed in any way<br />

to make them more usable? If they would use it, how long before they would be in a position to do<br />

so? What further work/investment would they need in-house?<br />

When asking the question it is important to understand the types <strong>of</strong> end-users. Those that are looking<br />

for new ideas are more likely to give positive answers than those that wait until something is fully<br />

proven and has been in the market a long time. This is important as the first group may be the ones to<br />

focus on at this stage in an investment program.<br />

It is also important to make sure that the target user is who you think it is. For example, who is the<br />

end user <strong>of</strong> research into new species with greater stiffness characteristics? Is it the grower, the<br />

sawmiller, the timber specifier/architect/engineer? Is it all <strong>of</strong> them and if so do they each have<br />

different ‘usability’ criteria?


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 25<br />

How are they going to use it?<br />

Once it has been ascertained that the end users are likely to use the outputs from the research the next<br />

question is how will they use it? This is critical for the usability <strong>of</strong> the outputs. With many research<br />

projects the last milestone is the production <strong>of</strong> a final report and/or a research paper. But is this usable<br />

by the end users? If the output is computer s<strong>of</strong>tware, is it user friendly? Will the end users need to<br />

undertake further adaptation work? Do they have the in-house skills to do that or should the research<br />

deliver a 'user' package as well?<br />

Will usability be monitored during the project?<br />

The usability <strong>of</strong> project outputs can <strong>of</strong>ten be improved during the project if users are kept engaged or<br />

asked for feedback at regular intervals. Build this into the project milestones, with ‘go-no-go’<br />

decision points. Waiting until the end can <strong>of</strong>ten mean that changes to project design that could have<br />

easily been made at an earlier stage, become major.<br />

Make sure that the users are giving real feedback by asking how will you use this in your business?<br />

Can we make it easier for you? How? Also make sure they are the ‘real’ users and not go betweens.<br />

There is the story <strong>of</strong> the industry project steering group who met every six months to discuss progress<br />

and were very supportive until at the end they were asked 'will you use this in your business' and most<br />

said 'no'! The concept was good but the cost-benefit <strong>of</strong> implementation was not there. Often when<br />

people are on a committee, they take a less accountable perspective than they would if they were<br />

assessing an investment using their own company’s money.<br />

How will intellectual property be identified and managed to optimise industry benefits?<br />

This is an issue that is not always handled well where there are multiple beneficiaries and where<br />

governments have contributed. The consequences <strong>of</strong> not handling it well and taking the easy 'public<br />

domain' route can mean that the attractiveness <strong>of</strong> the R&D outputs from a commercial perspective, are<br />

diminished and therefore the level <strong>of</strong> uptake and industry use may be lower than anticipated.<br />

R&D is the business <strong>of</strong> producing intellectual property (IP) and its management should be second<br />

nature to researchers, but our training does not do a good job in this area. All R&D produces IP. It<br />

takes many forms: some <strong>of</strong> it is directly protectable (eg PBR, patents, trade marks, copyright); some<br />

may need packaging into a 'system' before it can be protected; some can be protected through common<br />

law (eg trade secrets and confidential information) and other can be protected through contract. The<br />

important thing is to be clear why the research is being done, who for and how it will be used. Then a<br />

case by case 'structured' decision making process on IP management can be applied to each project, in<br />

the context <strong>of</strong> the portfolio <strong>of</strong> projects. Early disclosure may shut <strong>of</strong>f commercial avenues not only for<br />

one project but for others in the portfolio as well.<br />

If it is clear who the beneficiaries are then IP management/industry use plans can be developed to<br />

maximise benefits. Integrated protection can include more than one form eg trademark to protect an<br />

overall brand and send a message to the market about the product or service, together with copyright<br />

in manuals, brochures and publicity materials, Plant Breeder's Rights in new varieties and patents in<br />

new products or processes. Decisions in relation to who will have access to the IP and under what<br />

conditions are critical for maximising return on investment. For example, licensing the IP to particular<br />

groups, such as those who have contributed to the project, can provide incentive to invest/use, whereas<br />

putting the IP into the public domain gives access to everyone, which may reduce the commercial<br />

value.<br />

Are we making the best use <strong>of</strong> information already available?<br />

One aspect <strong>of</strong> industry focussed R&D that always troubles me is whether we are duplicating what has<br />

happened in the past and whether we are making the best use <strong>of</strong> the information already available.<br />

Good practice for any research project dictates that there be a literature search before a new project is<br />

started to ensure that the wheel is not being reinvented or impinging on someone else's intellectual<br />

property. What concerns me is that I don't know that this is done particularly well.<br />

The other aspect that bothers me is whether all the information in final reports for projects undertaken<br />

over the last 10 years is being used as well as it could be. When you see the number <strong>of</strong> reports<br />

produced and the little time that is given to reviewing them, particularly by industry, I sometimes ask


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 26<br />

should we stop where we are for 6-12 months and review/analyse these existing reports to see what we<br />

could use and what would be needed to make the outputs more usable – have we missed something?<br />

Can I design a project that will address the questions above and optimise the commercial usability?<br />

Traditionally projects are designed to meet broad research priorities and provide technical outputs.<br />

These technical solutions are then provided to the end user, only to discover that the fit with<br />

commercial reality is not always good. If the questions about why, who and how are answered before<br />

the project is designed, the usability <strong>of</strong> the outputs may be increased without jeopardising the scientific<br />

integrity <strong>of</strong> the project.<br />

Inclusion <strong>of</strong> an ‘industry use’ plan as part <strong>of</strong> the project with milestones related to ‘usability’ and<br />

‘market’ testing throughout the project rather than at the end can keep researchers and industry<br />

focussed on the work that is being done and the return that might be achieved.<br />

All <strong>of</strong> this may seem ‘motherhood’ to many, but experience has shown that a lot <strong>of</strong> our jointly funded<br />

research is based on available inputs rather than a thorough analysis <strong>of</strong> the why, who and how with<br />

design emphasis on usability and value add.<br />

PROMOTION AND MARKETING THE BENEFITS<br />

As FWPA is aware, as it seeks to move the industry forward through its 'Wood Naturally Better'<br />

campaign, you cannot promote a product unless you have sound scientific backing to substantiate the<br />

benefits. If that backing is in place, then a campaign is less likely to stall due to adverse response from<br />

competitors or the community. In addition the market research that monitors and evaluates and guides<br />

the direction <strong>of</strong> the campaign also delivers an insight into future market drivers and therefore R&D<br />

needs. You can't have one without the other and the integration <strong>of</strong> these two formerly separate<br />

activities can only have synergistic effects where 1+1=3 + .<br />

Along similar lines, the internationally endorsed (through the international Program for Endorsement<br />

<strong>of</strong> Forestry Certification Schemes - PEFC) <strong>Australia</strong>n Forestry Standards for Sustainable Forest<br />

Management and Chain <strong>of</strong> Custody can only strengthen the viability <strong>of</strong> forest and wood products in<br />

<strong>Australia</strong> and overseas, but again there cannot be a standard without solid scientific backing and<br />

ongoing research to keep our industry at the leading edge globally.<br />

This integration <strong>of</strong> R&D and generic marketing and promotion for the industry is a major step forward<br />

but its main benefits will occur when industry, through company engagement, uses the market<br />

intelligence to drive its strategic planning and from that focus and coordinate its investment in R&D<br />

for innovation.<br />

From an FWPA perspective, this is beginning, with a number <strong>of</strong> CEOs being appointed to the Board in<br />

late 2008. The role and focus <strong>of</strong> industry advisory groups is also being reviewed to obtain relevant<br />

industry engagement. The challenge is to engage the right people at the right level into the future.<br />

THE CHALLENGES<br />

To maximise the return on their joint investment in R&D, I see the challenges for our industry as:<br />

developing an industry strategic plan to identify where it sees itself being positioned in the<br />

next 5,10, 20 and 50 years to help focus joint R&D investment for the future<br />

identifying key market drivers where R&D could value add and assist with delivering the<br />

strategic outcomes<br />

taking a stronger role in the avenues available for industry feedback on R&D eg FWPA<br />

industry Advisory Groups, steering committees etc. If you can't answer 'yes' to the question<br />

'would I use that in my business' (assuming it is in the relevant target user group), then the<br />

investment should be questioned; similarly industry advisory groups should have a majority <strong>of</strong><br />

members from companies rather than industry associations – FWPA is your company and you<br />

should help it achieve your goals for you.<br />

encouraging young industry people to be engaged


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 27<br />

asking the questions set out above, or pushing others to do so for every new R&D investment<br />

with the project design as the last task, not the first; and not glossing over the questions –<br />

focussing on outcomes rather than inputs<br />

asking itself: if the return on investment from FWPRDC (now FWPA) is <strong>of</strong> the order <strong>of</strong> 11:1,<br />

as the Agtrans study indicates, supported by a similar result from all users <strong>of</strong> the RDC model,<br />

why is the industry investment (via levies) only about 0.2% <strong>of</strong> GVP when the <strong>Australia</strong>n<br />

Government will match up to 0.5% <strong>of</strong> GVP?; similarly why is the pulp and paper sector not<br />

taking advantage <strong>of</strong> this opportunity; and lastly<br />

could the return be greater if there was a less fragmented approach than that outlined by<br />

Turner and Lambert and is FWPA the glue that could help bring it together?<br />

I will end with the example provided at the recent CRCA Conference in Canberra earlier this year<br />

from the Roger Campbell, CEO <strong>of</strong> the Pork CRC (Campbell, 2009). In basic terms, the key outcome<br />

that the CRC is seeking from its investment in R&D is more $/kg <strong>of</strong> carcass weight. All projects are<br />

assessed against their ability to contribute to that outcome and they are regularly monitored to ensure<br />

they are continuing to meet it. If they are not performing against that target, they are either refocussed<br />

or discontinued. Not all organisations may want or be able to have such a single measure but it is an<br />

excellent example <strong>of</strong> the need to question what we are doing and why – is this contributing to our<br />

targeted outcomes and are we getting the best return from our investment?<br />

REFERENCES<br />

ABARE 2009, <strong>Australia</strong>n Forest and Wood Products Statistics, September and December quarters, 2008, p.13.<br />

http://www.abare.gov.au/publications_html/afwps/afwps_09/afwps_may09.pdf<br />

Campbell, R 2009, ‘Measuring Economic Impact’, presentation at Pathfinders: the Innovators Conference,<br />

Cooperative Research Centres Association Annual Conference, Canberra, 28 May 2009.<br />

http://www.crca.asn.au/conference/pdf/presentation/03_thursday/03_Murray_Room/1300_-<br />

_1600/Roger_Campbel_%20-_Economic_Impact.pdf<br />

Forest and Forest Products Committee 2008, Forest Research strategic Directions 2008-2011.<br />

http://www.forestscience.unimelb.edu.au/informing_government/rpcc/RPCC_Forest_Research_Priorities_<br />

Short.pdf<br />

FWPA 2008, Strategic Plan 2009-2013, p. 7.<br />

http://www.fwpa.com.au/Resources/About/5year/FWPA_Strategic_Plan_2009-2013.pdf<br />

Plantations for <strong>Australia</strong>: the 2020 Vision 2002, An industry/government initiative for plantation forestry in<br />

<strong>Australia</strong>. http://www.plantations2020.com.au/assets/acrobat/2020vision.pdf<br />

Rural R&D Corporations 2008, Measuring economic, environmental and social returns from Rural Research<br />

and Development Corporations’ investment, Report Summary.<br />

http://www.ruralrdc.com.au/WMS/Upload/Resources/Evaluation/Rural%20RDC%20Eval%20Summary%<br />

20low%20res.pdf<br />

The Macquarie Dictionary, 1981, Macquarie Library<br />

Turner, J and Lambert, M 2009, ‘Expenditure on forestry research and forest products research in <strong>Australia</strong><br />

2007-2008’, unpublished review for Forest and Wood Products <strong>Australia</strong>, March 2009. (final version<br />

should be on www.fwpa.com.au by September 2009.)


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 28<br />

ABSTRACT<br />

NATIVE FOREST MANAGEMENT OPTIONS FOR<br />

CLIMATE MITIGATION IN AUSTRALIA AND PNG<br />

Rodney Keenan 1<br />

Under the Kyoto Protocol, the use <strong>of</strong> forests to mitigate greenhouse gas emissions has<br />

been largely restricted to increasing forest carbon stocks through expanding the area <strong>of</strong><br />

forests compared with the area in 1990, and reducing emissions due to land clearing.<br />

Options to manage existing forests to reduce greenhouse gas emissions or increase<br />

carbon sequestration have been discussed for some time and were eventually included<br />

in the Kyoto Protocol on a voluntary basis. With the development <strong>of</strong> the national<br />

Carbon Pollution Reduction Scheme and a new international climate mitigation regime<br />

for the post-2012 period scheduled for agreement by December 2009, it is timely to<br />

review the options to manage existing native forests for climate change mitigation. For<br />

developed countries, forest management is likely to remain voluntary and broadly<br />

defined, including both emissions due to harvesting and sequestration in regrowth from<br />

all types <strong>of</strong> activities on managed forests. In developing countries, the potential<br />

inclusion <strong>of</strong> degradation under Reduced Emissions from Deforestation and Degradation<br />

(REDD) may mean that accounting is unbalanced, with emissions associated timber<br />

harvesting included but not regrowth sequestration. This paper analyses the situation for<br />

<strong>Australia</strong> and Papua New Guinea (PNG). Both countries currently have limited capacity<br />

to comprehensively report on carbon dynamics in managed native forests. In <strong>Australia</strong>,<br />

emissions from harvesting are more than <strong>of</strong>fset by managed regrowth, with emissions<br />

associated with wildfire likely to be the major factor impacting on carbon stocks. In<br />

PNG, harvesting is resulting in a net emission <strong>of</strong> carbon dioxide, even when regrowth<br />

sequestration is included. In PNG fire has limited impacts on carbon stocks. Options for<br />

reducing emissions are discussed. Establishing an agreed baseline against which to<br />

compare changes in emissions reductions remains a continuing policy and technical<br />

challenge.<br />

INTRODUCTION<br />

Vegetation and soils are major components <strong>of</strong> the global carbon cycle. The importance <strong>of</strong> effective<br />

management <strong>of</strong> these carbon sinks is recognised in the Framework Convention on Climate Change,<br />

which committed signatories to ‘promote sustainable management, and promote and cooperate in the<br />

conservation and enhancement, as appropriate, <strong>of</strong> sinks and reservoirs <strong>of</strong> greenhouse gases …<br />

including biomass, forests, oceans, and other terrestrial, coastal and marine ecosystems’ and other<br />

international agreements such as the Montreal Process on criteria and indicators <strong>of</strong> sustainable forest<br />

management (MIG 2008).<br />

In <strong>Australia</strong>, much <strong>of</strong> the focus has been on reducing greenhouse gas emissions from land clearing<br />

and increases in carbon storage in newly established forests. Comprehensive accounting <strong>of</strong> all carbon<br />

emissions and sinks has been proposed as a desirable long-term goal but there are still considerable<br />

risks and uncertainties associated with this approach (<strong>Australia</strong>n Government 2008).<br />

Globally, the amount <strong>of</strong> carbon stored in terrestrial ecosystems, and fluxes between ecosystems and<br />

the atmosphere, can change over time due to variation in climate and disturbances such as fires,<br />

storms, pests and diseases or human activities. Greenhouse gas emissions from deforestation are a<br />

significant proportion (18-20%) <strong>of</strong> total human emissions but there is also a substantial uptake in<br />

vegetation recovering from past clearing or disturbance, which varies considerably from year to year<br />

(Canadell et al. 2007).<br />

1<br />

Pr<strong>of</strong>essor Rodney J. Keenan, Department <strong>of</strong> Forest and Ecosystem Science, Melbourne School <strong>of</strong> Land and Environment,<br />

The University <strong>of</strong> Melbourne. Email: rkeenan@unimelb.edu.au.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 29<br />

Forests are part <strong>of</strong> the global carbon cycle through the natural processes <strong>of</strong> growth, disturbance, death<br />

and decay and human activities such as forest conversion, afforestation and utilisation <strong>of</strong> wood for<br />

energy, timber and other uses <strong>of</strong> wood and non-wood forest products. It is challenging to integrate all<br />

these components into a comprehensive assessment and policy framework for climate change<br />

mitigation.<br />

FORESTS AND CLIMATE CHANGE: THE POLITICS<br />

Inclusion <strong>of</strong> changes in carbon stocks in forests in Kyoto Protocol targets was controversial (Noble<br />

and Scholes, 2001, Keenan 2002). After considerable negotiation, the Protocol provided for Annex I<br />

countries to include change in carbon stocks due to afforestation, reforestation and deforestation and<br />

inclusion <strong>of</strong> eligible activities under Article 3.4 in broadly defined categories (forest management,<br />

cropland management, grazing land management and revegetation) at the discretion <strong>of</strong> the country.<br />

During the first commitment period, a Party that selected any or all <strong>of</strong> these activities needs to<br />

demonstrate that such activities have occurred since 1990 and are human-induced. Voluntary election<br />

<strong>of</strong> article 3.4 allowed Parties to leave out activities where there were methodological problems that<br />

meant there were high uncertainties related to net emissions or where the risk <strong>of</strong> emissions due to<br />

natural disturbances was potentially high. <strong>Australia</strong> chose not to include activities under Article 3.4,<br />

primarily due to risks <strong>of</strong> emissions associated with wildfire and drought.<br />

The 2007 Bali Action Plan put in place two negotiating processes, one addressing arrangements for<br />

Parties with commitments under the Kyoto Protocol beyond 2012 and another seeking global<br />

agreement for further co-operative action. Decisions under both processes are expected in<br />

Copenhagen in December 2009. Forests feature in both, with considerable debate over inclusion <strong>of</strong><br />

activities under Article 3.4 and for developing countries incorporating Reduced Emissions from<br />

Deforestation and Degradation (REDD). Measuring and accounting for changes in forest carbon<br />

stocks is a continuing challenge, as is the establishment <strong>of</strong> appropriate baselines. Adaptation <strong>of</strong><br />

societies to climate change, and achieving increased resilience and livelihood improvements, is a<br />

growing concern and aligning climate mitigation, poverty alleviation and biodiversity conservation<br />

objectives is a major policy challenge.<br />

In this paper, case studies from the Oceania region are considered to illustrate different issues<br />

associated with accounting for greenhouse gas emissions and removals due to forest management.<br />

<strong>Australia</strong> and PNG are near neighbours with quite different ecologies, indigenous cultures, European<br />

settlement histories, current land uses and stages <strong>of</strong> economic development. It has been proposed that<br />

the two countries become more deliberately linked in greenhouse gas emission reduction objectives,<br />

with reduced land-based emissions in PNG <strong>of</strong>fsetting emissions in other sectors in <strong>Australia</strong> (Garnaut<br />

2008). In international negotiations, the <strong>Australia</strong>n Government has suggested the inclusion <strong>of</strong><br />

changes in forest carbon stocks in the context <strong>of</strong> comprehensive accounting under ‘forest<br />

management’ but highlighted the need to factor out the unbalanced, episodic effects <strong>of</strong> events such as<br />

wildfire or drought. PNG has been an active promoter <strong>of</strong> the inclusion <strong>of</strong> REDD in future global<br />

arrangements for climate change mitigation.<br />

CARBON ACCOUNTING FOR FOREST MANAGEMENT<br />

In previous negotiations, accounting for greenhouse gas emissions and removals due to forest<br />

management in Article 3.4 considered defining activities as ‘broad’ or ‘narrow’. A broad definition<br />

(eg. forest management) required accounting for the effects <strong>of</strong> all practices on an area <strong>of</strong> land subject<br />

to the activity. A narrow definition would assess the effects <strong>of</strong> individually-defined practices (eg.<br />

fertilisation or species change). A broad approach was agreed, supported by land-based accounting.<br />

Activity-based approaches may re-enter the carbon accounting scene. For example, the recent<br />

Waxman-Markey Bill that recently passed the US House <strong>of</strong> Representatives provides for the<br />

inclusion <strong>of</strong> a long list <strong>of</strong> narrowly and broadly defined land management activities to meet proposed<br />

US emission reduction targets.<br />

Carbon pools specified for forest-based greenhouse gas accounting are: above and below ground<br />

living biomass, dead wood, litter and soil organic matter (IPCC 2006). Ideally, a framework for


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 30<br />

assessing GHG emissions associated with forest management involves (IPCC 2006, Penman et al<br />

2003):<br />

• Assessment <strong>of</strong> the forest area subject to the management activity.<br />

• Determining the extent <strong>of</strong> change in carbon stock associated with the activity.<br />

• Comprehensive accounting to include lands subject to past or present management.<br />

• Balanced accounting including all changes in carbon stocks.<br />

The extent <strong>of</strong> reduction in carbon stock can be estimated directly or by using emissions factors<br />

associated with indirect estimates <strong>of</strong> different activities. For example, reductions in carbon stock due<br />

to timber harvesting are <strong>of</strong>ten assessed using data on timber removals, because these statistics are<br />

more readily available than in-forest measures.<br />

Assessing the effect <strong>of</strong> change in practices or measures to reduce greenhouse gas emissions involves<br />

comparing the carbon stock change in one period (the baseline) with the change in the target period.<br />

1990 is the baseline year for the first commitment period <strong>of</strong> the Kyoto Protocol,. For afforestation or<br />

reforestation under Article 3.3 it was effectively assumed that there were no net emissions on cleared<br />

land in 1990. For deforestation, the 1990 emissions included the immediate loss <strong>of</strong> carbon associated<br />

land clearing and emissions due to decay <strong>of</strong> wood or soil carbon changes associated with clearing<br />

prior to 1990. For countries that included Forest Management under Article 3.4 accounting was on a<br />

‘gross-net’ basis with no 1990 baseline estimate because countries felt they could be penalised for<br />

imbalances in the age-class <strong>of</strong> their forest estate associated with past land use changes. Other<br />

measures were put in place to limit the extent to which countries could use this activity to meet<br />

emissions targets.<br />

Future requirements for accounting for greenhouse gas emissions in native forests are likely to be<br />

similar for forest degradation under new global arrangements and forest management under the<br />

Kyoto Protocol.<br />

Beyond 2012, a number <strong>of</strong> countries are arguing that a voluntary approach to Article 3.4 activities<br />

continues. Others have suggested that countries should at least demonstrate that carbon stocks in<br />

managed forests are being maintained (European Union 2008).<br />

Three alternative baseline options are currently under consideration: (i) maintaining a base year <strong>of</strong><br />

1990; (ii) using net emissions over a ‘base period’ covering a number <strong>of</strong> years to produce a reference<br />

carbon stock change; or (iii) developing a ‘forward-looking’ baseline based on business-as-usual.<br />

Assessing carbon dynamics over a longer period under option (ii) would reduce the effect <strong>of</strong><br />

imbalances in age-classes due to past human or natural disturbance, although that would require this<br />

period to cover a full cycle <strong>of</strong> plantation management (10-40 years) or even longer for cycles <strong>of</strong><br />

wildfire and regrowth. The challenges in the forward-looking baseline are determining the forward<br />

period, establishing a credible and verifiable baseline for that period and considering the potential<br />

impact <strong>of</strong> a carbon price.<br />

NATIVE FOREST HARVESTING IN AUSTRALIA<br />

Forests cover 149 million hectares, 147 million hectares <strong>of</strong> native forests and nearly 2 million<br />

hectares <strong>of</strong> forest plantations. The area <strong>of</strong> public forests zoned for multiple-use and available for<br />

harvesting declined from 11.4 M ha in 2000 to 9.4 M ha in 2006, and the area <strong>of</strong> public nature<br />

conservation reserves increased from 21.5 M ha to about 23 M ha over the same period, as a result <strong>of</strong><br />

processes to develop the conservation reserve system (MIG 2008).<br />

The carbon dynamics in intact native forests are complex. Forests vary widely in condition, with<br />

significant areas <strong>of</strong> old growth, with large trees and high carbon stocks (Raison et al. 2003, Keith et<br />

al. 2009). Others are in a regrowth condition or mixed-age class following past disturbance. There<br />

has been no systematic, repeated field-based inventory <strong>of</strong> <strong>Australia</strong>n native forests and forest carbon<br />

dynamics are generally poorly understood (Keenan 2002). Estimates <strong>of</strong> native forest timber removals<br />

are currently used as the basis for emissions reporting in native forest management, together with<br />

estimates <strong>of</strong> growth in regrowing forests following harvesting or other disturbance. Harvest rates are<br />

a function <strong>of</strong> the area <strong>of</strong> forest available for harvesting, estimated growth and market conditions.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 31<br />

Native forest harvest rates have declined by about 10% since the 1990s, to an average <strong>of</strong> about 9.2 M<br />

m 3 per year (ABARE 2009).<br />

The current national greenhouse gas inventory accounting approach uses national harvest figures to<br />

determine the quantity <strong>of</strong> carbon removed in timber harvest, combined with an emission factor<br />

associated with slash (currently about 0.9 times the timber removed, DCC 2009). Using these factors,<br />

a crude estimate <strong>of</strong> harvesting emissions is 5.68 MtC (20.8 Mt <strong>of</strong> CO2) or 2.3 t CO2 per cubic metre<br />

<strong>of</strong> timber removed. Some emissions from harvesting slash actually occur over time through decay,<br />

and a proportion <strong>of</strong> the timber removed is added to the wood products pool rather than the<br />

atmosphere. Emissions from burning firewood are estimated to be about 10 Mt <strong>of</strong> CO2 (MIG 2008).<br />

Thus, total emissions associated with timber harvesting and firewood removal are about 31 Mt CO2<br />

per year.<br />

Under a comprehensive assessment <strong>of</strong> forest management, these emissions are <strong>of</strong>fset by carbon<br />

sequestered in regrowth. Regrowth native forests were estimated to take up about 43.5 Mt CO2 per<br />

year in 2005 (MIG 2008). This suggests that carbon stocks in managed native forests are actually<br />

increasing by about 12.5 Mt <strong>of</strong> CO2 per year. The rate <strong>of</strong> increase is probably higher compared to a<br />

potential baseline period because <strong>of</strong> the reduced rate <strong>of</strong> timber harvesting, the shift from harvesting in<br />

old growth or mature forests to regrowth forests (with reduced slash emission factors). Also,<br />

estimated carbon uptake in regrowth does not include sequestration in a substantial area <strong>of</strong> regrowth<br />

forests that has been included in new conservation reserves and is not considered part <strong>of</strong> the<br />

‘managed’ forest estate. Emissions associated with firewood replace emissions that might otherwise<br />

occur due to heating with fossil fuels.<br />

In some local areas, long-term carbon stock reductions are occurring where old growth or mature<br />

forest is converted to regrowth through clearfelling or regrowth managed more intensively, with<br />

thinning and shorter rotations. Past timber harvesting or other disturbance may mean that carbon<br />

stocks in native forests are below their potential carbon carrying capacity (Roxburgh et al 2006) and<br />

it has been suggested that reducing native forest timber harvesting could result in long-term increases<br />

in native forest carbon stocks (Mackey et al. 2008). This is open to debate. A recent modelling study<br />

suggested that light selection harvesting could result in higher total forest carbon stock than in<br />

unharvested areas (Ranatunga et al. 2008).<br />

Additional reductions in emissions from native forest harvesting could be achieved by reducing the<br />

‘emissions factor’ associated with timber harvesting or increasing carbon stocks through regeneration<br />

and restoration <strong>of</strong> understocked areas. Reducing the emissions factor could be achieved by reducing<br />

the extent <strong>of</strong> clearfelling and conversion <strong>of</strong> old growth forest to even-aged regrowth forest, reducing<br />

the impact on retained stems during selection harvesting operations, retaining live standing trees in<br />

harvested areas, or maintaining a greater level <strong>of</strong> biomass in woody debris on sites after harvesting,<br />

perhaps through a reduction in intensity <strong>of</strong> slash burning and site preparation.<br />

Reducing removals from native forests may result in increased carbon stocks. However, this requires<br />

further analysis <strong>of</strong> sequestration capacity at different forest growth stages. Given that harvesting is<br />

now mostly <strong>of</strong> regrowth stands, harvesting at peak MAI may be the optimal short term carbon<br />

sequestration option, particularly if accounting includes storage in wood products. The overall effect<br />

<strong>of</strong> reducing native forest harvesting also requires assessment <strong>of</strong> ‘leakage’ associated with increased<br />

timber imports or more intensive utilisation <strong>of</strong> plantations and analysis <strong>of</strong> the cost <strong>of</strong> this emission<br />

mitigation measure compared to other options. Achieving maximum carbon potential will depend on<br />

effective management to maintain continued sequestration rates and consideration <strong>of</strong> the impacts <strong>of</strong><br />

wildfire. This is discussed in the next section.<br />

FIRE IN AUSTRALIAN NATIVE FORESTS<br />

Fire is a widespread feature <strong>of</strong> the <strong>Australia</strong>n landscape and much <strong>of</strong> the vegetation is adapted to or<br />

even dependent on fire and fire-related disturbances for regeneration and survival (Bradstock et al.<br />

2002). Wildfires have significant impacts on life, health, property, infrastructure and primary<br />

production systems (Whelan et al. 2006) and fire affects nutrient cycling and availability, forest<br />

productivity, vegetation composition, wildlife habitat, soil biota and hydrological functioning. Large-


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 32<br />

scale wildfires can occur as a result <strong>of</strong> deliberate (arson) or unintentional (eg. powerlines) human<br />

activities, or naturally through lightning strikes. Prescribed fire is widely used as a management tool,<br />

although this practice is not without controversy and the rate <strong>of</strong> burning in eastern <strong>Australia</strong> has<br />

declined in recent years after peaking in the early 1980s (Adams and Attiwill 2008).<br />

The incidence and severity <strong>of</strong> wildfire varies considerably from year to year and decade to decade,<br />

depending on climatic conditions, vegetation type, fuel loads and human actions, and there may be<br />

prolonged periods (75-150 years) between ‘stand replacing’ fires (McCarthy et al. 1999). The area <strong>of</strong><br />

forest impacted by fire is not well-monitored and fires vary greatly in extent, intensity and severity <strong>of</strong><br />

their impacts. Accounting for greenhouse gas emissions from fire is also problematical. ‘Emissions<br />

factors’ associated with different fire intensities have not been developed. Fire results in some<br />

emissions directly as CO2 but smoke is a complex mix <strong>of</strong> organic compounds, particulates and<br />

carbon monoxide. After a fire, a considerable amount <strong>of</strong> wood is left in the forest. In many forest<br />

types impacted trees can recover quickly. Dead standing trees decompose over time and some <strong>of</strong> the<br />

remaining biomass is converted to charcoal that may have long-term stability in the soil.<br />

In south eastern <strong>Australia</strong>, where much <strong>of</strong> the native forest timber harvest is now focused, there has<br />

been a relatively high incidence <strong>of</strong> large scale wildfires since 2000, with major events in the summers<br />

<strong>of</strong> 2002-03, 2006-07 and in February 2009, partly as a result <strong>of</strong> a prolonged period <strong>of</strong> below average<br />

rainfall. Over 2.6 M ha have been burnt in these events with significant consequences for forest<br />

carbon stocks. Estimates <strong>of</strong> carbon dioxide emissions associated with these fires range from 40 Mt<br />

for the first (MIG 2008) to 550 Mt for both (Attiwill and Adams 2008). The most recent events on 7<br />

February 2009 which resulted in the deaths <strong>of</strong> 173 people and the loss <strong>of</strong> over 2,000 homes may have<br />

resulted in carbon emissions <strong>of</strong> over 50-70 Mt CO2. This is about three times the emissions that have<br />

occurred due to timber harvesting over the past 10 years (excluding firewood emissions).<br />

If the fire interval is sufficiently infrequent, this emitted carbon will be sequestered over time in<br />

regrowth following the fires. However, successive fires in the same area may have long-term impacts<br />

on forest carbon. Higher fire frequency can lead to shifts in vegetation composition and reduced<br />

forest carbon stocks. Regular prescribed fire can reduce the intensity <strong>of</strong> wildfires, so the potential<br />

interactions are complex. Under future climate change scenarios, the frequency <strong>of</strong> severe fire weather<br />

days is projected to increase over the next 20-40 years (Hennessy 2007) with significant implications<br />

for forest composition, structure and carbon stocks.<br />

Research is required to support improved accounting <strong>of</strong> forest carbon dynamics (particularly soil<br />

carbon) associated with different types <strong>of</strong> fire to provide a better basis for incorporation <strong>of</strong> forest fire<br />

management in climate change policy (Attiwill and Adams 2008). The extent to which wildfires are<br />

considered human-induced and assessing the effect <strong>of</strong> management practices (such as prescribed<br />

burning) that aim to mitigate fire impacts will be key policy design issues.<br />

FOREST HARVESTING IN PAPUA NEW GUINEA<br />

Papua New Guinea has a wide variety <strong>of</strong> environments and forests are characterised by high species<br />

diversity. Human societies in PNG are also highly diverse, with over 700 different language groups<br />

and a large number <strong>of</strong> different cultural and ethnic groups in the population <strong>of</strong> about 6 million<br />

people, and many people are dependent on forests for their livelihoods.<br />

There are five primary drivers <strong>of</strong> forest change in PNG: agricultural conversion, fire, mining,<br />

subsistence agriculture and timber harvesting (Filer et al. 2009). Conversion to intensive agriculture<br />

has been relatively limited in PNG. About 120,000 ha <strong>of</strong> oil palm plantation has been established<br />

over 30 years <strong>of</strong> development and not all <strong>of</strong> this has resulted in forest conversion. Fire has been<br />

shaping PNG’s vegetation patterns through thousands <strong>of</strong> years. If the interval is not too frequent,<br />

forests generally recover from fire and the structure <strong>of</strong> forests has in some parts been determined by<br />

previous fires (Haberle et al. 2001). Mining has locally significant impacts on forest cover, particular<br />

in Western Province, where deposition <strong>of</strong> mine tailings from Ok Tedi and overbank flooding is<br />

continuing to cause the death <strong>of</strong> significant areas <strong>of</strong> riparian forest. Subsistence agriculture appears to<br />

take place largely in cycles <strong>of</strong> cultivation and regrowth fallow and does not appear to be impacting<br />

extensively on primary forest (Filer et al. 2009). Timber harvesting is therefore the activity with the


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 33<br />

greatest potential impact on forest carbon stocks. It has been suggested that harvesting is resulting in<br />

long-term degradation <strong>of</strong> forest carbon stocks and emissions <strong>of</strong> 75 – 85 Mt/year <strong>of</strong> CO2 and that there<br />

is little recovery <strong>of</strong> carbon stocks after harvesting (Shearman et al. 2008).<br />

Almost all land in PNG is under customary or tribal ownership. Forests have been seen as the basis<br />

for economic development for some time. Under the National Forest Plan, 11.9 million ha <strong>of</strong> land is<br />

identified as the production forest estate. Most <strong>of</strong> the forest subject to timber harvesting is allocated<br />

to processors and exporters through timber rights purchase and management agreements negotiated<br />

between the government and landowner groups, and approved by the National Forest Board. Most <strong>of</strong><br />

the timber removed is exported, and log exports rates have varied considerably, ranging between 1 to<br />

2 M m 3 per year over the last 10 years.<br />

Harvesting occurs largely in primary forests. It is selective, with the intensity <strong>of</strong> felling and<br />

associated damage varying widely with the density <strong>of</strong> merchantable species and the market<br />

requirements, and the skill <strong>of</strong> the forest operator. Some areas are permanently converted due to<br />

roading or inadequate regeneration. Average timber removals are 10-20 m 3 /ha (Keenan et al. 2005).<br />

Using a figure <strong>of</strong> 15 m 3 /ha combined with the reported annual log export volume results in an<br />

estimated total <strong>of</strong> about 3.2 M ha impacted by selective harvesting up to 2007.<br />

Analysis <strong>of</strong> 125 permanent sample plots across the country (Fox et al 2009) indicated that timber<br />

harvesting reduces the average carbon density in above-ground live biomass (AGLB) by about 33%,<br />

or about 40 tC/ha. Applying this stock reduction to the harvested area associated with timber<br />

removals gives average CO2 emissions associated with timber harvesting <strong>of</strong> about 26 Mt per year<br />

over the 10 years from 1999 to 2008 (including an allowance for total loss <strong>of</strong> carbon stock on areas<br />

subject to roading and conversion to gardens), or about 17 t CO2 per cubic metre <strong>of</strong> timber removed<br />

(7.5 times the <strong>Australia</strong>n value).<br />

The long-term impact on carbon stock in these forests will depend on their recovery capacity. Eighty<br />

nine <strong>of</strong> the above plots had measurement periods longer than about 5 years. About 76% showed an<br />

increase in basal area over the period (Yosi pers. comm.). Applying this percentage and carbon<br />

uptake figure <strong>of</strong> 2 tC/ha/year (Fox pers. comm.) to the cumulative area harvested results in an<br />

estimated rate <strong>of</strong> sequestration 12.8 Mt CO2/year.<br />

Thus, timber harvesting in primary forests in PNG is resulting in a net reduction in forest carbon<br />

stocks <strong>of</strong> about 13 Mt CO2/year when regrowth is included. Whether this stock reduction is long-term<br />

will depend on future timber harvesting or other human disturbances. The extent <strong>of</strong> future harvesting<br />

will depend on the residual stocking and rate <strong>of</strong> growth <strong>of</strong> merchantable species, the extent and<br />

location <strong>of</strong> the forest, future market conditions and, possibly, the value to the forest owners <strong>of</strong> any<br />

future avoided greenhouse emissions. Some accessible areas <strong>of</strong> are already being subjected to further<br />

cutting, either by larger companies or small-scale sawmilling.<br />

Reduced emissions could be achieved by reducing the rate <strong>of</strong> harvesting, reducing the impacts <strong>of</strong><br />

harvesting (Putz et al. 2008) or both. Incorporating these reductions into a payment mechanism such<br />

as REDD will require establishment <strong>of</strong> a baseline. This will vary considerably depending on the<br />

period chosen because <strong>of</strong> the large fluctuation in historical harvest levels (Bank <strong>of</strong> PNG 2008).<br />

Ongoing monitoring will be required to assess the whether stock reductions are long-term and<br />

degradation is occurring.<br />

DISCUSSION<br />

The state <strong>of</strong> knowledge <strong>of</strong> carbon dynamics across the <strong>Australia</strong>n and PNG landscapes is still highly<br />

uncertain. Research can improve capacity to quantify carbon fluxes across the landscape, but<br />

uncertainty is likely to remain high until comprehensive forest monitoring systems are established in<br />

both countries. In <strong>Australia</strong>, national accounting <strong>of</strong> greenhouse gas emissions due to deforestation is<br />

based on well-established system <strong>of</strong> remote sensing coupled with models <strong>of</strong> wood decay, soil carbon<br />

dynamics and regrowth sequestration (DCC 2009). Accurate accounting for carbon stock changes<br />

associated with forest management is less amenable to such approaches and requires a combination<br />

<strong>of</strong> remote sensing and ground-based measurements to verify carbon stock changes. The quantum <strong>of</strong><br />

emissions and removals associated with different types <strong>of</strong> activities are not insignificant and justify


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 34<br />

investment in improved monitoring, particular when it can be integrated with monitored timber<br />

stocks, vegetation composition and other values.<br />

A particular challenge is development <strong>of</strong> historical baselines. Establishment <strong>of</strong> a sound baseline<br />

using historical remote sensing data or aerial photographs requires considerable local knowledge to<br />

properly interpret past forest condition. To effectively incorporate changes in carbon stocks due to<br />

timber harvesting, fire or other forest management actions is likely to require the establishment <strong>of</strong> a<br />

retrospective baseline. Separating human-induced impacts (including indirect effects on forest<br />

growth such CO2-fertilisation) and legacy effects <strong>of</strong> pre-1990 activities (that have resulted in current<br />

forest age structures) from natural processes such as fire and inter-annual climatic variability will be<br />

a continuing scientific and policy challenge.<br />

A comprehensive accounting approach, that includes both losses due to disturbances and<br />

sequestration in regrowth, is important to provide balanced accounting and to provide incentives for<br />

effective establishment and management <strong>of</strong> regrowth, to restore degraded or failed areas and to<br />

maintain fully stocked stands.<br />

The interaction <strong>of</strong> timber harvesting and fire impacts is likely to become an increasingly important<br />

issue under future forest management and climate change scenarios. It has been argued that<br />

harvesting exacerbates the risk <strong>of</strong> fire in tropical forests but some types <strong>of</strong> harvesting activities may<br />

mitigate the effects <strong>of</strong> wildfire in temperate regions (Hurteau et al. 2008).<br />

Other forest values are also important. Maintaining variability in forest structure and understorey<br />

species can maintain carbon stocks and provide wildlife habitat benefits. Analysis <strong>of</strong> forest<br />

management effects on carbon stocks also should not stop at the forest boundary. The recent Fourth<br />

Assessment Report <strong>of</strong> the IPCC noted that carbon storage in wood products, replacing energy<br />

intensive building materials and fossil fuel emissions using bi<strong>of</strong>uels from wood can increase forest<br />

carbon benefits. We need to utilise extracted timber as efficiently as possible to further increase<br />

carbon stocks in wood products and replace emissions from fossil fuels. The optimal solution to<br />

forest carbon management needs to consider the whole carbon lifecycle (Lindner, M. et al. 2008).<br />

ACKNOWLEDGEMENTS<br />

Colin Filer, Bryant Allen, John MacAlpine provided valuable background on PNG. Figures on<br />

carbon stocks and dynamics in PNG have been the result <strong>of</strong> long-term collaboration with the Papua<br />

New Guinea Forest Research <strong>Institute</strong> funded by the <strong>Australia</strong>n Centre for International Agricultural<br />

Research (ACIAR). Julian Fox and Cossey Yosi provided data and other input from this work.<br />

REFERENCES<br />

ABARE 2009. Quarterly Forest Products Statistics. <strong>Australia</strong>n Bureau <strong>of</strong> Agriculture and Resource Economics,<br />

Canberra.<br />

Attiwill PM, Adams MA (2008) Harnessing forest ecological sciences in the service <strong>of</strong> stewardship and<br />

sustainability: a perspective from ‘down-under’. Forest Ecology and Management 256, 1636-1645.<br />

<strong>Australia</strong>n Government 2008. Land Use, Land-Use Change and Forestry (LULUCF) Sector Submission to the<br />

AWG-KP and AWG-LCA, UN Framework Convention on Climate Change.<br />

Bradstock, RA, Williams, JE and Gill, AM, (Eds) 2002, Flammable <strong>Australia</strong>: the Fire Regimes and<br />

Biodiversity <strong>of</strong> a Continent, Cambridge University Press, Cambridge.<br />

Canadell JG, Corinne Le Quéré, Michael R. Raupach, Christopher B. Field, Erik T. Buitehuis, Philippe Ciais,<br />

Thomas J. Conway, RA. Houghton, Gregg Marland (2007) Contributions to accelerating atmospheric<br />

CO2 growth from economic activity, carbon intensity, and efficiency <strong>of</strong> natural sinks. Proceedings <strong>of</strong> the<br />

National Academy <strong>of</strong> Science, 0702737104.<br />

DCC 2008. Agenda paper: International accounting rules relating to emissions from landuse: implications for<br />

scheme design’, <strong>Australia</strong>n Government Department <strong>of</strong> Climate Change, Canberra.<br />

http://www.climatechange.gov.au/emissionstrading/consultation/pubs/international-accounting-rules-<br />

19may2008.pdf (accessed 9 August 2009).<br />

DCC 2009. National Inventory Report 2007. Appendix 7.B. Harvested Native Forests. <strong>Australia</strong>n Government<br />

Department <strong>of</strong> Climate Change, Canberra.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 35<br />

Filer, C. Keenan, RJ, Allen, BJ and MacAlpine J. 2009. Deforestation and forest degradation in Papua New<br />

Guinea. Annals <strong>of</strong> Forest Science (in press)<br />

European Union 2008. Submission by France on behalf <strong>of</strong> The European Community and its member states:<br />

Land use, land-use change and forestry (AWG-KP). Submission to UN Framework Commission on<br />

Climate Change.<br />

Fox, JC, Yosi, CK, Oavika, F., Nimiago, P., Lavong, K. and Keenan, RJ 2009a. Assessing forest carbon in<br />

Papua New Guinea. Biotropica (in review).<br />

Garnaut, R, 2008. The Garnaut Climate Change Review: Final report. Cambridge University Press, Melbourne.<br />

Haberle, SG, Hope, GS and van der Kaars, S, 2001. Biomass burning in Indonesia and Papua New Guinea:<br />

Natural and human induced fire events in the fossil record. Palaeogeogr. Palaeoecol. 171(3/4): 259-268.<br />

Hennessy, K, Lucas, C, Nicholls, N, Bathols, J, Suppiah R, and Ricketts, J. 2005. Climate change impacts on<br />

fire-weather in south-east <strong>Australia</strong>. CSIRO, Melbourne.<br />

Hurteau M.D., Koch G.W., Hungate B.A. 2008. Carbon protection and fire risk reduction: towards a full<br />

accounting <strong>of</strong> forest carbon <strong>of</strong>fsets. Frontiers in Ecology and Environment 6, 493-498.<br />

IPCC 2006. Good practice guidance for land use, land-use change and forestry (edited by J. Penman, M.<br />

Gytarsky, T. Hiraishi, T. Krug, D. Kruger, R. Pipatti, L. Buendia, K. Miwa, T. Ngara, K. Tanabe and F.<br />

Wagner). <strong>Institute</strong> for Global Environmental Strategies (IPCC National Greenhouse Gas Inventories<br />

Programme), Kanagawa.<br />

Keenan, R.J. 2002. Historical vegetation dynamics and the carbon cycle: current requirements and future<br />

challenges for quantifying carbon fluxes in <strong>Australia</strong>n terrestrial ecosystems. Aust. J. Bot. 50:533-544.<br />

Keenan, RJ, Ambia, V, Brack, C, et al. 2005. Improved timber inventory and strategic forest planning in Papua<br />

New Guinea. Bureau <strong>of</strong> Rural Sciences, Canberra, and Forest Research <strong>Institute</strong>, Lae.<br />

Lindner, M., Green, T, Woodall, TW, Perry, CH, Nabuurs, GJ and Sanz, MJ 2008. Impacts <strong>of</strong> forest ecosystem<br />

management on greenhouse gas budgets. Forest Ecology and Management, 256: 191-193.<br />

McCarthy, MA, Gill, AM, and Lindenmayer, DB 1999. Fire regimes in mountain ash forests: evidence from<br />

forest age structure, extinction models and wildlife habitat. Forest Ecology and Management 124: 193-<br />

203.<br />

MIG 2008. <strong>Australia</strong>’s State <strong>of</strong> the Forest Report 2008. Montreal Implementation Group and Bureau <strong>of</strong> Rural<br />

Sciences, Canberra.<br />

Noble, I. and Scholes, R. 2001. Sinks and the Kyoto Protocol. Climate Policy, 1:5-25.<br />

Penman, J., Gytarsky, M., Hiraishi, T., Krug, T., Kruger, D., Pipatti, R., Buendia, L. Miwa, K. Ngara, T.<br />

Tanabe, K. and Wagner, F. (eds). Definitions and Methodological Options to Inventory Emissions from<br />

Direct Human-induced Degradation <strong>of</strong> Forests and Devegetation <strong>of</strong> Other Vegetation Types.<br />

Intergovernmental Panel on Climate Change and <strong>Institute</strong> for Global Environmental Strategies,<br />

Kanagawa.<br />

Putz, F.E., Zuidema, P.A., Pinard, M., Boot, R.G.A., Sayer, J.A., Sheil, D, Sist, P. Vanclay, J.K. 2008.<br />

Improved Tropical Forest Management for Carbon Retention. PLoS Biology 6:1368-1369.<br />

Raison, J, Keith, H, Barrett, D, Burrows B, and Grierson, P 2003. Spatial estimates <strong>of</strong> mature biomass. Tech.<br />

Rep. 44, <strong>Australia</strong>n Greenhouse Office, Canberra.<br />

Keith, H, Mackey, BG and Lindenmayer D. 2009. Re-evaluation <strong>of</strong> forest biomass carbon stocks and lessons<br />

from the world’s most carbon-dense forests. Proceedings <strong>of</strong> the National Academy <strong>of</strong> Science,<br />

0901970106.<br />

Ranatunga, K, Keenan, RJ Wullschleger, SD Post, WM and Tharp, ML 2008. Effects <strong>of</strong> harvest management<br />

practices on forest biomass and soil carbon in eucalypt forests in New South Wales, <strong>Australia</strong>:<br />

Simulations with the forest succession model LINKAGES. Forest Ecology and Management, 255:2407–<br />

2415.<br />

Roxburgh, SH, Wood, SW, Mackey, BG, Woldendorp, G and Gibbons, P 2006, Assessing the carbon<br />

sequestration potential <strong>of</strong> managed forests: a case study from temperate <strong>Australia</strong>, Journal <strong>of</strong> Applied<br />

Ecology, 43:1149–59.<br />

Shearman, PL, Bryan, JE, Ash, J, Hunnam, P, Mackey, B and Lokes, B, 2008. The state <strong>of</strong> the forests <strong>of</strong> Papua<br />

New Guinea: Mapping the extent and condition <strong>of</strong> forest cover and measuring the drivers <strong>of</strong> forest change<br />

in the period 1972-2002. University <strong>of</strong> Papua New Guinea, Port Moresby.<br />

Whelan, R, Kanowski, P. Gill, M. and Andersen, A. 2006. Living in a land <strong>of</strong> fire. prepared for the 2006<br />

<strong>Australia</strong>n State <strong>of</strong> the Environment Committee.<br />

http://www.environment.gov.au/soe/2006/publications/integrative/fire/effects-<strong>of</strong>-fire.html (accessed 10<br />

August 2009).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 36<br />

ABSTRACT<br />

AUSTRALIA’S PLANTATION VULNERABILITY<br />

TO CLIMATE CHANGE<br />

Jody 1 Bruce, Michael Battaglia, Libby Pinkard<br />

Predicting the likely impact <strong>of</strong> climate change on plantation productivity and<br />

security is fraught with uncertainty. While there is now some confidence as<br />

to the climate to 2030, models and scenarios beyond that period diverge. In<br />

addition we know little about how our plantation species will respond to<br />

elevated C02. Finally, the evidence we have to date suggests that climate<br />

change impacts will depend on the site conditions that present at a particular<br />

plantation locale. Changes in pest/pathogen distributions and the ability <strong>of</strong><br />

forests to recover from damage will depend on the inherent fertility <strong>of</strong><br />

individual sites and the future climate. Other disturbances such as fire<br />

introduce further uncertainty. We developed a series <strong>of</strong> surfaces across<br />

<strong>Australia</strong>’s plantation estate for Eucalyptus globulus, E. nitens, radiata pine<br />

and hybrid pine using a range <strong>of</strong> photosynthetic responses to elevated C02 in<br />

combination with varying fertility and future climates to investigate the<br />

changes in productivity and risk <strong>of</strong> drought death.<br />

This study presents an analysis that attempts to understand the spatial<br />

distribution <strong>of</strong> this uncertainty across <strong>Australia</strong>, identify on which parts <strong>of</strong><br />

the landscape do the assumptions we make matter, and establish broad trends<br />

<strong>of</strong> change and risk under different combinations <strong>of</strong> assumptions. This<br />

allowed us to determine where climate change is likely to be damaging,<br />

where neutral or beneficial and where we are uncertain without more<br />

information and detailed site by site analysis. Some adaptation strategies to<br />

minimize risks are also investigated.<br />

1<br />

CSIRO Sustainable Ecosystems & CRC for Forestry, Private Bag 12, Hobart, Tas, 7001. Ph- +61 3 6237 5616.<br />

Email: Jody.Bruce@csiro.au


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 37<br />

ENVIRONMENTAL MARKETS AND AUSTRALIAN <strong>FORESTRY</strong>: THE<br />

EMERGENCE OF A NEW PERIOD OF INDUSTRY DEVELOPMENT<br />

Paul Dargusch 1 , Steve Harrison 1 and John Herbohn 1<br />

ABSTRACT<br />

This paper investigates how carbon markets can influence change at an industry level<br />

by studying the case <strong>of</strong> the <strong>Australia</strong>n forest industries. The influence <strong>of</strong> carbon<br />

markets on the current structure <strong>of</strong> the <strong>Australia</strong>n forest industries is mostly evident in<br />

the emergence <strong>of</strong> 24 new firms that are primarily engaged in the business <strong>of</strong><br />

afforestation-based <strong>of</strong>fsets. Whilst these firms currently account for only a small<br />

amount <strong>of</strong> business (in terms <strong>of</strong> financial turnover and area <strong>of</strong> tree plantings) compared<br />

with the traditional constituents <strong>of</strong> the <strong>Australia</strong>n forest industries, their emergence is<br />

indicative <strong>of</strong> a changing strategic outlook for the industry in which carbon markets are<br />

likely to play a much more influential role. Fervent debate in <strong>Australia</strong> during the last<br />

two years over a proposed mandatory emissions trading scheme (which allows the use<br />

<strong>of</strong> <strong>of</strong>fsets generated through afforestation projects undertaken within <strong>Australia</strong>) has<br />

made the traditional constituents <strong>of</strong> the <strong>Australia</strong>n forest industries give special<br />

consideration to how carbon markets will affect their businesses. The case study<br />

highlights the important role that entrepreneurial first-movers play in evoking change<br />

in industries. The authors argue that regulated carbon market policy should therefore be<br />

designed to support the capacity <strong>of</strong> entrepreneurial firms to engage in carbon markets,<br />

particularly in the early stages <strong>of</strong> the introduction <strong>of</strong> a mandatory emissions trading<br />

scheme and particularly in terms <strong>of</strong> how policy influences the entrepreneurial firms’<br />

capacity to secure finance.<br />

INTRODUCTION<br />

Carbon markets are defined here as those markets associated with the trade <strong>of</strong> greenhouse gas<br />

emissions <strong>of</strong>fsets and regulator-issued permits (mandatory and voluntary) to emit greenhouse gases.<br />

Carbon markets are particularly interesting in the <strong>Australia</strong>n context because the current <strong>Australia</strong>n<br />

government has proposed legislation to implement a mandatory greenhouse gas emissions trading<br />

scheme (Carbon Pollution Reduction Scheme (CPRS)) commencing in 2011. The CPRS has received<br />

substantial attention in the <strong>Australia</strong>n mainstream media over the past two years and firms have begun<br />

to strategize to deal with the scheme’s pending introduction. The case <strong>of</strong> forest-related industries in<br />

<strong>Australia</strong> is also particularly interesting because the proposed CPRS allows <strong>of</strong>fsets generated from<br />

afforestation-based activities carried out within <strong>Australia</strong> to be used by organizations to meet their<br />

CPRS compliance obligations.<br />

The <strong>Australia</strong>n forest industries have had a turbulent history, characterized by a handful <strong>of</strong> notable<br />

political, market and investment-related events. It is convenient to consider the history <strong>of</strong> the<br />

<strong>Australia</strong>n forest industries within five broad periods (Figure 1). The way in which carbon markets<br />

have influenced the changing structure and strategic outlook <strong>of</strong> the <strong>Australia</strong>n forest industry – period<br />

5 - is discussed in the following sections <strong>of</strong> this paper. Evidence is presented in the form <strong>of</strong> an analysis<br />

<strong>of</strong> new types <strong>of</strong> businesses entering the industry. The paper concludes by discussing what this change<br />

might mean for the traditional constituents <strong>of</strong> the <strong>Australia</strong>n forest industries.<br />

1 School <strong>of</strong> Integrative Systems, St Lucia Campus, University <strong>of</strong> Queensland Email: p.dargusch@uq.edu.au


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 38<br />

Industry and Native Forests<br />

Small-scale<br />

harvesting for<br />

local markets<br />

1 st Period<br />

(up to 1960s)<br />

Industrial-scale<br />

native forest<br />

harvesting,<br />

mostly for<br />

export markets<br />

Industry and Plantations<br />

Plantation<br />

development<br />

mostly for<br />

research<br />

purposes<br />

Focus on<br />

avoiding<br />

overexploitation<br />

<strong>of</strong> native<br />

forests<br />

Substantial expansion in<br />

the s<strong>of</strong>twood plantation<br />

estate by state governments<br />

Industry and Government Forest Policy<br />

Focus on policies to<br />

support wood product<br />

supply security and<br />

national selfsufficiency<br />

2 nd Period<br />

(1960s-1990s)<br />

Principles <strong>of</strong> ecological sustainable development<br />

exercised through the National Forest Policy<br />

Statement and Regional Forest Agreements guide<br />

native forest management.<br />

3 rd Period<br />

(1990s)<br />

Substantial expansion <strong>of</strong><br />

the hardwood plantation<br />

estate by Managed<br />

Investment Scheme firm<br />

Focus on fostering<br />

hardwood plantation<br />

development using private<br />

sector funds<br />

4 th Period<br />

(2000s)<br />

Plantations for<br />

the primary<br />

purpose <strong>of</strong><br />

carbon<br />

sequestration<br />

Focus on<br />

climate change<br />

mitigation and<br />

adaptation<br />

5 th Period<br />

(emerging)<br />

Figure 1. History <strong>of</strong> the <strong>Australia</strong>n forest industries conceptualized as five periods <strong>of</strong> industry<br />

development. The three sub-headings pertain to the three categories <strong>of</strong> themes, namely<br />

(1) how <strong>Australia</strong>n forest industry has interacted with native forests, (2) how<br />

<strong>Australia</strong>n forest industry has interacted with plantations, and (3) the main focus <strong>of</strong><br />

<strong>Australia</strong>n government forest policy as it relates to industry development.<br />

CASE ANALYSIS OF CARBON MARKETS AND AUSTRALIAN FOREST INDUSTRIES<br />

There now appears to be an emerging fifth period in the evolution <strong>of</strong> the <strong>Australia</strong>n forest industries,<br />

characterised by how the industry engages in and is influenced by carbon markets. The proposed<br />

CPRS is a critical economic consideration for <strong>Australia</strong>. When implemented, it will constitute<br />

arguably the most far-reaching economic reform undertaken in the country for over 30 years. The<br />

proposed scheme has been intensely debated in media, political, industry and research circles in<br />

<strong>Australia</strong> for over two years, and as <strong>of</strong> July 1 st 2009, was at the stage <strong>of</strong> proposed legislation being<br />

deliberated by the lower and upper houses <strong>of</strong> the <strong>Australia</strong>n Parliament. Whether the legislation is<br />

passed into law is uncertain but its high pr<strong>of</strong>ile in mainstream debate has meant that many businesses<br />

and related stakeholders have given the scheme special consideration, and have begun strategizing and<br />

become increasingly engaged in actions in preparation for the scheme’s introduction.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 39<br />

Many <strong>of</strong> the 700 or so firms mandated to participate in the CPRS will hold substantial emissions<br />

liabilities 2 . These firms could manage the cost <strong>of</strong> their compliance obligations in various ways,<br />

including by buying permits via government-operated auctions, buying permits through the secondary<br />

market (including using various financial instruments such as derivates), increasing their ratio <strong>of</strong><br />

revenue to greenhouse gas emissions by implementing cleaner forms <strong>of</strong> production, restructuring their<br />

businesses so that less <strong>of</strong> their business falls under the coverage <strong>of</strong> the CPRS, or by acquiring <strong>of</strong>fsets.<br />

Given these options, it is likely that firms covered by the CPRS will be increasingly interested in<br />

buying or investing in greenhouse gas emissions <strong>of</strong>fsets because <strong>of</strong>fsets can <strong>of</strong>fer reduced investment<br />

risk and greater cost certainty<br />

Importantly, afforestation is recognized under the CPRS as the only legitimate option that firms can<br />

use to <strong>of</strong>fset their greenhouse gas emissions outside <strong>of</strong> their facility or corporate operational<br />

boundaries within <strong>Australia</strong> 3 . Permits created via afforestation projects registered under the CPRS will<br />

be bankable, tradable and extinguishable in the same way as government-issued permits.<br />

Consequently, there is likely to be substantial and growing interest over the next five years (coinciding<br />

with the proposed first commitment period <strong>of</strong> the CPRS) in afforestation projects developed within<br />

<strong>Australia</strong> that are registered under the CPRS. For example, a recent analysis by the <strong>Australia</strong>n Bureau<br />

<strong>of</strong> Agricultural and Resource Economics estimated that, given a carbon price under an enacted CPRS<br />

<strong>of</strong> AU$28/tCO2e, an additional 21.8 million ha <strong>of</strong> agricultural land in <strong>Australia</strong> would be become<br />

economically viable for tree planting (in the context <strong>of</strong> existing competing land uses) (Lawson et al.<br />

2008). If these potential areas were planted, they would equate to almost a 10 fold increase in the<br />

<strong>Australia</strong>n forest plantation estate.<br />

To be eligible for registration under the CPRS, <strong>Australia</strong>n-based afforestation projects must meet<br />

stringent eligibility criteria including Kyoto Protocol compliance relating to forest type and land-use<br />

history. However, not all existing afforestation projects within <strong>Australia</strong> that comply with these Kyoto<br />

Protocol provisions will be eligible for registration under the CPRS. Rather, only those currently<br />

registered under the NSW Greenhouse Gas Abatement Scheme (NSW GGAS) or the <strong>Australia</strong>n<br />

Commonwealth Government’s Greenhouse Friendly Program, will be eligible to be registered under<br />

the CPRS. To the best <strong>of</strong> the authors’ knowledge, this means that none <strong>of</strong> the existing plantings <strong>of</strong> the<br />

larger <strong>Australia</strong>n MIS firms will be eligible to be registered under the CPRS. In other words, the<br />

industrial-scale traditional forest growing constituents <strong>of</strong> the <strong>Australia</strong>n forest industries are unable to<br />

use their existing forest plantation estates to supply the demand for <strong>of</strong>fsets that is likely to occur if the<br />

CPRS is introduced. The CPRS-driven demand for <strong>of</strong>fsets would therefore need to be met by (1)<br />

existing constituents <strong>of</strong> the <strong>Australia</strong>n forest industries changing the focus <strong>of</strong> their business to include<br />

the production <strong>of</strong> <strong>of</strong>fsets, and (2) new entrants to the industry which choose to become directly<br />

involved in the business <strong>of</strong> afforestation-based <strong>of</strong>fsets to meet their CPRS compliance obligations or<br />

seek pr<strong>of</strong>it from the business <strong>of</strong> <strong>of</strong>fset production and trade.<br />

A number <strong>of</strong> <strong>Australia</strong>n companies have recently been established that are primarily involved in the<br />

business <strong>of</strong> afforestation-based emissions <strong>of</strong>fsets. Table 1 provides a summary <strong>of</strong> the characteristics <strong>of</strong><br />

these firms. Part A presents a summary <strong>of</strong> the 24 <strong>Australia</strong>n companies 4 that the authors identified<br />

were primarily in the business <strong>of</strong> afforestation-based <strong>of</strong>fsets as <strong>of</strong> February 25 th 2009 while Part B<br />

presents a summary <strong>of</strong> 11 <strong>of</strong> these 24 companies that were primarily in the business <strong>of</strong> managing<br />

2 Whilst most types <strong>of</strong> industries are included in the coverage <strong>of</strong> the CPRS (e.g. electricity generators, transport,<br />

manufacturing, mining, waste) firms that operate businesses in the ‘forestry industry’ can choose to engage or<br />

not to engage in the scheme and firms involved in agriculture are excluded.<br />

3 Firms covered by the <strong>Australia</strong>n CPRS will also be allowed to acquire <strong>of</strong>fsets from Clean Development<br />

Mechanism and Joint Implementation projects based in other countries.<br />

4 The trading names <strong>of</strong> these 24 companies are Carbon Balance Consulting, Carbon Planet, Carbon Reduction<br />

<strong>Institute</strong>, Carbonza, Green Pass, Ark Climate, <strong>Australia</strong>n Carbon Traders, Brokers Carbon, Canopy, Carbon<br />

Pool, Cool Planet, Emit Environmental Brokers, Green Pig, Auscarbon International, Carbon Conscious, Carbon<br />

Neutral, CarbonSmart (Landcare), CO2<strong>Australia</strong>, Elementree, Greenfleet, Greenhouse Balanced, Greening<br />

<strong>Australia</strong>, Offset Emissions and TreeSmart. The first 11 companies listed are those described in Part B <strong>of</strong> the<br />

Table as companies that were primarily in the business <strong>of</strong> managing afforestation-based <strong>of</strong>fset generation<br />

projects within <strong>Australia</strong>.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 40<br />

afforestation-based <strong>of</strong>fset generation projects within <strong>Australia</strong>. Note also that there are a number <strong>of</strong><br />

<strong>Australia</strong>n companies in addition to the 24 reviewed in Table 1 that are either primarily involved in the<br />

business <strong>of</strong> non-afforestation-based emissions <strong>of</strong>fsets or are also involved in the business <strong>of</strong><br />

afforestation-based <strong>of</strong>fsets but for whom the business <strong>of</strong> afforestation-based <strong>of</strong>fsets is secondary to<br />

other types <strong>of</strong> business activities. The nature and motivations <strong>of</strong> these additional firms is the subject <strong>of</strong><br />

another study currently being undertaken by the authors.<br />

Most <strong>of</strong> the 24 companies listed in Table 1 are young (22 <strong>of</strong> the 24 are less than 10 years old, with 15<br />

less than five years old), small (17 <strong>of</strong> the 24 companies had an annual turnover <strong>of</strong> less than AU$1<br />

million), privately-owned (19) and ‘for-pr<strong>of</strong>it’ (19). There is an approximately equal mix <strong>of</strong> firms that<br />

operate as traders, managers <strong>of</strong> afforestation-based emissions <strong>of</strong>fset generation projects, and advisors.<br />

All 13 firms that operate as either traders or advisors had an annual turnover <strong>of</strong> less than $AU1 million<br />

in 2007/2008 and were for-pr<strong>of</strong>it privately-owned companies with less than five employees.<br />

Interestingly, four <strong>of</strong> the five not-for-pr<strong>of</strong>it companies listed in Part A <strong>of</strong> Table 1 were in the business<br />

<strong>of</strong> managing afforestation-based <strong>of</strong>fset projects and these are some <strong>of</strong> the largest companies by<br />

turnover <strong>of</strong> the 24 companies reviewed (four <strong>of</strong> the not-for-pr<strong>of</strong>it companies turned over more than<br />

AU$1 million per annum). These data suggest that not-for-pr<strong>of</strong>it companies, although relatively small<br />

in number, currently play a disproportionately influential (and arguably market-leading) role in<br />

<strong>Australia</strong>n carbon markets, particularly amongst firms in the business <strong>of</strong> managing afforestation-based<br />

emissions <strong>of</strong>fsets projects.<br />

Table 1 Summary <strong>of</strong> characteristics <strong>of</strong> <strong>Australia</strong>n firms primarily engaged in the business <strong>of</strong><br />

afforestation-based emissions <strong>of</strong>fsets.<br />

Part A Summary characteristics <strong>of</strong> 24 <strong>Australia</strong>n companies primarily engaged in the business <strong>of</strong><br />

afforestation-based <strong>of</strong>fsets<br />

Age <strong>of</strong><br />

0 to 5 years 5 to 10 years 10 to 20 years >20 years<br />

company 15 (63%) 7 (29%) 1 (4%) 1 (4%)<br />

Commercial<br />

For-pr<strong>of</strong>it Not-for-pr<strong>of</strong>it<br />

purpose 19 (79%) 5 (21%)<br />

Commercial Trader or broker Project manager Advisor<br />

role 8 (33%) 11 (46%) 5 (21%)<br />

Ownership<br />

Private<br />

19 (79%)<br />

ASX listed<br />

1 (4%)<br />

Member based<br />

4 (17%)<br />

Clientele Governments Large companies Sees Individuals<br />

Primary 2 (8%) 13 (54%) 5 (21%) 4 (17%)<br />

Secondary 7 (29%) 0 4 (17%) 13 (54%)<br />

Are the traded<br />

Certified<br />

Not certified<br />

<strong>of</strong>fsets<br />

(NSWGGAS or ‘Climate Friendly) (NSWGGAS or ‘Climate Friendly)<br />

certified? 12 (50%) 12 (50%)<br />

Estimated<br />

turnover<br />

AU$5 million<br />

3 (12%)<br />

Part B Summary characteristics <strong>of</strong> 11 <strong>Australia</strong>n companies primarily engaged in the business <strong>of</strong><br />

managing afforestation-based emissions <strong>of</strong>fsets generation projects<br />

Age <strong>of</strong><br />

0 to 5 years 5 to 10 years 10 to 20 years >20 years<br />

company 4 (36%) 5 (46%) 1 (9%) 1 (9%)<br />

Commercial<br />

For-pr<strong>of</strong>it Not-for-pr<strong>of</strong>it<br />

purpose 7 (63%) 4 (37%)<br />

Clientele Governments Large companies Sees Individuals<br />

Primary 1 (9%) 7 (63%) 0 3 (28%)<br />

Secondary 4 (36%) 0 2 (18%) 5 (46%)<br />

Land<br />

Share farming Lease Purchase<br />

acquisition 2 (18%) 7 (64%) 2 (18%)<br />

Target land<br />

750 mm pa<br />

rainfall 11 (100%) 0<br />

Distance to port<br />

200 km<br />

11 (100%)


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 41<br />

Tree species<br />

planted<br />

Species traditionally used for wood<br />

production<br />

0<br />

Species not traditionally used for wood<br />

production<br />

11 (100%)<br />

Are the traded<br />

Certified<br />

Not certified<br />

<strong>of</strong>fsets<br />

(NSWGGAS or ‘Climate Friendly) (NSWGGAS or ‘climate friendly)<br />

certified? 7 (64%) 4 (36%)<br />

Estimated<br />

turnover<br />

AU$5 million<br />

3 (28%)<br />

Source: Websites and other company reports as at February 25 th 2009.<br />

About half <strong>of</strong> the primary clients <strong>of</strong> the 24 companies in Table 1 are large companies (13 <strong>of</strong> 24<br />

companies had large companies as their main clients) and close to the same proportion <strong>of</strong> the existing<br />

secondary clients were individuals (13 <strong>of</strong> the 24 companies had individuals as their second most<br />

common clients). Furthermore, only half <strong>of</strong> the 24 companies stipulated that the afforestation-based<br />

<strong>of</strong>fsets that they deal in must be certified under either the NSWGGAS or the Commonwealth<br />

Government’s Greenhouse Friendly program. These figures suggest that whilst many <strong>of</strong> the companies<br />

target large companies that will have a CPRS liability as clientele, there are a similar number that<br />

target either companies that do not have a CPRS liability or individuals that are perhaps more driven<br />

by the subjective personal values <strong>of</strong> <strong>of</strong>fsets as opposed to their CPRS compliance value.<br />

Part B <strong>of</strong> Table 1 reveals a number <strong>of</strong> interesting features <strong>of</strong> how the 11 companies that manage<br />

afforestation-based <strong>of</strong>fset projects go about the business <strong>of</strong> afforestation. All <strong>of</strong> the firms developed<br />

afforestation projects more than 200 km away from port facilities in semi-arid regions with annual<br />

rainfall typically less than 750 mm and all afforestation projects involved species not traditionally used<br />

for wood production but rather most used species endemic to the planting area. These differ<br />

substantially from traditional <strong>Australia</strong>n plantations.<br />

Future growth <strong>of</strong> this new ‘carbon market’ sector <strong>of</strong> the <strong>Australia</strong>n forest industries is likely to be<br />

fueled by demand for <strong>of</strong>fsets from firms wanting to meet their CPRS obligations. Demand would also<br />

arise from firms which, whilst not having a compliance obligation under the CPRS or NGER, are<br />

motivated to minimize their ‘carbon footprint’ for corporate legitimacy, ethical and marketing reasons.<br />

This market growth has implications for conventional <strong>Australia</strong>n forest enterprises, many <strong>of</strong> which<br />

may need or decide to adapt their existing capabilities to take advantage <strong>of</strong> new opportunities<br />

associated with carbon markets. Moreover, it is likely that government policies and protocols for the<br />

measurement, verification and commercialization <strong>of</strong> <strong>of</strong>fsets from afforestation projects will be refined<br />

and made more ‘industry friendly’. As well, issues such as how to best include and manage soil carbon<br />

in carbon markets and how to best integrate trees and agricultural crops in agr<strong>of</strong>orestry and<br />

silvopastural systems, are being included in the policy debate.<br />

DISCUSSION AND CONCLUDING REMARKS<br />

The entrance <strong>of</strong> the 24 new firms primarily in the business <strong>of</strong> afforestation-based <strong>of</strong>fsets into the<br />

<strong>Australia</strong>n forest industries is arguably the most notable indication <strong>of</strong> how carbon markets are<br />

influencing change in the industry. These new firms are typically entrepreneurial; their business<br />

activities are closely aligned with principles <strong>of</strong> ecologically sustainable development, and most are<br />

young and small-scale. Such features are characteristic <strong>of</strong> so-called eco-preneurial firms (a concept<br />

coined by Blue, 1990). Some research (e.g. Isaak, 2002; Schaper, 2002; Dixon and Clifford, 2007) has<br />

indicated that eco-preneurial firms can invigorate industry-wide change towards more sustainable<br />

business practices by introducing ‘innovation, adaptation and new ideas’ (Schaper, 2002, p. 27). It<br />

seems likely that the 24 new eco-preneurial entrants into the <strong>Australia</strong>n forest industries are positioned<br />

to invigorate similar change in their industry. However the proposed CPRS and related climate change<br />

mitigation policies in <strong>Australia</strong> contain relatively few provisions that are specifically targeted towards<br />

supporting the development <strong>of</strong> these eco-preneurial new entrants. The authors believe that the<br />

effectiveness <strong>of</strong> regulated carbon market policy in achieving climate change mitigation objectives<br />

could be greatly enhanced by including more effective provisions that more specifically support the<br />

business development <strong>of</strong> eco-preneurial firms (particularly in relation to the capacities <strong>of</strong> those firms<br />

to secure finance).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 42<br />

Some <strong>of</strong> the 24 new entrants (including Greening <strong>Australia</strong> and Landcare, which are amongst the<br />

largest <strong>of</strong> the 24 firms by turnover) have been in the business <strong>of</strong> planting trees for more than 10 years<br />

but have only recently started to engage in carbon markets as a means <strong>of</strong> revenue generation. For most<br />

<strong>of</strong> their existence, Greening <strong>Australia</strong> and Landcare have utilized funding from corporate-based<br />

philanthropic sources or from the <strong>Australia</strong>n Commonwealth government, to undertake extensive tree<br />

planting for ecosystem rehabilitation. With much <strong>of</strong> this public funding no longer available, and with<br />

increased interest in carbon markets, many <strong>of</strong> these organisations are now making the transition to<br />

business models with greater reliance on non-government funding. How successful these firms are in<br />

making this transition will be an important factor in how the <strong>Australia</strong>n forest industry changes in<br />

response to carbon markets over the next 10 years.<br />

Three <strong>of</strong> the four largest new entrants into the <strong>Australia</strong>n forest industries that are in the business <strong>of</strong><br />

afforestation-based <strong>of</strong>fsets are not-for-pr<strong>of</strong>it companies. Whether these companies would be able to<br />

raise enough capital or attract sufficient appropriately skilled staff to enable a rapid expansion <strong>of</strong> their<br />

business operations in the event demand for <strong>of</strong>fsets increased dramatically is questionable. The authors<br />

are concurrently undertaking a more detailed survey <strong>of</strong> these not-for-pr<strong>of</strong>it firms to understand better<br />

their capacity constraints and expansion intentions. But regardless <strong>of</strong> these capacity constraints, the<br />

question needs to be asked whether not-for-pr<strong>of</strong>it companies are actually better suited to the<br />

production <strong>of</strong> <strong>of</strong>fsets in carbon markets than pr<strong>of</strong>it-driven alternatives?<br />

Carbon markets are embedded in broader notions <strong>of</strong> sustainability and firms which have a compliance<br />

obligation under a regulated carbon market such as the CPRS are generally expected by stakeholders<br />

to not only seek the lowest cost means <strong>of</strong> compliance, but also to comply in a manner that contributes<br />

broader community benefits and enhances the sustainability <strong>of</strong> their business practices.<br />

It follows that perhaps not-for-pr<strong>of</strong>it companies are better placed than their pr<strong>of</strong>it-driven<br />

contemporaries to provide <strong>of</strong>fset products that, whilst perhaps produced on a smaller and more costly<br />

scale, provide preferred sustainability benefits, such as the ability to engage the community in tree<br />

planting projects, smaller-scale plantings that more effectively enhance complementary ecological<br />

services, and better socio-economic outcomes at a regional scale. In this sense, government-owned<br />

organizations may also be well suited to afforestation-based <strong>of</strong>fset production. Furthermore, it may be<br />

preferable for some firms which have a compliance obligation under the CPRS to engage in the<br />

business <strong>of</strong> producing <strong>of</strong>fsets themselves (i.e. growing the trees), particularly in cases where those<br />

firms have some knowledge <strong>of</strong> land use and forestry, or have access to land suitable for afforestation<br />

projects. The authors are currently involved in research investigating the nature and motivations <strong>of</strong> a<br />

number <strong>of</strong> firms which are engaged in these types <strong>of</strong> activities.<br />

While this paper has focussed on the CPRS, voluntary carbon markets can also play an important role<br />

in industry change. Indeed, multiple factors are likely to influence change at an industry level,<br />

including the emergence <strong>of</strong> both regulated and voluntary carbon markets, dynamics <strong>of</strong> ecopreneurship,<br />

dynamics <strong>of</strong> industry economics and competition, and processes <strong>of</strong> organisational change<br />

in larger, traditional industry constituents. But the case <strong>of</strong> the <strong>Australia</strong>n forest industries certainly<br />

does indicate that carbon markets are a prominent factor in industry change.<br />

The case study has highlighted how regulated carbon market policy provisions relating to <strong>of</strong>fsets from<br />

forests can invigorate eco-preneurial business activity, and there is reason to believe that ecopreneurial<br />

activity can act as a catalyst for wider industry change that supports better climate change<br />

mitigation outcomes.<br />

This finding has relevance to the design <strong>of</strong> other regulated carbon markets, particularly those that<br />

might include provisions for <strong>of</strong>fsets from forests, as is the case with proposed regulated carbon market<br />

policy included in the US Clean Energy and Security Act 2009. The effectiveness <strong>of</strong> carbon market<br />

policy should be enhanced if it included provisions to address the specific business development needs<br />

<strong>of</strong> eco-preneurial firms. Furthermore, it seems likely that not-for-pr<strong>of</strong>it and government organisations<br />

may be better suited to the production <strong>of</strong> forest-related <strong>of</strong>fsets, given their ability to better support<br />

complementary social and ecological services.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 43<br />

The overall benefits <strong>of</strong> regulated carbon market policy could be increased if policy included provisions<br />

supporting the development <strong>of</strong> these types <strong>of</strong> organisations as well. Whilst the authors believe that the<br />

most pending types <strong>of</strong> assistance that eco-preneurial firms need are measures that support the firm’s<br />

ability to secure finance, the authors also acknowledge that a complex mix <strong>of</strong> factors influence ecopreneurship.<br />

The effectiveness <strong>of</strong> carbon market policy would be increased if policy decision-making<br />

was supported by a better understanding <strong>of</strong> the dynamics <strong>of</strong> eco-preneurship. Therefore, future<br />

research on carbon markets should investigate what Beveridge and Guy (2005, p. 672) described as<br />

the ‘processes and practices <strong>of</strong> emergence, negotiation and innovation’ <strong>of</strong> eco-preneurship.<br />

ACKNOWLEDGEMENTS<br />

We would like to thank Neil Byron and Ian Ferguson for their helpful comments during the<br />

development <strong>of</strong> this paper.<br />

REFERENCES<br />

ABARE 2008. <strong>Australia</strong>n Forest and Wood Product Statistics, March and June quarters 2008. <strong>Australia</strong>n<br />

Bureau <strong>of</strong> Agriculture and Resource Economics, Canberra.<br />

Ajani J. 2008. <strong>Australia</strong>'s transition from native forests to plantations: the implications for woodchips, pulpmills,<br />

tax breaks and climate change. Agenda: A Journal <strong>of</strong> Policy Analysis and Reform, 15(3): 3-10.<br />

Beveridge, R. and Guy, S. 2005. The rise <strong>of</strong> the eco-preneur and the messy world <strong>of</strong> environmental innovation.<br />

Local Environment, 10(6): 665-676.<br />

Blue, R. 1990. Ecopreneuring: Managing for Results. Scott Foresman, London.<br />

Dargavel, J. 1995. Fashioning <strong>Australia</strong>’s Forests. Oxford University Press, Oxford<br />

Dargusch, P 2008 ‘Understandings <strong>of</strong> sustainable corporate governance by <strong>Australia</strong>n managed investment<br />

schemes and some implications for small-scale forestry in <strong>Australia</strong>’, Small-scale Forestry, 7(1): 67-75.<br />

Dixon, S. and Clifford A. 2007. Ecopreneurship - a new approach to managing the triple bottom line. Journal <strong>of</strong><br />

Organizational Change Management 20(3): 326-345.<br />

Herbohn, J. and Harrison, S. 2004. The evolving nature <strong>of</strong> small-scale forestry in <strong>Australia</strong>. Journal <strong>of</strong> Forestry<br />

102(1): 42–47<br />

Isaak R. 2002. The making <strong>of</strong> the ecopreneur. Greener Management International 38(2): 81-91.<br />

Lawson, K, Burns, K, Low, K, Heyhoe, E and Ahammad, H 2008, Analysing the economic potential <strong>of</strong> forestry<br />

for carbon sequestration under alternative carbon price paths, ABARE, Canberra.<br />

NAFI 2008. Submission on CPRS Green Paper. National Association <strong>of</strong> Forest Industries, Canberra.<br />

Routley, R. and Routley, V. 1975. The Fight for the Forests – the Takeover <strong>of</strong> <strong>Australia</strong>n Forests for Pines,<br />

Woodchips and Intensive Forestry. RSBS, ANU, Canberra.<br />

Schaper, M. 2002. The essence <strong>of</strong> ecopreneurship. Greener Management International 38(2): 26-30.<br />

Slee, B. 2001. Resolving production-environment conflicts: the case <strong>of</strong> the Regional Forest Agreement Process<br />

in <strong>Australia</strong>. Forest Policy and Economics 3(1-2_:17–30.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 44<br />

ABSTRACT<br />

LESSONS LEARNED FROM BUILDING ESTATE-SCALE<br />

CARBON ACCOUNTING SYSTEMS<br />

Scott Arnold 1<br />

In 2008 the Victorian Department <strong>of</strong> Sustainability and Environment embarked on a<br />

project to construct a set <strong>of</strong> carbon accounts for approximately 7.5 million hectares <strong>of</strong><br />

public land in Victoria. The land comprised mostly National Park and State Forest<br />

tenures but also included a range <strong>of</strong> other crown land tenures. The area was stratified on<br />

the basis <strong>of</strong> historical event data and geographic region. The large volumes <strong>of</strong> historical<br />

data were simplified and used to determine the stock trajectories for each stratum. The<br />

initial data sets resided in a GIS and a database system was developed to pass data to and<br />

from FullCAM, where carbon stocks were calculated. The carbon stocks were most<br />

heavily influenced by wildfire events, principally as a result <strong>of</strong> the very large areas that<br />

are sometimes burnt. Harvesting events had a greater impact on a per-hectare basis but<br />

the relatively small areas involved limited their influence.<br />

INTRODUCTION<br />

The Department <strong>of</strong> Sustainability and Environment (DSE) manages approximately 7.5 million<br />

hectares <strong>of</strong> public land across Victoria under a range <strong>of</strong> tenures. Parks and Reserves (including<br />

National Parks, State Parks and Regional Parks) and State forests constitute the vast majority <strong>of</strong> that<br />

land, being 3.6 million hectares and 3.3 million hectares respectively. Public land represents one third<br />

<strong>of</strong> the total area <strong>of</strong> Victoria.<br />

Interest in the carbon stock held on that public land has increased in recent years and that interest has<br />

accelerated with the Commonwealth Government’s proposed Carbon Pollution Reduction Scheme<br />

(CPRS).<br />

Land Carbon Project<br />

In December 2008 DSE commenced a project, known as the Land Carbon Project, aimed at<br />

quantifying the amount <strong>of</strong> carbon stored on publicly managed land. The stated aims <strong>of</strong> the Land<br />

Carbon Project were to:<br />

• Provide an indicative set <strong>of</strong> carbon accounts for Victoria’s publicly managed land,<br />

• Increase the Department’s understanding <strong>of</strong> fluxes in carbon stocks, and<br />

• Inform the policy development process in areas relating to managing carbon as an asset.<br />

A Project Team <strong>of</strong> three staff was assembled in order to deliver a product by the end <strong>of</strong> April 2009.<br />

DELIVERABLES<br />

The Land Carbon Project Team were faced with a situation where there was very limited<br />

understanding <strong>of</strong> exactly what was expected in terms <strong>of</strong> deliverable products. Aside from the clear<br />

requirement to produce a ‘map <strong>of</strong> the carbon stocks’ the balance <strong>of</strong> the deliverable products were not<br />

enunciated in any detail.<br />

Given the ‘open’ nature <strong>of</strong> the project, the Land Carbon Project Team sought to clarify project<br />

expectations through a series <strong>of</strong> formal and informal meetings. Two key themes emerged from the<br />

meetings:<br />

1. The project must use data that, where possible, was <strong>of</strong> Victorian origin, and<br />

1 Department <strong>of</strong> Sustainability and Environment. Level 3, 8 Nicholson Street, PO Box 500, East Melbourne Vic 3002,<br />

Ph: (03) 9637 8520 Email: scott.arnold@dse.vic.gov.au


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 45<br />

2. The impact <strong>of</strong> management events such as harvesting and burning must be identifiable.<br />

These themes played a significant role in shaping the project. The Land Carbon Project Team were<br />

now charged with using DSE data to construct a set <strong>of</strong> carbon accounts that described both total stocks<br />

and stock changes as a result <strong>of</strong> harvesting and burning on publicly managed land in Victoria.<br />

LAND CARBON STRUCTURE<br />

With a more refined project scope in place, the structure <strong>of</strong> the Land Carbon Project could be defined.<br />

Figure 1 illustrates the main components <strong>of</strong> the Land Carbon Project.<br />

Figure 1: Land Carbon Project structure<br />

GIS<br />

• Tree Cover<br />

• Harvesting<br />

• Fire<br />

• IBRA<br />

Map<br />

Products<br />

Database<br />

• Link to FullCAM<br />

• Format results<br />

Project<br />

Report<br />

FullCAM<br />

• Templates<br />

• Plot Files<br />

• Carbon models<br />

Analyses<br />

• Total stock<br />

• Impact <strong>of</strong> events<br />

Geographic Information System<br />

Carbon accounts are data-oriented products that, depending on their level <strong>of</strong> sophistication, can require<br />

large amounts <strong>of</strong> input data. Almost all <strong>of</strong> the data used in the Land Carbon Project was spatial data<br />

stored in the DSE geographic information system (GIS) library.<br />

Each <strong>of</strong> the GIS data inputs could be categorised as being either biophysical or administrative. The<br />

biophysical data described the events or characteristics that influence the carbon stocks on site and<br />

included harvesting extent, fire history, vegetation type and geographic region. The administrative<br />

data served two purposes; one was to define the extent <strong>of</strong> publicly managed land in Victoria and the<br />

other was to provide a basis for determining how the total carbon stock was distributed among the<br />

various State government land management agencies (eg: Catchment Management Authorities<br />

(CMAs), Local Government Areas).<br />

Database System<br />

After the input data had been prepared in the GIS, additional formatting work was required before the<br />

data could be passed on to the carbon model. A Micros<strong>of</strong>t Access 2003 database was built to create<br />

files in a format that allowed data to be interpreted by the carbon model.<br />

This same database was also used to receive all the outputs <strong>of</strong> the carbon model and to present that<br />

data in a format that allowed the results to be joined back to the GIS for map production and further<br />

analysis.<br />

FullCAM<br />

FullCAM is a stock balance carbon model for forest and agricultural land uses developed by the<br />

Commonwealth Department <strong>of</strong> Climate Change (DCC), formerly known as the <strong>Australia</strong>n Greenhouse<br />

Office (AGO). FullCAM was chosen as the carbon model for the Land Carbon Project. This decision<br />

was made on the basis <strong>of</strong> allowing the project to leverage the credibility that FullCAM has while also<br />

maximising the ‘compatibility’ <strong>of</strong> project estimates with estimates contained in the national accounts.<br />

The results <strong>of</strong> FullCAM simulations were imported into the database for further analyses and linkage<br />

back to the GIS.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 46<br />

Map Products<br />

In order to generate map products, a series <strong>of</strong> tables from the database were linked to the GIS and from<br />

there maps showing the distribution <strong>of</strong> any variable, at any point in time, could be made. Figure 2<br />

shows an early draft map <strong>of</strong> the distribution <strong>of</strong> total onsite carbon as at 2005, with the impact <strong>of</strong> the<br />

2003 alpine fire event clearly evident.<br />

Figure 2: Sample map product.<br />

Analyses<br />

In order to minimise the number <strong>of</strong> calculations in FullCAM, the decision was made not to include any<br />

administrative data in the actual carbon model and only the biophysical data was used to describe the<br />

publicly managed land. Once the results <strong>of</strong> FullCAM were linked to the GIS it was then possible to<br />

intersect the carbon data with any administrative data layer to produce both maps and tabular data<br />

breaking down the FullCAM results.<br />

As an example, the layer <strong>of</strong> total onsite carbon stocks could be unioned with a layer <strong>of</strong> all the CMAs<br />

to calculate stocks for each CMA.<br />

Project Report<br />

The Project Team maintained a very strong commitment to ‘document as you go’, generating a series<br />

<strong>of</strong> procedural documents that served to record all the decisions made and actions taken for each<br />

discrete task. The task <strong>of</strong> preparing the Project Report was greatly simplified by the availability <strong>of</strong> the<br />

procedural documents.<br />

POPULATING THE MODEL<br />

The first stage in producing the carbon accounts was to assemble all the input data required to<br />

populate a model <strong>of</strong> the target estate. This task involved using the GIS to prepare layers that described<br />

harvesting history, fire history, forest cover, Interim Biogeographical Regions <strong>of</strong> <strong>Australia</strong> (IBRA) and<br />

the extent <strong>of</strong> Victoria’s publicly managed land.<br />

While almost all <strong>of</strong> the GIS processes were relatively unaffected by the number <strong>of</strong> records being<br />

processed, the Project Team was acutely aware that once data was passed to FullCAM then the<br />

number <strong>of</strong> records would have a huge impact as there is no capacity within FullCAM to ‘batch’


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 47<br />

operations. If a particular process needed to be applied to each record (or Plot as they are called in<br />

FullCAM) then that process would have to occur manually – one at a time. For this reason there was a<br />

very strong emphasis on simplifying the data when working with the GIS layers.<br />

Harvesting History<br />

The DSE GIS library contains several layers that contain data on the extent and timing <strong>of</strong> harvest<br />

events dating back to the early 1930’s. Four key decisions were made in order to reduce the sheer<br />

volume <strong>of</strong> harvesting data.<br />

Firstly, it was decided that only the most recent harvesting event would be included in the model.<br />

Secondly, it was decided to aggregate harvesting events into decades. Thirdly, it was decided that<br />

only one type <strong>of</strong> harvesting event would be modelled. The alternative was to model thinning, group<br />

selection and clearfall events and to do so would triple the size <strong>of</strong> the overall model.<br />

Finally, it was decided that all records less than 10ha would be discarded. Again, this decision was<br />

taken to reduce the number <strong>of</strong> records in the harvesting layer and the figure <strong>of</strong> 10ha was chosen after<br />

examining the distribution <strong>of</strong> areas (see Figure 3). Removing small areas reduced the number <strong>of</strong><br />

records to less than 9,800 records while removing only 2% <strong>of</strong> the area harvested.<br />

Figure 3: Distribution <strong>of</strong> harvesting records by area.<br />

No attempt was made to include planned future harvesting events in the model.<br />

Fire History<br />

One <strong>of</strong> the management events that was clearly identified as essential when determining the Land<br />

Carbon Project deliverables was the impact <strong>of</strong> planned burning on carbon stocks. The DSE GIS<br />

library contained several layers that differentiated between planned burns and wildfires and contained<br />

data going back to the 1930’s.<br />

As with harvesting data, a process was undertaken to simplify the fire history data<br />

Forest Cover<br />

Victoria’s publicly managed land comprises both forested and non-forested areas. The presence or<br />

absence <strong>of</strong> forest was assumed to be the single most significant influence on carbon stocks.<br />

The DSE GIS library contained a layer that represented the distribution <strong>of</strong> tree cover as at 1995. Areas<br />

described as non-forested in 1995 were assumed to have been non-forest throughout the model.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 48<br />

Interim Biogeographical Regions <strong>of</strong> <strong>Australia</strong><br />

The IBRA layer (Natural Resource Management Ministerial Council, 2004) was used to assist in<br />

differentiating between areas with similar event histories but occurred in locations with distinctly<br />

different climatic, soil and vegetation characteristics.<br />

The IBRA layer was sourced from a Commonwealth Department <strong>of</strong> the Environment, Water, Heritage<br />

& the Arts website (http://www.environment.gov.au/metadataexplorer/explorer.jsp).<br />

Publicly Managed Land Extent<br />

The final GIS layer used to populate the model was one that would define the extent <strong>of</strong> the area to be<br />

included in the model. The relevant layer <strong>of</strong> the DSE GIS, containing data on all publicly managed<br />

land in Victoria was used. This layer also required simplification prior to being included in the spatial<br />

data model.<br />

Combining Layers<br />

Once the five GIS layers were in place they were unioned to create the Input Layer. Critically, the<br />

Input Layer was created as a multipart layer, meaning that all polygons with the same attribute set<br />

were stored in the attribute table as a single record. By using a multipart layer and the processes to<br />

simplify the harvesting, fire and public land data, the total number <strong>of</strong> records was kept to 822.<br />

Each record in the Input Layer then had its area, in hectares, calculated as well as the latitude and<br />

longitude, in decimal degrees, <strong>of</strong> the polygon’s centroid. These three values along with the record<br />

identifier would be passed to FullCAM.<br />

USING FULLCAM<br />

FullCAM is a highly complicated collection <strong>of</strong> carbon models and this paper considers FullCAM only<br />

from the perspective <strong>of</strong> a s<strong>of</strong>tware user. FullCAM uses a proprietary file format as the basis for using<br />

the model.<br />

There is no capacity to ‘batch’ processes in order to treat a number <strong>of</strong> Plots in the same fashion. This<br />

is unfortunate for users wishing to generate Plot level data, as was the case with the Land Carbon<br />

Project where we required data for each Plot to be joined to the GIS in order to compile map products<br />

and perform spatial analyses (such as dividing carbon stocks by CMAs).<br />

Of equal significance is the inability to directly access the back-end database in instances where<br />

changes need to be made to large numbers <strong>of</strong> plots. Even a change that took only 30 seconds for each<br />

Plot represented almost a full day’s work when applied to all 822 Plots.<br />

Using Templates<br />

The Research edition <strong>of</strong> FullCAM supports the use <strong>of</strong> templates to set up Plots that hold settings and<br />

values that users wish to apply to a number <strong>of</strong> Plots. For example, template Plots could hold settings<br />

for a series <strong>of</strong> events and then all new Plots created from that template would all have the same event<br />

series. The Land Carbon Project used this capacity to ensure that all Plots with the same treatment<br />

histories were the same. Figure 4 illustrates the concept <strong>of</strong> templates.<br />

Configuring Templates<br />

The configuration <strong>of</strong> the template Plots was the most challenging aspect <strong>of</strong> the entire Land Carbon<br />

Project. All estimates <strong>of</strong> carbon stocks and their fluxes were driven by the way both the vegetation<br />

and the events applied to the sites were modelled.<br />

Plots in FullCAM, quite literally, have hundreds <strong>of</strong> variables that the user can adjust. Wherever<br />

practical, the Land Carbon Project Team accepted the default values <strong>of</strong> FullCAM and made changes<br />

only where deemed necessary.<br />

Each harvesting event is followed by a post-harvest burn event. The post-harvest burn event is<br />

assumed to impact on 80% <strong>of</strong> the harvested area. The settings used to describe combustion rates and<br />

transfers for the trees and debris are shown in Figures 5a and 5b respectively.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 49<br />

Figure 4: Using templates in FullCAM.<br />

Text File<br />

Template<br />

Plot Files<br />

Plot Area Lat Long Harvesting Fire<br />

Plot1<br />

Plot2<br />

Plot<br />

Plot2<br />

146.<br />

914.<br />

Area<br />

146.<br />

914.<br />

-36.22<br />

-35.92<br />

Figure 5a and 5b: Post-harvest burn impact on trees and debris.<br />

Lat<br />

-35.92<br />

Wildfire events were assumed to impact 80% <strong>of</strong> the nominated burnt area. The settings used to<br />

describe combustion rates and transfers for the trees and debris are shown in Figures 6a and 6b<br />

respectively. These settings are designed to represent a ‘typical’ wildfire in terms <strong>of</strong> severity.<br />

Figure 6a and 6b: Wildfire impact on trees & debris.<br />

70<br />

Plot<br />

Plot1<br />

Area<br />

Post-harvest Burn Impact on Trees<br />

85<br />

10<br />

Wildfire Impact on Trees<br />

12<br />

5<br />

18<br />

Lat<br />

-36.22<br />

Combusted<br />

To Debris<br />

Alive<br />

Combusted<br />

To Debris<br />

Alive<br />

+<br />

=<br />

155.729<br />

154.836<br />

Long<br />

Long<br />

155.729<br />

154.836<br />

39<br />

58.5<br />

Harvesting<br />

1965<br />

Harvesting<br />

1965<br />

1965<br />

Fire<br />

1985<br />

Fire<br />

1985<br />

1985<br />

Post-harvest Burn Impact on Debris<br />

1<br />

Wildfire Impact on Debris<br />

1.5<br />

60<br />

40<br />

Combusted<br />

To Soil<br />

Unburnt<br />

Combusted<br />

To Soil<br />

Unburnt


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 50<br />

Planned burn events were modelled to be more mild fires with lower combustion and transfer rates for<br />

both trees and debris. Planned burns are assumed to impact 50% <strong>of</strong> the nominated burnt area.<br />

During the development <strong>of</strong> the burning events (both wildfire and planned burns), a lack <strong>of</strong> stand<br />

response after the event was detected. This issue was raised with the DCC who advised that fire<br />

events in FullCAM did not contain any capacity for the stand to respond to the disturbance and that<br />

stand response was only modelled for harvesting events.<br />

To overcome this lack <strong>of</strong> stand response to burning, each burn event was immediately followed by a<br />

harvest event. The harvest event removed no volume but it’s presence ‘alerted’ the model to the<br />

changed stand circumstances that were present after the burn event and FullCAM subsequently<br />

modelled a stand response (even though it was, in theory, responding to the wrong event).<br />

Templates for each <strong>of</strong> the 247 different combinations <strong>of</strong> events were constructed. Managing the<br />

timing <strong>of</strong> events in FullCAM was important, as the chronological sequence <strong>of</strong> events impacted on the<br />

relative size <strong>of</strong> each pool at the time an event occurs.<br />

Event Timing<br />

FullCAM <strong>of</strong>fers two options for managing the timing <strong>of</strong> events; events can be scheduled by either<br />

absolute dates or dates relative to the start <strong>of</strong> simulation. Both options have their pros and cons and<br />

the Land Carbon Project chose to use relative dates for all events.<br />

The Plots were configured to start simulation at ‘year 0’ (corresponding to the calendar year 1900) and<br />

run for 150 years. For example, a Plot with harvesting in the 1940’s and a wildfire in the 1960’s<br />

would have the harvesting scheduled for year 45 and the wildfire scheduled for year 65.<br />

After examining the model output it became clear that the model needed a longer period <strong>of</strong> time to<br />

reach stable levels <strong>of</strong> carbon in the soil pools. The starting time for all plots was then brought back<br />

400 years to allow the model to fully stabilise before any events were scheduled. Figure 7 shows how<br />

long soil carbon takes to stabilise.<br />

Figure 7: Soil carbon stocks vs. tree carbon stocks.<br />

Estate Disturbance<br />

Figure 7 also shows the impact <strong>of</strong> an event scheduled to occur in 1860. This event was included to<br />

provide disturbance <strong>of</strong> the estate prior to events occurring. Omitting this ‘early’ event meant that<br />

FullCAM was assuming all stands were untouched, fully mature forests, when in fact almost all <strong>of</strong> the<br />

forests in the model have been burnt or otherwise disturbed in the past.<br />

All reporting was confined to the period 1930 to 2050 because that includes the period for which there<br />

is event data (1935 to 2005) as well as a period over which stock trends could be seen. All stock<br />

estimates for earlier periods were discarded.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 51<br />

Model Validation<br />

One <strong>of</strong> the FullCAM model outputs is total stem volume. The model estimates <strong>of</strong> stem volume were<br />

compared with stem volume estimates derived from field measurements as a means <strong>of</strong> verifying the<br />

performance <strong>of</strong> FullCAM.<br />

A set <strong>of</strong> 18 Plots were selected for validation against State Forest Resource Inventory (SFRI) data.<br />

The SFRI data was collected between 1999 and 2002 and the data was updated in 2006. The<br />

validation was performed against the FullCAM estimate <strong>of</strong> stem volume for the year 2006.<br />

Figure 8 shows the initial plot <strong>of</strong> FullCAM and SFRI data, indicating that FullCAM was under<br />

estimating stem volume.<br />

Figure 8: Initial plot <strong>of</strong> FullCAM and SFRI data.<br />

NCAT modelled VolStems m3 ha-1<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

y = 0.3038x + 81.646<br />

R 2 = 0.1575<br />

0<br />

0 100 200 300 400 500 600<br />

SFRI VolStems m3 ha-1<br />

Although the regression statistics suggested that the FullCAM estimates should be tripled, this was not<br />

pursued. Instead, the FullCAM estimates were modified by introducing a Type 2 treatment (Snowdon,<br />

2002) immediately after the first disturbance event. The coefficient <strong>of</strong> change for the Type 2<br />

treatment was 1.2. This value was chosen because it allowed stands to return to their pre-disturbance<br />

volumes without exceeding them. Introducing the Type 2 treatment reduces the intercept from 81.6 to<br />

34.8 and increases the R2 from 0.15 to 0.21.<br />

The Project Team attempted to improve the regression statistics by adjusting the forest type according<br />

to the corresponding IBRA region (up to this stage all plots had been modelled as Medium Eucalypt<br />

forest). This did not lead to any improvement and the original forest typing was retained.<br />

Selecting Model Outputs<br />

FullCAM <strong>of</strong>fers approximately 800 model outputs. The Land Carbon Project Team had to decide<br />

which model outputs to keep and which to ignore. To minimise the likelihood <strong>of</strong> making changes to<br />

the outputs (thereby necessitating significant volumes <strong>of</strong> work), the Land Carbon Project took the<br />

approach <strong>of</strong> including any value that was thought ‘might’ be <strong>of</strong> interest. As a result, 86 outputs were<br />

selected and this set <strong>of</strong> outputs was propagated to all Plots.<br />

Simulating Plots<br />

Each plot was simulated for the standard simulation period and the standard list <strong>of</strong> outputs selected.<br />

The model results were then saved as a Micros<strong>of</strong>t Excel file and the files were named after the Plot<br />

they represented.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 52<br />

UNCERTAINTY<br />

When providing estimates <strong>of</strong> modelled forest resource data it is common to provide an estimate <strong>of</strong> the<br />

level <strong>of</strong> uncertainty associated with those estimates. The normal practice is to calculate the standard<br />

deviation or coefficient <strong>of</strong> variation for a set <strong>of</strong> estimates <strong>of</strong> a single parameter such as stand volume.<br />

The Land Carbon Project is rather more complex, making use <strong>of</strong> data from various sources in a model<br />

system that is essentially a collection <strong>of</strong> other models (FullCAM makes use <strong>of</strong> elements <strong>of</strong> RothC,<br />

3PG and CAMFor). There is also the factor <strong>of</strong> scale, the reliability <strong>of</strong> the estimates <strong>of</strong> carbon stocks<br />

may be acceptable at the State level but unacceptable at the regional or stand level.<br />

Estimating Uncertainty<br />

Two methods <strong>of</strong> estimating uncertainty were applied. The first was to make use <strong>of</strong> the Sensitivity<br />

Analysis function <strong>of</strong> FullCAM and the second was for the Project Team to make an informed estimate<br />

<strong>of</strong> uncertainty (at the State level).<br />

In order to perform Sensitivity Analysis, a licensed copy <strong>of</strong> Palisade’s @Risk Developer’s Kit is<br />

required. During Sensitivity Analysis <strong>of</strong> four plots with different event histories, the values for key<br />

model drivers, such as proportion <strong>of</strong> canopy removed in harvesting, were systematically varied over<br />

thousands <strong>of</strong> simulations and the results were presented as a distribution <strong>of</strong> possible carbon stocks.The<br />

results <strong>of</strong> Sensitivity Analysis for each <strong>of</strong> the four selected Plots are shown in Table 2.<br />

Table 2: Sensitivity Analysis in FullCAM<br />

Plot Type PLE (90% confidence) Total area <strong>of</strong> plots with the same<br />

event history<br />

Harvesting and fire 24 467,317<br />

Harvesting and no fire 17 213,576<br />

No harvesting and fire 18 4,874,142<br />

No harvesting and no fire 18 1,696,618<br />

The Sensitivity Analysis did not include uncertainty associated with the model parameters and the list<br />

<strong>of</strong> elements included in the analysis was not exhaustive. The quantity <strong>of</strong> uncertainty associated with<br />

the model parameters within FullCAM was not pursued.<br />

By comparison, the Land Carbon Project Team’s estimate <strong>of</strong> uncertainty, which was settled prior to<br />

performing the FullCAM Sensitivity Analysis, was a PLE <strong>of</strong> 40% at 90% confidence.<br />

PROCESSING RESULTS<br />

After all analysis in FullCAM was complete, there were 822 Micros<strong>of</strong>t Excel files that contained<br />

results in a particular format. That format could not be imported directly into the Land Carbon<br />

Database and so some intermediate formatting steps were required.<br />

Once the results <strong>of</strong> FullCAM simulation were stored in the Land Carbon Database a series <strong>of</strong> queries<br />

were run to create a set <strong>of</strong> tables that would be joined to the GIS. Each <strong>of</strong> those tables contained data<br />

for a single FullCAM output as a time series with one record in the table for each Plot. Structuring the<br />

tables in this manner allowed these tables to be joined to the Input Layer <strong>of</strong> the GIS.<br />

All FullCAM output was expressed in total tonnes <strong>of</strong> carbon whereas tonnes <strong>of</strong> carbon per hectare<br />

were considered more suitable for map production. During the process <strong>of</strong> creating the new tables, all<br />

total carbon values were converted to per hectare values.<br />

Some additional processing was required to separate emissions attributable to planned burning or postharvest<br />

burning from those emissions from wildfire. The objective was to be able to report on<br />

anthropogenic and non-anthropogenic emissions.<br />

Joining to the GIS<br />

Once the set <strong>of</strong> GIS join tables had been created they were added to the ArcMap project file so that<br />

they were available to be joined to the Input Layer. Once a result table had been joined to the Input<br />

Layer it was possible for the user to select data from any year and use that as the basis for displaying


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 53<br />

the result in a map. By saving maps from successive years as Portable Document Format (PDF) files,<br />

it was possible to create a ‘movie sequence’ showing changes in carbon stocks over time.<br />

Estate Analysis<br />

Information about the whole <strong>of</strong> Victoria’s publicly managed land was analysed in the Land Carbon<br />

Database where sets <strong>of</strong> queries were used to create tabular summaries for inclusion in the Project<br />

Report. Tabular summaries <strong>of</strong> total carbon stocks, stocks for individual pools, emissions by source<br />

and removals from the atmosphere were all created.<br />

Individual sets <strong>of</strong> accounts for ‘areas <strong>of</strong> interest’, such as State forest tenure were created using the<br />

GIS to ‘clip’ out those areas. Maps and tabular summaries for those areas were generated using the<br />

same methods as applied to the entire publicly managed land estate.<br />

RESULTS<br />

The results <strong>of</strong> the Land Carbon Project were not publicly available at the time <strong>of</strong> preparing this paper.<br />

CONCLUSIONS<br />

The provision <strong>of</strong> a set <strong>of</strong> carbon accounts led to an immediate request for further information and in<br />

that regard these initial accounts have served to sharpen focus on the management <strong>of</strong> carbon stocks on<br />

Victoria’s publicly managed land.<br />

There are clear policy implications for fire management and the heavy influence <strong>of</strong> fire events on total<br />

stocks generally support the notion that <strong>Australia</strong> would be unwise to sign up to Article 3.4 <strong>of</strong> the<br />

Kyoto Protocol.<br />

The Land Carbon Project Team acknowledges the limitations <strong>of</strong> these accounts but remain confident<br />

that they have met the objectives for which they were created.<br />

REFERENCES<br />

Natural Resource Management Ministerial Council (2004), Directions for the National Reserve System –<br />

A Partnership Approach, <strong>Australia</strong>n Government, Department <strong>of</strong> the Environment and Heritage,<br />

Canberra, ACT<br />

Snowdon, P. 2002. Modelling Type 1 and Type 2 growth responses in plantations after application <strong>of</strong><br />

fertiliser or other silvicultural treatments. Forest Ecology and management 163 (2002) 229-244


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 54<br />

CARBON STORES AND BIOENERGY SUPPLIES:<br />

OPPORTUNITIES TO DEVELOP SUSTAINABLE PLANTATION<br />

INDUSTRIES IN LOW RAINFALL AGRICULTURAL LANDSCAPES<br />

John Tredinnick 1 and Jodie Mason 1<br />

ABSTRACT<br />

There is now a global focus on climate change that provides opportunities to develop<br />

financially viable industries on <strong>Australia</strong>n agricultural lands. The opportunities will be<br />

assisted by new policies and regulation favouring low carbon fuels and sustainable land<br />

management practices. Planting <strong>of</strong> mallee crops on cleared land in central and southern<br />

<strong>Australia</strong> is ideally structured to meet these objectives. The mallee system focuses on the<br />

planting <strong>of</strong> belts <strong>of</strong> trees across the landscape, fully integrated with cropping and livestock<br />

production, while providing an alternative income source and providing public and private<br />

environmental benefits. Mallees therefore create a unique opportunity for the development<br />

<strong>of</strong> sustainable products, including wood products and bioenergy, as a dedicated carbon sink,<br />

or even providing all products from the same area <strong>of</strong> land. Current forest carbon policy<br />

frameworks support the creation <strong>of</strong> carbon forests, however these policies don’t allow<br />

accounting for the use <strong>of</strong> the harvested wood products, nor is there a direct benefit to the<br />

grower from the displacement <strong>of</strong> fossil fuels when biomass is harvested for energy creation.<br />

This paper suggests that the greenhouse and regional socio-economic benefits <strong>of</strong> a<br />

harvesting regime that leads to the substitution <strong>of</strong> fossil fuels or the retention <strong>of</strong> carbon in<br />

wood products can be greater than those from dedicated carbon sinks, and the greenhouse<br />

benefits more permanent.<br />

INTRODUCTION<br />

Sustainable and secure fuel and energy supplies will need to be developed in the future. There are<br />

very promising opportunities for meeting this demand from energy crops such as mallee eucalypts<br />

planted in the lower rainfall regions <strong>of</strong> <strong>Australia</strong>. Although there is some concern that energy crops<br />

will compete for land for the production <strong>of</strong> food, the mallee system can be structured to balance the<br />

demands for both energy and food production.<br />

For the industry to develop, the trees will need to provide at least an equivalent financial return to<br />

existing farm enterprises. Strategies for industry development also must seek to overcome three key<br />

impediments to the development <strong>of</strong> new industries:<br />

• The technologies and markets associated with the processing <strong>of</strong> mallees into industrial<br />

products on a large scale are not yet developed to the stage where any single product, or<br />

combination <strong>of</strong> products, provides a clear basis for economically viable development;<br />

• If clear opportunities were currently available, current mallee plantings are not at a<br />

sufficient scale for a processing industry to be developed; and<br />

• The demand for mallees as part <strong>of</strong> a dedicated carbon sink is far greater than that for<br />

industrial production. Under current policy settings this could result in suitable land<br />

being “locked up” on a non-harvest regime for more than 70 years.<br />

The first two impediments result in a “chicken and egg” scenario where the commitment from<br />

processors towards developing an industry is tempered by the lack <strong>of</strong> resource in the short term, and<br />

potential growers are looking for clear evidence <strong>of</strong> future pr<strong>of</strong>itability before they develop tree crops.<br />

1 URS Forestry, Level 3, 20 Terrace Road, East Perth, WA 6004, <strong>Australia</strong>. Email: john_tredinnick@urscorp.com


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 55<br />

This paper draws on work undertaken as part <strong>of</strong> the development <strong>of</strong> the Oil Mallee Industry<br />

Development Plan for Western <strong>Australia</strong> (URS 2008) to summarise the silvicultural regimes that have<br />

potential to maximise productivity. It also describes the potential for mallee eucalypts to produce a<br />

number <strong>of</strong> commercial products and the potential economic returns from these products. It then<br />

illustrates how the impediments to industry development might be overcome by capitalising on the<br />

opportunities that can be provided by carbon trading. Such policy developments have the potential to<br />

provide a way forward for the industry that includes both carbon sequestration and the production <strong>of</strong><br />

energy.<br />

RESOURCE DEVELOPMENT<br />

In NSW and Victoria a eucalyptus oil industry based on blue mallee (Eucalytpus polybractea) in<br />

native woodlands has existed for nearly 100 years. These traditional industries have also commenced<br />

their own planting to secure and expand the resource.<br />

Mallee plantings in the WA wheat belt region began in the early 1990s, based on the concept <strong>of</strong><br />

developing a commercially viable tree crop to reduce agricultural salinity on the necessary scale<br />

(Bartle and Shea 2002). Woody crops like mallee, with their industrial products potential, were also<br />

seen as an option to diversify farm business and regional economies in the face <strong>of</strong> the long term<br />

decline in the terms <strong>of</strong> trade for agricultural products (Bartle 2006).<br />

The WA Department <strong>of</strong> Conservation and Land Management (CALM) 2 coordinated the first broad<br />

scale investment and establishment <strong>of</strong> oil mallee plantings in the period 1994-96 during which a total<br />

<strong>of</strong> six million mallee trees were planted (Bartle 2006). By 1995 there was a substantial interest in<br />

furthering industry development and a body <strong>of</strong> growers formed the Oil Mallee Association (OMA)<br />

(Bartle and Shea, 2002, OMA 2007). In 1997 the Oil Mallee Company (OMC) was established by the<br />

OMA to form a commercial arm for the oil mallee industry in WA (OMA 2007). Significant<br />

investment into mallee plantings for carbon sequestration in WA began in 2003, when the OMC<br />

established 1,000 ha <strong>of</strong> plantings across 24 farms for voluntary <strong>of</strong>fsets purchased by a Japanese power<br />

producer, General Environmental Technos Co Ltd (previously Kansai).<br />

Both the OMA and OMC are still in operation today, furthering the expansion and diversification <strong>of</strong><br />

the industry. Other companies that are focussed on the economic benefits <strong>of</strong> carbon trading have also<br />

driven an expansion <strong>of</strong> the resource base over the previous 12 months. This is likely to continue as a<br />

result <strong>of</strong> the current Federal Government’s commitment to a Carbon Pollution Reduction Scheme<br />

(CPRS) by 2011 has provided further impetus for the mallee industry in NSW and this is likely to<br />

follow in other states.<br />

Integration with Agriculture<br />

Mallees are readily combined with large scale annual cropping (Bartle et al 2007). However, to make<br />

a real contribution to improving agricultural productivity mallees have to be complementary to, not<br />

just compatible with, annual plant agriculture. Although water tends to be the most limiting resource<br />

in southern <strong>Australia</strong>n, current farm management systems allow a proportion <strong>of</strong> rainfall to run-<strong>of</strong>f or<br />

infiltrate below root zones and to be lost from production, leading to rising water tables and dryland<br />

salinity. A real complementary role for mallee is achieved when some <strong>of</strong> this water is captured.<br />

Carefully designed narrow belts <strong>of</strong> deep rooted perennials like mallee are being used for this purpose.<br />

Robertson et al (2006) and Sudmeyer and Goodreid (2007) have shown that belts <strong>of</strong> Eucalyptus<br />

species can develop a broad (up to 20 m either side) and deep (beyond 10 m) zone <strong>of</strong> dewatered soil<br />

that can theoretically be used as a sink for lateral flows and leakage from the adjacent annual plant<br />

agriculture (Figure 1).<br />

In the evaluation <strong>of</strong> mallee performance the interactions between mallee belts and adjacent annuals<br />

must also be taken into account. The strong lateral root systems <strong>of</strong> mallee in belts can impose<br />

competition on adjacent crop or pasture.<br />

2 Now the Department <strong>of</strong> Environment and Conservation (DEC)


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 56<br />

Source: Huxtable and Bartle (2007)<br />

Figure 1: Water capture by mallees<br />

Optimising Yields<br />

The rate <strong>of</strong> growth <strong>of</strong> mallee plantings is affected by a number <strong>of</strong> variables. Apart from species and<br />

provenance, almost all <strong>of</strong> the factors are linked to water availability. These variables include climate<br />

(rainfall, temperature and evaporation); soil (depth, texture and structure); and the quality <strong>of</strong> site<br />

preparation and management – particularly weed and pest control.<br />

The most common belt configuration has been four row belts with a width <strong>of</strong> 2 metres (m) between<br />

rows. The effective belt width to the drip line <strong>of</strong> the trees under this configuration is 10 m 3 . The<br />

distance between mallee belts allows integration with agricultural practices and typical farm plans<br />

have the belts 40-100 m apart, depending upon farm machinery requirements.<br />

Current analysis by DEC (Bartle pers comm. 2008) shows that the effectiveness <strong>of</strong> mallee in water use<br />

means that economic factors will push down planting density and row number. It appears likely that a<br />

two row or even a one row belt may be the economic optimum from the biomass yield perspective, but<br />

other considerations like operational efficiency, cost, aesthetics or shelter benefit may work against<br />

this. Mallee water use effectiveness will also push design <strong>of</strong> belt layout into configurations that will<br />

facilitate better interception <strong>of</strong> lateral water flows.<br />

Huxtable and Bartle (2007) report that actual yields <strong>of</strong> above ground biomass range from 5-10<br />

bdmt 4 /ha/annum based on measurements at ages 7 to 11 years. Using these growth increments to<br />

adjust outcomes from Cooper et al (2005), Huxtable and Bartle (2007) suggest that productivity will<br />

lie in the range 5-15 bdmt/ha/annum over the rainfall range 350-550 mm pa. It should be noted that<br />

these yields have been standardised such that they are based on two row belts (i.e. where four row<br />

belts had been established only the two outer rows have been included in the productivity estimates).<br />

If a given area is to be established using four row belts the overall productivity is expected to be<br />

around 25% lower. Block plantings will not enable the same benefits to be concentrated in the area<br />

planted and may only achieve 30-50% <strong>of</strong> the productivity <strong>of</strong> two row belts over the area planted. This<br />

is shown diagrammatically in Figure 2.<br />

3<br />

This width meets <strong>Australia</strong>’s current definition <strong>of</strong> Kyoto compliant forests under Article 3.3 <strong>of</strong> the Kyoto Protocol<br />

4 Bone dry metric tonnes


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 57<br />

Two row belts<br />

Assuming 7 bdmt/ha/annum <strong>of</strong> growth, yield at age 5 is expected<br />

to be:<br />

• 35 bdmt per net hectare planted<br />

• 1.75 bdmt per gross (paddock) hectare, assuming 50% <strong>of</strong><br />

soils suitable and 10% <strong>of</strong> suitable soils planted.<br />

Source: URS (2008)<br />

Figure 2: Examples <strong>of</strong> mallee planting configurations and yields<br />

Four row belts<br />

Assuming 5.25 bdmt/ha/annum <strong>of</strong> growth (75% <strong>of</strong> two row<br />

productivity), yield at age 5 is expected to be:<br />

• 26.25 bdmt per net hectare planted<br />

• 1.3 bdmt per gross (paddock) hectare, assuming 50% <strong>of</strong><br />

soils suitable and 10% <strong>of</strong> suitable soils planted.<br />

Block planting<br />

Assuming 2.8 bdmt/ha/annum <strong>of</strong> growth (40% <strong>of</strong> two row<br />

productivity), yield at age 5 is expected to be:<br />

• 14 bdmt per net hectare planted<br />

• 7 bdmt per gross (paddock) hectare, assuming 50% <strong>of</strong> soils<br />

suitable and all suitable soils planted.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 58<br />

USES AND POTENTIAL PRODUCTS<br />

A range <strong>of</strong> potential processing options and markets exists for mallees. The commercial viability <strong>of</strong><br />

these options is dependent on a number <strong>of</strong> factors, including:<br />

• Costs <strong>of</strong> production from planting to processing;<br />

• Demand for products;<br />

• Prices that can be achieved in markets; and<br />

• Achieving necessary resource requirements (scale) for market development.<br />

Historically, there has been a significant focus on processing mallees to produce multiple products.<br />

Integrated wood processing (IWP) was considered as a production option in WA for its potential to<br />

efficiently direct specific biomass fractions to higher value products while directing residues to lower<br />

value uses such as bioenergy and eucalyptus oil, and achieve a commercially viable industry. Western<br />

Power (now Verve Energy) completed construction <strong>of</strong> a trial IWP plant in 2006 to undertake<br />

operational scale testing. During the trial phase the plant produced activated carbon from the wood<br />

fraction <strong>of</strong> the mallees and Eucalyptus oil from the leaves, while all residues, including spent leaf and<br />

waste heat, were used for bioelectricity (the production <strong>of</strong> electricity). The trial indicated that<br />

production <strong>of</strong> activated carbon, eucalyptus oil and energy from mallee biomass is feasible.<br />

Verve Energy is now investigating the potential to attract investment in the construction and operation<br />

<strong>of</strong> a larger plant and to replicate the plant in other areas <strong>of</strong> the WA wheat belt and other parts <strong>of</strong><br />

<strong>Australia</strong>. The prospects for integrated processing are well developed and appear promising. The<br />

potential scale <strong>of</strong> the mallee resource is such that a range <strong>of</strong> additional markets should be considered<br />

as either stand alone industries or as part <strong>of</strong> an integrated production process. Table 1 provides a<br />

summary these products and markets.<br />

Table 1: Summary <strong>of</strong> market opportunities<br />

Product Opportunities and constraints<br />

Bioelectricity is the production<br />

<strong>of</strong> renewable electricity from<br />

solid biomass through direct<br />

combustion <strong>of</strong> wood, co-firing<br />

with other fuel sources, or via<br />

production <strong>of</strong> gaseous and liquid<br />

fuels.<br />

Carbon The carbon dioxide<br />

removed from the atmosphere<br />

and stored in biomass is<br />

recognised as a carbon sink<br />

which <strong>of</strong>fsets greenhouse gas<br />

emissions and thereby helps to<br />

mitigate global climate change.<br />

Wood pellets are produced by<br />

grinding wood material into<br />

small particles, then<br />

compressing the material to bind<br />

the wood together. Energy is<br />

produced upon combustion.<br />

• Strong market opportunities exist in <strong>Australia</strong> for bioelectricity production. Small scale co-firing is<br />

being implemented around <strong>Australia</strong> and there are currently several proposals to develop dedicated<br />

bioenergy plants.<br />

• Opportunities for renewable electricity in <strong>Australia</strong> are expected to grow as a result <strong>of</strong> government<br />

policies, particularly the Commonwealth Government’s Mandatory Renewable Energy Target<br />

(MRET) scheme. The increased renewable energy target <strong>of</strong> 20% by 2020 being introduced by the<br />

Federal Government is likely to be a catalyst for significant investment in renewable energy<br />

infrastructure<br />

• The placement and size <strong>of</strong> bioenergy plants can be linked closely to the electricity grid in an area<br />

with excess network transmission capacity, and must have sufficient long term feedstock supplies<br />

within viable transport distances.<br />

• Distributed energy supply can also be provided to isolated projects.<br />

• The Federal Government established the Department <strong>of</strong> Climate Change and water (DCC) on 3<br />

December 2007. Shortly thereafter, the <strong>Australia</strong>n Government <strong>of</strong>ficially ratified the Kyoto<br />

Protocol, and in doing so committed to limiting its national greenhouse gas emissions to 8% above<br />

1990 levels during the first Kyoto period <strong>of</strong> 2008 to 2012. The Government also committed to<br />

implement the CPRS in 2011.<br />

• Principal markets for wood pellets are in Europe, Canada, Japan and the United States. There is<br />

increasing interest from <strong>Australia</strong>n investors to export wood pellets. Tests have shown that pellets<br />

manufactured from mallee will meet the standards required for the export market.<br />

• Export represents the clearest opportunity for <strong>Australia</strong>n wood pellets in the short term,<br />

notwithstanding the significant transport distances and increasing domestic production in Europe.<br />

• Prices for renewable electricity in <strong>Australia</strong> may increase and potentially support a domestic market.<br />

• Domestic heating could also provide a market for pellets as a replacement for firewood as the pellets<br />

have proven energy efficiency and environmental advantages.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 59<br />

Product Opportunities and constraints<br />

Activated carbon is a form <strong>of</strong><br />

charcoal that is valued for its<br />

ability to purify liquids and<br />

gases. It can be produced from<br />

wood through carbonisation and<br />

activation processes.<br />

Composite wood products such<br />

as Medium Density Fibreboard<br />

(MDF), particleboard and<br />

Engineered Strand Lumber<br />

(ESL) are used extensively in a<br />

range <strong>of</strong> applications in the<br />

construction, manufacturing and<br />

furniture sectors.<br />

Eucalyptus oil is produced from<br />

the leaves <strong>of</strong> trees and used as an<br />

ingredient in a range <strong>of</strong> products.<br />

Charcoal presents two<br />

opportunities: industrial carbon<br />

which is used in the steel<br />

industry as a substitute for<br />

coking coal or in the non-ferrous<br />

industry as a specialty reductant,<br />

and biochar which is used as a<br />

soil additive.<br />

Lignocellulosic ethanol is a<br />

liquid bi<strong>of</strong>uel produced from<br />

wood by breaking down biomass<br />

to isolate the component sugars,<br />

which are then processed to<br />

produce ethanol. Its most likely<br />

immediate application is in the<br />

blended fuel product – E10.<br />

Biomass to Liquid fuels are<br />

produced when biomass<br />

undergoes a gasification process<br />

and gases are then synthesised<br />

into a range <strong>of</strong> liquid fuels such<br />

as methanol and synthetic diesel.<br />

Bio-oil is produced through a<br />

process <strong>of</strong> pyrolysis where<br />

woody biomass is heated in the<br />

absence <strong>of</strong> air. It is a<br />

combustible product and has<br />

potential as a substitute for fossil<br />

fuels in a range <strong>of</strong> applications.<br />

Technology has recently been<br />

commercialised overseas.<br />

ECONOMIC MODELLING<br />

• Opportunities exist to replace imports <strong>of</strong> activated carbon into <strong>Australia</strong>, particularly for use in the<br />

gold industry and water filtration systems where it is used in the recovery process.<br />

• Feasibility will depend on production costs, but the quality <strong>of</strong> the product that can be produced from<br />

mallees has proven to be suitable.<br />

• <strong>Australia</strong> is a significant exporter <strong>of</strong> MDF and consumption <strong>of</strong> composite wood products is likely to<br />

continue to increase in <strong>Australia</strong>. Tests have shown that mallee is a suitable input for the production<br />

<strong>of</strong> MDF.<br />

• Global markets for engineered wood products such as ESL are expanding and there are strong<br />

opportunities to supply the ESL market based on growing international demand for this relatively<br />

new product.<br />

• Most <strong>of</strong> <strong>Australia</strong>’s eucalyptus oil consumption is based on imports from China and there appears to<br />

be an attractive opportunity for domestically produced oil to replace imports.<br />

• There are strong precedents to the use <strong>of</strong> naturally occurring terpenes (such as those found in<br />

Eucalyptus oil) as industrial solvents and cleaners, albeit at significantly lower prices than current<br />

uses <strong>of</strong> Eucalyptus oil.<br />

• Eucalyptus oil is currently a niche product and large scale sales have the potential to significantly<br />

lower prices and reduce the viability <strong>of</strong> the industry.<br />

• It will be very difficult for industrial carbon produced from biomass to compete with the current<br />

coking coal industry, where supply is well established and low prices for coal are highly competitive.<br />

However, specialised activities requiring wood charcoal for reduction could provide a market for<br />

certain fractions <strong>of</strong> mallee.<br />

• Development <strong>of</strong> the biochar product is being undertaken, including trials to demonstrate the benefits<br />

<strong>of</strong> biochar as a soil enhancer and partial replacement for fertiliser. Other financial incentives,<br />

including <strong>of</strong>ficial recognition <strong>of</strong> biochar as a tradeable carbon sink, will increase the potential <strong>of</strong> the<br />

market.<br />

• Lignocellulosic ethanol <strong>of</strong>fers several advantages over traditional ethanol production from<br />

agricultural feedstocks (such as sugar cane and grain) in that it utilises a potentially cheaper<br />

feedstock and has a better energy balance as a result <strong>of</strong> lower greenhouse gas emissions during<br />

production. Lignocellulosic ethanol is also less likely to compete with human and agricultural food<br />

markets.<br />

• The lignocellulosic process is yet to be commercialised but is currently undergoing intensive research<br />

and development overseas. Several commercial prototypes <strong>of</strong> the technology are planned over the<br />

next 3-5 years.<br />

• Biomass to Liquid fuels present an opportunity to supply <strong>Australia</strong>’s diesel market. Growing<br />

demand and increasing prices for diesel will increase the viability <strong>of</strong> synthetic fuels.<br />

• Individual technologies are well understood, but integrated production requires further development<br />

and is a current constraint to the industry.<br />

• Markets for bio-oil are limited by the fact that it cannot currently be used as a fuel in existing forms<br />

<strong>of</strong> transportation. However, it can be used as a boiler fuel or in gas turbines for power generation.<br />

• Research and development is being undertaken to allow blending with hydrocarbon fuels which is<br />

hoped to broaden the overall market for the product.<br />

This section describes a model that was developed to allow the assessment <strong>of</strong> oil mallee production<br />

alternatives. The model was developed using principles <strong>of</strong> benefit-cost analysis to provide a<br />

discounted cash flow analysis <strong>of</strong> biomass production. The model allows comparison between


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 60<br />

locations and areas with different growth potential, and between biomass production options at<br />

different scales. A 30 year timeframe has been used and a real, pre-tax discount rate <strong>of</strong> 7% has been<br />

assumed. Further details <strong>of</strong> the assumptions are outlined in URS (2008).<br />

Case study examples have been developed in the south west <strong>of</strong> Western <strong>Australia</strong> to provide a<br />

comparison between rainfall zones and areas adjacent to a port, rail or major electrical transmission<br />

line. The characteristics <strong>of</strong> the four examples are described in Table 2.<br />

Table 2: Characteristics <strong>of</strong> case study examples<br />

Region<br />

Rainfall<br />

(mm)<br />

Characteristics Potential industries<br />

On<br />

SWIS 1<br />

Near a<br />

port<br />

Near other<br />

biomass<br />

residue<br />

sources<br />

Pellet<br />

export<br />

Small bioenergy<br />

plant<br />

(5MW)<br />

Medium<br />

bio-energy<br />

plant<br />

(25MW)<br />

Engineered<br />

wood<br />

products<br />

Eucalyptus<br />

oil product<br />

Activated<br />

carbon<br />

products<br />

1 500-600 Yes Yes Yes Co-product Co-product<br />

2 400-500 No Yes Yes Co-product Co-product<br />

3 300-400 Yes No No Co-product Co-product<br />

4 300-400 No No No Co-product Co-product<br />

Very good potential Good potential Some potential<br />

Note 1. South West Interconnected System. The SWIS consists <strong>of</strong> nearly 88,000 kilometres <strong>of</strong> powerlines stretching from Kalbarri in the<br />

north to Kalgoorlie in the east and south to Albany.<br />

Table 3 shows the value <strong>of</strong> returns from biomass over a range <strong>of</strong> factory gate prices for mallee<br />

biomass and the value <strong>of</strong> carbon sequestered.<br />

Table 3: Annualised on-farm returns from biomass ($/ha pa)<br />

Factory gate value ($/gmt)<br />

Region 1 -$34.02 $35 $40 $45 $50 $55<br />

$0 -$3 $21 $44 $67 $90<br />

$20 $39 $62 $85 $108 $131<br />

Carbon value ($/t CO2e) $40 $80 $103 $126 $150 $173<br />

$60 $121 $145 $168 $191 $214<br />

$80 $163 $186 $209 $232 $255<br />

Region 2 $10.99<br />

$0 -$8 $10 $28 $46 $64<br />

$20 $24 $42 $60 $78 $96<br />

Carbon value ($/t CO2e) $40 $56 $74 $92 $110 $128<br />

$60 $88 $106 $124 $142 $160<br />

$80 $120 $138 $156 $174 $192<br />

Region 3 -$53.93<br />

$0 -$41 -$28 -$15 -$2 $11<br />

$20 -$18 -$5 $8 $21 $34<br />

Carbon value ($/t CO2e) $40 $5 $18 $31 $44 $57<br />

$60 $28 $41 $54 $67 $80<br />

$80 $51 $64 $77 $90 $103<br />

Region 4 -$64.11<br />

$0 -$36 -$23 -$10 $3 $16<br />

$20 -$13 $0 $13 $26 $39<br />

Carbon value ($/t CO2e) $40 $10 $23 $36 $49 $61<br />

$60 $33 $46 $59 $72 $84<br />

$80 $56 $69 $82 $95 $107


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 61<br />

The estimates <strong>of</strong> returns are presented as annualised values which are based on the net present value <strong>of</strong><br />

returns from biomass plantings harvested regularly over a thirty year period. These returns should be<br />

compared against returns from agriculture or other land uses that might be achieved in each rainfall<br />

zone.<br />

In the analysis, the value <strong>of</strong> CO2e impacts directly on the value that the resource has as a carbon <strong>of</strong>fset.<br />

In the case <strong>of</strong> wood fibre being supplied to a domestic energy producer, the value <strong>of</strong> CO2e will also<br />

have an impact on the factory gate value, as will the value <strong>of</strong> RECs and the value <strong>of</strong> any capacity<br />

credits 5 . The potential interaction between markets for CO2e, RECs and capacity credits is complex<br />

(see Garnaut 2008). As a general guide it can be assumed that as the value <strong>of</strong> CO2e increases, so will<br />

the factory gate price that can be paid by an energy producer for the mallee biomass.<br />

In making comparisons between agricultural returns and the returns from tree crops under various<br />

values <strong>of</strong> CO2e, it needs to be noted that there are also likely to be changes in agricultural production<br />

values that result from carbon trading if agriculture becomes a covered sector under the CPRS. These<br />

impacts are likely to be in the form <strong>of</strong> liabilities incurred by agricultural production and could make<br />

any advantages <strong>of</strong> tree crops more favourable.<br />

PROMOTING CARBON SEQUESTRATION AND BIOMASS PRODUCTION<br />

To a large extent, policy initiatives that support development <strong>of</strong> new products based on the mallee<br />

resource are already in train, particularly the federal Government’s proposed increase in renewable<br />

energy targets and the development <strong>of</strong> a national CPRS. Investors, processors and energy producers<br />

are expected to respond to these initiatives. Hence any proposed changes to policy need to be<br />

focussed on the development <strong>of</strong> a mallee resource that is available to be utilised by project proponents<br />

as technologies develop over the next five years.<br />

To do this, it is necessary to break the “chicken and egg” scenario <strong>of</strong> industry development versus<br />

resource development that is described in the introduction to this paper. Carbon trading provides the<br />

circuit breaker, but current carbon trading settings do not encourage optimal establishment <strong>of</strong> mallees<br />

for biomass production and carbon sequestration from the same area <strong>of</strong> land. This can be changed<br />

through initiatives that recognise that the objective <strong>of</strong> minimising greenhouse gas emissions can be<br />

achieved by both harvesting trees for the production <strong>of</strong> bioenergy or wood products and by creating a<br />

carbon <strong>of</strong>fset credit from these products as well as the standing tree biomass.<br />

Under proposed guidelines for the CPRS, forest owners who “opt-in” will be <strong>of</strong>fered units for the<br />

carbon sequestered on their land that are equivalent to the average amount <strong>of</strong> carbon in standing<br />

timber over the sequestration period after allowing for periodic harvesting. In the same way that there<br />

is increasing support for the inclusion <strong>of</strong> carbon stored in harvested wood products as <strong>of</strong>fset credits,<br />

there must also be potential for the production <strong>of</strong> woody biomass for uses where it displaces fossil<br />

fuels, or the use <strong>of</strong> biochar as a soil additive, to generate an <strong>of</strong>fset credit for the grower.<br />

Figure 3 describes the process <strong>of</strong> sequestration via fossil fuel displacement. The figure shows that<br />

there are ongoing greenhouse benefits from the biomass production process, in the same way that<br />

carbon can be sequestered in a dedicated sink. Furthermore, the greenhouse benefits <strong>of</strong> a harvesting<br />

regime that leads to the substitution <strong>of</strong> fossil fuels can be greater (per hectare planted), and more<br />

permanent, than the development <strong>of</strong> tree crops as dedicated carbon sinks.<br />

In proposing this benefit to mallee growers it is acknowledged that the use <strong>of</strong> woody biomass as a<br />

substitute for fossil fuels in energy production is likely to be recognised as a carbon neutral source <strong>of</strong><br />

energy that reduces the potential liability <strong>of</strong> energy producers. It could be expected that these benefits<br />

will be recognised financially through market prices that are influenced by the value <strong>of</strong> RECs, the<br />

penalties to energy producers for carbon emissions under a future CPRS, or as a combination <strong>of</strong> the<br />

impacts <strong>of</strong> both factors. However, values under this implicit market will be dependent on the<br />

interaction <strong>of</strong> a number <strong>of</strong> supply/demand factors and energy producers will not necessarily pass on<br />

5<br />

A capacity credit is a measure <strong>of</strong> an electric power generator’s expected or actual contribution to meeting system reliability<br />

goals.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 62<br />

the full market value <strong>of</strong> carbon to the grower.<br />

Figure 3: Cumulative sequestration via fossil fuel displacement versus non-harvest sinks<br />

tC/ha<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0 10 20 30 40 50<br />

Time (years)<br />

Recognition for fossil fuel displacement or wood products<br />

Above ground biomass (Harvest)<br />

Litter / debris<br />

Below ground biomass (Harvest)<br />

Change in soil carbon<br />

--- Non- harvest regime<br />

Source: URS (2008)<br />

Note: Figure and scale is indicative only and not intended to represent actual rates <strong>of</strong> sequestration<br />

Furthermore, without the direct recognition <strong>of</strong> fossil fuel displacement as an <strong>of</strong>fset credit, large<br />

emitters <strong>of</strong> carbon dioxide (such as oil and gas producers) that establish mallees for <strong>of</strong>fset credits will<br />

need to forego their accumulated credits at the time <strong>of</strong> harvest. They will then need to buy another<br />

<strong>of</strong>fset on the market with the proceeds <strong>of</strong> sales from the biomass to meet their obligations under any<br />

future emissions cap. This exposes the organisation to a price risk associated with future carbon<br />

trading that they were seeking to avoid through the early action <strong>of</strong> planting mallees.<br />

CONCLUSIONS<br />

The development <strong>of</strong> an <strong>Australia</strong>n CPRS should encourage parties wishing to establish <strong>of</strong>fsets via tree<br />

crops to also participate in a future industry based on biomass harvesting. Under such a framework<br />

the resource owner would retain carbon credits at the time <strong>of</strong> harvest that are equivalent to the net<br />

greenhouse benefit from replacing fuels that release more CO2e in the production <strong>of</strong> energy or<br />

commodities.<br />

The advantages <strong>of</strong> such an initiative are significant and numerous:<br />

• Investors in carbon sinks would have an incentive to keep their options open for future<br />

harvesting;<br />

• A mallee resource would be created as a result <strong>of</strong> the demand for carbon <strong>of</strong>fsets that also provides<br />

the basis for a future processing industry;<br />

• Periodic harvesting controls the height and spread <strong>of</strong> the mallee belt, keeps them in the more<br />

physiologically active young coppice stage <strong>of</strong> growth and moderates the competition imposed on<br />

adjacent crops and pastures;<br />

• In addition to environmental benefits <strong>of</strong> mallees, there is also the potential for new jobs and<br />

industries and associated socio-economic benefits in regional communities as a result <strong>of</strong><br />

processing industries that would not arise from dedicated sinks; and<br />

• An increase in <strong>of</strong>fsets/ sequestration per hectare planted.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 63<br />

ACKNOWLEDGMENTS<br />

URS Forestry would like to thank the following individuals and groups for their assistance, comments<br />

and contributions towards the development <strong>of</strong> the Oil Mallee Industry Development Plan for Western<br />

<strong>Australia</strong>:<br />

• Simon Dawkins (Oil Mallee Association)<br />

• Tim Emmott (Oil Mallee Association)<br />

• Paul Brennan (Forest Products Commission)<br />

• John Bartle (Department <strong>of</strong> Environment and Conservation)<br />

• Col Stucley (Enecon)<br />

• Jim Bland (Enecon)<br />

This Oil Mallee Industry Development Plan for Western <strong>Australia</strong> was funded by the <strong>Australia</strong>n<br />

Government and the Government <strong>of</strong> Western <strong>Australia</strong> through the National Action Plan for Salinity<br />

and Water Quality<br />

REFERENCES<br />

Bartle, J.R. (2006). New Non-Food Crops and Industries for <strong>Australia</strong>n Dryland Agriculture. Paper presented at<br />

the Green Processing Conference, Newcastle, NSW, 5-6 June 2006.<br />

Bartle J.R. and Shea S. (2002). Development <strong>of</strong> mallee as a large-scale crop for the wheat belt <strong>of</strong> WA. In<br />

'<strong>Australia</strong>n Forest Growers 2002 National Conference: Private Forestry - Sustainable accountable and<br />

pr<strong>of</strong>itable'. Albany, WA, <strong>Australia</strong> pp. 243-250.<br />

Bartle J.R., Olsen G., Cooper D. and Hobbs T. (2007). Scale <strong>of</strong> biomass production from new woody crops for<br />

salinity control in dryland agriculture in <strong>Australia</strong>. International Journal <strong>of</strong> Global Energy, Issues 27 2:<br />

115-137.<br />

Cooper, D., Olsen, G. and Bartle, J.R. (2005). Capture <strong>of</strong> agricultural surplus water determines the productivity<br />

and scale <strong>of</strong> new low-rainfall woody crop industries. <strong>Australia</strong>n Journal <strong>of</strong> Experimental Agriculture<br />

45:1369-1388.<br />

Garnaut R. (2008). Garnaut climate change review. Draft report June 2008. Report commissioned by the<br />

Commonwealth <strong>of</strong> <strong>Australia</strong>.<br />

Huxtable, D. and Bartle, J. (2007). .Predicting mallee biomass yield in the WA wheat belt. An interim report on<br />

investigations, Department <strong>of</strong> Environment and Conservation, February 2007.<br />

OMA (2007). The WA mallee story. A situation analysis prepared for the national mallee workshop, Oil Mallee<br />

Association, Canberra, 2007.<br />

Robinson N, Harper RJ, Smettem KRJ (2006). Soil water depletion by Eucalyptus spp integrated into dryland<br />

agricultural systems. Plant and Soil. 286:141-151<br />

Sudmeyer RA and Goodreid A (2007). Short rotation woody crops: a prospective method for phytoremediation<br />

<strong>of</strong> agricultural land at risk <strong>of</strong> salinisation in southern <strong>Australia</strong>. Ecological Engineering. 29: 350-361<br />

URS (2008). Oil Mallee Industry Development Plan for Western <strong>Australia</strong>. Report prepared for Forest Products<br />

Commission, Western <strong>Australia</strong> and the Oil Mallee Association <strong>of</strong> WA.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 64<br />

GIS-BASED CLASSIFICATION, MAPPING AND VALUATION<br />

OF ECOSYSTEM SERVICES IN PRODUCTION LANDSCAPES:<br />

A CASE STUDY OF THE GREEN TRIANGLE REGION<br />

OF SOUTH-EASTERN AUSTRALIA<br />

Himlal Baral 1* , Sabine Kasel 1 , Rodney Keenan 2 , Julian Fox 1 , Nigel Stork 3<br />

ABSTRACT<br />

This paper presents a GIS-based approach for classification, mapping and valuation <strong>of</strong><br />

selected ecosystem services using market and non-market valuation techniques. First, we<br />

identified and compiled a variety <strong>of</strong> spatial and non-spatial data and developed a land<br />

cover typology <strong>of</strong> the study area into a GIS environment. Second, we estimate the annual<br />

flow <strong>of</strong> economic value <strong>of</strong> each service using various economic valuation techniques. We<br />

found that the economic value <strong>of</strong> market ecosystem services such as timber and carbon<br />

was relatively straightforward. However the quantification and valuation <strong>of</strong> non-market<br />

services such as biodiversity was complicated. Finally, we produced an annual flow <strong>of</strong><br />

total economic value <strong>of</strong> sub-catchment G8 <strong>of</strong> the Lower Glenelg Basin, south-eastern<br />

<strong>Australia</strong> using the spatial economic valuation technique. We expect that this work will<br />

highlight research avenues to advance the ecosystem services framework as an<br />

operational basis in plantation dominant production landscapes in <strong>Australia</strong> and<br />

elsewhere.<br />

INTRODUCTION<br />

Forested ecosystems provide a variety <strong>of</strong> benefits for human beings – such as food, fibre, flood<br />

protection, clean water, and clean air. The benefits human populations derive, directly or indirectly,<br />

from ecosystem functions are labelled as ecosystem services (Costanza et al. 1998). According to this<br />

definition the goods and services are derived from the ‘functions’ and are utilised by ‘people’. The<br />

goods and services produced by the ecosystem are not always complementary, because the production<br />

<strong>of</strong> certain goods and services <strong>of</strong>ten results in the depletion or degradation <strong>of</strong> other goods and services.<br />

For example, timber extraction can affect visual quality, water quality and recreation. However, with<br />

proper management, there is a potential to integrate a variety <strong>of</strong> goods and services in production<br />

landscapes. The nature and typology <strong>of</strong> ecosystem services has been extensively elaborated elsewhere<br />

(de Groot et al. 2002; Boyd & Banzhaf 2007; Fisher & Kerry 2008; Wallace 2008; Fisher et al. 2009).<br />

The process <strong>of</strong> identifying, classifying, quantifying and mapping ecosystem services at the landscape<br />

level is increasingly recognised as an essential prerequisite for the efficient allocation <strong>of</strong> natural<br />

resources (Heal et al. 2005). Estimating the economic value <strong>of</strong> ecosystem services can play an<br />

important role in conservation planning and ecosystem-based management (Plummer 2009; Stenger et<br />

al. 2009). Conversely, a lack <strong>of</strong> economic valuation can potentially conceal the importance <strong>of</strong> such<br />

resources. However, economic valuation <strong>of</strong> ecosystem services requires up-to-date and reliable<br />

information and considerably better understanding <strong>of</strong> the landscapes that provide such services (Troy<br />

& Wilson 2006).<br />

We present a framework for classifying and mapping ecosystem services for a production landscape<br />

and assess the value <strong>of</strong> both market and non-market goods and services in a Geographic Information<br />

1<br />

Department <strong>of</strong> Forest and Ecosystem Science, University <strong>of</strong> Melbourne, 500 Yarra Boulevard, Richmond, Victoria, 3121,<br />

<strong>Australia</strong>. Email: h.baral@pgrad.unimelb.edu.au<br />

2<br />

Department <strong>of</strong> Forest and Ecosystem Science, University <strong>of</strong> Melbourne, Water Street, Creswick, Victoria, 3363, <strong>Australia</strong>.<br />

3<br />

Department <strong>of</strong> Resource Management and Geography, University <strong>of</strong> Melbourne, 500 Yarra Boulevard, Richmond, Victoria,<br />

3121, <strong>Australia</strong>.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 65<br />

System (GIS). This provides a baseline estimate <strong>of</strong> the ecosystem services value (ESV) <strong>of</strong> selected<br />

ecosystem services, such as timber, carbon, water regulation and biodiversity.<br />

CLASSIFICATION, MAPPING AND ECONOMIC VALUATION<br />

GIS as an ecosystem services mapping tool<br />

Ecosystem services are not homogenous across the landscape and their supply changes through time<br />

(Fisher et al. 2009). For this reason, ecosystem services are best expressed and most easily studied at<br />

particular spatial and temporal scales (MEA 2003). GIS provides land managers with a tool for<br />

quantifying and mapping the values <strong>of</strong> multiple ecosystem services across landscapes for improved<br />

resource planning and decision planning. Moreover, the valuation <strong>of</strong> ecosystem services, and<br />

understanding <strong>of</strong> how these resources interact, also provides an indication <strong>of</strong> trade <strong>of</strong>fs and synergies.<br />

This paper will demonstrate how GIS can be used to analyse disparate data to generate spatially<br />

explicit results within a production landscape.<br />

Economic valuation <strong>of</strong> ecosystem services<br />

The knowledge and recognition <strong>of</strong> the importance <strong>of</strong> ecosystem services in providing benefits to<br />

society, and as the basis for the sustainable functioning <strong>of</strong> natural systems, have grown in recent years.<br />

Despite their importance, ecosystem services are yet to be incorporated into decision making (Chan et<br />

al. 2006). Many <strong>of</strong> these goods and services are not valued on markets and there is a gap between<br />

market valuation and the economic value <strong>of</strong> many ecosystem services (Stenger et al. 2009).<br />

Ecosystem goods and services can be divided into two broad categories: (i) the provision <strong>of</strong> direct<br />

market goods or services such as timber, pulp and carbon; and (ii) the provision <strong>of</strong> non-market goods<br />

or services, which include biodiversity and habitat for plant and animal life (Wilson & Hoehn 2006;<br />

Mertz et al. 2007). Economic valuation <strong>of</strong> the former is straightforward while the latter is more<br />

complicated and controversial (Wilson & Hoehn 2006; Stenger et al. 2009). However, resource<br />

managers and policy analysts involved in protecting and managing natural resources must make<br />

decisions which involve multiple trade-<strong>of</strong>fs in allocating resources. These are mainly economic<br />

decisions and based either explicitly or implicitly on the value society places on services. To this end,<br />

economic valuation provides a useful tool for justifying set priorities and programs, policies, or actions<br />

that protect or restore ecosystem and associated services.<br />

Ecosystem valuation techniques<br />

Market price method: This method estimates the economic value <strong>of</strong> ecosystem goods or services that<br />

are bought and sold in commercial markets. This method can be used to value changes in either the<br />

quantity or quality <strong>of</strong> a good or service (Wilson & Hoehn 2006). In this study, this method was used to<br />

estimate the economic values <strong>of</strong> timber/pulp and carbon sequestration.<br />

Non-market valuation methods: Values for many ecosystem goods and services are not readily<br />

captured in market transactions, and thus require non-market valuation methods, such as travel cost<br />

method, hedonic approach, and contingent valuation (Wilson & Hoehn 2006; Stenger et al. 2009).<br />

Environmental value transfer: Value transfer is an accepted economic methodology which obtains<br />

an estimate for the economic value <strong>of</strong> non-market goods or services through work conducted at<br />

another site or group <strong>of</strong> sites (Troy & Wilson 2006). The ‘transfer’ itself refers to the application <strong>of</strong><br />

economic values and other information from the original ‘study site’ to a ‘policy site’. This technique<br />

was used in this study to estimate the economic value <strong>of</strong> biodiversity and associated services.<br />

METHODOLOGY<br />

Study area<br />

The study location lies within the Green Triangle region <strong>of</strong> southern <strong>Australia</strong>, some 300 km west <strong>of</strong><br />

Melbourne, Victoria, and covers approximately 305 km 2 (Figure 1). The area is known as subcatchment<br />

G8 <strong>of</strong> the Glenelg Hopkins Catchment. This area was selected because <strong>of</strong> its concentration<br />

<strong>of</strong> commercially valuable hardwood and s<strong>of</strong>twood plantations where the focus is on production<br />

forestry with the provision <strong>of</strong> a variety <strong>of</strong> environmental services.<br />

The area has a variable annual rainfall <strong>of</strong> approximately 800 mm and mean annual temperature <strong>of</strong> 8 ºC<br />

(min) to 19 ºC (max). The altitudinal gradient ranges from 8–198 m above sea level. The major land


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 66<br />

use in G8 is production forestry including both hardwood (Eucalyptus globulus) and s<strong>of</strong>twood (Pinus<br />

radiata) plantations. (Refer .Figure 3).There are also some large areas <strong>of</strong> uncleared crown land<br />

(reserved forest) and limited grazing <strong>of</strong> sheep and cattle. Few social networks and services are<br />

available in sub-catchment G8 with Digby the only township, with a population <strong>of</strong> about 50 people.<br />

Community based groups include Rifle Downs Landcare, Miakite-Grassdale Tree Group, and the<br />

Smokey River Land Management Group.<br />

Fig. 1 Location <strong>of</strong> the study area – sub-catchment G8, Lower Glenelg Basin within The Green<br />

Triangle region <strong>of</strong> south-eastern <strong>Australia</strong>. Shading represents areas <strong>of</strong> high to low relief<br />

according to a digital terrain model.<br />

Methods<br />

The approach is based on a combination <strong>of</strong> market-based and non-market valuation techniques and has<br />

six key steps: (i) define the study area, identify the data requirement and compile data; (ii) develop<br />

land cover typology; (iii) select ecosystem services for economic valuation; (iv) map land cover and<br />

associated ecosystem services; (v) quantify goods and services and calculate economic value; and (vi)<br />

assign economic value in GIS and calculate total ecosystem service value (Figure 2).<br />

Fig. 2 Conceptual framework for the method <strong>of</strong> classification, mapping and economic<br />

valuation <strong>of</strong> ecosystem services used in this study.<br />

The first step was to identify both spatial and attribute data requirements, and then compile the data<br />

from a variety <strong>of</strong> sources. Datasets included: (i) plantation GIS data, which includes location, plant<br />

year, species, previous land use; (ii) ecological vegetation class (EVC); (iii) location <strong>of</strong> threatened


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 67<br />

flora and fauna; (iv) a digital terrain model; and (v) topographic data such as contours, roads and<br />

watercourses. EVCs are systems <strong>of</strong> classifying native vegetation types based on differences in broad<br />

landscape features and environmental regimes (DNRE, 1997). EVC and plantation statistics provide<br />

an important overview <strong>of</strong> land cover types <strong>of</strong> the study area as vegetation is a highly visual component<br />

<strong>of</strong> the landscape and is an important part <strong>of</strong> timber resources as well as flora and fauna habitats. This<br />

study used the EVC dataset developed by the Victorian Department <strong>of</strong> Sustainability and<br />

Environment, and plantation GIS data available from the forestry companies represented in the study<br />

area (ITC, Timbercorp, Hancock Victorian Plantations, Wollybutt, Midway Afforestation, Great<br />

Southern Plantations, Green Triangle Plantation Forests, and Auspine).<br />

The second step entails development <strong>of</strong> land cover typology for the study area and starts with a<br />

preliminary analysis <strong>of</strong> available GIS data and EVCs. Land cover types from EVC and plantation data<br />

present in the study area are listed, a range <strong>of</strong> ecosystem services produced from each land cover type<br />

are identified and possible means <strong>of</strong> quantification and economic valuation techniques are reviewed.<br />

Ecosystem services identified in step two are screened in the third step in order to determine which are<br />

most relevant to the planning and decision making, and to set priorities for further assessment and<br />

valuation. Given the limited time and resources, it was not possible to assess in detail all the<br />

ecosystem services and assign economic value.<br />

In the fourth step, a land cover map is created using analysis tools in GIS which combine input layers<br />

from the diverse data sources to produce the final land cover map. Although significant variation<br />

within each land cover type may exist due to various factors, such as slope, aspect, and soil type, this<br />

level <strong>of</strong> detail was not measured in this preliminary study. Once each land cover type is finalised, then<br />

goods and services provided by each land cover type is quantified and economic value is assigned<br />

(Step 5). The economic value <strong>of</strong> ecosystem services was summed and cross-tabulated by each good<br />

and service and land cover type. Finally, a total spatial economic value map was produced (Step 6) by<br />

transferring ecosystem services value calculated in Step 5 using an ArcGIS 9.2 (from ESRI Inc.). The<br />

equation below illustrates the calculation <strong>of</strong> total ecosystem services value (ESV) in GIS environment<br />

(Troy and Wilson 2006).<br />

n<br />

V ( ES ) = ∑ A( LU )· V ( ES<br />

i i<br />

k = 1<br />

)<br />

ki ……(i)<br />

Where A(LUi) = area <strong>of</strong> land use/cover type (i), and V(ESki) = annual value per unit area <strong>of</strong> ecosystem<br />

service type (k) generated by each land cover type (i)<br />

Ecosystem services identified and valuation for this study<br />

Although there are large numbers <strong>of</strong> ecosystem services produced by the sub-catchment G8 (timber or<br />

fibre, carbon sequestration and climate regulation, water regulation, biodiversity and associated<br />

services, erosion control, salinity mitigation, tourism, recreation and cultural value), this study only<br />

focused on four goods and services – timber, carbon, water, and biodiversity and associated services.<br />

Timber: Plantation GIS data available from various forestry companies were used to identify and map<br />

the areas for timber production. Timber yield per hectare was calculated using a generic growth model<br />

(Fig. 4) available from Farm Forestry Toolbox (FFT) (Version 5.0, Private Forests Tasmania). Current<br />

timber value per unit area was estimated by multiplying the current stumpage price for hardwood and<br />

s<strong>of</strong>twood logs and for pulpwood. Current market prices (2009 as a base year) were reviewed from<br />

Product Disclosure Statements (PDS) issued by various Managed Forestry Investment Scheme (MIS)<br />

forestry projects. The accuracy <strong>of</strong> growth and yield calculated from FFT was not validated with actual<br />

growth and yield data due to high level <strong>of</strong> confidentiality <strong>of</strong> inventory data.<br />

Carbon: A generic Carbon Sequestration Predictor (Version 3.1, New South Wales Department <strong>of</strong><br />

Primary Industries) was used to estimate average carbon sequestration per hectare per year. Plantation<br />

GIS data were used to classify sites and determine the area and age <strong>of</strong> plantation. The average market<br />

value <strong>of</strong> current CO2 trading in <strong>Australia</strong>n and international markets was used to calculate carbon<br />

value.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 68<br />

Biodiversity: Classifying, measuring and valuing biodiversity was a challenging issue for a number <strong>of</strong><br />

reasons – (i) it was complicated by the wide spectrum <strong>of</strong> biotic scales at which biodiversity operates,<br />

ranging from the molecular specific to the ecosystem level; (ii) even for a given level <strong>of</strong> diversity, e.g.,<br />

presence <strong>of</strong> certain threatened flora or fauna, there is no well established and agreed means for<br />

defining, measuring and valuing biodiversity; and (iii) a number <strong>of</strong> different indicators have been<br />

proposed which neither provide consistent nor comparable results on which to base general<br />

interpretations (Bene & Doyen 2008).<br />

As detailed mapping and valuation <strong>of</strong> biodiversity was not within the scope <strong>of</strong> this study, we were left<br />

with two alternatives – (i) transfer biodiversity values from other studies; or (ii) use an approximate<br />

dollar value employed by the <strong>Australia</strong>n and various State government initiatives to conserve native<br />

vegetation on private land, such as ‘bush tender’ and ‘habitat hectare’ in Victoria, ‘biodiversity<br />

banking’ in New South Wales, ‘NatureAssist’ in Queensland, and ‘levy/biodiversity <strong>of</strong>fset package’ in<br />

South <strong>Australia</strong>. The second option was used due to lack <strong>of</strong> comparable studies to transfer appropriate<br />

values for this site. The estimated value was transferred to each land cover type <strong>of</strong> native remnant<br />

vegetation.<br />

Water: Estimating accurate water use by plantations and various vegetation types is also a difficult<br />

issue. It is generally agreed that forest plantations use such a large quantity <strong>of</strong> water that downstream<br />

flow may be affected. Following harvest, water quantity increases but the quality may decrease where<br />

sediment discharge in the overland flow reaches stream channels. 3-PG, a simple, process-based forest<br />

growth model, developed by Landsberg and Waring (1997), was used estimate annual plantation water<br />

use per unit area. The average water harvesting charge per megalitre available from the Independent<br />

Pricing and Regulatory Tribunal (IPART) NSW was used to estimate the cost <strong>of</strong> water utilised by<br />

forest plantations.<br />

ESV transfer<br />

The ecosystem service value <strong>of</strong> selected ecosystem services, estimated by market and non-market<br />

valuation techniques,, was transferred to the GIS and the total economic value <strong>of</strong> selected ecosystem<br />

services was calculated using equation (i) above.<br />

RESULTS<br />

The landcover map <strong>of</strong> sub-catchment G8 is dominated by ‘heathy woodlands’ (27.4 percent) followed<br />

by ‘plain woodlands’ (19.3 percent). Hardwood and s<strong>of</strong>twood plantations cover approximately 18<br />

percent <strong>of</strong> the sub-catchment (Table 1 and Figure 3).<br />

Table 1. Land cover typologies (and proportion <strong>of</strong> area occupied) for sub-catchment G8, Lower<br />

Glenelg basin<br />

• Heathy woodlands (27.4%)<br />

• Plain woodlands (19.3%)<br />

• Agriculture/pasture (18%)<br />

• E. globulus plantation (10.2%)<br />

• Herb-rich woodlands (8.8%)<br />

• P. radiata plantation (7.3%)<br />

• Lowland forests (3.9%)<br />

• Riparian or swampy scrubs (2.4%)<br />

• Heathlands (1.9%)<br />

• Riverine grassy woodlands (0.5%)<br />

• Wetlands (0.3%)<br />

The values <strong>of</strong> the selected ecosystem services – timber, carbon, biodiversity and water – have<br />

distinctly different spatial distributions, although some areas are <strong>of</strong> high value for multiple services<br />

while others are low value (Fig. 5). For example, the preliminary assessment and economic valuation<br />

<strong>of</strong> these ecosystem services indicate plantations represent the highest economic value per unit area.<br />

This is mainly due to the relatively detailed valuation <strong>of</strong> consumptive services available from<br />

commercial plantation areas. The areas with higher timber yields also produce more carbon which<br />

further increases the economic value per unit area.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 69<br />

Fig. 3 Land cover map <strong>of</strong> sub-catchment G8, Lower Glenelg Basin.<br />

Fig. 4. Site productivity (top) and mean annual increment (bottom) for Pinus radiata (4a, b) and<br />

Eucalyptus globulus (4c, d) used to estimate timber yield ha -1 ,(where, MDH = mean<br />

dominant height, MDDob = mean dominant diameter over bark and MAI = mean<br />

annual increment.)


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 70<br />

Fig. 5 Annual flow <strong>of</strong> selected Ecosystem services <strong>of</strong> sub-catchment G8, Lower Glenelg Basin,<br />

the large light grey area indicates ESV less then $1350 ha -1 yr -1 or no data available at<br />

this stage.<br />

The value <strong>of</strong> native vegetation is relatively low due to our use <strong>of</strong> the conservation value based on<br />

various government initiatives to conserve native vegetation in private land (Section). This does not<br />

reflect the true value <strong>of</strong> all native vegetation and the value <strong>of</strong> a number <strong>of</strong> other services such as<br />

erosion control and water regulation are not assessed in this study.<br />

The land cover map (Fig 3) was based on the most recent, updated plantation and native vegetation<br />

datasets. However, improved estimates for growth and yield, together with improved conservation<br />

values, will permit further refinement <strong>of</strong> our valuation <strong>of</strong> ecosystem services for sub-catchment G8.<br />

DISCUSSION<br />

The quantification and valuation <strong>of</strong> timber and carbon were relatively straightforward due to readily<br />

available datasets. Similarly, the accuracy level is relatively high because prices are well-defined, and<br />

real data and accepted economic techniques were used. However, some issues, such as the true<br />

economic value <strong>of</strong> goods or services, may not be fully reflected in market transactions which are<br />

subject to market imperfections, seasonal variations and unforeseen impacts on markets.<br />

Other services which are not readily bought and sold in the market place, such as biodiversity and<br />

water quantity and quality, are difficult to quantify and value. The complexity and uncertainty<br />

underlying the functioning <strong>of</strong> biodiversity contribute to the difficulty in assessing such services (Bene<br />

& Doyen 2008). Although it is difficult to put an accurate price tag on such services, it is important to<br />

estimate the value <strong>of</strong> these ecosystem services because <strong>of</strong> the growing environmental market.<br />

While the current global financial downturn threatens markets for timber resources, eco-products are<br />

gaining in market popularity due to perceived environmental challenges, such as drought, declining<br />

supplies <strong>of</strong> drinking water, and loss <strong>of</strong> biodiversity (Metz et al. 2007). The economic valuation <strong>of</strong><br />

ecosystem services provides better understanding <strong>of</strong> these forms <strong>of</strong> natural capital, and this is critical<br />

to our ability to plan and manage it over long term.<br />

The critical underlying assumption <strong>of</strong> the value transfer approach is that the economic value <strong>of</strong><br />

ecosystem goods or services at the study site can be inferred with sufficient accuracy from the analysis<br />

<strong>of</strong> existing valuation studies. Clearly, as the level <strong>of</strong> information increases within the source literature,<br />

the accuracy <strong>of</strong> the value transfer improves. For this reason we are currently compiling recent forest<br />

ecosystem valuation studies in <strong>Australia</strong> as most existing datasets are outdated.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 71<br />

We are also in the process <strong>of</strong> accessing CABALA (CArbon BALAnce), a more advanced model for<br />

predicting forest growth and carbon sequestration in plantations and managed forests developed by<br />

CSIRO, and anticipate that tree growth and carbon estimation will improve as a result.<br />

CONCLUSIONS<br />

This study presented a simple GIS-based process <strong>of</strong> classification, mapping and valuation <strong>of</strong><br />

ecosystem goods and services at the landscape scale. Rather than a single methodological approach,<br />

this study used a number <strong>of</strong> tools to estimate ESV using market and non-market based valuation<br />

approaches in a spatially explicit manner. Although the study is still in its early stages and does not<br />

account for all non-market ecosystem services, it demonstrates an important contribution <strong>of</strong> such<br />

services at a landscape level. The approach and methodology can be used for more precise future<br />

mapping for all ecosystem goods and services, including agriculture, pasture and other forest goods<br />

and services.<br />

ACKNOWLEDGEMENTS<br />

The research was a part <strong>of</strong> PhD study (H. Baral) funded by The University <strong>of</strong> Melbourne and The<br />

CRC for Forestry. The GIS data were provided mostly by the Victorian Department <strong>of</strong> Sustainability<br />

and Environment, while plantation GIS data were supplied by forestry companies through the South<br />

East Resource Information Centre. Additional geospatial data were available from the Bureau <strong>of</strong><br />

Meteorology, Geosciences <strong>Australia</strong>, and the Glenelg Hopkins Catchment Management Authority.<br />

REFERENCES<br />

Bene, C & Doyen, L 2008, 'Contribution values <strong>of</strong> biodiversity to ecosystem performances: A viability<br />

perspective', Ecological Economics, 68 (1-2), 14-23.<br />

Boyd, J & Banzhaf, S 2007, 'What are ecosystem services? The need for standardized environmental accounting<br />

units', Ecological Economics, 63 (2-3), 616-626.<br />

Chan, KMA, Shaw, M, Cameron, DR, Underwood, E & Daily, G 2006, 'Conservation Planning for Ecosystem<br />

Services', Plos Biology, 4 (11), e379.<br />

Costanza, R, d'Arge, R, de Groot, R, Farber, S, Grasso, M, Hannon, B, Limburg, K, Naeem, S, O'Neill, RV,<br />

Paruelo, J, Raskin, RG, Sutton, P & van den Belt, M 1998, 'The value <strong>of</strong> ecosystem services: putting the<br />

issues in perspective', Ecological Economics, 25 (1), 67-72.<br />

de Groot, RS, Wilson, MA & Boumans, RMJ 2002, 'A typology for the classification, description and valuation<br />

<strong>of</strong> ecosystem functions, goods and services', Ecological Economics, 41 (3), 393-408.<br />

Fisher, B & Kerry, T 2008, 'Ecosystem services: Classification for valuation', Biological Conservation, 141 (5),<br />

1167-1169.<br />

Fisher, B, Turner, RK & Morling, P 2009, 'Defining and classifying ecosystem services for decision making',<br />

Ecological Economics, 68 (3), 643-653.<br />

Heal, GM, Barbier, EB & Boyle, KJ 2005, Valuing Ecosystem Services: Toward Better Environmental Decision-<br />

Making, Washington, DC.<br />

MEA 2003, Ecosystems and Human Well-being: A framework for Assessment, Island Press, Washington DC.<br />

Mertz, O, Ravnborg, HM, Lovei, G, Nielsen, I & Konijendijk, CC 2007, 'Ecosystem services and biodiversity in<br />

developing countries', Biodiversity and Conservation, 16 (10), 2729-2737.<br />

Metz, B, Davidson, O, Bosch, P, Dave, R, and Meyer, L 2007, Climate Change 2007: Mitigation <strong>of</strong> Climate<br />

Change, Cambridge University Press, London.<br />

Plummer, ML 2009, 'Assessing benefit transfer for the valuation <strong>of</strong> ecosystem services', Frontiers in Ecology<br />

and the Environment, 7 (1), 38-45.<br />

Stenger, A, Harou, P & Navrud, S 2009, 'Valuing environmental goods and services derived from the forests',<br />

Journal <strong>of</strong> Forest Economics, 15 (1), 1-14.<br />

The former DNRE,1997, Victoria’s Biodiversity: Directions in Management, Department <strong>of</strong> Natural Resources<br />

and Environment, East Melbourne.<br />

Troy, A & Wilson, MA 2006, 'Mapping ecosystem services: Practical challenges and opportunities in linking<br />

GIS and value transfer', Ecological Economics, 60 (2), 435-449.<br />

Wallace, K 2008, 'Ecosystem services: Multiple classifications or confusion?', Biological Conservation, 141 (2),<br />

353-354.<br />

Wilson, MA & Hoehn, JP 2006, 'Valuing environmental goods and services using benefit transfer: The state-<strong>of</strong>the<br />

art and science', Ecological Economics, 60 (2), 335-342.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 72<br />

SELECTING SPECIES FOR CARBON SEQUESTRATION UNDER<br />

CLIMATE CHANGE SCENARIOS IN SUBTROPICAL QUEENSLAND<br />

David Lee 1, 2 , John Huth 1 , David Osborne 1 , Bruce Hogg 1<br />

ABSTRACT<br />

Optimal matching <strong>of</strong> species to sites is required for a sustainable hardwood plantation<br />

industry in the subtropics. This paper reports on the performance and adaptation <strong>of</strong> 60<br />

taxa (species, provenances and hybrids) across two rainfall zones and a range <strong>of</strong> soil types<br />

in southern Queensland. Specifically, taxa performance is compared across five<br />

replicated taxa-site matching trials at age six years. Three trials are in a 1000 mm rainfall<br />

zone <strong>of</strong> the Wide Bay near Miriam Vale and two in a drier (c. 750 mm) rainfall zone near<br />

Kingaroy in the South Burnett.<br />

In the higher rainfall zone, the taxa with the fastest growth in the three trials at age six<br />

years are Corymbia citriodora subsp. variegata Woondum provenance ranked 1 st 6 th and<br />

5 th , E. longirostrata Coominglah provenance ranked 3 rd 2 nd and 3 rd respectively; and two<br />

sources <strong>of</strong> E. grandis Copperlode provenance (ranked 4 th and 1 st ) and SAPPI seed orchard<br />

(ranked 6 th and 4 th ) planted in only two <strong>of</strong> the three trials. Similarly in the lower rainfall<br />

zone, E. grandis and its hybrids appear promising from the growth data; however, this<br />

excellent early growth has not continued in either rainfall zones, with these taxa now<br />

(over eight years old) showing signs <strong>of</strong> stress and deaths. Based on trial results in these<br />

two rainfall zones, the taxa that appear the most promising for sustainable plantation<br />

development with high average annual volume indexes and low incidence <strong>of</strong> borer attack<br />

is Corymbia citriodora subspecies variegata (6.7 m³/ha). E. grandis and E. longirostrata<br />

both have better average annual volume indexes (8.2 m³/ha and 7.4 m³/ha respectively)<br />

but are highly susceptible to borer attack. The current and long term productivity and<br />

sustainability <strong>of</strong> plantation forestry in these rainfall zones is discussed. Further, the<br />

implications <strong>of</strong> predicted climate change (particularly reduced rainfall) on growing trees<br />

for fibre production and carbon sequestration are explored.<br />

INTRODUCTION<br />

The expansion <strong>of</strong> commercial hardwood plantations in Queensland has been dramatic over the past<br />

decade from 400 ha in 1997 to 49400 ha in 2007 (ABARE, 2008). This has come at a time when there<br />

has been an escalation in the price <strong>of</strong> land in the higher rainfall zones (>1000 mm) and therefore a<br />

reduction <strong>of</strong> the amount <strong>of</strong> land available for economically viable plantation establishment. This has<br />

resulted in hardwood plantation growers expanding into lower rainfall zones which have traditionally<br />

been regarded as marginal for production forestry. The problem faced by the plantation growers is<br />

that, while there is potentially a large number <strong>of</strong> species and inter-specific hybrids that may grow in<br />

these regions, there is little long-term growth, carbon sequestration, wood quality and pest tolerance<br />

data available on which to base species selection and estimate potential plantation productivity and<br />

pr<strong>of</strong>itability. To address these problems 150 taxa (species, provenance and hybrid) trials have been<br />

established over the past 12 years on a range <strong>of</strong> site types across Queensland and northern New South<br />

Wales.<br />

This paper presents the results from five large trials (11952 trees) in south-east Queensland at age six<br />

years. Three trials are in a moderately high rainfall zone (1000–1200 mm) in the Wide Bay region<br />

near Miriam Vale, and two are in the lower rainfall zone (750–800 mm) in the South Burnett region,<br />

near Kingaroy (Figure 1).<br />

1<br />

Horticulture and Forestry Science, Queensland Primary Industries & Fisheries, LB 16 Fraser Road, Gympie Qld 4570,<br />

<strong>Australia</strong>.<br />

2<br />

Faculty <strong>of</strong> Science, Health and Education, University <strong>of</strong> the Sunshine Coast, Maroochydore DC Qld 4558, <strong>Australia</strong>. Email:<br />

dlee@usc.edu.au.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 73<br />

MATERIAL AND METHODS<br />

Study sites<br />

The five trials were established over two years. Two trials in the Wide Bay were planted in 2000<br />

(trials 506a and 506b) and a third (trial 506c) was planted in 2001. Both trials in the South Burnett<br />

(523c and 523d) were planted in 2000 (Table 1, Figure 1). All trials were planted on sites that were<br />

being used by commercial hardwood plantation growers. Site details are presented in Table 1. Before<br />

planting each trial site was deep ripped, mounded and sprayed with pre-plant herbicide. After<br />

planting, weed free conditions were maintained in the planting row until the end <strong>of</strong> the following wet<br />

season.<br />

Table 1. Site descriptions for the five taxa trials.<br />

Trial <strong>Australia</strong>n Soil<br />

Classification 1<br />

Wide Bay trials (higher rainfall)<br />

506a Red Chromosol to Yellow<br />

Dermosol<br />

506b Red Chromosol to Yellow<br />

Dermosol<br />

506c Yellow Chromosol and<br />

Black Vertosol<br />

Latitude<br />

(°S)<br />

Longitude<br />

(°E)<br />

Mean annual<br />

rainfall (mm)<br />

Elevation<br />

(m asl)<br />

Average annual<br />

rainfall during<br />

trial (mm) 2<br />

24.60 151.64 1033 65 877 low<br />

Frost<br />

risk<br />

24.60 151.64 1033 55 877 medium<br />

24 38 151 52 1138 90 868 medium<br />

South Burnett trials (lower rainfall)<br />

523c Red Ferrosol 26.36 151.72 786 440 598 high<br />

523.d Red Dermosol 26.55 151.73 760 480 594 high<br />

1 Isbell (1996)<br />

2 Interpolated daily observations from the SILO database see: http://www.longpaddock.qld.gov.au/silo/ (Jeffrey et al., 2001)<br />

Fertiliser was applied to all trials shortly after planting. Trials in the Wide Bay were fertilised with a<br />

low dose <strong>of</strong> fertiliser (2.1 kg/ha <strong>of</strong> nitrogen, 3.7 kg/ha <strong>of</strong> phosphorus, 2 kg/ha <strong>of</strong> potassium and 3.8<br />

kg/ha <strong>of</strong> sulphur, plus trace elements). The very low levels <strong>of</strong> elements were used by the plantation<br />

owner as a base dressing as these trials were established on ex pasture sites that had a high number <strong>of</strong><br />

fertiliser applications over past years (exact details are not known). The trials in the South Burnett<br />

were given a higher dose <strong>of</strong> fertiliser (72.6 kg/ha <strong>of</strong> nitrogen, 89.5 kg/ha <strong>of</strong> phosphorus, 58.6 kg/ha <strong>of</strong><br />

potassium and 8.6 kg/ha <strong>of</strong> sulphur plus trace elements.<br />

The rainfall over the period since planting has been much lower than the long-term average with the<br />

Wide Bay trials receiving 84 per cent and the South Burnett trials 76 per cent respectively (Table 1).<br />

Experimental design<br />

Each trial was established as a randomised incomplete block design with three replications. In each<br />

trial, one or more provenances <strong>of</strong> best-bet species and or hybrids (Eucalyptus, Corymbia) as well as<br />

traditional s<strong>of</strong>t wood plantation species <strong>of</strong> south-east Queensland (Araucaria and Pinus) were tested.<br />

Allocation <strong>of</strong> taxa to trials was based on expert opinion <strong>of</strong> the potential <strong>of</strong> the taxa as well as plant<br />

availability. Details <strong>of</strong> the 60 taxa planted in the trials and their origins are given in Table 2; this table<br />

also details the abbreviations used to describe the taxa throughout this paper (e.g. acmen -neer). Six<br />

taxa: ccc -glad, ccv -brooy, ccv –woon, g × c -dend, g × u -csir and longi -coom were planted across<br />

all trial sites. Plot sizes varied with site due to seedling and land availability. All measure plots had<br />

isolations <strong>of</strong> the same taxa planted around the outside <strong>of</strong> the plot. Trials 506a and 506b were planted<br />

approximately 800 m apart, on the same property with 32-tree measure plots (4 rows × 8 trees). Trial<br />

506c was established with 40-tree measure plots (4 rows × 10 trees) and trials 523c and 523d with 20tree<br />

measure plots (4 rows × 5 trees). Trials in the 506 series were planted at 794 stems per hectare<br />

whereas the two trials in the 523 series were planted at 1111 stems per hectare. Trials were un-thinned<br />

at the time <strong>of</strong> the latest measure.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 74<br />

Measurement and assessments<br />

Survival was assessed and total height and diameter at breast height over bark were measured for all<br />

surviving trees in each trial at age six years. As no specific volume equations were available for these<br />

species a volume index (VI) was calculated using the following formula:<br />

VI (m³) = [1/3 tree height (m) × basal area at 1.3 m (m 2 )]<br />

In trials 506a, 506b and 506c the incidence <strong>of</strong> borer attack was also assessed at age six years (presence<br />

or absence <strong>of</strong> longicorn beetles Phoracantha spp. or giant wood moths Endoxyla cinerea).<br />

Figure 1. Location <strong>of</strong> trial sites in the Wide Bay (series<br />

506) and Burnett (series 523) regions <strong>of</strong> subtropical<br />

Queensland referred to in the paper.<br />

Statistical analysis<br />

Analysis <strong>of</strong> variance <strong>of</strong> the data, for each<br />

trial, was carried out using the Genstat<br />

Version 9.2 s<strong>of</strong>tware package with a<br />

randomised complete block design<br />

model:<br />

Yijk = μ + Ti + Rj + TRij+ Eijk<br />

where, Yij = the observation on the i th<br />

taxa in the j th replicate; μ = overall mean;<br />

Ti = the effect <strong>of</strong> the i th taxa; Rj = the<br />

effect if the j th replicate, TRij is the<br />

interaction between the i th taxa and j th<br />

replicate and Eijk the random error<br />

associated <strong>of</strong> the ijk th observation. This<br />

model was considered appropriate as<br />

most taxa (species and hybrids) did not<br />

have nested taxonomic groupings<br />

(multiple provenances or seedlots).<br />

Arcsine transformation <strong>of</strong> binary data<br />

(survival and borer incidence) was<br />

performed, but did not alter the result <strong>of</strong><br />

the analysis. Post hoc Fischer’s<br />

Protected least significant difference test<br />

was applied if significant differences<br />

were detected by ANOVA. Means are<br />

reported, and treatment differences or<br />

interactions were regarded as significant<br />

at P


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 75<br />

significant attack (16 to 100 per cent; Table 3). E. grandis and its hybrids were heavily attacked by<br />

borer with incidence ranging from 44 to 100 per cent. E. urophylla and its hybrids were also highly<br />

susceptible to borer attack with incidence ranging from 89 to 100 per cent. Mean survival <strong>of</strong> the six<br />

taxa represented in all trials was similar, ranging from 79 to 84 per cent. Across the trials survival <strong>of</strong><br />

the taxa ranged from 6 to 98 per cent.<br />

The average VI <strong>of</strong> the six taxa planted across all the trials (ccc –glad, ccv –brooy, ccv –woon, g × c –<br />

dend, g × u –csir and longi –coom) was calculated to provide an indication <strong>of</strong> site potential. Based on<br />

this, trial 506a had the lowest site potential at age six (24.5 m 3 /ha) followed by 523d (26.0 m 3 /ha),<br />

506b (28.5 m 3 /ha), 506c (48.4 m 3 /ha) with trial 523c exhibiting the best site potential at age six (55.3<br />

m 3 /ha; Table 3).<br />

Performance <strong>of</strong> taxa in trials in the high rainfall region<br />

Volume index in trial 506a at age six years ranged from 2.8 m³/ha for uro -mt.egon to 28.5 m³/ha for<br />

ccv -woon, which along with seven other taxa formed a group that had significantly better growth than<br />

the other 21 taxa in the trial (Table 3). Survival ranged from 19 per cent for uro -mt egon to 92 per<br />

cent for g × c -dend. Incidence <strong>of</strong> borer attack ranged from 0 per cent for ccv -woon, ccv -brooy, cloez<br />

-home, cloez -mungy and sphaero -bla to 100 per cent for uro -mt ergon. Incidence <strong>of</strong> borers in E.<br />

grandis and E. grandis hybrids ranged from 50 to 97 per cent while that <strong>of</strong> E. longirostrata ranged 16<br />

to 41 per cent across the three trials.<br />

Trial 506b was on the same property as trial 506a but planted in a lower position in the landscape<br />

(creek flats) 800 m away. In this trial at age six years VI ranged from 0.1 m³/ha for sphaero -bla to<br />

45.6 m³/ha for grandis -cop, which along with longi -coom and g × r -csir had significantly larger<br />

volumes than the rest <strong>of</strong> the taxa in the trial. Survival ranged from 6 per cent for sphaero -bla to 91<br />

per cent for the g × r -csir seedlot. Incidence <strong>of</strong> borer attack ranged from 0 per cent for ccv -brooy to<br />

100 per cent for the g × r -csir seedlot. Borer attack in E. grandis and its hybrids ranged from 44 to<br />

100 per cent.<br />

Six year VI in trial 506c was generally higher than the other two trials in the higher rainfall zone (as<br />

indicated above in the section on site potential). In this trial age six years VI ranged from 9.0 m³/ha<br />

for sphaero -bla to over 90 m³/ha for grandis -koo and grandis -dpi, which had a significantly larger VI<br />

than the rest <strong>of</strong> the taxa in the trial. Survival ranged from 28 per cent for maid -bondi to 98 per cent<br />

for pch. Borer incidence in trial 506c was lower than the other two trials in the higher rainfall zone<br />

ranging from 0 per cent for 12 taxa including the six Corymbia taxa to just over 90 per cent for grandis<br />

-koo and g × u -csir.<br />

Performance <strong>of</strong> taxa in trials in the low rainfall region<br />

Thirty four taxa were planted in trial 523c. Age six VI ranged from 2.6 m³/ha for hoop to 84.3 m³/ha<br />

for saligna -clo. At age six years this latter clone, along with grandis-wed, grandis-dpi and g × r-csir<br />

seedlots (71.3, 72.1 and 74.4 m³/ha respectively) had significantly higher VI than the other taxa in the<br />

trial. The VI for E. grandis and its hybrids at age six years was in the range 41.2 to 74.4 m³/ha. The<br />

VI <strong>of</strong> the Corymbia species and hybrids in this trial ranged from 38.7 to 57.6 m³/ha. Survival ranged<br />

from 32 per cent for u hybrid -mix to 98 per cent for siderox -ural. Borer attack was not recorded at<br />

this site or the other site in the South Burnett but anecdotal observations indicated that borers are an<br />

issue for plantations in these regions.<br />

Trial 523d had 28 taxa planted. Six year VI ranged from 0.2 m³/ha for hoop to 33.0 m³/ha for maid -<br />

bolaro which did not have a significantly better VI than 12 other taxa in the trial. The VI <strong>of</strong> the<br />

Corymbia species ranged from 23.2 to 31.7 m³/ha and for E. grandis and its hybrids the range was<br />

17.5 to 32.0 m³/ha. Survival ranged from 43 per cent for hoop to 98 per cent for the g × c -dend<br />

clones.<br />

Trends across the five trials are given in the discussion.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 76<br />

Table 2. Details <strong>of</strong> seedlots and clones planted in the trials. If the seedlot number is followed by an ‘a’ it came from<br />

the <strong>Australia</strong>n Tree Seed Centre, otherwise it came from the Forestry Tree Seed Centre, Queensland. The<br />

details in the taxa column are used in Table 3 to describe the taxa. This table is sorted by the taxon<br />

abbreviation.<br />

Species / hybrid Provenance Taxon abbreviation<br />

Seedlot /<br />

Clones identity<br />

No. seed<br />

parents<br />

Latitude<br />

(°S)<br />

Longitude<br />

(°E)<br />

Eucalyptus acmenoides Neerdie acmen -neer 10804 / 15874 >7 25.98 152.77 70<br />

E. acmenoides Woolgoolga acmen -woolg 15534a 8 30.12 153.17 80<br />

E. argophloia Ballon argo -ballon 5520 18 26.50 150.92 350<br />

Corymbia hybrids QPIF 1 CP 2 families C hybrid – 12 – – –<br />

C. citriodora subsp. citriodora Gladstone ccc -glad 20016a 8 – – –<br />

C. citriodora subsp. citriodora Yeppoon ccc -yepp 4971 / 11246 >7 23.10 150.73 30<br />

C. citriodora subsp. variegata Brooyar ccv -brooy 10248 12 26.17 152.50 90<br />

C. citriodora subsp. variegata Woondum ccv -woon 5567 21 26.25 152.82 40<br />

C. henryi Lockyer ch -lock 10250 10 27.47 152.28 150<br />

C. henryi Nerang ch -ner 10257 11 27.98 153.32 100<br />

E. cloeziana Home cloez -home 4363 8 26.05 152.70 220<br />

E. cloeziana Mungy cloez -mungy 10823 20 25.28 151.35<br />

E. cloeziana South African CSO 3 cloez -saso – – – – –<br />

E. dunnii SAPPI 4 dunnii -sapp 10513 unknown – – –<br />

E. grandis Wedding Bells grandis -wed 10542 18 30.17 153.12 100<br />

E. grandis Copperlode grandis -cop 5970 28 16.97 145.67 425<br />

E. grandis Mixed selects grandis -dpi – 5 – – –<br />

E. grandis Koorlong grandis -koo 20261r unknown – – –<br />

E. grandis SAFCOL 5 grandis -saf 10520 >10 – – –<br />

E. grandis SAPPI grandis -sap 10519 unknown – – –<br />

E. grandis × E. camaldulensis CSIR 6 g × c -csir 11307 unknown – – –<br />

E. grandis ×E. camaldulensis Dendros clones g × c -dend – 9 clones – – –<br />

E. grandis ×E. pellita QPIF CP families g × p -mix 2000 g × p 15 – – –<br />

E. grandis ×E. resinifera CSIR g × r -csir 5991 / 11304 10 – – –<br />

E. grandis ×E. tereticornis CSIR g × t -csir 10502 / 11306 unknown – – –<br />

E. grandis ×E. tereticornis CSIR and DPI&F CP g × t -mix – 10 – – –<br />

E. grandis ×E. urophylla Dendros clones g × u -clone – 6 clones – – –<br />

E. grandis ×E. urophylla CSIR g × u -csir 11302 / 11301 unknown – – –<br />

E. grandis hybrid mix g × t, g × u and g × p hybrid -mix – unknown – – –<br />

Araucaria cunninghamii FPQ 7 CSO seed hoop – unknown – – –<br />

E. longirostrata Coominglah longi -coom 19312a / 11153 5 24.92 151.00 400<br />

E. macarthurii SAPPI mac -sappi 10854 unknown – – –<br />

E. globulus subsp. maidenii Bolaro maid -bolaro 18728a 78 35.67 150.03 380<br />

E. globulus subsp. maidenii Bondi maid -bondi 19454a 24 37.18 149.46 480<br />

E. moluccana Ravenshoe mol -ravens 5696 10 17.60 145.48 913<br />

Pinus caribaea var. hondurensis FPQ Kennedy CSO pch pch unknown – – –<br />

E. pellita Kuranda pell -kur 5081 15 16.75 145.57 440<br />

E. pellita Ellerbeck SSO 8 PNG-IJ pell -png 5952 26 – – –<br />

E. pilularis Beerburrum pil -beer 248 9 26.93 152.88 159<br />

E. pilularis Deongwar pil -deong 5504 15 27.31 152.25 550<br />

Pinus elliottii var. elliottii ×<br />

P. caribaea var. hondurensis FPQ best-F2 hybrid pine hybrid<br />

CSO families<br />

bulk unknown – – –<br />

E. reducta Ravenshoe reduct -rave 272 >10 17.55 145.45 980<br />

E. resinifera Beerburrum resin -beer 13981a 10 26.95 152.83 100<br />

E. resinifera Dorrigo resin -dorr 13976a 11 – – –<br />

E. resinifera Ravenshoe resin -ravens 314 unknown 17.55 145.45 980<br />

E. saligna subsp. saligna Dendrotech – clone 56 saligna -clo – 1 – – –<br />

E. siderophloia Mt Mee siderop -mtmee 17388a 4 27.02 152.70 350<br />

E. sideroxylon Uralla siderox -ural 20394a 9 30.63 151.50 895<br />

E. sphaerocarpa Blackdown sphaero -bla 391_392 – 23.83 149.08 760<br />

E. tereticornis Rundle teret -rund 10456 7 23.63 151.00 30<br />

E. tereticornis Zimbabwe SSO teret -zimb so 10872 unknown – – –<br />

E. urophylla hybrids u × p, u × d and u × g × c u hybrid -mix – unknown – – –<br />

E. urophylla Dongmen, China uro -dongmen 5979 10 – – –<br />

E. urophylla Sumatra provenances 9 uro -mix 1121 20 – – –<br />

E. urophylla Mt Egon Flores uro -mt egon 14531a 10 08.63 122.45 515<br />

E. urophylla Sumatra uro -sum 11152 unknown – – –<br />

E. urophylla Uhak Wetar uro -uhak 17836i 10 7.57 126.50<br />

E. urophylla × E. camaldulensis CSIR u × c 10508 – – – –<br />

E. urophylla × E. pellita SAPPI u × p -sappi 10551-1–7 5 – – –<br />

E. urophylla × E. tereticornis CSIR u × t -csir 10505 6 – – –<br />

Altitude<br />

(m asl)<br />

1 QPIF = Queensland Primary Industries and Fisheries, 2 CP = control pollination, 3 CSO = clonal seed orchard, 4 SAPPI = Sappi Forest<br />

Products, 5 SAFCOL = South African Forestry Company Ltd, 6 CSIR = Council for Scientific and Industrial Research, South Africa, 7 FPQ =<br />

Forestry Plantations Queensland, 8 SSO = seedling seed orchard, 9 Provenances are: Tele, Aek Nauli, Samosir and Harbinsaran.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 77<br />

DISCUSSION<br />

The large scale establishment <strong>of</strong> commercial hardwood plantations in Queensland’s subtropics only<br />

commenced in 1997. The selection <strong>of</strong> taxa to grow in these plantations was initially based on overseas<br />

experience and on utilisation <strong>of</strong> the species traditionally harvested from the native stands (Lee et al.,<br />

1999). For example, the establishment <strong>of</strong> E. grandis and E. grandis × E. urophylla and E. grandis × E.<br />

camaldulensis hybrids was based on data from South Africa and Brazil (e.g. Eldridge 1993; Verryn<br />

2000) and the planting <strong>of</strong> E. dunnii in China (e.g. Wang et al., 1999; Arnold and Luo, 2002). As a<br />

result some <strong>of</strong> these hardwood plantations have failed. The trial series discussed here were established<br />

to provide a better basis for taxa selection.<br />

The taxa planted in these trials showed significant differences in VI, survival and borer incidence at<br />

age six years. The average rainfalls during this growth period for the three Wide Bay trials were very<br />

similar ranging from 868 to 877 mm (which is approximately 15 per cent below the long-term<br />

average). Nevertheless, soil moisture conditions and frost risk would have varied with trial position in<br />

the landscapes and soil types thus providing variable conditions over which to assess taxa potential.<br />

Moreover, the droughty conditions provided test environments simulating those being predicted for<br />

Queensland’s subtropics in the future as a consequence <strong>of</strong> climate change. These conditions would<br />

have favoured taxa better able to tolerate drought conditions. In these three trials (506a, 506b, 506c)<br />

the taxa showing consistently high VI at age six were ccv -woon (ranked 1 st , 6 th and 5 th ) and longi -<br />

coom (ranked 3 rd , 2 nd and 3 rd ; Table 3). In 506a and 506b, grandis -cop ranked 4 th and 1 st and grandissap<br />

ranked 6 th and 4 th respectively. In these trials a bulk <strong>of</strong> the nine commercial g × c -dend clones<br />

ranked 8 th , 11 th and 7 th . The other commercial taxa planted in the region, dunnii -sap, ranked 5 th and<br />

10 th respectively in trials 506a and 506b. Other taxa that performed well on at least one site in this<br />

region are grandis -koo and grandis-dpi ranked 1 st and 2 nd in 506c, g × r -csir ranked 3 rd in 506b and<br />

ccv -brooy which ranked 2 nd and 8 th in 506a and 506c respectively. When borer attack is taken into<br />

account, the stand-out taxa across the three trials is ccv -woon. This result confirms other results (Lee<br />

et al., 2001; 2005) on this taxa which is generally only surpassed by Corymbia hybrids (Lee 2007, Lee<br />

et al., 2009) in terms <strong>of</strong> growth across a range <strong>of</strong> sites in Queensland’s subtropics.<br />

In the South Burnett the average annual rainfall during the period reported in this paper was between<br />

590 and 600 mm, 23 per cent below the long-term average. In these two trials (523c and 523d) the<br />

taxa with a consistent high volume index at age six years were: grandis -dpi (ranked 3 rd and 4 th ),<br />

grandis -wed (ranked 4 th and 7 th ), g × c -dend (ranked 8 th and 2 nd ), maid -bolaro (ranked 7 th and 1 st ),<br />

and ccv -woon (ranked 11 th and 5 th ). In 523c taxa that had high VI in only in this trial were saligna -<br />

clo (ranked 1 st ) g × r -csir (ranked 2 nd ), maid -bondi (ranked 5 th ) and grandis -koo (ranked 6 th ).<br />

Similarly, in 523d the taxa with high VI were: ch -lock (ranked 3 rd ), ch -ner (ranked 6 th ) and ccv -<br />

brooy (ranked 8 th ).<br />

Since our measurement in 2006 and 2007, E. grandis and its hybrids, E. dunnii, E. globulus subsp.<br />

maidenii and E. pellita have suffered considerable mortality (up to 25 per cent), possibly due to water<br />

stress and borers. It is expected that at age 10 the performance <strong>of</strong> these taxa will show marked<br />

reductions in volume yield on a per hectare basis. This projection is similar to findings for E. globulus<br />

in West <strong>Australia</strong> where early-age performance did not foreshadow later-age performance (Walsh et<br />

al., 2008). White et al. (2003) warns the susceptibility to drought is an important consideration<br />

particularly in low rainfall regions where rainfall is highly variable. For many parts <strong>of</strong> Queensland the<br />

long-term average rainfall might appear to be relatively high (e.g. 1000–1200 mm in the Wide Bay<br />

region) but it is highly variable and species such as E. grandis, E. globulus subsp. maidenii, E. saligna<br />

and E. dunnii which have water use strategies similar to E. globulus may be at risk.<br />

Indications <strong>of</strong> taxa most suitable for various products across all trials<br />

A sustainable hardwood plantation industry in south-east Queensland requires optimal choice <strong>of</strong><br />

species in relation to sites (landscape positions, soils) and changing climate conditions. Regarding the<br />

latter, a five per cent reduction in rainfall is predicted by 2030 in Queensland’s subtropics followed by<br />

at least a 10 per cent reduction by 2070 (Anon. Office <strong>of</strong> Climate Change, QLD 2008). Along with


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 78<br />

this, there are predicted increases in temperature, more intense severe weather events and higher<br />

evapotranspiration rates. The trials have been subjected to a period <strong>of</strong> lower than average rainfall.<br />

Therefore taxa performance reported here should be a good indicator <strong>of</strong> their potential when grown<br />

under the more extreme climatic conditions as a result <strong>of</strong> climate change.<br />

While it is important to match species to sites it is probably even more important to understand which<br />

species with suitable wood properties can be grown across a wide range <strong>of</strong> sites and climatic zones.<br />

The stability <strong>of</strong> growth <strong>of</strong> ccv -woon and longi -coom across all trials confirms the early promising<br />

potential <strong>of</strong> these species, identified in Queensland by Lee et al. (2001) and Gardner et al., (2001;<br />

2007) in South Africa. Although ccv -woon has low annual VI increments at age six (


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 79<br />

Impact <strong>of</strong> site on taxa performance<br />

The growth variation within the three sites in the higher rainfall zone is thought to be related to soil<br />

types and topographic position. Trials 506a and 506b were established within the same plantation;<br />

with 506b sited on a lower slope position. While the soil types are fairly similar, the lower slope<br />

position was expected to have a higher moisture status, which may explain the small growth<br />

advantage. Trial 506c is on a different property and is located on a broad alluvial area with deeper<br />

soils. This site has a 70 per cent VI advantage (using the mean <strong>of</strong> the six common taxa) over 506b.<br />

The better growth on this site can probably be explained by deeper and possibly more fertile soils and<br />

access to more water than that available in 506a and 506b.<br />

The two trials in the lower rainfall zone are on similar soils; however, there is overall a 113 per cent<br />

advantage in VI for trees at site 523c site over those in experiment 523d (using the mean <strong>of</strong> the six<br />

common taxa). While soil and site differences may account for some <strong>of</strong> this variation, anecdotal<br />

evidence indicates that there are shallow water tables throughout the area and it is possible that there is<br />

underground water available to trees in 523c.<br />

CONCLUSION<br />

The taxa with the best VI at age six years in these trials were: ccv –woon (28.5 m³/ha in 506a), for<br />

grandis -cop (45.6 m³/ha in 506b), grandis -koo (97.6 m³/ha in 506c), for saligna -clo (84.3 m³/ha in<br />

523c) and for maid -bolaro (33.0 m³/ha in 523d). This equates to annual VI increments 4.7, 7.6, 16.3,<br />

14.1, and 5.5 m³ /ha /annum respectively. If the better annual VI increments could be achieved on a<br />

broad-scale, then the success <strong>of</strong> the hardwood plantations in this region would be assured.<br />

Unfortunately, selecting sites where higher growth rates are achieved is a hit and miss affair the southeast<br />

Queensland. The Symphyomyrtus species (E. dunnii, E. grandis and its hybrids, E. longirostrata,<br />

E. globulus subsp. maidenii and E. saligna) are all highly susceptible to borer attack so they are<br />

unlikely to supply high-value solid wood products. Further these species have water-use strategies<br />

similar to that <strong>of</strong> E. globulus subsp. globulus so the variable rainfall pattern in south-east Queensland<br />

could put these species at risk. If the product is solid wood then the taxa that meet this criteria in<br />

Queensland’s subtropics are select provenances Corymbia citriodora subsp. variegata (e.g. ccv –<br />

woon), the Corymbia hybrids and possible E. cloeziana. If carbon sequestration is the objective then<br />

select provenances <strong>of</strong> species with good growth and high densities such as E. longirostrata and C.<br />

citriodora subsp. variegata should be considered.<br />

ACKNOWLEDGEMENTS<br />

We thank many staff at Department <strong>of</strong> Employment, Economic Development and Innovation and its<br />

predecessor departments: Cliff Raddatz, Murray Johnson, and Alan Ward helped in the establishment<br />

<strong>of</strong> the trials and Tony Burridge validated the measure and assessment data and prepared Figure 1.<br />

Thanks are given to Integrated Tree Cropping and Forest Enterprises <strong>Australia</strong> for provision <strong>of</strong> land<br />

and with help in the establishment and measurement <strong>of</strong> the trials. We thank Garth Nikles and Mike<br />

Shaw for a helpful prepublication review <strong>of</strong> the paper.<br />

REFERENCES<br />

ABARE. (2008). <strong>Australia</strong>n forest and wood products statistics. March and June quarters 2008<br />

(Canberra: <strong>Australia</strong>n Bureau <strong>of</strong> Agricultural and Resource Economics), pp. 82.<br />

Anon. (2008). Climate change in Queensland what the science is telling us (Brisbane: The State <strong>of</strong><br />

Queensland. Environmental Protection Agency Office <strong>of</strong> Climate Change), pp. 36.<br />

Arnold, R., and Luo, J. (2002). Cold Tolerant Plantation Eucalypts for South Central China. In<br />

International Symposium on Eucalyptus Plantations - Guangdon, P.R. China (China), pp. 1-<br />

13.<br />

Atyeo, W.J., Hopewell, G.P., and McGavin, R.L. (2008). Evaluation <strong>of</strong> wood characteristics <strong>of</strong><br />

tropical post-mid rotation plantation Eucalyptus cloeziana and E. pellita: Part (d) Veneer and<br />

plywood potential. (Melbourne: Forest and Wood Products <strong>Australia</strong>, Limited), pp. 47.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 80<br />

Bailleres, H., Hopewell, G.P., and McGavin, R.L. (2008). Evaluation <strong>of</strong> wood characteristics <strong>of</strong><br />

tropical post-mid rotation plantation Eucalyptus cloeziana and E. pellita: Part (c) Wood<br />

quality and structural properties (Melbourne: Forest and Wood Products <strong>Australia</strong>, Limited),<br />

pp. 65.<br />

Boland, D.J., Brooker, M.I.H., Chippendale, G.M., Hall, N., Hyland, B.P.M., Johnson, R.D., Kleinig,<br />

D.A., McDonald, M.W., and Turner, J.D. (2006). Forest Tree <strong>of</strong> <strong>Australia</strong>. (Melbourne:<br />

CSIRO Publishing).<br />

Eldridge, K., Davidson, J., Harwood, C., and van Wyk, G. (1993). Eucalypt domestication and<br />

breeding. (Oxford: Clarendon Press).<br />

Gardner, R.A.W. (2001). Alternative eucalypt species for Zululand : Seven year results <strong>of</strong> site :<br />

species interaction trials in the region. Southern African Forestry Journal 190, 79-88.<br />

Gardner, R.A.W., Little, K.M., and Arbuthnot, A. (2007). Wood and fibre productivity <strong>of</strong> promising<br />

new eucalypt species for coastal Zululand, South Africa. <strong>Australia</strong>n Forestry 70, 37-47.<br />

Isbell, R.F. (1996). The <strong>Australia</strong>n soil classification. (Melbourne, Victoria: CSIRO Publishing).<br />

Jeffrey, S.J., Carter, J.O., Moodie, K.M., and Beswick, A.R. (2001). Using spatial interpolation to<br />

construct a comprehensive archive <strong>of</strong> <strong>Australia</strong>n climate data. Environmental Modelling and<br />

S<strong>of</strong>tware 16, 309-330.<br />

Lawson, S. (2003a). Susceptibility <strong>of</strong> eucalypt species to attack by the giant wood moth (Endoxyla<br />

cinerea) in Queensland. Hardwoods Queensland report No 7. Queensland Forestry Research<br />

<strong>Institute</strong>, Agency for Food and Fibre Sciences, Queensland Department <strong>of</strong> Primary Industries,<br />

pp. 7.<br />

Lawson, S. (2003b). Susceptibility <strong>of</strong> eucalypt species to attack by longicorn beetles (Phoracantha<br />

spp.) in Queensland. Hardwoods Queensland report No 10. Queensland Forestry Research<br />

<strong>Institute</strong>, Agency for Food and Fibre Sciences, Queensland Department <strong>of</strong> Primary Industries,<br />

pp. 11.<br />

Lee, D.J. (2007). Achievements in forest tree improvement in <strong>Australia</strong> and New Zealand 2:<br />

Development <strong>of</strong> Corymbia species and hybrids for plantations in eastern <strong>Australia</strong>. <strong>Australia</strong>n<br />

Forestry 70, 11-16.<br />

Lee, D.J., Nikles, D.G., and Walker, S.M. (1999). Strategy and progress on the genetic improvement<br />

<strong>of</strong> hardwood timber species in Queensland. In Proceedings <strong>of</strong> the 11th <strong>Australia</strong>n Plant<br />

Breeding Conference (Adelaide, <strong>Australia</strong>: CRC for Molecular Plant Breeding), pp. 38-40.<br />

Lee, D.J., Nikles, D.G., and Dickinson, G.R. (2001). Prospects <strong>of</strong> eucalypt species, including<br />

interspecific hybrids from South Africa, for hardwood plantations in marginal subtropical<br />

environments in Queensland, <strong>Australia</strong>. Southern African Forestry Journal 190, 89-94.<br />

Lee, D.J., Huth, J.R., Brawner, J., and Dickinson, G.R. (2009). Comparative Performance <strong>of</strong> Corymbia<br />

Hybrids and Parental Species in Subtropical Queensland and Implications for Breeding and<br />

Deployment. Silvae Genetica, (in press).<br />

Lee, D.J., Debuse, V.J., Pomroy, P.C., Robson, K.J., and Nikles, D.G. (2005). Developing genetically<br />

adapted tree varieties for marginal areas <strong>of</strong> northern <strong>Australia</strong> (Barton, Canberra: Rural<br />

Industries Research and Development Corporation), pp. 58.<br />

Listyanto, T. Glencross, K., Nichols, J.D., Schoer, L. and Harwood, C. (submitted) Performance <strong>of</strong><br />

eight eucalypt species and interspecific hybrid combinations at three sites in northern New<br />

South Wales, <strong>Australia</strong>. <strong>Australia</strong>n Forestry (submitted).<br />

Verryn, S.D. (2000). Eucalyptus hybrid breeding in South Africa. In Hybrid breeding and genetics <strong>of</strong><br />

forest trees. Proceedings <strong>of</strong> QFRI/CRCSPF Symposium, D.G. Nikles, ed (Noosa,<br />

Queensland, <strong>Australia</strong>: Department <strong>of</strong> Primary Industries, Brisbane), pp. 191-199.<br />

Walsh, P.G., Barton, C.V.M., and Haywood, A. (2008). Growth and carbon sequestration rates at age<br />

ten years <strong>of</strong> some eucalypt species in the low- to medium-rainfall areas <strong>of</strong> New South Wales,<br />

<strong>Australia</strong>. <strong>Australia</strong>n Forestry 71, 70-77.<br />

Wang, G., Arnold, R.J., Gardiner, C.A., Zhang, J., and Wu, Z. (1999). Seed source variation for<br />

growth in Eucalyptus dunnii; results from trials in south central China. <strong>Australia</strong>n Forestry<br />

62, 120-127.<br />

White, D.A., Battaglia, M., Macfarlane, C., Mummery, D., McGrath, J., F., and Beadle, C.L. (2003).<br />

Selecting species for recharge management in Mediterranean southwestern <strong>Australia</strong>: some<br />

ecophysiological considerations. Plant and Soil 257, 283-293.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 81<br />

Table 3 Performance means and differences among 60 taxa in five trials for six year volume index, survival and borer incidence 1 .<br />

Definitions <strong>of</strong> taxa are provided in Table 2.<br />

Volume index at age six years (m³/ha) Survival at age six years (%) Borer incidence at age six years (%)<br />

Taxa 506a 506b 506c 523c 523d 506a 506b 506c 523c 523d 506a 506b 506c<br />

acmen -neer 9.1 abc 7.1 ab 82.3 hijkl 64.6 bcd 87.5 efgh 3.8 a 3.8 ab 0.8 a<br />

acmen -woolg 13.9 ab 85.8 efgh 0.0 a<br />

argo -ballon 36.7 cde 16.7 cdef 85.0 ghijkl 82.8 efg<br />

Corymbia hybrid 38.7 cdef 41.1 abc<br />

ccc –glad 2 15.4 cdef 15.3 bcde 28.5 abcdef 45.0 efghi 23.2 efghijkl 84.4 jkl 62.5 bcd 91.7 fgh 85.0 ghijkl 80.0 defg 1.0 a 6.0 ab 0.0 a<br />

ccc -yepp 10.5 abc 39.9 cdefg 46.5 efghij 28.3 hijklm 71.9 ghi 83.3 efgh 85.0 ghijkl 70.0 bcde 1.4 a 0.0 a<br />

ccv -brooy 27.6 gh 26.0 ghij 47.0 efgh 53.1 fghijk 28.8 hijklm 88.5 kl 86.5 gh 94.2 gh 71.7 defgh 76.7 cdef 0.0 a 0.0 a 0.0 a<br />

ccv -woon 28.5 h 31.9 ijkl 57.6 ghi 57.6 hijklmn 29.7 jklm 86.5 jkl 86.5 gh 88.3 efgh 95.0 jkl 95.0 fg 0.0 a 1.0 a 0.0 a<br />

ch -lock 35.6 bcdefg 45.0 efghi 31.7 lm 74.2 cdefg 76.7 efghij 83.3 efg 0.0 a<br />

ch -ner 39.3 cdefg 55.6 ghijkl 29.3 ijklm 77.5 defgh 86.7 ghijkl 83.3 efg 0.0 a<br />

cloez -home 22.4 fgh 15.5 bcdef 47.1 efghij 26.2 ghijklm 86.5 jkl 72.9 cdefg 85.0 ghijkl 80.0 defg 1.1 a 1.4 a<br />

cloez -mungy 22.5 fgh 50.1 efghijk 19.9 defgh 86.5 jkl 82.8 ghijkl 58.3 abc 0.0 a<br />

cloez -saso 23.3 fgh 87.5 kl 0.0 a<br />

dunnii -sapp 24.9 gh 25.3 fghij 86.5 jkl 79.2 efgh 39.6 cd 42.7 bc<br />

grandis -wed 71.3 mnop 28.9 hijklm 86.7 ghijkl 70.0 bcde<br />

grandis -cop 25.6 gh 45.6 m 70.8 gh 80.2 efgh 97.5 fg 93.0 d<br />

grandis -dpi 94.5 j 72.1 nop 30.4 klm 92.5 fgh 96.7 kl 85.0 efg 64.4 ef<br />

grandis -koo 97.6 j 62.0 klmno 26.0 ghijklm 86.7 efgh 83.3 ghijkl 70.0 bcde 90.6 g<br />

grandis -saf 57.4 hijklm 23.2 efghijkl 93.3 ijkl 70.0 bcde<br />

grandis -sap 24.8 gh 33.8 jkl 69.8 fg 62.5 bcd 96.9 fg 95.1 d<br />

g × c -csir 11.5 abcd 58.3 ef 81.9 ef<br />

g × c -dend 23.6 fgh 22.8 efghi 50.8 fghi 59.5 ijklmn 32.0 lm 91.7 l 82.3 efgh 89.2 efgh 95.0 jkl 98.3 g 50.0 d 44.1 bc 29.7 bc<br />

g × p -mix 62.9 hi 82.5 efgh 53.9 de<br />

g × r -csir 21.4 fgh 39.7 klm 74.4 op 21.9 defghijk 77.1 ghijk 90.6 h 88.3 hijkl 81.1 defg 85.8 efg 63.9 cd<br />

g × t -csir 20.6 efgh 21.6 efgh 41.2 defg 83.3 ijkl 75.0 defg 58.3 cde 75.4 e 89.9 d<br />

g × t -mix 21.1 fgh 28.6 hij 45.8 cd 64.6 bcd 96.5 fg 95.2 d<br />

g × u -clone 57.6 ghi 54.9 ghijkl 86.7 efgh 86.7 ghijkl 86.1 fg<br />

g × u -csir 24.4 gh 33.1 jkl 35.8 bcdefg 57.5 hijklm 17.5 cdefg 57.3 de 70.8 cdef 56.7 bcd 60.0 cdef 56.7 abc 95.1 fg 100.0 d 90.7 g<br />

hoop 2.6 NA 0.2 a 83.3 ghijkl 42.8 a<br />

hybrid -mix 28.9 abcdef 55.8 bcd 63.1 de<br />

longi -coom 27.4 gh 41.6 lm 71.0 i 59.0 ijklmn 24.6 fghijklm 88.5 kl 85.4 gh 83.6 efgh 83.3 ghijkl 66.7 bcde 41.0 cd 35.1% abc 16.1 ab<br />

mac -sappi 50.1 efghijk 71.7 defgh


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 82<br />

Table 3 cont.<br />

Volume index at age six years (m³/ha) Survival at age six years (%) Borer incidence at age six years (%)<br />

Taxa 506a 506b 506c 523c 523d 506a 506b 506c 523c 523d 506a 506b 506c<br />

maid -bolaro<br />

26.1 abcde 61.1 jklmno 33.0 m 53.3 bc 83.3 ghijkl 85.0<br />

efg 15.3<br />

ab<br />

maid -bondi 11.8 a 68.7 lmno 25.9 fghijklm 28.3 a 73.3 defgh 61.7 abcd 42.0 cd<br />

mol -ravens 36.2 bcde 10.6 bc 90.0 hijkl 80.0 defg<br />

pch 38.7 cdefg 98.3 h 0.0% a<br />

pell -kur 21.6 fgh 28.8 hij 83.3 ijkl 84.4 fgh 27.4 bc 36.8 abc<br />

pell -png 19.4 defg 17.6 cdefg 82.3 hijkl 68.8 cde 82.6 efg 60.4 cd<br />

pil -beer 22.2 abcd 46.0 efghi 17.2 cdefg 77.5 defgh 86.7 ghijkl 55.0% ab 0.0 a<br />

pil -deong 12.6 a 39.9 def 12.8 bcd 50.0 ab 53.9 bcd 45.0 a 0.0 a<br />

pine hybrid 25.8 abcde 94.2 gh 0.0 a<br />

reduct -rave 8.3 abc 17.4 abc 52.1 b 70.8 bcdef 3.5 ab 1.1 a<br />

resin -beer 38.1 cdefg 20.9 defghij 80.8 efgh 75.0 bcdef 4.2 a<br />

resin -dorr 42.5 defgh 43.8 efgh 20.2 defghi 79.2 efgh 75.0 efghi 81.7 defg 1.1 a<br />

resin -ravens 20.4 efgh 30.9 hijk 87.5 kl 89.6 h 21.9 b 34.7% abc<br />

saligna -clo 84.3 p 93.3 ijkl<br />

siderop -mtmee 38.0 cde 22.4 efghijk 98.3 l 93.3 fg<br />

siderox -ural 24.3 abc 15.0 cde 68.3 defg 86.7 efg<br />

sphaero -bla 11.9 bcde 0.1 a 9.0 a 21.9 ab 4.8 ab 82.3 hijkl 6.3 NA 68.3 bcde 78.3 fghijk 43.3 a 0.0 a 5.5 ab 0.0 a<br />

teret -rund 18.7 defg 86.5 gh 46.5 c<br />

teret -zimb so 13.1 bcde 84.4 fgh 40.3 abc<br />

u hybrid -mix 20.3 a 32.2% a<br />

uro -dongmen 6.7 abc 30.2 ab 88.5 efg<br />

uro -mix 5.0 ab 8.1 abc 30.2 ab 28.1 a 99.8 g 99.8 d<br />

uro -mt egon 2.8 a 18.8 a 99.7 g<br />

uro -sum 6.3 ab 9.0 abcd 34.4 bc 28.1 a 99.9 g 95.6 d<br />

uro -uhak 6.7 abc 50.0 de 89.7 efg<br />

u × c 58.3 hijklmn 71.1% defgh<br />

u × p -sappi 19.4 defg 75.0 ghij 88.5 efg<br />

u × t -csir 16.5 bcdefg 60.4 bc 89.6 d<br />

Mean six<br />

common taxa 2 24.5 28.5 48.4 55.3 26.0 82.8 79.0 83.9 81.7 78.9 31.2 31.0 22.8<br />

1<br />

Taxa differences were significant at the P


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 83<br />

ABSTRACT<br />

RESPONSES TO A DRYING CLIMATE<br />

IN THE NORTHERN JARRAH FOREST<br />

Frank Batini 1<br />

I am deeply concerned at the health <strong>of</strong> the jarrah forest ecosystem. Some <strong>of</strong> the symptoms<br />

measured over the past three decades include: a change to rainfall patterns, lower average<br />

rainfall, a fall <strong>of</strong> several metres in regional water tables, a reduction in streamflows by seventy<br />

five percent, a summer drying <strong>of</strong> streams that were once considered as perennial, consequent<br />

changes to stream biota and the death <strong>of</strong> old, large trees on shallower soils.<br />

These changes are due, in part, to climate change, but also to other hydrologic factors. In the<br />

past ten years it now takes fifty percent more rain for some streams to commence to flow, and,<br />

for an equivalent rainfall, the stream now yields only one-third <strong>of</strong> the previously recorded flow.<br />

Why is this so?<br />

Forest management has also changed. Large areas are now protected as reserves and are not<br />

available for silvicultural operations. There has been a reduction in the frequency <strong>of</strong> prescribed<br />

burning. Bauxite mining has altered the hydrology <strong>of</strong> the rehabilitated mine-pits. All these<br />

changes compound the climatic effects.<br />

However the situation is retrievable, if we have the will as a community to implement alternate<br />

policies and management practices. Research trials have shown that thinning, the utilisation <strong>of</strong><br />

bi<strong>of</strong>uels and regular prescribed burning can increase water tables and streamflow. I estimate that<br />

sound forest management on 100,000 hectares <strong>of</strong> jarrah forest in the high rainfall zone would<br />

benefit forest health, increase water tables, improve stream flow and aquatic biodiversity as well<br />

as producing 80-100MW <strong>of</strong> power annually from bi<strong>of</strong>uels and increasing yield into the surface<br />

dams by between 30 and 50 GL each year, at a fraction <strong>of</strong> the cost <strong>of</strong> desalination.<br />

INTRODUCTION<br />

As a forester and scientist with a lifetime <strong>of</strong> experience in the jarrah forests north <strong>of</strong> Collie, I am<br />

deeply concerned for the future for this ecosystem. Let me share my concerns with you.<br />

• In 1975 the streamflow into dams that are part <strong>of</strong> the Integrated Water Supply System<br />

averaged 350 billion litres. Since that year it has fallen to one-quarter and in 2006 to<br />

one-sixth <strong>of</strong> that amount.<br />

• In Wungong Brook the Vardi road gauging station has been operating for 21 years.<br />

Five <strong>of</strong> the six lowest flows recorded there have occurred in the last eight years.<br />

• Until recently, this Brook was considered to be perennial. In 2006, a very dry year, the<br />

stream ceased to flow for about 12 weeks and even the larger pools were dry. The<br />

stream also ceased to flow in the following summer, even though the 2007 rainfall was<br />

above average.<br />

• Data published by Alcoa scientists (Croton and Reed 2007) show that while water<br />

yields from experimental catchments in the northern jarrah forest subjected to bauxite<br />

mining initially increased (during mining), they are then significantly reduced<br />

following rehabilitation. The decline between 2001 and 2005 in four experimental<br />

catchments ranged between 31 and 67 mmpa, or between 29 and 58 percent <strong>of</strong> the<br />

average streamflow. In another near-by catchment where most <strong>of</strong> the rehabilitation is<br />

younger, the reduction in yield is currently much smaller, 4mmpa and 5 percent <strong>of</strong><br />

flow.<br />

1 Consultant in the Management <strong>of</strong> Natural Resources. Email: fbatini@bigpond.net.au


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 84<br />

• These changes in flow volumes and flow patterns <strong>of</strong> forested streams in bauxite<br />

mining areas are replicated across the jarrah forest, and have already had measurable<br />

effects on stream fauna. Data collected by University <strong>of</strong> Western <strong>Australia</strong> Dr Storey<br />

show significant shifts in faunal assemblages, differences in species richness and<br />

abundance (Storey 2007)<br />

• Regional water-tables have fallen by as much as 12 metres in the past 30 years (Batini<br />

2004). If the specific coefficient <strong>of</strong> the soil is taken as 8 percent, this equates to a drop<br />

<strong>of</strong> one metre <strong>of</strong> water, or 10 000 kilolitres per hectare, or a staggering 4500 billion<br />

litres from the forested catchments that feed into the Integrated Water Supply Scheme.<br />

• The key hydrologic process operating is evapotranspiration by trees and understorey.<br />

In the jarrah forest this can account for 90–95 percent <strong>of</strong> the annual rainfall ( Doley<br />

1967, Greenwood et al 1985).<br />

• The forest is now using more water than is available through rainfall. It is essentially<br />

‘mining’ the reserves that are stored in the soil pr<strong>of</strong>ile and water table.<br />

• Data provided to me by the Department <strong>of</strong> Environment and Conservation (DEC) on<br />

‘Soil dryness index’ show that the upper layers <strong>of</strong> soil, the litter, heavy fuels and living<br />

vegetation are becoming progressively drier.<br />

• Jarrah trees have died from drought stress on shallow soils close to exposed rock<br />

surfaces. This is not unusual but what is most concerning now is that some <strong>of</strong> these<br />

trees are very large, probably 120–150 years old. There is a similar situation in the<br />

wandoo forest, where large old trees are dying back from the crowns ― an obvious<br />

symptom <strong>of</strong> drought. Having survived previous droughts, why are they dying now?<br />

• Hydrologists and water supply engineers agree that the reductions in water yields<br />

appear to follow a step-like process, which is then difficult to reverse. For example, at<br />

the Cobiac gauging station, a rainfall <strong>of</strong> 967 mm in 2006 now yields only one-third the<br />

flow <strong>of</strong> a similar rainfall 10 years ago. This stream now needs about 300mm <strong>of</strong> rain to<br />

begin to flow, whereas 10 years ago it only needed 200mm. In that time the vegetation<br />

has not changed, but water tables have fallen by about 3–5 m on upland sites and 1–2<br />

m on lower slopes (Wungong Whispers 2009).<br />

• Most water movement in the jarrah forest is by ‘interflow’ that occurs within a metre<br />

or so <strong>of</strong> the soil surface. These flows raise the shallow watertable, saturating the<br />

swamps and near-stream zones that then act as ‘roaded’ catchments. Any rain that hits<br />

these wetted areas cannot penetrate into the soil and will then flow into the stream. A<br />

fall in the shallow watertable means that the ‘sponge’ that needs to be filled before any<br />

streamflow can occur is much larger. It is this ‘disjunction’ that may explain these<br />

anomalous results.<br />

You may consider that all <strong>of</strong> these effects are solely due to ‘climate change’ and thus outside <strong>of</strong> our<br />

direct control. This is not so. While lower rainfall is a key driver <strong>of</strong> change, other factors must also be<br />

considered. Let us examine how the forest has changed since 1829.<br />

The jarrah forest in 2008 is very different to that <strong>of</strong> 1829 when settlement at Perth began. Early<br />

photographs show that the original forest was dominated by old, large, mature to overmature trees. It<br />

had a sparser understorey and was probably burnt on a very regular cycle by the Noongar people, and<br />

by lightning-caused fires (Burrows and Abbot 2003). There was no logging, dieback, bauxite mining<br />

or roads. Access by rail and road began in the 1880s and by the 1920s all <strong>of</strong> the Wungong catchment<br />

and much <strong>of</strong> the western jarrah forest had been logged. Most then burnt under wildfire conditions, but<br />

the forest is very resilient and regenerated successfully, from coppice, lignotubers and seed.<br />

The logging and burning changed the structure and composition <strong>of</strong> this forest. A smaller number <strong>of</strong><br />

large, old trees were replaced by large numbers <strong>of</strong> smaller, younger trees. Many areas were again<br />

logged, some as many as four times since 1880. The current forest is now dominated by regrowth, the<br />

oldest up to 120 years <strong>of</strong> age. The density <strong>of</strong> this regrowth (as measured by basal area) is not much<br />

different from that <strong>of</strong> the virgin forest but there are now at least twice as many, smaller trees on each


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 85<br />

hectare (Inventory data supplied by DEC). For a given density, smaller trees have a much greater<br />

‘sapwood area’ and it appears that they are able to maintain higher transpiration rates than larger trees<br />

(C McFarlane CSIRO). While most areas regenerated well, the proportion <strong>of</strong> sheoak (Allocasuarina<br />

fraseriana) also increased on some sites. There are now thickets <strong>of</strong> sheoak on sites where only large<br />

jarrah stumps remain.<br />

Dieback disease was introduced into the western jarrah forest in about the 1920s and was well<br />

established in the Wungong catchment by the 1940s, when the first aerial photos became available (<br />

Batini 1973). The disease spread rapidly during the wetter period between 1950 and 1970 and is now<br />

found over hundreds <strong>of</strong> thousands <strong>of</strong> hectares. Bauxite mining began in the forest in 1964 and now,<br />

some 40 years later, over 13000 hectares have been mined and rehabilitated. The rehabilitated minepits<br />

have been deep ripped to encourage infiltration, are planted densely with tree and understorey<br />

species and are engineered to retain water.<br />

Fire regimes have also changed, from the very frequent burning by the Noongar, through a period <strong>of</strong><br />

attempted fire exclusion in the 1920–40s, to prescribed burning on 5–7 year cycles during the 1960–<br />

1990s to a cycle now averaging 10–12 years over the forest area (DEC pers. comm.).<br />

From this discussion about these landscape-wide changes, the following five factors now play a more<br />

significant role in affecting streamflow:<br />

1. Tree density. If the forest has too many trees, it will transpire more and cause a drop in<br />

water-tables. The forest has always been fully stocked (i.e. stocked to the limit <strong>of</strong> its water supply)<br />

but, since the rainfall has decreased, it is now overstocked with younger regrowth resulting from<br />

past commercial operations. It now contains a high proportion <strong>of</strong> smaller trees that not only<br />

transpire more than larger trees, but are only suitable for use as re-constituted wood, as bio-fuels<br />

or bio-energy, and are currently not able to be widely utilised. Data show that the canopy cover <strong>of</strong><br />

the original stand can be reached or exceeded by regrowth at a very early age, between 10 and 20<br />

years <strong>of</strong> age (Stoneman 1988, Grigg and Grant 2009).<br />

2. Understorey density. The understorey can account for up to 30–40 percent <strong>of</strong> the total<br />

evapotranspiration (Greenwood et al 1985, Marshall et al 1994). If the understorey is denser, it<br />

will transpire more and cause a drop in water-tables. As a result <strong>of</strong> changes to prescribed burning<br />

rotations, lengthening these from about 5–7 years (Forests Department) to about 10–12 years<br />

(DEC, pers. comm.), I believe that the understorey is now denser than it has been in the past. In<br />

addition, some stream areas are left unburnt for “habitat”.<br />

3. Bauxite mining. If bauxite mining has altered the sub-surface, interflow patterns and the<br />

rehabilitation is too dense, transpiration will increase and run<strong>of</strong>f will reduce. By removing the<br />

bauxitic layer, deep ripping, ensuring that any surface run<strong>of</strong>f is captured within pits, and by<br />

seeding densely with native shrub and tree species, mining has severely reduced streamflow in<br />

some catchments (Croton and Reed 2007). This is not a criticism <strong>of</strong> Alcoa World Alumina. These<br />

practices were mutually-agreed to between the mining company and State Agencies as providing<br />

appropriate environmental outcomes, especially the protection <strong>of</strong> water quality, but are probably<br />

more suited to rainfall conditions <strong>of</strong> 30–40 years ago.<br />

4. Larger reserves. Many areas <strong>of</strong> the forest that are now overstocked have been allocated either a<br />

‘formal’ or ‘informal’ reserve status. Some stream reserves are now 400 metres wide, yet these<br />

are the key “source areas” for streamflow. In addition, large Fauna Habitat Zones are now<br />

required at regular spacing within State forest (Conservation Commission 2004) An example is<br />

the Wungong catchment, where one-third <strong>of</strong> the total area is not available for any form <strong>of</strong><br />

silviculture and prescribed burning is only done at intervals <strong>of</strong> as much as 10–12 years —<br />

possibly longer.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 86<br />

5. Dieback disease. Dieback disease is the exception. By killing many species in both the<br />

overstorey and understorey, dieback disease has in fact increased the water yield from affected<br />

areas. (Batini et al 1980) Also, many <strong>of</strong> the dieback areas are close to streamlines and contribute<br />

to enlarging the ‘source area’ for flow. However, this contribution has decreased in recent years as<br />

many areas have now been actively replanted or seeded and others are regenerating naturally.<br />

These factors have major negative implications on water yield and management will need to be<br />

changed on the 100,000 hectares <strong>of</strong> higher rainfall, forested catchment that I believe should be<br />

designated with a priority <strong>of</strong> ‘water supply’ and managed for this purpose. I believe that the concepts<br />

<strong>of</strong> Ecologically Sustainable Forest Management and <strong>of</strong> Multiple Use allow for the identification <strong>of</strong><br />

some areas, such as Nature Reserves where conservation <strong>of</strong> biodiversity is a priority, and other areas<br />

where water production should be a priority, leading to different management regimes.<br />

Is it all doom and gloom? No, I don’t think so. The situation is retrievable, at least in part, if we are<br />

willing to work together as true partners in the management <strong>of</strong> these forests rather than as ideological<br />

competitors for ‘this piece <strong>of</strong> turf’ or as protectors <strong>of</strong> the ‘status quo’. Land management policies that<br />

were developed at a time when the hydrologic regime was quite different need to be re-examined and<br />

many will need to be changed.<br />

We also need to give considerable thought to re-defining our land use priorities in the jarrah forest.<br />

Just as Nature Reserves rightly have a priority for conservation, and National Parks a priority for<br />

conservation and visitor use, the higher rainfall areas <strong>of</strong> State forest within catchment areas (whether<br />

surface or groundwater) should have a priority for the sustained production <strong>of</strong> good quality water,<br />

while recognising other values such as biodiversity ( Forests Department 1978, 1980). The State’s<br />

current investment in our dams, pipelines and infrastructure in these areas now exceeds a billion<br />

dollars. In addition, each hectare <strong>of</strong> forest on surface water catchments in the higher rainfall zone can<br />

yield an average <strong>of</strong> 100 mm per annum <strong>of</strong> high quality drinking water valued at between $1000 and<br />

$1500. I believe that this yield can be increased by 30- 50 percent by good forest management.<br />

The priorities for drinking water catchments should be set by the Department <strong>of</strong> Water (DoW), as I<br />

believe that they are not sufficiently well articulated in the current Forest Management Plan 2004-<br />

2013 nor given adequate recognition by either the Conservation Commission (responsible for forest<br />

planning) or DEC (responsible for the management <strong>of</strong> these forests). This is understandable, given<br />

that that Agency’s prime functions are environmental protection and the conservation <strong>of</strong> biodiversity.<br />

I see that there is a clear need for a very strong leadership role here from the Department <strong>of</strong> Water.<br />

The key ecosystem drivers in the jarrah forest are energy (from the sun), nutrients and water. We have<br />

plenty <strong>of</strong> the first but the other two are both limiting factors. In my view there is now too much<br />

emphasis on research into individual species (eg the quokka, threatened flora, Carnaby cockatoo) and<br />

too little thought given to the impact that landscape-wide processes have on nutrient and water<br />

availability and on forest health. Let me now discuss some <strong>of</strong> these.<br />

Landscape wildfire vs regular prescribed burning<br />

There is minimal direct surface run<strong>of</strong>f from forested catchments in WA, except after wildfire.<br />

Following the 27,700 hectare-2005 Hills wildfire, the soil was bare and hydrophobic. The pattern <strong>of</strong><br />

flow observed at the gauging station and field checks showed severe surface run<strong>of</strong>f and massive<br />

amounts <strong>of</strong> erosion and nutrient loss during the 2005 winter (Batini and Barrett 2007). This pattern is<br />

not observed after a cooler prescribed burn: run-<strong>of</strong>f is increased, but this is not accompanied by soil<br />

damage and erosion.<br />

We also know that although water yield increases temporarily after a wildfire, by as much as 200<br />

percent after the “Hills fire <strong>of</strong> 2005”, it then decreases for many years due to massive regeneration. In<br />

contrast, a statistical analysis for the Water Corporation <strong>of</strong> two large catchments showed that regular<br />

prescribed burning increased water yield by at least 20 percent for 12–18 months after each burn (C<br />

Terry pers com). The more frequent the burn, the greater the increase. However, the DEC has some


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 87<br />

data from two small experimental catchments that do not show an increase in yield after prescribed<br />

burning ( Stoneman pers com).<br />

It is estimated that the 2005 wildfire killed between 1.5 and 2.3 million trees. Many were large, habitat<br />

trees. Regular prescribed burning protects the jarrah forest by allowing these wildfires to be contained<br />

sooner.<br />

Silviculture<br />

A forest ecosystem that is managed with appropriate silviculture will not only yield sustainable<br />

products for our use, but will also increase the water available for use by the remaining trees, the<br />

understorey plants, the soil organisms, the stream flora and stream fauna, possibly by as much as 80–<br />

100 mmpa. Eventually more water will also flow into dams and be available for use by human beings.<br />

There are considerable research data from the jarrah forest thinning trials during the 1980’s-1990’s to<br />

support these statements. (Bari and Ruprecht 2003, Marshall and Chester 1992, Ruprecht et al 1991,<br />

Ruprecht and Stoneman 1993, Sch<strong>of</strong>ield et al 1989 and Stoneman 1993).<br />

Data from the 1980’s show that in regrowth forests there is competition for water and nutrients at<br />

basal area densities underbark <strong>of</strong> about 15 m 2 /ha, say about 20m 2 /ha overbark (Stoneman et al 1996).<br />

With the reductions in rainfall, this value could now be between 10-12 m 2 /ha. Some recent modelling<br />

by J Croton (pers com) shows that, with current rainfall regimes, water tables and streamflows cannot<br />

be maintained if basal areas exceed 10m 2 /ha. However, in the high rainfall zone, the mean basal area<br />

<strong>of</strong> 35 m 2 /ha is well in excess <strong>of</strong> these values. Thinning will substantially increase the growth rate on<br />

the remaining trees (Stoneman et al 1996), and will not greatly alter the composition <strong>of</strong> the<br />

understorey (Mattiske pers com). To maintain the increases in water and timber yields, coppice and<br />

seedling regrowth will need to be controlled. In a time <strong>of</strong> declining rainfall, the silviculture and<br />

utilisation <strong>of</strong> regrowth forests on water catchments needs to be increased, not reduced, as is being<br />

advocated by some interest groups, including the Conservation Commission.<br />

Utilisation<br />

DEC records show that most <strong>of</strong> the jarrah forest on drinking water catchments has been cut over at<br />

least twice, some three or four times. We now have a forest structure which is dominated by smaller<br />

trees that are not large enough to produce sawlogs and can only be utilised on the scale available for<br />

producing re-constituted wood products, bio-energy or second generation bio-fuels. Preliminary<br />

calculations show that the resource <strong>of</strong> bio-energy in the 100,000 hectares <strong>of</strong> State forest within the<br />

western, high-rainfall zone <strong>of</strong> surface water catchments could produce 80-100MW <strong>of</strong> electricity<br />

annually for at least 20-25 years (J Clarke Forest Products Commission, pers com). It is likely that any<br />

move by Government in this direction will be strongly opposed by the environmental movement.<br />

However there are really only two other choices: to carry out non-commercial thinning, at a<br />

considerable cost ($400-1000/ha), and leave the dead trees in the forest; or do nothing and let the<br />

forest sort itself out in the longer term, by dying from drought. This option, preferred by some in the<br />

environmental movement, may take decades to have full effect and in the meantime the stream<br />

environments and the catchment water supply system will continue to decline.<br />

Forest health<br />

The Forest Management Plan prepared by the Conservation Commission is based on the concept <strong>of</strong><br />

Ecologically Sustainable Forest Management. A forest ecosystem where water and nutrients are not<br />

limiting will be a healthier ecosystem. There are data from the 1980s, when rainfall was considerably<br />

higher, that show that the growth rates <strong>of</strong> jarrah are severely reduced at densities in excess <strong>of</strong> 20 m 2 in<br />

basal area, due to competition for these scarce resources. Over 85 percent <strong>of</strong> the forest now exceeds<br />

this density.<br />

Significant ill-health and extensive drought deaths are a real possibility in the future. Trees such as<br />

jarrah and marri are very long-lived organisms (150 to 300+ years) and have therefore developed<br />

multiple strategies to cope with threatening processes such as unusual weather, fire, climatic change


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 88<br />

and attack by insects and fungi. These processes are dynamic; tree crowns may partially recover, and<br />

then be affected once again.<br />

However, because <strong>of</strong> their great age, these older trees may not necessarily be in balance with the<br />

current environment, especially at a time <strong>of</strong> rapid change. Visiting foresters <strong>of</strong>ten remark at the<br />

current sorry state <strong>of</strong> jarrah crowns.<br />

Drought strategies all use less water and therefore mean that a lesser amount <strong>of</strong> photosynthetic<br />

material is produced. Where a forest is subjected to chronic drought or nutrient stress it is also much<br />

less able to resist attack by insect pests or fungal disease. Active response to infection or insect attack<br />

requires mobilisation <strong>of</strong> defence mechanisms that need additional photosynthetic products. These new<br />

leaves are <strong>of</strong>ten much more palatable to insect grazers leading to an increase in decline and further<br />

stress on the tree.<br />

Where the negative impacts on the trees are sustained over long periods (some years or decades) by a<br />

chronic climatic change, drought or continuous insect or fungal attack, the availability <strong>of</strong> food<br />

resources becomes limiting, defence mechanisms are severely stretched and may then fail. External<br />

influences that would have been dealt with by the tree in a normal situation may now lead to severe<br />

debility and tree death.<br />

.<br />

While many elements <strong>of</strong> the jarrah forest will survive, there will be considerable ecological change<br />

with a shift to the hardier, more drought tolerant plants as well as the suites <strong>of</strong> animals and insects that<br />

prefer these plants. While jarrah also grows in lower rainfall areas it is more open, lower in height,<br />

smaller in girth and has a lower basal area. Significant shifts in plant populations can occur over a 30–<br />

40 year period. The biota that are likely to be most negatively affected in the early phases are found<br />

within the stream ecosystems. These negative effects can already be measured (Storey 2007).<br />

In the past 30 months, the Water Corporation <strong>of</strong> Western <strong>Australia</strong> has been funding fieldwork for a<br />

large- scale experiment in the Wungong water-supply catchment that incorporates commercial logging<br />

by the Forest Products Commission, as well as non-commercial thinning and widespread prescribed<br />

burning by the Department <strong>of</strong> Environment and Conservation (DEC) as contractors. Twenty<br />

independent research/monitoring programs are also funded. The trial is anticipated to last for 12 years<br />

and cost $20 million (Water Corporation 2005, www.watercorporation.com.au/wungong).<br />

OTHER STATES<br />

Despite sustained drought and water restrictions in other mainland capital cities, I am not aware <strong>of</strong> any<br />

other Water Authority in <strong>Australia</strong> that is actively managing its forested catchments over large areas<br />

with prescribed fire and silviculture in order to maintain streamflow and biodiversity. In fact, quite the<br />

opposite it appears.<br />

There is limited prescribed burning, <strong>of</strong>ten leading to recent, widespread wildfire in several States<br />

(Victoria, NSW, ACT), including major water-supply catchments. The consequences <strong>of</strong> these largescale<br />

wildfires on reductions in yield as the forest recovers will be considerable and are still to be felt.<br />

Most <strong>of</strong> the major water-supply catchments are still “closed” to logging and there is controversy even<br />

over the logging <strong>of</strong> small areas <strong>of</strong> State forest such as those in the Thompson dam catchment in<br />

Victoria ( M Poynter pers com).<br />

CONCLUSION<br />

I believe we do have a choice. I can accept that Nature Reserves and National Parks in the jarrah<br />

forest be managed conservatively, without silviculture and with longer fire intervals, but the<br />

consequences <strong>of</strong> this management option must be closely monitored. These areas will then provide<br />

our baseline, or reference areas, in terms <strong>of</strong> the impacts <strong>of</strong> this method <strong>of</strong> management on forest<br />

health, biodiversity, stream health and water yield. However, this “experiment ”is being conducted by


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 89<br />

the Conservation Commission over one million hectares <strong>of</strong> former State forest, with minimal<br />

monitoring <strong>of</strong> the consequences.<br />

I do not accept that similar management practices, focussing on the priority <strong>of</strong> conservation <strong>of</strong><br />

biodiversity, should also apply to all <strong>of</strong> the remaining State forest, particularly State forest in the<br />

higher rainfall area <strong>of</strong> water catchments. In some <strong>of</strong> these areas, for example the 13000 ha Wungong<br />

catchment, I propose the need to be able to trial (and monitor) alternative management systems that<br />

favour streams, water yield, timber use, biodiversity and forest health without the imposition <strong>of</strong><br />

unreasonable environmental constraints by the Conservation Commission, the EPA, the DEC or the<br />

environmental movement. What are some <strong>of</strong> these “ unreasonable” constraints?<br />

• The Conservation Commission refuses to permit thinning operations below the approved level<br />

<strong>of</strong> 15 m 2 /ha, even though current data suggest that this level is too high to maintain water tables<br />

and streams.<br />

• The Conservation Commission and DEC refuse to reduce the interval between prescribed burns<br />

to between 4-6 years required for control <strong>of</strong> understorey, preferring an 8-10 year cycle, citing<br />

vague “biodiversity concerns”<br />

• The Conservation Commission refuses to reduce the width <strong>of</strong> stream buffers, some <strong>of</strong> which are<br />

now 400 m wide and therefore negate the effects <strong>of</strong> any thinning upslope. A suitable width <strong>of</strong><br />

buffer could well be 30m either side <strong>of</strong> the streamside vegetation.<br />

• The Conservation Commission will not allow any non-commercial thinning to occur within the<br />

existing wide buffers. This operation could be carried out on foot, targeting only the smaller<br />

trees.<br />

• The DEC has refused a proposal for a trial logging to be conducted under “moist soil”<br />

conditions within areas previously mined and rehabilitated. The reason given is “soil damage”,<br />

but without a trial, this assumption cannot be tested.<br />

• The DEC has refused a proposal to keep some <strong>of</strong> the rehabilitated areas that could be clearfelled<br />

open for between 4 and 6 years to allow time for water tables to build up. The reason<br />

given is that the current guidelines require that rehabilitation must commence within 30 months.<br />

• The Government’s Policy is to allow the use <strong>of</strong> bi<strong>of</strong>uels, but only from plantation residue and<br />

not from native forest. This is due to pressure from the conservation lobby. However this leads<br />

to excessive costs, waste <strong>of</strong> material and visual impact.<br />

If these constraints can be overcome, we will then have at least two, possibly more, very different<br />

management systems available for comparison. Given what we know now, which is likely to be the<br />

more robust in a time <strong>of</strong> climate change?<br />

ACKNOWLEDGEMENTS<br />

The Water Corporation <strong>of</strong> Western <strong>Australia</strong>.<br />

This paper was first presented at a public seminar organised by the Department <strong>of</strong> Water, October 15 th<br />

2008. The Abstract was published by DoW in the seminar’s proceedings. The paper was expanded<br />

subsequently. I have benefited from discussions with colleagues, especially K Barrett, A Reed, J<br />

Bradshaw and R Underwood.<br />

BIBLIOGRAPHY<br />

Bari M A and J K Ruprecht ( 2003) Water yield response to land use change in south-west Western<br />

<strong>Australia</strong> Department <strong>of</strong> Environment Report No SLUI 31<br />

Batini F E (1973) Jarrah Dieback- a disease <strong>of</strong> the jarrah forest <strong>of</strong> Western <strong>Australia</strong>,. Bulletin Forests<br />

Department Western <strong>Australia</strong> No 84 pp45.<br />

Batini FE, Black R E, Byrne J and Clifford P J ( 1980) An examination <strong>of</strong> the effects <strong>of</strong> land use<br />

changes in catchment condition on water yield in the Wungong catchment Western <strong>Australia</strong><br />

Aust For Res10: 29-3


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 90<br />

Batini F E and K Barrett ( 2007)-Monitoring the effects <strong>of</strong> wildfire on water, vegetation and<br />

biodiversity. The Forester 50, 3, p21.<br />

Batini F E ( 2004) Health <strong>of</strong> wandoo ( Eucalyptus wandoo) in the Helena catchment in relation to<br />

depth and reductions in water tables. Report to the Wandoo Recovery Group.<br />

Burrows, N. and Abbott I. (2003) Fire in south-west Western <strong>Australia</strong>: synthesis <strong>of</strong> current<br />

knowledge, management implications and new research directions. Fire in ecosystems <strong>of</strong><br />

south-west Western <strong>Australia</strong>; impacts and management. CALM, 2003, Perth, WA.<br />

Conservation Commission <strong>of</strong> Western <strong>Australia</strong> (2004) Forest Management Plan 2004-2013<br />

Croton J T and Amanda J Reed (2007) Hydrology and bauxite mining on the Darling Plateau.<br />

Restoration Ecology 15,4,40-47.<br />

Doley, D (1967) Water relations <strong>of</strong> Eucalyptus marginata Sm under natural conditions. J.Ecol.55:597-<br />

614.<br />

Forests Department <strong>of</strong> Western <strong>Australia</strong> (1978). A Perspective for Multiple Use Planning in the<br />

Northern Jarrah Forest.<br />

Forests Department <strong>of</strong> Western <strong>Australia</strong> (1980). Land Use Management Plan Northern Jarrah Forest<br />

Management Priority Areas.<br />

Greenwood, E.A. N., Klein, L., Beresford, J. D., Watson, G. D. and Wright, K. D. (1985) Evaporation<br />

from the understorey in the jarrah (Eucalyptus marginata Don ex Sm) forest, south-western<br />

<strong>Australia</strong>. Journal <strong>of</strong> Hydrology 80: 337-349.<br />

Grigg A H and C D Grant ( 2009)- Overstorey growth response to thinning, burning and fertiliser in<br />

10-13 –year-old rehabilitated jarrah ( Eucalyptus marginata ) forest after bauxite mining in<br />

south-western <strong>Australia</strong>. Aust For 72, no 2, pp 80-86.<br />

Marshall, J. K., Chester, G. W. and Colquhoun, I. J. (1994) Water use by rehabilitation vegetation.<br />

Alcoa <strong>of</strong> <strong>Australia</strong>, Report No 94/28, 12p.<br />

Marshall J K and G W Chester (1992) Effect <strong>of</strong> forest thinning on jarrah ( Eucalyptus marginata)<br />

water uptake. Land and Water Research News no 14.24-27<br />

Ruprecht, J.,Sch<strong>of</strong>ield, N., Crombie, D.,Vertessy, R. and Stoneman, G. (1991). Early hydrological<br />

response to intense forest thinning in south-western <strong>Australia</strong>. Journal <strong>of</strong> Hydrology 127: 261-<br />

277.<br />

Ruprecht, J. and Stoneman, G. (1993) Water yield issues in the jarrah forest <strong>of</strong> south-western<br />

<strong>Australia</strong>. Journal <strong>of</strong> Hydrology 150: 369-391.<br />

Sch<strong>of</strong>ield, N. J,. Stoneman, G. L. and Loh, I. C. (1989) Hydrology <strong>of</strong> the jarrah forest. In B. Dell, J.<br />

Havel and N. Malaczjuk (eds). The jarrah forest-A complex mediterranean ecosystem, pp 179-<br />

201. Kluwer Academic Publishers, Dordreck.<br />

Stoneman G L, P W Rose and H Borg (1988) –Recovery <strong>of</strong> Forest Density After Intensive Logging in<br />

the Southern Forest <strong>of</strong> Western <strong>Australia</strong>. Technical report no 19. Department <strong>of</strong> Conservation<br />

and Land Management.<br />

Stoneman G L (1993) Hydrological response to thinning a small jarrah ( Eucalyptus marginata)<br />

catchment . Journal <strong>of</strong> Hydrology 150; 393-407.<br />

Stoneman G L, D S Crombie, K Whitford, F J Hingston, R Giles, C C Portlock, J H Galbraith, and G<br />

M Dimmock- (1996) Growth and water relations <strong>of</strong> Eucalyptus marginata (jarrah) stands in<br />

response to thinning and fertilisation. Tree Physiology 16, 267-274<br />

Storey A ( 2007) Wungong catchment trial project. Aquatic fauna biodiversity assessment. Report to<br />

the Water Corporation<br />

Water Corporation (2005)- Wungong Catchment Environment and Water Management project.<br />

Water Corporation( 2009)-Wungong Whispers vol 8, p2-3.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 91<br />

ABSTRACT<br />

THE NATIONAL <strong>FORESTRY</strong> MASTERS PROGRAM:<br />

A NEW ERA OF COLLABORATION<br />

IN PROFESSIONAL <strong>FORESTRY</strong> EDUCATION<br />

Lyndall Bull 1 , Peter Kanowski 1<br />

Both competition and cooperation have characterised <strong>Australia</strong>n forestry education over the<br />

past century. The recent establishment <strong>of</strong> the National Forestry Masters Program, a<br />

collaborative framework for graduate pr<strong>of</strong>essional forestry education, is the most significant<br />

cooperative endeavour in <strong>Australia</strong>n forestry education since the establishment <strong>of</strong> the<br />

<strong>Australia</strong>n Forestry School in 1926. We argue that the NFMP is a logical means for sustaining<br />

and advancing pr<strong>of</strong>essional forestry education in the larger context <strong>of</strong> the <strong>Australia</strong>n higher<br />

education sector, and that a strategy which identifies and enables the roles and responsibilities<br />

<strong>of</strong> all key actors with significant interests in forestry education is necessary to ensure that<br />

pr<strong>of</strong>essional forestry education in <strong>Australia</strong> survives and prospers.<br />

INTRODUCTION<br />

Recognition <strong>of</strong> the need for pr<strong>of</strong>essional forestry education in <strong>Australia</strong> dates back to 1887, when the<br />

Indian Forest Service Conservator, F.D’A Vincent, recommended establishment <strong>of</strong> a forestry school in<br />

Victoria (Roche and Dargarvel, 2008). It is now nearly 100 years since the Victorian School <strong>of</strong><br />

Forestry was founded in 1910. Negotiation between the states about the establishment <strong>of</strong> a national<br />

forestry school continued until its establishment in 1926 (Carron 1985), and exemplified the mix <strong>of</strong><br />

competition and collaboration between the state agencies and universities that continued to<br />

characterise <strong>Australia</strong>n forestry education. The recent development <strong>of</strong> the National Forestry Masters<br />

Program (NFMP – see http://www.forestry.org.au/masters) is the most explicit example <strong>of</strong><br />

collaboration in forestry education since the establishment <strong>of</strong> the <strong>Australia</strong>n Forestry School. This time<br />

round, at least, the initiative includes Victoria. The current shortage <strong>of</strong> forestry graduates (Figure 1) in<br />

<strong>Australia</strong> is now recognised within the sector as a pressing problem.<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

1996<br />

1997<br />

1998<br />

1999<br />

Source: NAFI/A3P 2006, Figure 9.9<br />

2000<br />

2001<br />

2002<br />

2003<br />

2004<br />

2005<br />

<strong>Australia</strong>n National University<br />

University <strong>of</strong> Melbourne<br />

Southern Cross University<br />

University <strong>of</strong> Queensland<br />

1 Fenner School <strong>of</strong> Environment & Society, College <strong>of</strong> Medicine, Biology and Environment, The <strong>Australia</strong>n National<br />

University, Canberra ACT 0200 <strong>Australia</strong> and National Forestry Masters Program, <strong>Australia</strong>.<br />

Total


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 92<br />

Figure 1: University forestry graduates 1996-2005<br />

The 2006 forestry sector skills report (NAFI/A3P 2006) documented the strong demand for<br />

pr<strong>of</strong>essional forestry skills and graduates, with around 70% <strong>of</strong> forest growing and management<br />

organisations reporting a shortage <strong>of</strong> foresters 2 . The lack <strong>of</strong> suitable candidates to fill vacancies in the<br />

<strong>Australia</strong>n forest sector has forced many organisations to recruit internationally to fill pr<strong>of</strong>essional<br />

vacancies. Whilst the globalisation <strong>of</strong> forestry employment has many benefits for employees,<br />

employers and the pr<strong>of</strong>ession, it also has direct and indirect costs, and the recent reliance <strong>of</strong> the<br />

<strong>Australia</strong>n forestry sector on international recruitment suggests substantial local market failure in the<br />

extent <strong>of</strong> forester demand – supply imbalance. These trends in undergraduate enrolment are also<br />

evident in North America and many European countries (Kanowski 2008).<br />

From the start <strong>of</strong> <strong>Australia</strong>n forestry education until the 1980s, most students studying forestry in<br />

<strong>Australia</strong> were supported by state forestry agencies or other scholarships. Nor were there many<br />

alternative environment-focused degrees until the late 1980s. As a result, the number <strong>of</strong> forestry<br />

students in <strong>Australia</strong> grew to a peak in the 1980s, and has declined subsequently. There are various<br />

reasons to which this decline can be attributed; some reflect broader societal changes, such as the<br />

urban shift in <strong>Australia</strong>’s population, and an attendant decline in interest in pr<strong>of</strong>essions associated with<br />

rural and regional <strong>Australia</strong>. <strong>Australia</strong>n agricultural science education faces a similar challenge in this<br />

respect (Pratley and Leigh 2008). Other reasons may also be in common with agriculture, such as<br />

community perceptions <strong>of</strong> a dumb, sunset industry rather than a smart, innovative one. Unpleasant and<br />

misinformed though they may be, such perceptions are not necessarily entirely without foundation –<br />

for example, the proportion <strong>of</strong> the agriculture, forestry and fishing sector with a degree qualification,<br />

at around 7% in 2004, remains well below that for competitor sectors <strong>of</strong> the economy - eg 17% for<br />

mining and 24% for services (Productivity Commission 2005, Chapter 5).<br />

These changes are taking place in the context <strong>of</strong> diminished national investment in education, at least<br />

in relative terms. Recent reviews <strong>of</strong> <strong>Australia</strong>n innovation and education (Commonwealth <strong>of</strong> <strong>Australia</strong><br />

2008, Cutler 2008) recognised that while higher education lies at the heart <strong>of</strong> <strong>Australia</strong>’s research and<br />

innovation system, <strong>Australia</strong> is falling behind other countries in investment in its higher education<br />

system.<br />

One <strong>of</strong> the consequences <strong>of</strong> the diminished investment in higher education, and <strong>of</strong> the university<br />

funding models that have prevailed over the past decade, has been the increasing inability <strong>of</strong><br />

universities to sustain programs with low student numbers. This is particularly problematic for<br />

pr<strong>of</strong>essional degrees such as forestry, which require breadth across a range <strong>of</strong> topics and problem<br />

solving skills (Brown 2003).<br />

In response to these factors, five <strong>Australia</strong>n universities <strong>of</strong>fering forestry education - <strong>Australia</strong>n<br />

National University, Southern Cross University, University <strong>of</strong> Melbourne, University <strong>of</strong> Queensland<br />

and University <strong>of</strong> Tasmania - cooperated to initiate the <strong>Australia</strong>n National Forestry Masters Program<br />

(http://www.forestry.org.au/masters/), with seed funding over 3 years <strong>of</strong> $1.56 million from the<br />

<strong>Australia</strong>n Government. The NFMP was modelled in part on the relatively new European masters<br />

programs, including two in forestry (SUFONAMA and SUTROFOR), which are predicated on student<br />

mobility. The seed funding has been used principally to employ program convenors at most<br />

participating universities, and to provide c. 70 student mobility scholarships. Political support from<br />

forest sector peak bodies was instrumental in securing the seed funding, and support from the <strong>Institute</strong><br />

<strong>of</strong> <strong>Foresters</strong>, the CRCs for Forestry and Bushfire, Greening <strong>Australia</strong>, and other forest sector<br />

organisations has been fundamental in delivering the program.<br />

2 The range <strong>of</strong> skills shortages reported by the sector incorporates a number <strong>of</strong> skills sets, both pr<strong>of</strong>essional (e.g<br />

Roberts 2007) and vocational. Given the focus <strong>of</strong> the NFMP, this paper’s focus is limited to tertiary education<br />

for pr<strong>of</strong>essional foresters.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 93<br />

THE NATIONAL <strong>FORESTRY</strong> MASTERS PROGRAM<br />

The National Forestry Masters Program is a coordinating framework linking graduate coursework<br />

degrees already <strong>of</strong>fered at the five participating universities. Students enrol at one <strong>of</strong> the participating<br />

universities, and follow its degree rules, but can access courses <strong>of</strong>fered by the other partners. The<br />

program thus <strong>of</strong>fers students access to the best available teaching, field experience, industry and<br />

research opportunities across <strong>Australia</strong>. It also encourages the development <strong>of</strong> pr<strong>of</strong>essional networks,<br />

and links students to forestry in the Asia-Pacific region, by requiring students to participate in two<br />

joint courses, one <strong>of</strong> which is conducted abroad.<br />

The NFMP’s collaboration at the postgraduate coursework level, rather than the undergraduate level,<br />

reflects a number <strong>of</strong> factors: new institutional initiatives, such as the “Melbourne Model”, which focus<br />

pr<strong>of</strong>essional education at the graduate level; the continuing challenges <strong>of</strong> recruiting students into<br />

undergraduate forestry programs; and opportunities for encouraging a more diverse group <strong>of</strong> students,<br />

including those with first degrees in entirely different topic areas, and those with pr<strong>of</strong>essional<br />

experience both within and outside the forest sector. Nor is the approach new for <strong>Australia</strong>n forestry<br />

education – it emulates, in a contemporary context, that which was used at the <strong>Australia</strong>n Forestry<br />

School.<br />

The NFMP therefore plays an important complementary role to the undergraduate forestry programs<br />

still <strong>of</strong>fered at the <strong>Australia</strong>n National University and Southern Cross University. In addition to<br />

attracting a different cohort <strong>of</strong> students, the NFMP is helping to maintain courses – such as forest<br />

operations – which undergraduate numbers alone cannot sustain. While it may be the case that a twoyear<br />

graduate program cannot deliver experience and learning identical to that <strong>of</strong> a traditional fouryear<br />

undergraduate degree, it is also the case that pr<strong>of</strong>essionally-oriented masters are becoming a<br />

common means – both internationally and within <strong>Australia</strong> - <strong>of</strong> delivering pr<strong>of</strong>essional education from<br />

the basis <strong>of</strong> a more generalist undergraduate degree, or one in a different topic area. Nor is it prudent<br />

for the <strong>Australia</strong>n forestry sector to seek to continue to rely on the graduates <strong>of</strong> undergraduate forestry<br />

degrees, unless either the numbers <strong>of</strong> students attracted to undergraduate forestry programs increases<br />

substantially, or educational policy changes to recognise the need for specific, greater investment to<br />

sustain specialist undergraduate programs such as forestry.<br />

The NFMP has been successful in its goals <strong>of</strong> both attracting new candidates to the pr<strong>of</strong>ession and<br />

assisting the further development <strong>of</strong> individuals already working within the sector. To date, more than<br />

50 students are participating in the program. Current NFMP students:<br />

• include individuals from a wide range <strong>of</strong> backgrounds - including information<br />

technology, landscape architecture, natural resource management,<br />

telecommunications and physics;<br />

• range in age from the mid 20s to the mid 50s, and;<br />

• are based in five states and territories.<br />

Most NFMP courses are <strong>of</strong>fered as two-week blocks, to facilitate student mobility and enable<br />

participation <strong>of</strong> pr<strong>of</strong>essionals already in employment. There are both pedagogical advantages and<br />

disadvantages to this model, as with any other, but it is increasingly common at all levels <strong>of</strong> tertiary<br />

education. The learning challenges presented by compression <strong>of</strong> the course into a concentrated period<br />

can largely be addressed by pre- and post-contact activities, and by thoughtful structuring <strong>of</strong> the<br />

contact time; an associated challenge for many students is finding sufficient time pre- and post-contact<br />

to prepare adequately, and to complete assessment requirements.<br />

Other NFMP activities<br />

The NFMP has also served as a vehicle for collaboration between universities to deliver other<br />

educational programs. In 2008/9, the NFMP consortium was funded under the <strong>Australia</strong>n<br />

Government’s Asia Pacific Forestry Skills and Capacity Building Program (APFSCBP), administered<br />

by the Department <strong>of</strong> Agriculture, Fisheries and Forestry, to deliver 3 forestry training programs in the<br />

Asia – Pacific region in 2008/09. These were:


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 94<br />

• Certification training – delivered to c. 40 participants from across the region at Deramakot<br />

Forest, Sabah;<br />

• Participation in NFMP coursework and a work placement in <strong>Australia</strong> - for students from<br />

Malaysia, Laos and the Philippines; and<br />

• Leadership training for c.10 emerging forestry leaders in the Asia-Pacific region, conducted in<br />

Japan in association with the British Council’s Climate Cool program.<br />

The <strong>of</strong>fshore components <strong>of</strong> the APFSCBP have also been open to small numbers <strong>of</strong> NFMP students,<br />

providing them with additional learning and networking opportunities.<br />

Next steps for the NFMP<br />

The <strong>Australia</strong>n Government seed funding for the NFMP will cease at the end <strong>of</strong> 2009, although some<br />

funds will carry over through 2010. The current NFMP model is predicated on student mobility, and it<br />

is hard to envisage any collaborative model that does not include some level <strong>of</strong> student mobility. Staff<br />

mobility, while possible in principle and already a small part <strong>of</strong> the NFMP, is less attractive because it<br />

does not deliver efficiency gains in terms <strong>of</strong> either class size or lecturer workload. Consequently, the<br />

central challenges for the NFMP in the near and medium terms are to secure sufficient funding to<br />

sustain a minimum level <strong>of</strong> provision <strong>of</strong> mobility scholarships, and to continue recruiting activities for<br />

both graduate and undergraduate forestry programs. The <strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, in<br />

conjunction with the NFMP partner universities, has taken the lead in the former, through the<br />

establishment <strong>of</strong> a forestry education trust fund. Supporting that trust fund, and developing and<br />

sustaining effective recruiting strategies, requires an effective partnership across the whole <strong>Australia</strong>n<br />

forest sector.<br />

One way <strong>of</strong> thinking about this partnership is to identify the respective primary and shared roles and<br />

responsibilities <strong>of</strong> the key actors concerned with forestry education. The principal <strong>of</strong> these are<br />

summarised in Table 1.<br />

The structure represented in Table 1 is intended to emphasize that all interested parties have<br />

complementary and important roles in supporting forestry education. As discussed at the IFA Forestry<br />

Education Summit in May 2008, and at subsequent meetings, these roles need to be coordinated within<br />

an overall strategy for fostering forestry education. This strategy would recognise the<br />

interdependencies between both activities supporting forestry education and the roles <strong>of</strong> the different<br />

actors.<br />

It is imperative that those parties committed to sustaining and developing forestry education – the IFA,<br />

the universities, and the forest sector businesses who employ foresters and agencies which have<br />

responsibility for advancing the sector’s development – agree and give effect to a forestry education<br />

strategy. We have at most until mid-2010 to do this if the NFMP is to be sustained.<br />

The consequences <strong>of</strong> not agreeing and implementing such a strategy are likely to be that the<br />

cooperative mechanism represented by the NFMP will collapse – not for lack <strong>of</strong> goodwill, but for lack<br />

<strong>of</strong> resources to enable it to continue. Should the NFMP not continue, forestry education will revert to<br />

being the responsibility <strong>of</strong> individual universities. Unless the policy settings and funding for higher<br />

education change dramatically to favour forestry, the ultimate consequence is likely to be progressive<br />

loss <strong>of</strong> capacity for forestry education at both undergraduate and graduate levels, an enhanced risk that<br />

forestry education in <strong>Australia</strong> will end, and further deterioration in the availability <strong>of</strong> forestry<br />

pr<strong>of</strong>essionals with skills relevant to <strong>Australia</strong>n forestry.<br />

Conversely, sustaining a collaborative model should allow participating universities to evolve their<br />

contributions to the NFMP, and forestry education more generally, to reflect both their strengths and<br />

the strategic directions <strong>of</strong> their institutions, and minimise the risk <strong>of</strong> loss <strong>of</strong> forestry education capacity<br />

nationally. It will also provide the vehicle for continuing collaborative engagement with forestry<br />

education and training in and for the Asia-Pacific region, and more widely, which will itself further<br />

support <strong>Australia</strong>n capacity in forestry education.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 95<br />

Table 1. Principal roles and responsibilities <strong>of</strong> actors associated with forestry education<br />

Activity<br />

Partner<br />

Student recruitment<br />

Scholarship funding<br />

Pr<strong>of</strong>essional development<br />

and networking<br />

Ensure program meets<br />

needs <strong>of</strong> sector<br />

Pr<strong>of</strong>essional body Universities<br />

Engage membership in<br />

effective recruiting<br />

Continue to take lead on<br />

behalf <strong>of</strong> the sector<br />

Facilitate student<br />

membership, networking<br />

activities, and mentoring<br />

Provide fora for connecting<br />

students and employers (eg<br />

advertising within<br />

newsletters, web portal)<br />

Active role in communicating<br />

members’ views<br />

Promote forestry degrees<br />

within overall recruiting<br />

strategy<br />

Pursue funding for forestry<br />

scholarships within the<br />

scholarships and<br />

endowment portfolio<br />

Actively engage sector in<br />

course delivery and extracurricula<br />

events<br />

Engage with industry<br />

events<br />

Establish mechanisms to<br />

enable external partners to<br />

contribute to curriculum<br />

review and development<br />

Forest sector businesses,<br />

agencies and bodies<br />

Promote forestry as a<br />

pr<strong>of</strong>ession as part <strong>of</strong> the<br />

overall business and<br />

communication strategy;<br />

encourage and allow relevant<br />

staff to pursue degrees<br />

Recognise the need for<br />

scholarship funding from the<br />

sector, and contribute to<br />

scholarship pool<br />

Provide vacation work<br />

placements and internships.<br />

Be responsive to requests for<br />

engagement at universities<br />

Support employees to<br />

undertake pr<strong>of</strong>essional<br />

development opportunities<br />

within NFMP<br />

Contribute constructively to<br />

curriculum review and<br />

development, cognisant <strong>of</strong><br />

constraints within universities<br />

Government (DEEWR &<br />

other relevant departments)<br />

Recognise and support forestry<br />

as a sector <strong>of</strong> national<br />

importance and critical skills<br />

shortage<br />

Recognise the need for<br />

scholarship funding in the<br />

sector; contribute to scholarship<br />

pool; ensure tax regulations<br />

encourage corporate and<br />

individual support<br />

Support for international<br />

exchange programs and<br />

placements<br />

Support program and<br />

curriculum development to<br />

meet sector needs.<br />

95


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 96<br />

CONCLUSION<br />

The NFMP has proved in its initial phase to be a feasible and effective model for delivery <strong>of</strong><br />

forest education in the broader context <strong>of</strong> the contemporary <strong>Australia</strong>n higher education<br />

system. It was established and has succeeded to date because <strong>of</strong> the commitment <strong>of</strong> many<br />

partners in the forest sector; it needs their continuing commitment and support to be<br />

sustained. The cooperative model represented by the NFMP <strong>of</strong>fers the <strong>Australia</strong>n forestry<br />

sector the best strategy for the continued delivery and further development <strong>of</strong> specialist<br />

forestry education, at both graduate and undergraduate levels, as the basis for meeting the<br />

sector’s needs for pr<strong>of</strong>essional foresters. The development and implementation <strong>of</strong> a strategy<br />

based on the respective roles and responsibilities <strong>of</strong> key actors with interests in forestry<br />

education is the next critical step in sustaining forestry education; we have only a little time to<br />

complete this task.<br />

REFERENCES<br />

Brown, N. 2003. “A Critical Review <strong>of</strong> Forestry Education” Bioscience Education E-journal, 1-4.<br />

http://www.bioscience.heacademy.ac.uk/journal/vol1/beej-1-4.aspx<br />

Carron, LT. 1985. A history <strong>of</strong> forestry in <strong>Australia</strong>. ANU Press. 355 p.<br />

Commonwealth <strong>of</strong> <strong>Australia</strong>. 2008. “Review <strong>of</strong> <strong>Australia</strong>n Higher Education” Department <strong>of</strong><br />

Education, Employment and Workplace Relations, Canberra, <strong>Australia</strong>.<br />

http://www.deewr.gov.au/he_review_finalreport.<br />

Cutler, T. 2008. “Report on the Review <strong>of</strong> the National Innovation System” Department <strong>of</strong> Innovation,<br />

Industry, Science and Research, Canberra, <strong>Australia</strong>.<br />

http://www.innovation.gov.au/innovationreview/Pages/home.aspx<br />

Kanowski, P. 2008. “Centennial challenges: pr<strong>of</strong>essional forestry education in Canada and the USA in<br />

2007, and learnings for <strong>Australia</strong>; Forest and Wood Products <strong>Australia</strong> 2007 Dennis Cullity<br />

Fellowship Report. Forest and Wood Products <strong>Australia</strong>, Melbourne <strong>Australia</strong><br />

National Association <strong>of</strong> Forest Industries and <strong>Australia</strong>n Plantation Products and Paper Industry<br />

Council. 2006. “Wood and Paper Products Industry Skills Shortage Audit”<br />

Pratley, J and Leigh, R. 2008. “Agriculture in decline at <strong>Australia</strong>n Universities” Global Issues<br />

Paddock Action. Proceedings <strong>of</strong> the 14th <strong>Australia</strong>n Agronomy Conference. September 2008,<br />

Adelaide South <strong>Australia</strong>.<br />

Productivity Commission. 2005. “Trends in <strong>Australia</strong>n Agriculture,” Research Paper, Canberra.<br />

Roberts, R. 2007. “The role <strong>of</strong> graduates and their education for the <strong>Australia</strong>n wood processing sector<br />

– results from a forest industry survey” Forest and Wood Products Research and Development<br />

Corporation, Melbourne, <strong>Australia</strong>.<br />

Roche, M and Dargavel, J. 2008. “Imperial Ethos, Dominions Reality: Forestry Education in New<br />

Zealand and <strong>Australia</strong>, 1910–1965” Environment and History. 14: 523-43.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 97<br />

Summary <strong>of</strong> Presentation<br />

EDUCATION AND TRAINING PATHWAYS<br />

Antoinette Hewitt 1<br />

ForestWorks’ core business is facilitating skills development for the forest, wood, paper and timber<br />

industry workforce. ForestWorks has Industry Skills Council (ISC) status giving it responsibility for<br />

the development <strong>of</strong> nationally consistent competency standards and qualifications.<br />

Skills <strong>Australia</strong>’s recent review <strong>of</strong> the governance <strong>of</strong> the Vocational Education and Training (VET)<br />

sector, and the Council <strong>of</strong> <strong>Australia</strong>n Governments’ (COAG) push to integrate workforce<br />

development, employment and education priorities, have seen a move towards focusing on unity<br />

within the tertiary education and training area consisting <strong>of</strong> the two sectors: VET and Higher<br />

Education.<br />

Government is driving a future in which pathways between the two sectors will be increasingly<br />

seamless with multiple entry and exit points.<br />

In the push for economic and social sustainability and the pivotal role that forests will play, it is<br />

increasingly important that forests are managed by a skilled workforce. Alarmingly, trends in<br />

<strong>Australia</strong> show that the numbers <strong>of</strong> pr<strong>of</strong>essional foresters are declining to the extent that a skills<br />

shortage exists and, with the current small number <strong>of</strong> university enrolments, this is unlikely to ease in<br />

the foreseeable future.<br />

Over the past year, ForestWorks has been engaging with universities and industry on this issue in<br />

order to see if ForestWorks can in some way contribute to the efforts to overcome a shortage <strong>of</strong> people<br />

skilled in Forest management. This has involved participation in the national <strong>Foresters</strong>’ Summit in<br />

Canberra, and more recently with discussions with various universities and industry. In the future<br />

these discussions must also involve the IFA.<br />

ForestWorks is now looking at developing a strategy to expand the traditional pathways into Forester<br />

activities by including existing workers who may be or have been studying in VET programs. Such<br />

inclusion would boost the potential number <strong>of</strong> graduates eligible for entry to the Forestry Masters<br />

Program. Part <strong>of</strong> this strategy will be to explore whether VET outcomes can be delivered via the<br />

university-based Masters program allowing additional, funded, VET students to become part <strong>of</strong> the<br />

program and thereby increasing the likelihood <strong>of</strong> maintaining a critical mass <strong>of</strong> students at the<br />

universities with forestry courses.<br />

1 Manager - Industry Skills Council Project, ForestWorks, 559A Queensberry Street<br />

North Melbourne,Vic 3051. Ph: +61 9321 3512 Email: ahewitt@forestworks.com.au


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 98<br />

<strong>FORESTRY</strong> IS A HANDY WORD BUT NOT THE RIGHT ONE<br />

FOR ATTRACTING STUDENTS.<br />

Chris Weston and Katherine Whittaker 1<br />

ABSTRACT<br />

Building interest in bachelor qualifications in forest science has proven difficult over the<br />

last decade, with graduate forester numbers recently falling, and now well short <strong>of</strong> forest<br />

sector demand. This paper relates our experience over the last two years in attracting<br />

students to the Master <strong>of</strong> Forest Ecosystem Science at The University <strong>of</strong> Melbourne and<br />

degree-level forest science in general. Although job and career prospects, teaching mode,<br />

course location and financial support are all important drivers <strong>of</strong> student choice <strong>of</strong> degree<br />

– an overriding factor is the need to appeal to prospective students’ sense <strong>of</strong> idealism and<br />

desire to contribute to big picture issues that are globally significant. In this paper we<br />

explore this theme and present our vision for the steps to a future well populated with<br />

forest science graduates.<br />

INTRODUCTION<br />

It's hard to believe that a nation <strong>of</strong> over 20 million people with more than150 million hectares <strong>of</strong><br />

forest, including 40 million hectares <strong>of</strong> productive open forest formations and almost 2 million<br />

hectares <strong>of</strong> plantation forests, would struggle to maintain a forest-specific degree in tertiary education.<br />

Yet it's true, and the problem has exercised the minds <strong>of</strong> our best exponents <strong>of</strong> forest disciplines in<br />

universities and in the forest sector more broadly for over a decade (Kanowski, 2006).<br />

Analysing the problem <strong>of</strong> low student numbers in forestry and the causes is difficult, as there are few<br />

sources <strong>of</strong> hard data to draw on other than student numbers, forest sector job numbers and the nature<br />

<strong>of</strong> existing tertiary forestry courses (NAFI and A3P, 2006).<br />

Here we consider the key elements <strong>of</strong> the problem by first making a brief overview <strong>of</strong> the<br />

development <strong>of</strong> university-level forest education nationally, followed by discussing our experience in<br />

attracting students to forest science over the last two years. Our focus here is at degree level –<br />

bachelor and master qualifications – rather than diploma qualifications.<br />

We propose that awareness <strong>of</strong> forest science, the role <strong>of</strong> foresters, and the importance and role <strong>of</strong> the<br />

forest sector is poorly developed in the general community and that forestry as a catch all for forest<br />

disciplines is a poor choice <strong>of</strong> term for attracting students to bachelor and coursework master degrees.<br />

It seems the term forestry neither inspires interest nor conveys the broader relevance <strong>of</strong> forest<br />

disciplines to society and thus to a new audience <strong>of</strong> potential students.<br />

<strong>FORESTRY</strong> DISCIPLINES - have a long heritage in <strong>Australia</strong>n universities<br />

Turning briefly to the history <strong>of</strong> forestry education we see that forestry has been a part <strong>of</strong> university<br />

curricula in <strong>Australia</strong> since the University <strong>of</strong> Adelaide established a Forestry Department in 1911. The<br />

forestry disciplines developed at Adelaide were eventually transferred to the <strong>Australia</strong>n Forestry<br />

School in 1927 under the control <strong>of</strong> the Commonwealth Forestry Bureau, and reformed as the<br />

Forestry Department <strong>of</strong> the ANU in 1965.<br />

In Victoria, the Forests Act <strong>of</strong> 1907 prescribed an examination system for the training <strong>of</strong> pr<strong>of</strong>essional<br />

foresters leading to the first intake <strong>of</strong> forestry diploma students in 1910 at Creswick (Ferguson and<br />

Youl, 1998). Select graduates from Creswick entered a Bachelor <strong>of</strong> Science at The University <strong>of</strong><br />

Melbourne (UM) where a Bachelor <strong>of</strong> Science (Forestry) was eventually formalized in 1943.<br />

Two aspects <strong>of</strong> this history <strong>of</strong> forest education are relevant here. The first is that forest science<br />

degrees emerged in the university system as a continuation <strong>of</strong> diploma level forestry qualifications.<br />

1 School <strong>of</strong> Forest and Ecosystem Science, University <strong>of</strong> Melbourne. Tel: 53214103 Email: Weston@unimelb.edu.au


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 99<br />

We think this legacy highlights the need to adopt a different name for tertiary or degree level<br />

qualifications to distinguish them from the more applied and technical diploma forestry qualifications.<br />

A general lack <strong>of</strong> clarity around the distinction between diploma and degree qualified foresters is<br />

likely a contributing factor to declining interest in forestry degrees at university. The second aspect is<br />

that both State and Federal government funds external to the university system have been crucial to<br />

the development <strong>of</strong> forest science disciplines and the training <strong>of</strong> foresters. Unless student patronage <strong>of</strong><br />

forest courses at universities increases very soon, these external funds will be required again to shore<br />

up forest disciplines.<br />

Today forest science disciplines persist across half a dozen or so universities and each <strong>of</strong> them has<br />

challenges in maintaining sufficient capacity to <strong>of</strong>fer comprehensive forest science programs. The<br />

National Forestry Masters Program (NFMP) has begun the difficult task <strong>of</strong> seeking to establish a<br />

practical and viable way to <strong>of</strong>fer comprehensive forest science education where the expertise is drawn<br />

from across the participating universities.<br />

The whole forest sector has a stake in the success <strong>of</strong> the NFMP and its drive to support and sustain<br />

both undergraduate and postgraduate coursework in forestry – as set out by Lyndall Bull, Peter<br />

Kanowski and the NFMP partners elsewhere in these proceedings. It is crucial that the forest sector<br />

and the university partners in the NFMP agree on how to move this initiative from seed funding by<br />

government to an ongoing basis. These aspects are discussed in the NFMP paper.<br />

ENROLMENTS IN FOREST SCIENCE DEGREES – the Melbourne experience<br />

Here we briefly review the experience at The University <strong>of</strong> Melbourne to understand how enrolments<br />

in forest science have changed over the last few decades.<br />

Graduates from Melbourne's Bachelor <strong>of</strong> Science (Forestry), created as a degree in 1943, have<br />

numbered between about 2 and 20 students from the 1950’s until the early 1980’s. Throughout the<br />

1980's the Bachelor <strong>of</strong> Forest Science (BForSci) attracted between 15 and 25 students each year; first<br />

year enrolments did not exceed 30 students until the early 1990’s. Enrolments increased in 1994 with<br />

the introduction <strong>of</strong> combined degrees (five-year with Science or Commerce), reaching a peak <strong>of</strong> about<br />

60 students into the 1997 class. The peak years were 1993 to 1997, when between 50 and 62 students<br />

commenced a degree in forest science.<br />

Because first year completions were about 10 to 15 less than commencing students during these high<br />

enrolment years, the maximum number <strong>of</strong> students entering second year at Creswick has always been<br />

less than 48 and has only exceeded 40 in two years (1995 and 1997).<br />

BForSci enrolments, including combined degree students, declined to about 22 in 2002 and to around<br />

10 to 12 annually over 2003 to 2007, when the degree was discontinued at Melbourne in a university<br />

level overhaul <strong>of</strong> undergraduate course <strong>of</strong>ferings.<br />

To summarize, annually the number <strong>of</strong> students continuing to 2nd year <strong>of</strong> the BForSci rose from<br />

around 10 to 20 in the 1980s to about 40 in the mid 1990s and then fell back to 8 to 10 in the years<br />

from 2005 to 2007.<br />

We can see from this history <strong>of</strong> enrolment in forest science that student numbers have never been<br />

large – Melbourne has graduated about 1000 BForSc (or equivalent) students over the last 65 years.<br />

The peak interest in forest science at the University <strong>of</strong> Melbourne coincided with a relatively fleeting<br />

popularity <strong>of</strong> double degrees in the mid 1990s. The 5-year forest science/science or commerce double<br />

degrees likely lost patronage due to the rising costs <strong>of</strong> university education for students, combined<br />

with the increasing tendency to work part-time to support tertiary studies.<br />

Undoubtedly a rise in the <strong>of</strong>fering <strong>of</strong> natural resource management and environmental science degrees<br />

throughout the 1990s had a significant impact on forest science enrolments. These degrees were<br />

generally <strong>of</strong> 3 years and more accessible than forest degrees which were 4 years duration and for most<br />

students incurred travel and living away from home costs.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 100<br />

FROM BACHELOR TO MASTER – the forest discipline at the University <strong>of</strong> Melbourne<br />

from 2008 and the National Forestry Masters Program<br />

In shifting from discipline-specific degrees <strong>of</strong> 4 or 5 years to a broader 3-year general undergraduate<br />

degree followed by 2-year graduate programs (The Melbourne Model), the University <strong>of</strong> Melbourne<br />

moved the forest science discipline to graduate school.<br />

The last intake to the Bachelor <strong>of</strong> Forest Science degree in 2007 has been accompanied by the<br />

introduction <strong>of</strong> a two-year graduate coursework program in 2008 – the Master <strong>of</strong> Forest Ecosystem<br />

Science (MFES).<br />

Melbourne's shift <strong>of</strong> forest science to graduate coursework has provided a strong impetus to the<br />

National Forestry Masters Program (NFMP) as shown by the enrolment <strong>of</strong> more than 80% <strong>of</strong> NFMP<br />

students in the Master <strong>of</strong> Forest Ecosystem Science.<br />

The new coursework masters has been named Master <strong>of</strong> Forest Ecosystem Science to emphasize the<br />

central understanding <strong>of</strong> forests as systems in developing students’ understanding <strong>of</strong> forest disciplines<br />

and to appeal more broadly to potential students. Many <strong>of</strong> the 45 MFES students who have<br />

commenced in 2008 and 2009 have commented that they were initially attracted to the ecosystem<br />

component <strong>of</strong> the degree. Our experience in establishing the MFES has shown that gaining the interest<br />

<strong>of</strong> potential students is the first crucial step to increasing enrolments.<br />

Once engaged in investigating our course, website, or in conversation, the key aspects <strong>of</strong> a forest<br />

ecosystem science degree can be explained. Once the students have studied a foundation subject that<br />

outlines the importance <strong>of</strong> forests to human welfare both globally and in <strong>Australia</strong> they come to<br />

appreciate the importance <strong>of</strong> more traditional forestry disciplines and to go on to choose these subjects<br />

in their course.<br />

The structure and content <strong>of</strong> the MFES, especially the first few subjects, is very important to drawing<br />

in students who are new to the discipline area, and showing them the relevance <strong>of</strong> forests and forest<br />

disciplines to big picture problems facing society.<br />

In becoming wise to the relevance <strong>of</strong> forest sciences the students can then choose to study any or each<br />

<strong>of</strong> technical, social and policy or economic aspects <strong>of</strong> forestry. So far we have found that students new<br />

to forestry relish the opportunity to study across a broad range <strong>of</strong> forest subjects, the niche<br />

specializations <strong>of</strong>fered, and to take advantage <strong>of</strong> national and international study options.<br />

STIMULATING STUDENT INTEREST IN FOREST SCIENCE<br />

Searle and Bryant (2009) surveyed students in the Bachelor <strong>of</strong> Science (Forestry) at ANU in the late<br />

1990s and from responses concluded that general lack <strong>of</strong> public awareness <strong>of</strong> forestry and the<br />

relevance <strong>of</strong> forest disciplines was central to declining student numbers. Their paper describes well the<br />

declining interest in courses labelled forestry and their conclusions accord with our experience at<br />

Melbourne.<br />

We think it is unfruitful to persist with the term forestry in the naming <strong>of</strong> university degrees and in<br />

promoting pr<strong>of</strong>essional forester qualifications. Forestry in a degree title or in the promotional text and<br />

dialogue conveys a narrow industrial focus that does not reflect broader course content and the full<br />

range <strong>of</strong> career possibilities known to arise from a degree in forest science.<br />

The forestry moniker thus fails to capitalise on interest in new and emerging employment<br />

opportunities for which forest science pr<strong>of</strong>essionals are well qualified. Attracting student interest in<br />

forest science qualifications is our biggest challenge.<br />

To illustrate our key points we summarized in Table 1 our strategy in promoting the Master <strong>of</strong> Forest<br />

Ecosystem Science.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 101<br />

Table 1. A summary <strong>of</strong> our key messages and actions in both promoting the Master <strong>of</strong> Forest<br />

Ecosystem Science and maintaining students in the course.<br />

1. We have emphasized the role <strong>of</strong> forests in key environmental issues facing society<br />

Climate change – forests central to carbon cycling, C sequestration and bi<strong>of</strong>uels<br />

Water – forested catchment management and protection crucial for water security<br />

Fire – knowledge <strong>of</strong> forests and fire central to community response to rising fire risk<br />

Landscape restoration – is a big part <strong>of</strong> future farming landscapes and land management<br />

2. We have made explicit the distinctive features <strong>of</strong> studying in the Master <strong>of</strong> Forest Ecosystem<br />

Science relative to more general environmental science degrees<br />

Study based on a core <strong>of</strong> forest subjects with a wide choice <strong>of</strong> electives<br />

Two week intensive teaching allows for study during short breaks from work<br />

Strong emphasis on field experience and related practical exercises<br />

The course retains a good mix <strong>of</strong> traditional forest science subjects such as silviculture, wood<br />

science, forest operations and forest inventory – these subjects are rarely available elsewhere<br />

and are a distinctive feature <strong>of</strong> the course<br />

Part-time study is possible and students can start in the program at any stage throughout the year<br />

The course structure and learning and teaching mode results in strong cohort and collaborative<br />

learning experience<br />

Strong forest sector employer connections linked to internship study opportunities and research<br />

project work.<br />

Good opportunities for developing a pr<strong>of</strong>essional network including international study and<br />

employment opportunities<br />

3. We emphasize the financial incentives<br />

NFMP scholarships are attractive to students and important in differentiating forest masters<br />

from more general environmental masters courses<br />

The availability <strong>of</strong> commonwealth supported places and eligibility for Centrelink payments<br />

Students eligible for financial assistance from the University <strong>of</strong> Melbourne's Graduate Access<br />

program<br />

A pr<strong>of</strong>essional coursework degree in contrast to many more general environmental management<br />

degrees<br />

THE UNIVERSITY ENVIRONMENT – challenges in maintaining forestry academics and their<br />

disciplines<br />

A key driver <strong>of</strong> the NFMP – as discussed in Lyndall Bull and Peter Kanowski's contribution to this<br />

conference – is the decline in forestry disciplines in universities. Broadly, academics are selected and<br />

maintained in universities based on potential teaching and research demand in their area <strong>of</strong> expertise,<br />

and on their record and perceived ability to maintain a strong original research output as demonstrated<br />

by publication in international journals.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 102<br />

Academics who score highly on ability to attract teaching income, research higher degree students and<br />

external research grants, survive and prosper in the system. With dwindling student teaching loads,<br />

forest academics are challenged to maintain relevance in competitive university environments.<br />

In <strong>Australia</strong> we have already reached the point at which comprehensive forestry programs are difficult<br />

to mount with the existing academic staff. Universities have dealt with this by maintaining a small<br />

core <strong>of</strong> forestry academic staff and have contracted external expertise to cover teaching in essential<br />

forestry disciplines. The NFMP <strong>of</strong>fers a better solution to this problem as it has the potential to<br />

strengthen the position <strong>of</strong> existing academic appointments.<br />

CONCLUSIONS – a changing context for tertiary forest education in <strong>Australia</strong><br />

Degree qualified foresters carry, develop and contribute knowledge and skills essential for the social,<br />

economic and environmental wellbeing <strong>of</strong> our society. The problem is that you have to be one or know<br />

one to recognize the critical role <strong>of</strong> foresters' skills in society. All foresters have a key role in<br />

educating the community as a basis for attracting prospective students into tertiary forest education.<br />

<strong>Foresters</strong> can take advantage <strong>of</strong> the rapid emergence <strong>of</strong> climate change in the public consciousness,<br />

and the many challenges it poses that link forests with community safety and wellbeing – such as<br />

management <strong>of</strong> fire risk, water security in forest catchments, and growth and protection <strong>of</strong> carbon<br />

sinks for greenhouse gas reduction. This provides a good opportunity to promote forester skills.<br />

Knowledge and expertise in these areas can be linked to the more traditional understanding <strong>of</strong> foresters<br />

roles — e.g. in the stewardship <strong>of</strong> forests for biodiversity conservation, wood for high value timber<br />

and paper products. This strategy will help the community to understand the crucial role <strong>of</strong> foresters<br />

and lead to interest in forest science in education.<br />

Our key point, repeated throughout this paper, is that forestry is not the right term for the forest<br />

pr<strong>of</strong>ession and forest educators to adopt in the ongoing campaign to attract new students – as it<br />

conveys a narrow industrial concept <strong>of</strong> foresters' skills. It is time to universally embrace new language<br />

such as forest sector instead <strong>of</strong> forest industry, forest ecosystem services or benefits instead <strong>of</strong> forest<br />

products, and forest science instead <strong>of</strong> forestry.<br />

As pr<strong>of</strong>essional foresters we should redouble our efforts to speak with the broader community in<br />

language that leads community attitudes and shows the immediate relevance <strong>of</strong> foresters to the<br />

pressing issues <strong>of</strong> the times from practical through to policy level, as well as reinforce the broad range<br />

<strong>of</strong> employment available. In doing so we will be appealing to the idealism <strong>of</strong> potential students and<br />

attracting their patronage.<br />

Finally, in reaching out to a broader audience we must look ahead and speak in positives rather than<br />

look back and search among the dark perceptions <strong>of</strong> forestry – there is too much at stake for <strong>Australia</strong><br />

as a nation if we loose our forest disciplines in tertiary education.<br />

REFERENCES<br />

Searle S, Bryant C (2009) Why students choose to study for a forestry degree and implications for the forestry<br />

pr<strong>of</strong>ession. <strong>Australia</strong>n Forestry 72: 71-79.<br />

NAFI and A3P (2006) Wood and paper products industry skills shortage audit. NAFI and A3P 249 pp.<br />

http://www.nafi.com.au/skills/ (accessed July 2009).<br />

Kanowski PJ (2006) Forestry education – where are the students and what should we do? <strong>Australia</strong>n Forestry 4:<br />

241-241.<br />

Ferguson IS, Youl R (1998) Forestry at Creswick and the University, 1910.<br />

http://www.landfood.unimelb.edu.au/dean/book2/index.html (accessed July 2009)


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 103<br />

ABSTRACT<br />

TRAINING OPERATIONS STAFF IN THE FIELD -<br />

CHALLENGES AND OPPORTUNITIES<br />

Matthew Doig 1<br />

The last decade has seen major changes in the role <strong>of</strong> Forest Operations staff.<br />

They are now expected to undertake and supervise tasks with a greater degree<br />

<strong>of</strong> complexity and responsibility than in the past. <strong>Foresters</strong> now face major<br />

challenges in training these people in the various aspects <strong>of</strong> forest management.<br />

A project to develop a series <strong>of</strong> operationally based Capability Statements and<br />

associated Assessment Tools covering key areas <strong>of</strong> forest operations is<br />

examined. This examination shows how these Capabilities and Tools ensure a<br />

consistency <strong>of</strong> approach and clear directions for trainers.<br />

The paper highlights the following:<br />

• The greater attention needed to meet Operations Staff training<br />

needs.<br />

• Industry based Capability Standards can help meet that need.<br />

• Training can be better targeted, and validated, through such<br />

capabilities, together with Assessment Tools.<br />

• Capabilities and Tools can complement and strengthen existing<br />

training activities.<br />

INTRODUCTION<br />

Forestry, pr<strong>of</strong>essionally and operationally, has undergone significant changes over the last<br />

twenty years. This has been reflected in such things as the technology employed and in the<br />

way Forestry work itself is undertaken. A 1949 promotional brochure for the School <strong>of</strong><br />

Forestry Creswick lists what foresters will learn and apply in the course <strong>of</strong> their career. Some<br />

<strong>of</strong> the tasks mentioned on this brochure are now very much in the province <strong>of</strong> operations staff,<br />

many <strong>of</strong> whom do not hold any qualifications in Forestry and who formerly operated at the<br />

sub-pr<strong>of</strong>essional level under direction from foresters. As this situation becomes more<br />

common the need for better training for this emerging group assumes greater urgency.<br />

Providing training for operations staff continues to be a significant challenge for Forestry<br />

organisations, particularly due to the limited options such people have in terms <strong>of</strong> external<br />

programs focused on them. This training response can be better achieved via a process <strong>of</strong><br />

identifying industry or even company specific capabilities and developing assessment tools<br />

with which to validate learning. This can be illustrated with the example <strong>of</strong> VicForests which<br />

has responsibility for the development <strong>of</strong> many operations staff and which has sought to<br />

ensure a consistent level <strong>of</strong> training via such an approach.<br />

VicForests is a leading native forest timber harvesting and sales business in Victoria,<br />

<strong>Australia</strong>. According to its mission statement it seeks to ensure “a key role in ensuring the<br />

economic prosperity <strong>of</strong> Victoria’s timber industry while ensuring operations are sustainable in<br />

the long term.” VicForests prides itself on “following strict sustainable forest management<br />

systems and standards that are globally recognized” (VicForests 2008). Like all major forestry<br />

organisations, such a statement has implications for clearly defining the required capabilities<br />

<strong>of</strong> staff, addressing them through appropriate training and ensuring that skill and knowledge<br />

sets which support certification <strong>of</strong> its products and operations are clearly validated.<br />

1 Melliodora Solutions, 13 Hamel Street, Hampton, Vic. Ph: 0412081199 Email: mattdoig@bigpond.net.au.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 104<br />

VicForests identified the need to develop and implement its own practical, operationally<br />

based Capability Statements and supporting Assessment Tools, which would support its<br />

Certification process and ongoing adherence to <strong>Australia</strong>n Forestry standards. The<br />

compliance-based nature <strong>of</strong> much <strong>of</strong> the work VicForests undertakes further highlighted the<br />

need for staff capabilities to be comprehensively validated.<br />

Existing training packages at the VET level were not considered comprehensive enough to<br />

cover Forestry Operations in a stand alone sense and did not provide the depth to cover the<br />

needs <strong>of</strong> operations staff. A decision was made to draft the Capabilities in a similar format to<br />

Competency Standards used within the VET system, but with an emphasis on being practical,<br />

widely used and operationally based. Complementary assessment tools were also required to<br />

be developed along these lines. Technical input was provided by employees within VicForests<br />

who had current responsibility for training operations staff in the various areas covered by the<br />

capability.<br />

How then are operations staff classified? Such staff can include employees who are subpr<strong>of</strong>essional,<br />

i.e. who do not possess qualifications in forestry, but who are nonetheless<br />

increasingly doing work formerly completed by qualified foresters. Anecdotal comments<br />

suggest that the vision <strong>of</strong> an operations staff member who had a patch in the field and walked<br />

this as part <strong>of</strong> his work is no longer the case and in many instances they are now actually<br />

<strong>of</strong>fice bound and carrying out supervision <strong>of</strong> contractors or other junior staff.<br />

So how has the role <strong>of</strong> Operations staff changed over time? Whilst <strong>Foresters</strong> still retain<br />

control over key areas, there is now more significant delegation <strong>of</strong> duties to the subpr<strong>of</strong>essional<br />

groups. Experience with students who undertook the old Diploma in Forestry at<br />

The School <strong>of</strong> Forestry, Creswick, provided insights into the expanded range <strong>of</strong> responsibility<br />

being delegated to Operations staff. One observation was that everything in forestry had<br />

moved up a rung, with contractors carrying out a lot <strong>of</strong> work formerly performed by<br />

operations staff, and the operations staff doing a lot <strong>of</strong> the work formerly done by foresters.<br />

The title <strong>of</strong> “Forest Supervisor”, which many organisations now use, serves to remind us <strong>of</strong><br />

the expanded role <strong>of</strong> operations people generally.<br />

Whilst there remains a strong need for industry and company focused training for this group,<br />

their actual opportunities for training beyond their organisations are somewhat limited. The<br />

former Diploma in Forestry provided an opportunity for these people to acquire some<br />

pr<strong>of</strong>essional recognition, and the theoretical and practical foundations important to the sort <strong>of</strong><br />

work they were doing. The course provided opportunities for people who wished to pursue<br />

careers as technical supervisors in the forestry and forest industries and as such was more<br />

vocationally oriented.<br />

One feature <strong>of</strong> the course was the large number <strong>of</strong> students who were drawn from Forestry<br />

Operations within companies across <strong>Australia</strong> and who were sponsored by their employers.<br />

There was an acknowledgement that this sort <strong>of</strong> training was unique and also complementary<br />

to the sort <strong>of</strong> work they would be expected to do. The block release nature <strong>of</strong> the course also<br />

assisted in them getting time <strong>of</strong>f work to attend, particularly for those from interstate.<br />

With the demise <strong>of</strong> this program, there are fewer options for this group to receive training<br />

outside <strong>of</strong> their own organisations. This makes it more likely that training <strong>of</strong> this group is<br />

going to have to be provided “in house” by existing staff. As part <strong>of</strong> certification, forestry<br />

organisations must be able to show that appropriate training <strong>of</strong> staff has been carried out and<br />

ensure that this is done properly against correct performance measures.<br />

This necessity to train means that a framework must be used to ensure that such training<br />

covers all necessary areas <strong>of</strong> skill and knowledge. This raises a clear need for such training to<br />

be conducted against standards and measures. Whilst these exist in the form <strong>of</strong> various<br />

technical notes, publications and guidelines, they are not always user friendly. Hence,<br />

VicForests desire to ensure that their training <strong>of</strong> operations staff be conducted against a<br />

standardised set <strong>of</strong> capabilities reflecting current practice. VicForests quest is to ensure


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 105<br />

standards are adhered to via reference to Capability Statements. In the VET sector there are<br />

many thousands <strong>of</strong> nationally endorsed competencies which are grouped together into<br />

qualifications. Whilst many are generic in nature and rely on further interpretation down to a<br />

certain level, there are <strong>of</strong>ten specific things missing which are felt to be essential to a<br />

company being able to meet its compliance requirements. For this reason VicForests sought<br />

its own specific capabilities with which they can confidently undertake staff development.<br />

The Capability Statements developed for VicForests have been designed to broadly reflect the<br />

design <strong>of</strong> Competency standards in the VET system. The key structure should equate with<br />

overall VET standards but should also reflect detail based on the level <strong>of</strong> operations with<br />

which the company is involved. The capability statements themselves need to closely match<br />

the suite <strong>of</strong> tasks required to be carried out by operations staff in support <strong>of</strong> the overall<br />

company mission. For this reason they must highlight quite clearly the areas <strong>of</strong> skill and<br />

knowledge required and how this is to be applied in a range <strong>of</strong> contexts.<br />

In the case <strong>of</strong> the VicForests capabilities, the foundation for this was the Native Forest<br />

Silvicultural Guidelines produced by the then Department <strong>of</strong> Conservation and Natural<br />

Resources. These guidelines already contained excellent information on the knowledge and<br />

application <strong>of</strong> skills required within these areas <strong>of</strong> silvicutural activity. Whilst these served as<br />

a useful guide, the process involved in using them to deliver training was not well defined.<br />

The information they contain is sequentially based, and there is a lot <strong>of</strong> detail, but they were<br />

more <strong>of</strong> a reference and key points were not easily extracted for training purposes. For this<br />

reason the guides were used as the basis for the capability statements which were themselves<br />

based on the key result areas described in the guidelines.<br />

A Capability Statement and the first element relating to Browsing Management is provided<br />

below.<br />

BROWSING MANAGEMENT CAPABILITY STATEMENT<br />

Description<br />

This statement specifies the capability required to successfully manage the browsing<br />

<strong>of</strong> regeneration and reforestation sites and undertake appropriate risk assessment<br />

and monitoring activities necessary to mitigate the effects <strong>of</strong> browsing on company<br />

operations.<br />

Application <strong>of</strong> Capability<br />

<strong>Foresters</strong> within Vic Forests need to understand the impact <strong>of</strong> browsing on<br />

regeneration and reforestation operations and the necessary steps required to protect<br />

these operations from predation from a range <strong>of</strong> animals. This will include skills<br />

involved in the identification <strong>of</strong> animal species and trees involved, assessing types <strong>of</strong><br />

damage and planning, and implementing and assessing various control measures.<br />

This capability is required across selected coupe sites experiencing browsing<br />

pressure.<br />

Element 1: Assess Browsing activities<br />

Performance Measures<br />

1.1 Use knowledge <strong>of</strong> pertinent forest types and coupe locations to determine<br />

areas likely to be affected by browsing<br />

1.2 Investigate history <strong>of</strong> sites to determine browsing impact and likely affect on<br />

future operations<br />

1.3 Be able to identify the key browsing animals and the species <strong>of</strong> tree most<br />

susceptible to browsing predation


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 106<br />

Evidence Guidelines for Supervisor/assessor<br />

The following list provides guidance to the Supervisor/Assessor <strong>of</strong> the key evidence<br />

categories required to be matched in order to determine that the capability has been<br />

achieved. To demonstrate achievement against this capability, employees must be<br />

able to provide evidence that they have an appropriate knowledge <strong>of</strong> underlying<br />

principles relating to Browsing Management as well as an ability to apply this<br />

knowledge via necessary procedures and techniques. The Assessment Tool located<br />

at section 2 provides the Supervisor/Assessor with the necessary means to obtain<br />

this evidence via questioning, observed activities, projects etc.<br />

Knowledge Requirements:<br />

Forests types; Browsing animal types; Susceptible tree species; Browsing control<br />

methods<br />

Relevant codes <strong>of</strong> practice and instructions<br />

Skills and attributes:<br />

Ability to interpret key principles, codes <strong>of</strong> practice, instructions and written<br />

procedures<br />

Ability to apply the above in a practical context or in support <strong>of</strong> other operations<br />

High level communication skills to relate knowledge to others<br />

Analytical skills to apply techniques and procedures in a range <strong>of</strong> contexts<br />

Research/Investigation, planning and self-organisation skills<br />

Field identification <strong>of</strong> plants and animals<br />

Application <strong>of</strong> browsing risk assessment methods<br />

Practical examples <strong>of</strong> Evidence:<br />

Presentation <strong>of</strong> relevant sections <strong>of</strong> guides, instructions and procedures to support<br />

actions<br />

Demonstration <strong>of</strong> plant animal identification techniques in a field setting<br />

Record <strong>of</strong> application <strong>of</strong> browsing risk assessment method<br />

Examples <strong>of</strong> completed and verified remedial actions relating to Browsing<br />

Management<br />

Oral and written explanations <strong>of</strong> browsing control methods applied in different<br />

situations<br />

Resources Required for Evidence Collection<br />

Natural Forest Silviculture Guideline No 7<br />

Code <strong>of</strong> Forest Practices<br />

Copy <strong>of</strong> browsing risk assessment sheets<br />

Coupe plans<br />

Field location for practical demonstration <strong>of</strong> skills associated with Browsing<br />

Management<br />

Equipment used in browsing control<br />

In the above example, the Capability has been interpreted from the Natural Forest Silviculture<br />

guidelines and the following key sections developed. There is a general description which<br />

provides an overview <strong>of</strong> the capability and what the knowledge and skills within contribute to<br />

in terms <strong>of</strong> completing work place tasks. This is designed to enable a manager to get a<br />

snapshot <strong>of</strong> what the capability enables staff to do and how this fits in a particular location or<br />

context. There is also a brief guide on how this capability might be applied and a description<br />

<strong>of</strong> its application in terms <strong>of</strong> a particular location or context. The capability is divided into<br />

elements which enable its key parts to be described. This is important both in terms <strong>of</strong> being<br />

able to show what the capability actually this is required.<br />

Each element has key Performance Measures which give a clear indication <strong>of</strong> what a staff<br />

member has to be able to do, know and think in order to demonstrate the appropriate skills,<br />

knowledge and attitude to carry this out in a work setting. The emphasis on skills, knowledge<br />

and attitude is particularly important in specifying what the people carrying out the operations<br />

to an acceptable standard must be able to demonstrate. These are provided in a sequential and


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 107<br />

easily understood format so that a manager can look quickly at the required measures for each<br />

element.<br />

An Evidence Guide is included to assist the manager in the process <strong>of</strong> meeting each <strong>of</strong> the<br />

performance measures through submission <strong>of</strong> appropriate evidence. This gives listings under<br />

the headings <strong>of</strong> Knowledge Requirements, Skills and Attributes and Practical examples <strong>of</strong><br />

Evidence, and can be used as part <strong>of</strong> an assessment process as a guide for both the<br />

training/assessing manager and the staff member.<br />

To assist the manager’s use <strong>of</strong> these lists there is also a guide to the resources required for the<br />

evidence collection. This includes various internal and external documents such as the Native<br />

Forest Silviculture Guideline and legislation, as well as pertinent equipment. This is to ensure<br />

that the capabilities can be tested in the appropriate setting and the manager has recourse to<br />

these resources whilst planning and conducting assessment.<br />

Research undertaken as part <strong>of</strong> the Capability project has shown that whilst the overall<br />

standard <strong>of</strong> field training is good, it has tended to be more dependent on the style and whims<br />

<strong>of</strong> individual trainers rather than adhering to a set formula in terms <strong>of</strong> what each person needs<br />

to know and be able to do. Training in the operational setting does need to suit local<br />

conditions and requirements but still must be able to be delivered in such a way that it is seen<br />

to be adhering to major forestry and management principles, which in themselves are a key<br />

part <strong>of</strong> appropriate compliance and certification. Staff will only be well trained if they are<br />

trained correctly and this training can be tested to the satisfaction <strong>of</strong> management against<br />

objective results.<br />

This testing process has traditionally been left to managers who would mark <strong>of</strong>f someone’s<br />

development against internal performance management plans. The capabilities developed<br />

specifically for VicForests are able to be used across a range <strong>of</strong> operations in different<br />

locations and contexts, and provide better confirmation <strong>of</strong> the skills and knowledge being<br />

attained for the process <strong>of</strong> operational effectiveness. Testing <strong>of</strong> these capabilities had to be<br />

easily achievable by managers who, whilst well versed in their areas <strong>of</strong> expertise, were not<br />

necessarily experienced in the training function. VicForests also addressed this by arranging<br />

training for selected staff in units drawn from the TAA40104 Certificate IV in Training and<br />

Assessment. This was done prior to the Capability project to ensure that more staff<br />

responsible for training in the field have the necessary skills in training and assessment, which<br />

makes their background and ability in Forestry more easily deployable to the task <strong>of</strong> the<br />

ongoing development <strong>of</strong> staff.<br />

Whilst the training <strong>of</strong> trainers has been an important step, there will always be, due to<br />

transfers, changing roles and attrition, a group <strong>of</strong> people responsible for the training <strong>of</strong><br />

operations staff who have not had the benefit <strong>of</strong> this kind <strong>of</strong> development. In this instance,<br />

there is a great need for both the Capability Statement and associated Assessment Tool to be<br />

as complementary and self explanatory as possible. For this reason the language on the<br />

Capability statement was written to reflect the “jargon” <strong>of</strong> Forest Operations whilst at the<br />

same time being relatively easy to follow by people newer to the industry who were both<br />

training and being trained.<br />

In order to facilitate this process, a complementary and easy to use Assessment Tool was<br />

developed concurrently with the Capability statement. This is designed to be used by anyone<br />

with responsibility for training (see below). The tool commences with instructions to assist<br />

the Manager who is responsible for the assessment <strong>of</strong> the particular capability. This includes<br />

not only advice on the process <strong>of</strong> assessment but also measures to ensure that staff with<br />

existing skills and knowledge pertinent to the capability could produce this in lieu <strong>of</strong> having<br />

to undertake some or all <strong>of</strong> the Assessment tasks. This will also speed up the process, where<br />

the need to test against the capability remains but potentially avoids time consuming<br />

questioning, demonstrations etc.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 108<br />

The assessor is taken through the three recommended forums for the Capability (in this case<br />

Browsing Management) which are used to collect evidence <strong>of</strong> the staff member’s capacity to<br />

carry out all aspects.<br />

The interview, questioning and activities are all carried out under each element <strong>of</strong> the<br />

capability so that the assessor can be confident that evidence is collected against the whole<br />

capability and the staff member can also follow this process sequentially. Of critical<br />

importance to the process is the need to carefully record the results <strong>of</strong> the Assessment against<br />

the Capabilities Elements and Performance Measures. The Assessment Tools final section is a<br />

recording grid which mirrors the Capability Statement but also includes the methods <strong>of</strong><br />

assessment used, evidence submitted by the staff member and a confirmation that the<br />

capability was demonstrated against every element. In this way nothing can be left out and the<br />

trainer can feel confident in signing <strong>of</strong> the staff member.<br />

BROWSING MANAGEMENT ASSESSMENT TOOL<br />

Instructions for Assessor<br />

Depending on the nature <strong>of</strong> the duties required from the staff member you may need to<br />

assess against some or all <strong>of</strong> the elements <strong>of</strong> the Capability Statement. If you normally work<br />

with the applicant you will be able to complete the record <strong>of</strong> evidence progressively over an<br />

extended period. To undertake assessment you will need to see the candidate in a one on<br />

one capacity to enable questioning to be undertaken, skills to be demonstrated and any<br />

records <strong>of</strong> past work experience to be examined against the Capability. In some cases the<br />

staff member may be able to demonstrate that they already have the required capability or<br />

parts <strong>of</strong> it and this will need to be verified via past work records and supervisor verification.<br />

You should also make arrangements to observe the candidate undertaking tasks and<br />

applying knowledge in a practical setting.<br />

Evidence to be gathered<br />

The Assessor should aim to collect three types <strong>of</strong> evidence as part <strong>of</strong> the verification process<br />

against the Capability. One or more types are required against each performance measure<br />

and should be written beside the measure in the record <strong>of</strong> evidence column. Evidence will be<br />

collected via:<br />

Interview<br />

The candidate should undertake an initial interview with the assessor to determine their<br />

current duties experience. The process <strong>of</strong> assessment and evidence collection should<br />

also be explained and any previous work experience verified and recorded.<br />

Questioning:<br />

Questions are used to collect evidence about the candidate’s knowledge. These must be<br />

answered without help although the assessor may assist by clarifying questions.<br />

Activities:<br />

Candidates will be observed undertaking activities specified under each element <strong>of</strong> the<br />

assessment tool and demonstrating various skills in the workplace. The assessor will<br />

record all observations and demonstrations.<br />

Element 1: Assess Browsing activities<br />

Interview:<br />

1. What kinds <strong>of</strong> experience have you had in assessing the impact <strong>of</strong> browsing animals<br />

on regenerating Eucalypt forests?<br />

2. Can you produce verifiable records <strong>of</strong> past activities/achievements in this element?


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 109<br />

Questions:<br />

1. How does the forest type and location <strong>of</strong> a coupe influence the extent to which it is<br />

susceptible to browsing damage?<br />

2. In what ways can the history <strong>of</strong> a site be useful in determining future browsing<br />

potential?<br />

3. Describe some <strong>of</strong> the key browsing species within your region and the impact they<br />

have on various trees.<br />

Observation <strong>of</strong> Activities/Demonstration <strong>of</strong> Skills:<br />

1. Produce a copy <strong>of</strong> the relevant guidelines and instructions relating to Browsing<br />

Management. Show the assessor the pertinent sections relating to the management<br />

<strong>of</strong> biodiversity.<br />

2. In a field location, identify evidence <strong>of</strong> predation on production tree species.<br />

3. Using evidence such as tracks and scat, determine which browsing species are<br />

present within a coupe. ASSESSMENT RECORD<br />

Element 1: Assess Browsing activities<br />

Date <strong>of</strong> Assessment: Methods <strong>of</strong><br />

Assessment<br />

1.1 Use knowledge <strong>of</strong> pertinent forest types and coupe<br />

locations to determine areas likely to be affected by<br />

browsing<br />

1.2 Investigate history <strong>of</strong> sites to determine browsing<br />

impact and likely affect on future operations<br />

1.3 Be able to identify the key browsing animals and<br />

the species <strong>of</strong> tree most susceptible to browsing<br />

predation<br />

Interview<br />

Questioning<br />

Observation<br />

Demonstration<br />

Evidence<br />

supplied<br />

Capability<br />

demonstrated<br />

Yes/No<br />

Assessor Comments:<br />

________________________________________________________________<br />

________________________________________________________________<br />

________________________________________________________________<br />

________________________________________________________________<br />

________________________________________________________________<br />

________________________________________________________________<br />

________________________________________________________________<br />

Assessor Signature: ________________________________________<br />

Name: ___________________________________________________<br />

Workplace Coach/Supervisor Signature: ________________________<br />

Name: ___________________________________________________<br />

Date: Recommendation<br />

C / NYC


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 110<br />

The Capability Statements and Assessment Tools currently being trialed by VicForests are<br />

designed to support rather than replace conventional training. The time honoured process <strong>of</strong><br />

learning on the job that both graduates and other field workers must go through can be<br />

expected to continue, but it will now be conducted within a more identifiable framework<br />

provided by the Capabilities. The major change required will relate to the use <strong>of</strong> the<br />

Capability Statements to ensure that the training undertaken covers all required areas <strong>of</strong><br />

performance and that trainees are in a position at the end <strong>of</strong> this demonstrate achievement <strong>of</strong><br />

the Capability against all the required tasks on the Assessment Tool. Because the Capabilities<br />

were derived in large part from the DSE instructions already used by the company, they<br />

should not represent a large departure from current and accepted practices and if anything,<br />

should make the process <strong>of</strong> drawing necessary training approaches simpler as the Capabilities<br />

still sub head the key skill and knowledge areas.<br />

The trialing process at VicForests is seeking comment from trainers in the field on how well<br />

they have been able to apply training from the Capability Statements and how this has fitted<br />

in with the Assessment Tools. The project was careful not to be too proscriptive in terms <strong>of</strong><br />

the way training should be conducted, as this will obviously vary somewhat due to locations,<br />

personnel and work contexts, but having a clear structure and starting point as well as<br />

definitive final results will help the process greatly.<br />

The extent to which the Capabilities and Assessment Tools are easily used will be revealed by<br />

field trials currently being carried out by VicForests. This involves a number <strong>of</strong> managers<br />

using the newly developed format to carry out development activities with their own<br />

Operations Staff.<br />

Whilst it could not be expected to be a flawless process, it is anticipated that such an approach<br />

will boost an area <strong>of</strong> Forestry training in major need <strong>of</strong> attention. There is also a continuing<br />

challenge <strong>of</strong> how the sub-pr<strong>of</strong>essional groups can achieve recognition beyond validation<br />

against company Capabilities, but being able to undertake a formal process like this is a step<br />

in the right direction.<br />

REFERENCES<br />

VicForests <strong>Australia</strong> 2009, viewed 26 th February 2009,<br />

http://www.vicforests.com.au/<br />

Mark Poynter, Peter Fagg, 2005. Native Forest Silviculture Guideline No 7, Department <strong>of</strong><br />

Sustainability and Environment,


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 111<br />

PRACTICAL USE OF 3P SAMPLING IN FOREST INVENTORY<br />

Phil West 1<br />

ABSTRACT<br />

Before starting an inventory <strong>of</strong> a forest area, broad-scale information is <strong>of</strong>ten available<br />

about the forest area. Commonly, this is obtained through remote sensing, using aerial<br />

photography, airborne laser scanning or satellite imagery. In combination with ground<br />

measurement <strong>of</strong> a sample <strong>of</strong> stands, this information is used to determine the mean and<br />

confidence limit for whatever is being assessed in the inventory (stand sequestered carbon<br />

say); sampling with probability proportional to size, model-based sampling and stratified<br />

random sampling are techniques used in such cases. When this sort <strong>of</strong> prior information is<br />

not available, sampling with probability proportional to prediction (3P sampling) may<br />

serve a similar purpose; the alternative is simple random sampling, which is generally a<br />

far less efficient sampling technique. Previously, 3P sampling has required that each and<br />

every individual sampling unit in the entire population be visited during the sampling<br />

process; this has restricted its use in inventory to small populations only. Recent<br />

developments have removed this requirement. This paper describes those developments<br />

and applies 3P sampling to example forest populations.<br />

INTRODUCTION<br />

Forest inventory generally involves the selection <strong>of</strong> a sample <strong>of</strong> stands from throughout the total forest<br />

area being included in the inventory (that is, from throughout the forest ‘population’). In each <strong>of</strong> those<br />

sample stands, measurements are made on the ground <strong>of</strong> whatever it is it is desired to assess in the<br />

inventory, perhaps wood volume or the amount <strong>of</strong> carbon sequestered in the tree biomass <strong>of</strong> the forest.<br />

These sample data are then used to estimate the stand mean for the population, <strong>of</strong> whatever is being<br />

assessed, together with a confidence limit about that mean. When the mean and its confidence limit are<br />

multiplied by the area <strong>of</strong> the entire forest, an estimate is obtained <strong>of</strong> the amount available across the<br />

entire forest <strong>of</strong> whatever is being assessed.<br />

In forest inventory today, additional information about the forest is frequently obtained through remote<br />

sensing, using techniques such as aerial photography, airborne laser scanning or satellite imagery. This<br />

can provide certain measures <strong>of</strong> the forest at any point across its entire area, at least to the resolution<br />

possible with the remote sensing device; such measures might be the average height <strong>of</strong> the trees, their<br />

crown cover or some measure <strong>of</strong> the productive capacity <strong>of</strong> the forest at any point. These measures<br />

may be correlated, at least partially, with whatever it is that is being assessed in the inventory. For<br />

example, it would be expected that the amount <strong>of</strong> carbon sequestered at any point in the forest would<br />

be larger if the trees were taller at that point.<br />

Variables which are correlated with whatever is being assessed in inventory are known as ‘covariate’<br />

variables. When values <strong>of</strong> covariates are available for points right across the entire forest area, those<br />

values may be used to improve the ‘efficiency’ <strong>of</strong> the inventory. That is, they may allow use <strong>of</strong><br />

inventory methods which produce a lower confidence limit <strong>of</strong> the estimate <strong>of</strong> the mean with the same<br />

sample size. Given the substantial expense involved in forest inventory, a method which provides a<br />

more efficient result is clearly desirable. Inventory techniques such as ‘sampling with probability<br />

proportional to size’, ‘model-based sampling’ or ‘stratified random sampling’ all make use <strong>of</strong> such<br />

covariate information to provide a more efficient inventory. Descriptions <strong>of</strong> these techniques are<br />

available in texts on forest inventory (eg Schreuder et al. 1993, Shiver and Borders 1996, West 2009).<br />

1 SciWest Consulting, 16 Windsor Court, Goonellabah, NSW 2480 and School <strong>of</strong> Environmental Science and Management,<br />

Southern Cross University, Lismore, NSW 2480. Email: pwest@nor.com.au. Web: http://www.nor.com.au/users/pwest


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 112<br />

However, suppose that such covariate information is not available, whether from remote sensing or<br />

from any other source. Remote sensing is expensive (involving things like aircraft hire or purchase <strong>of</strong><br />

satellite imagery) and usually requires that specialists be employed to extract the required covariate<br />

information and make it available in a readily accessible form. Smaller forest owners may not be able<br />

to afford the expense, or, perhaps it is desired to undertake a preliminary inventory <strong>of</strong> a forest area, on<br />

a budget which precludes such expense.<br />

Under these circumstances, the most basic method by which a sample is selected in inventory, ‘simple<br />

random sampling’ may be used; in that type <strong>of</strong> sampling, any point in the entire forest area has the<br />

same chance <strong>of</strong> being included as part <strong>of</strong> the sample. An option is to use ‘sampling with probability<br />

proportional to prediction’ (<strong>of</strong>ten abbreviated as ‘3P sampling’). This does not require that covariate<br />

information be available from the entire inventory area. However, it has the potential to provide an<br />

estimate from the inventory which is just as efficient as that which might be obtained when the<br />

covariate information is available.<br />

Whilst 3P sampling has been used to some extent in forest inventory in the past, some restrictions in<br />

the way it has been practised has limited its usefulness generally. Recent developments are removing<br />

those restrictions. This paper describes those developments and considers some practical applications<br />

<strong>of</strong> the method.<br />

3P SAMPLING<br />

The 3P sampling method was invented in the 1960s by the American forest biometrician L.R.<br />

Grosenbaugh. Shiver and Borders (1996) and Iles (2003) give some <strong>of</strong> the background <strong>of</strong> how the<br />

method developed. In America particularly, it seems to have had some use in forest inventory.<br />

The method requires that a suitable covariate variable be identified, a variable which can be quickly<br />

and easily measured at any point in the forest. However, there is no need to have available values for<br />

that covariate variable measured right across the entire forest population. Instead, 3P sampling<br />

requires that the covariate be measured only for those sampling units (that is, points in the population<br />

which might constitute one <strong>of</strong> the inventory sample points) which are actually being considered for<br />

inclusion in the sample and only at the time the sample is actually being selected and measured in the<br />

field.<br />

Traditionally, the covariate variable used in 3P sampling has been a visual estimate in the field, by the<br />

person doing the sampling, <strong>of</strong> whatever is being assessed in the inventory (stand sequestered carbon,<br />

say); as long as the sampler has some reasonable experience <strong>of</strong> the forest concerned, such estimates<br />

should correlate quite well with the variable being assessed and so be suitable as covariate values.<br />

However, the present author has recognised that any other quickly and easily measured variable,<br />

which correlates to a reasonable extent with the variable being assessed, can be used as the covariate.<br />

Sample selection<br />

Before starting the selection <strong>of</strong> the sampling units to be included in a 3P sample, the sampler must<br />

obtain an estimate <strong>of</strong> the maximum and minimum values <strong>of</strong> the covariate which will be encountered<br />

anywhere in the population. These would usually be determined during a preliminary survey <strong>of</strong> the<br />

population. A decision has to be made also as to what sample size is desired for the inventory.<br />

Given these decisions, the sampler then sets out to select the 3P sample. In the field, the sampler visits<br />

a sampling unit. Whilst the choice <strong>of</strong> which sampling units are visited should be made objectively, the<br />

order in which they are visited is not particularly important. However, by the time the sample selection<br />

has been completed, the sampler should have visited sampling units spread widely throughout the<br />

entire population.<br />

As each sampling unit is visited, the sampler obtains the value <strong>of</strong> the covariate variable for it. This<br />

may be done by the ‘estimation by eye’ process, as discussed above, or by any other measurement<br />

method. The important thing is that the covariate value should be obtained quickly and easily, to<br />

minimise the time involved in doing so.<br />

Once the covariate value has been obtained for a sampling unit, the sampler selects a value chosen at<br />

random from within the range <strong>of</strong> the minimum and maximum values assumed for the covariate. To


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 113<br />

obtain random values in the field, the sampler might carry a calculator or laptop computer or might<br />

simply refer to a list <strong>of</strong> such values, printed before sampling started.<br />

If the covariate value is greater than or equal to the random value chosen for that sampling unit, the<br />

sampling unit is then included in the sample. The measurement crew accompanying the sampler would<br />

then measure the actual value, for that sampling unit, <strong>of</strong> the variable that is being assessed ultimately<br />

in the inventory.<br />

If the covariate value is less than the random value, that sampling unit is simply ignored and the<br />

sampler moves on to the next sampling unit. This process continues until the required number <strong>of</strong><br />

sampling units has been selected to make up the sample.<br />

The disadvantage <strong>of</strong> 3P sampling is that many sampling units may be visited and rejected from the<br />

sample. If an inventory is being carried out over a large forest area, a lot <strong>of</strong> time is usually spent by the<br />

sampler and the measuring crew moving around the forest to visit sampling units. In the case <strong>of</strong> 3P<br />

sampling, a lot <strong>of</strong> time and effort can be spent getting to a sampling unit, only to have it rejected<br />

immediately from the sample.<br />

However, it is not this disadvantage which has rendered 3P sampling to only have limited use in forest<br />

inventory to date. Over the years since it was introduced, a more or less ‘standard’ protocol (in the<br />

terminology <strong>of</strong> West 2005) has been developed by American foresters as to how 3P sampling should<br />

be done (e.g. Shiver and Borders 1996, Avery and Burkhart 2002, Iles 2003). In particular, this<br />

protocol has required that each and every sampling unit in the entire population be visited in the field<br />

and a covariate value be obtained for each. Of course this is quite impractical if the population is very<br />

large, with many thousands <strong>of</strong> sampling units. As well, this protocol has considered that the only way<br />

<strong>of</strong> obtaining the covariate value for any sampling unit is ‘estimation by eye’ <strong>of</strong> the variable being<br />

assessed ultimately by the inventory.<br />

It has now been recognised (West 2005, 2009) that there is an alternative way <strong>of</strong> viewing 3P sampling.<br />

This no longer requires that a covariate value be obtained for each and every sampling unit. Instead, it<br />

is necessary to obtain a covariate value only for those sampling units which are visited during the<br />

sample selection process. It recognises also that any variable is suitable as a covariate variable, not just<br />

an ‘estimated by eye’ variable, as long as the variable is quick and easy to measure and has some<br />

reasonable degree <strong>of</strong> correlation with whatever variable is being assessed ultimately by the inventory.<br />

The removal <strong>of</strong> both these constraints from the 3P sampling process renders the method generally<br />

more practical for use in forest inventory.<br />

The only disadvantage <strong>of</strong> West’s approach to 3P sampling is that the whole process is rendered invalid<br />

if, during the selection <strong>of</strong> the 3P sample, a sampling unit is encountered with a covariate value outside<br />

the range within which it was assumed they would lie. This problem is allowed for in the standard<br />

protocol.<br />

Calculation <strong>of</strong> results<br />

The key to West’s approach to 3P sampling is recognition that it is in fact a form <strong>of</strong> ‘sampling with<br />

variable probability <strong>of</strong> selection’. This means that different sampling units within the population are<br />

assigned different probabilities (that is, different chances) <strong>of</strong> being selected as part <strong>of</strong> the sample.<br />

Provided the probabilities are assigned appropriately, this can lead to a more efficient form <strong>of</strong><br />

sampling than simple random sampling.<br />

The properties and mathematical approach to sampling with varying probability <strong>of</strong> selection are well<br />

known (eg Schreuder et al. 1993). Earlier work on 3P sampling did not use this approach and unique<br />

mathematical treatments were developed for it (see Shivers and Borders 1996, pp. 303-305 and Avery<br />

and Burkhart 2002, pp. 262-263, as reiterated by West 2005).<br />

Suppose a forest population contains a total <strong>of</strong> N sampling units, any <strong>of</strong> which might be included in a<br />

sample <strong>of</strong> size n drawn from it. Suppose that cx and cm are the maximum and minimum, respectively,<br />

values, which are assumed to occur anywhere in the population, <strong>of</strong> a covariate variable. When 3P<br />

sampling is done, as described above, to select from a population a sample <strong>of</strong> size n, West (2005,


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 114<br />

2009) recognised that the i th (i=1…n) <strong>of</strong> those sampling units had a probability pi that it was included<br />

in the sample, where<br />

pi = (nv/N)(ci−cm)/(cx−cm) (1)<br />

where ci is the covariate value measured on that sampling unit and nv is the total number <strong>of</strong> sampling<br />

units which were visited in the 3P sampling process before the final n were selected in the sample.<br />

Suppose that a value yi <strong>of</strong> the variable that is being assessed in the inventory was then measured on the<br />

i th selected sampling unit. Following the theory for sampling with variable probability <strong>of</strong> selection, the<br />

estimate from the sample data <strong>of</strong> the population mean (YM) is then (as adapted from Eqs. 3.7 and 3.9<br />

<strong>of</strong> Schreuder et al. 1993)<br />

YM = [Σi=1...n (yi/pi)]/N (2)<br />

and <strong>of</strong> its variance (VM) is<br />

VM = {1/(2N 2 )}Σi,j=1...n, i≠j{[(pipj−pij)/pij][yi/pi−yj/pj] 2 } (3)<br />

where pij (j=1…n) is the probability that both sampling units i and j were included in the sample,<br />

where<br />

pij = pipjN(n−1)/[n(N−1)] (4)<br />

If it is now assumed that the population size is much greater than the sample size, that is, N»n, then<br />

with some algebraic manipulation <strong>of</strong> Eqs. (1−4), it can be shown that YM and VM can be approximated<br />

very closely as<br />

and<br />

YM ≈ [(cx−cm)/nv] Σi=1...n[yi/(ci−cm)] (5)<br />

VM ≈ {[cx−cm] 2 /[2(n−1)nv 2 ]} Σi,j=1...n, i≠j{[yi/(ci−cm)]−[yj/(cj−cm)]} 2 (6)<br />

In practice for most forest inventories, it is perfectly reasonable to assume that N»n, that is, to assume<br />

that only a relatively small sample is taken from a large forest population. Eqs. (5) and (6) are then <strong>of</strong><br />

particular interest, because neither involves the population size, N. This means that it is possible to<br />

apply 3P sampling without knowing the total population size, a matter overlooked by West (2005).<br />

However, it is clear from Eqs. (5) and (6) that the estimates <strong>of</strong> both the mean and its variance are<br />

functions <strong>of</strong> cx and cm, the maximum and minimum values <strong>of</strong> the covariate which were assumed to<br />

occur in the population. Thus, it will be necessary to consider carefully what values are to be used for<br />

those before setting out to take a 3P sample; that issue is considered below.<br />

The estimate <strong>of</strong> the confidence limit <strong>of</strong> the population mean (CM) may be determined as<br />

CM = t√VM<br />

(7)<br />

where t is Student’s t, for whatever probability level is desired and with (n−1) degrees <strong>of</strong> freedom.<br />

PRACTICAL EXAMPLES OF 3P SAMPLING<br />

The application <strong>of</strong> 3P sampling in practical forest inventory will be considered using data collected<br />

from two plantation forests. The first was a small (1 ha in area) planting <strong>of</strong> a mixture <strong>of</strong> rainforest<br />

species, established in 1997 on a farm a few kilometres from Alstonville (28°50’S, 153°26’W) in<br />

subtropical, north-eastern NSW. At the time <strong>of</strong> measurement the plantation was unthinned and had a<br />

stocking density <strong>of</strong> about 880 stems/ha.<br />

The second was a compartment (19 ha in area) <strong>of</strong> the 1996 Pinus radiata plantations <strong>of</strong> Forestry SA,<br />

established near Wandilo (37°43’S, 140°44’W) in south-eastern SA. The plantation had been thinned<br />

at 9 years <strong>of</strong> age to a residual stocking density <strong>of</strong> about 790 stems/ha. Both plantations were measured<br />

when they were 11 years old.<br />

A simple random sample <strong>of</strong> plots was selected from each plantation. Plots averaged about 0.01 ha in<br />

size in the mixed-rainforest plantation and about 0.02 ha in the P. radiata plantation. In each plot, the


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 115<br />

diameters at breast height over bark <strong>of</strong> all the trees in the plot were measured. The data were used to<br />

determine both the stand basal area over bark and the amount <strong>of</strong> carbon sequestered in the stand total<br />

tree biomass (above- and below-ground) <strong>of</strong> each plot. For the mixed-rainforest plantation, tree<br />

biomasses were estimated using the individual tree biomass function, for plantations <strong>of</strong> this nature,<br />

developed by Specht and West (2003); this function predicts total oven-dry biomass <strong>of</strong> a tree (bT,<br />

tonne) as<br />

bT = 0.000364d 2.00 (8)<br />

where d (cm) is stem diameter at breast height over-bark (cm). For the P. radiata plantation, tree<br />

above-ground biomasses were estimated using the individual tree biomass function for P. radiata in<br />

<strong>Australia</strong> (Snowdon et al. 2000, Table 1.4); this function predicts above-ground oven-dry biomass <strong>of</strong><br />

a tree (bA, tonne) as<br />

bA = 0.000118d 2.25 (9)<br />

Once this function had been used to estimate stand above-ground biomass (BA, tonne/ha), an<br />

additional function (Snowdon et al. 2000, Table 3.6) was then used to convert the above-ground<br />

biomass to a stand total oven-dry biomass (BT, tonne/ha) as<br />

BT = BA + 0.677BA 0.712 .(10)<br />

In both plantations, it was then assumed that 50% <strong>of</strong> oven-dry biomass was carbon (West 2009). A<br />

summary <strong>of</strong> the inventory data for both plantations is given in Table 1.<br />

These data were used as follows, to simulate results which might be obtained if 3P sampling was used<br />

instead <strong>of</strong> simple random sampling in these plantations. Firstly, the data sets were simply repeated<br />

many times to create, in effect, data collected from a very much larger simple random sample <strong>of</strong> the<br />

plantations. Stand basal area was used as the covariate variable, for selection <strong>of</strong> a 3P sample from this<br />

larger data set. Values were chosen for cx and cm, the maximum and minimum covariate values which<br />

were assumed to occur in the population. Then, a 3P sample <strong>of</strong> size n=20 was selected from these data<br />

using the procedures described above. If this was a real inventory, it might be envisaged that a quick<br />

and easy way to measure the covariate variable would be to take a point sample, as any point in the<br />

forest was being considered for inclusion in the sample. The number <strong>of</strong> plots which had to be<br />

considered (that is, ‘visited’), before this sample size was reached, was counted to give a value for nv.<br />

The estimates <strong>of</strong> the mean, variance and confidence limit were then determined from this 3P sample<br />

using Equations. (5-7).<br />

Table 2 shows an example <strong>of</strong> how this 3P sample selection was done, using the random sample data<br />

from the mixed-rainforest plantation. In this instance, values <strong>of</strong> cx=26.2 m 2 /ha and cm=2.2 m 2 /ha were<br />

assumed. As each plot was considered, a random value was selected within the range 2.2−26.2 m 2 /ha;<br />

the values selected are shown in the third column <strong>of</strong> the table. If the stand basal area <strong>of</strong> the plot<br />

exceeded the random value, the plot was included in the 3P sample; selected plots are indicated with a<br />

‘y’ in the fourth column <strong>of</strong> the table. In this case, nv=38 plots were ‘visited’ to obtain a sample size <strong>of</strong><br />

n=20; it is data for only those 38 plots which are shown in the table. The stand total carbon <strong>of</strong> the<br />

selected plots is shown in the last column <strong>of</strong> the table.<br />

Table 1. Summary <strong>of</strong> inventory data from two plantation forest blocks, using simple random<br />

sampling. Shown are sample sizes and the minimum−mean−maximum <strong>of</strong> stand basal<br />

area and estimates <strong>of</strong> total amount <strong>of</strong> carbon sequestered in the above- and below-<br />

ground biomass <strong>of</strong> the trees.<br />

Mixed-rainforest Pinus radiata<br />

Sample size 20 24<br />

Stand basal area (m 2 /ha) 3−14−20 18−24−30<br />

Stand total carbon (tonne/ha) 7−33−47 35−45−59


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 116<br />

Table 2. An example illustrating the selection <strong>of</strong> a 3P sample from a simple random sample<br />

from the mixed-rainforest plantation. In this case, it was assumed cx=26.2 m 2 /ha and<br />

cm=2.2 m 2 /ha. Plots included finally in the 3P sample are marked with a ‘y’ in the<br />

fourth column.<br />

Plot<br />

Stand basal<br />

area (m 2 /ha)<br />

Random<br />

value<br />

Included in 3P<br />

sample (y)<br />

Stand total carbon<br />

(tonne/ha)<br />

1 3.2 14.7<br />

2 18.0 19.2<br />

3 16.3 14.8 y 37.8<br />

4 13.3 7.1 y 30.9<br />

5 19.3 19.7<br />

6 3.2 11.5<br />

7 14.8 10.0 y 34.2<br />

8 18.6 10.2 y 43.0<br />

9 19.8 19.1 y 46.0<br />

10 10.1 4.8 y 23.4<br />

11 12.0 7.8 y 27.8<br />

12 12.4 22.3<br />

13 18.6 15.1 y 43.0<br />

14 9.9 18.2<br />

15 17.5 2.4 y 40.5<br />

16 12.4 19.1<br />

17 13.8 23.3<br />

18 14.6 23.8<br />

19 14.8 13.0 y 34.2<br />

20 3.2 16.9<br />

21 8.3 8.6<br />

22 13.3 10.6 y 30.9<br />

23 8.3 8.4<br />

24 10.1 12.6<br />

25 11.8 12.8<br />

26 15.6 2.8 y 36.0<br />

27 16.3 11.9 y 37.8<br />

28 15.6 17.1<br />

29 19.3 20.6<br />

30 18.0 3.1 y 41.8<br />

31 11.8 10.9 y 27.4<br />

32 10.1 7.2 y 23.4<br />

33 19.8 4.4 y 46.0<br />

34 20.1 13.4 y 46.7<br />

35 15.6 25.7<br />

36 10.1 25.2<br />

37 17.5 5.5 y 40.5<br />

38 18.6 12.9 y 43.0<br />

The 20 plots selected in this process were then used to determine the inventory results from the 3P<br />

sample, using Equations (5-7). The results are shown in the first column <strong>of</strong> values in Table 3. The<br />

corresponding results obtained from a simple random sample with n=20 are shown in the second<br />

column. Clearly, 3P sampling led to a much smaller 95% confidence limit than simple random<br />

sampling (0.7 tonne/ha <strong>of</strong> carbon as opposed to 2.7 tonne/ha). Similar results are shown for the P.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 117<br />

radiata plantation in the last two columns <strong>of</strong> Table 3. Again, 3P sampling was far more efficient than<br />

simple random sampling.<br />

For both the mixed-rainforest and P. radiata plantation data, similar results for 3P sampling were then<br />

obtained many times, but using different values <strong>of</strong> cx and cm each time to obtain different estimates <strong>of</strong><br />

the means and variances. It follows from Equations (5) and (6) that these will vary, depending on the<br />

values used for cx and cm. The effect <strong>of</strong> this on estimates <strong>of</strong> the confidence limit from the 3P samples is<br />

shown in Figure 1.<br />

Table 3. Estimates from inventory <strong>of</strong> the mean stand carbon amount, and its 95% confidence<br />

limit, for two plantation forest blocks using either 3P sampling or simple random<br />

sampling.<br />

(a)<br />

95% Confidence limit<br />

3<br />

2<br />

1<br />

0<br />

Mixed rainforest Pinus radiata<br />

3P<br />

sample<br />

Simple<br />

random<br />

sample<br />

3P<br />

sample<br />

Simple<br />

random<br />

sample<br />

cx (m 2 /ha) 26.2 - 32.6 -<br />

cm (m 2 /ha) 2.2 - 0 -<br />

n 20 20 20 20<br />

nv 38 - 26 -<br />

Mean carbon (tonne/ha) 34.3 32.3 47.0 46.7<br />

95% confidence limit (tonne/ha) 0.7 2.7 0.2 3.2<br />

3.2<br />

2.2<br />

1.0<br />

10 30 50 70<br />

cx-cm<br />

(b)<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

12.6<br />

9.0<br />

0 20 40 60 80 100<br />

Figure 1. Change in size <strong>of</strong> 95% confidence limit (tonne/ha) from 3P sampling as the range<br />

(cx−cm) changes for data from the (a) mixed-rainforest and (b) P. radiata plantations.<br />

The three solid lines ( ______ ) shown on each graph are for three different values <strong>of</strong> cm,<br />

which are marked adjacent to the line. The confidence limit from a simple random<br />

sample is shown also (- - -). In all cases, the sample size was n=20.<br />

95% Confidence limit<br />

cx-cm<br />

0


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 118<br />

Firstly, these results suggest that there is generally a slight tendency for the estimate <strong>of</strong> the confidence<br />

limit from 3P sampling to increase as the range within which the covariate values are assumed to lie<br />

(cx−cm) increases. Far greater than this is the effect on the number <strong>of</strong> sampling units which must be<br />

visited to obtain the requisite sample size <strong>of</strong> n=20; the number visited increased approximately in<br />

direct proportion to the width <strong>of</strong> the range, rising to well over 100 in both plantation types, at the<br />

largest values <strong>of</strong> (cx−cm) used here (data not shown). That is to say, the narrower the range chosen for<br />

(cx−cm), the fewer sampling units will have to be visited to obtain the 3P sample. However, the<br />

narrower the range, the greater the likelihood <strong>of</strong> encountering a sampling unit <strong>of</strong> which the covariate<br />

value lies outside the chosen range, a circumstance which, as discussed above, invalidates the present<br />

3P sampling process.<br />

What is particularly evident in the results <strong>of</strong> Fig. 1 is that the value chosen for cm has a very large<br />

effect on the estimate <strong>of</strong> the confidence limit from 3P sampling. In these two example cases, the closer<br />

the value chosen for cm approached the actual minimum stand basal area in the data set (3.2 m 2 /ha in<br />

the mixed-rainforest plantation and 18.0 m 2 /ha in the P. radiata plantation), the larger was the estimate<br />

<strong>of</strong> the confidence limit. In the P. radiata plantation, once a value <strong>of</strong> cm greater than about 10 m 2 /ha<br />

was used, the estimate <strong>of</strong> the confidence limit from the 3P sample was greater than that from simple<br />

random sampling.<br />

It will require further study <strong>of</strong> the theory developed here to understand why the value chosen for cm<br />

has such large effects on the estimate <strong>of</strong> the confidence limit from the 3P sample. Ultimately, it will be<br />

desirable to develop some system for choosing cm and cx, through some preliminary survey <strong>of</strong> the<br />

population before 3P sampling starts, so that the 3P sampling leads to the most efficient estimate<br />

possible <strong>of</strong> the desired population characteristic.<br />

CONCLUSIONS<br />

The results here suggest that 3P sampling does have potential as a sampling method for forest<br />

inventory, when suitable covariate information from the entire forest population is not available.<br />

Recognition <strong>of</strong> 3P sampling as a case <strong>of</strong> sampling with varying probability <strong>of</strong> selection has allowed it<br />

to be applied without the need to obtain, in the field, covariate information for each and every<br />

sampling unit in the population. In earlier consideration <strong>of</strong> the theory <strong>of</strong> 3P sampling, this limitation<br />

rendered the technique practical for use only in very small forest populations.<br />

However, the efficiency <strong>of</strong> 3P sampling as applied here was highly dependent on the value assumed to<br />

be the minimum covariate value which occurred in the population. Further theoretical studies may<br />

show exactly how and why this is so and may <strong>of</strong>fer ways to select the value to render the application<br />

<strong>of</strong> 3P sampling at its most efficient. Until such studies are undertaken, it may require preliminary<br />

simple random sampling <strong>of</strong> the population under consideration and use <strong>of</strong> simulation studies, similar<br />

to those done here, to determine empirically, for any population under consideration, the most<br />

appropriate minimum covariate value to use.<br />

ACKNOWLEDGEMENTS<br />

I am indebted to Messrs John Gough Snr and Jnr for access to their mixed-rainforest plantation and to<br />

Forestry SA for access to their P. radiata plantation. The data used here were collected by students in<br />

the 2007 and 2008 ‘Measuring Trees and Forests’ undergraduate classes <strong>of</strong> Southern Cross University.<br />

REFERENCES<br />

Avery, TE and Burkhart, HE 2002, Forest measurements, 5th edn, McGraw-Hill, New York.<br />

Gifford, RM 2000, Carbon contents <strong>of</strong> above-ground tissues <strong>of</strong> forest and woodland trees, National Carbon<br />

Accounting System Technical Report No 22, <strong>Australia</strong>n Greenhouse Office, Canberra.<br />

Iles, K 2003, A sampler <strong>of</strong> inventory topics, Kim Iles & Associates Ltd, Nanaimoa, British Columbia.<br />

Schreuder, HT, Gregoire, TG & Wood GB 1993, Sampling methods for multiresource forest inventor, Wiley,<br />

New York.<br />

Shiver, BD & Borders, BE 1996, Sampling techniques for forest resource inventory, Wiley, New York.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 119<br />

Snowdon, P, Eamus, D, Gibbons, P, Khanna, PK, Keith, H, Raison, RJ & Kirschbaum, MUF 2000, Synthesis <strong>of</strong><br />

allometrics, review <strong>of</strong> root biomass and design <strong>of</strong> future woody biomass sampling strategies, National<br />

Carbon Accounting System Technical Report No 17, <strong>Australia</strong>n Greenhouse Office, Canberra.<br />

Specht, A and West, PW 2003, ‘Estimation <strong>of</strong> biomass and sequestered carbon on farm forest plantations in<br />

northern New South Wales, <strong>Australia</strong>’, Biomass & Bioenergy, vol. 25, pp. 363–79.<br />

West, PW 2005, ‘An alternative approach to selecting a 3P sample’, Paper to a meeting, Western Forest<br />

Mensurationists, Hilo Hawaii, (Available at http://www.growthmodel.org/wmens/m2005/ west(a).pdf).<br />

West, PW 2009, Tree and forest measurement, 2nd edn, Springer, Berlin.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 120<br />

ESTIMATING FOREST CHARACTERISTICS IN<br />

YOUNG VICTORIAN ASH REGROWTH FORESTS<br />

USING FIELD PLOTS AND AIRBORNE LASER SCANNING DATA<br />

Andrew Haywood 1,2 and Michael Sutton 1<br />

ABSTRACT<br />

Airborne laser scanning data have the ability to measure the vertical and horizontal<br />

structure <strong>of</strong> forest vegetation. The aim <strong>of</strong> this study was to investigate how metrics<br />

derived from laser scanning data could be used in simple regression models for<br />

estimation <strong>of</strong> eucalypt top height, basal area and stems per hectare on 20m by 20m field<br />

plots. The study was located in the Central Forest Management Area in Victoria. The<br />

target population was ash dominated regrowth forests aged 20 to 60 years old. A linear<br />

regression function was able to provide precise estimates <strong>of</strong> eucalypt top height (R 2 =<br />

0.87; RMS error = 3.8 m) using a single height percentile variable. On the strength <strong>of</strong><br />

these results it should be possible to predict top height with a precision that is close to<br />

traditional field measurement methods. Regression estimates <strong>of</strong> eucalypt basal area<br />

were less precise (R 2 = 0.56; RMS error = 14.69 m 2 ) than the top height model. The<br />

model included both a height percentile and intensity variable. Regression modelling<br />

was unable to provide a reasonable estimate (R 2 = 0.41) <strong>of</strong> eucalypt stems per hectare<br />

using height percentiles, intensity and canopy structure variables. This model should be<br />

used with caution. Basal area and height models were applied across the target regrowth<br />

stands and used to estimate future timber volume from these areas.<br />

INTRODUCTION<br />

There is a need for accurate and up-to-date information on resource information for sustainable forest<br />

management <strong>of</strong> Victorian native forests. The Statewide Forest Resource Inventory (SFRI) program has<br />

delivered comprehensive mapping and sawlog volume information for mature and older regrowth<br />

forests (DNRE, 2000). However, the stand mapping associated with this process is based on aerial<br />

photography now in excess <strong>of</strong> 15 years old. In addition, the SFRI field sampling focused on stands <strong>of</strong><br />

1939 and older age classes. The younger regrowth stands were not sampled and are currently assumed<br />

to have the same productivity as the sampled 1939 (and older) stands. Following the Victorian fires <strong>of</strong><br />

2003 and 2006/07 there is an increased need for resource information to aid sustainable forest<br />

management. In particular, improved resource information on the younger (un-sampled) regrowth<br />

stands is required.<br />

Previous research has shown that laser scanning data can be used to provide accurate estimates <strong>of</strong><br />

forest structure variables that are currently measured using manual field collection and aerial photo<br />

interpretation methods (Naesset 1997, Hyyppä et al. 2000, Persson et al. 2002, Holmgren 2003).<br />

Forest estimates (<strong>of</strong> height, basal area and stems per hectare), using laser data, are <strong>of</strong>ten based on<br />

statistical measures derived from the distribution <strong>of</strong> laser point data. Magnussen and Boudewyn (1998)<br />

showed that for a given plot size and canopy structure, a certain percentile in the laser height<br />

distribution exists that corresponds to the canopy height <strong>of</strong> interest (i.e. top height). Other researchers<br />

have also noted that the inclusion <strong>of</strong> measures <strong>of</strong> canopy characteristics derived from the laser height<br />

distribution, in combination with selected laser height percentiles, have proven useful for estimating<br />

tree density (Naesset 2002), timber volume (Nelson et al. 1984; Means et al. 1999; Naesset & Økland<br />

2002) and identifying species (Donoghue et al., 2007).<br />

1<br />

Victorian Department <strong>of</strong> Sustainability and Environment, PO Box 500., East Melbourne, Victoria, 3002, <strong>Australia</strong><br />

2<br />

Corresponding author: E-mail: Andrew.Haywood@dse.vic.gov.au


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 121<br />

Much <strong>of</strong> this work has been focused on estimating forest attributes at a local scale. The development<br />

<strong>of</strong> new airborne laser scanning instruments allows forest information to be measured in three<br />

dimensions with precision over moderately large areas at low unit cost. The data derived from such<br />

laser scanning surveys are increasingly being used operationally over large areas, both internationally<br />

(Nilsson, 1996; Næsset, 2002; Watt, 2005; Holmgren & Wallerman, 2006) and domestically in<br />

<strong>Australia</strong> (Turner, 2007; Turner 2008).<br />

This project investigates the potential <strong>of</strong> airborne scanning and field data to provide improved younger<br />

ash regrowth resource information for the Central Forest Management Area (FMA) in Victoria,<br />

<strong>Australia</strong>.<br />

MATERIALS AND METHODS<br />

Study area<br />

The study area is located in the Central Forest Management Area (FMA) in Victoria (see Figure 1).<br />

The Central FMA covers some 290 000 hectares <strong>of</strong> public land located north <strong>of</strong> the Great Dividing<br />

Range, south <strong>of</strong> the Goulburn River and between the Mt. Disappointment forest near Broadford on the<br />

Hume Highway and the upper reaches <strong>of</strong> the Big River south <strong>of</strong> Lake Eildon. Of the forested land, one<br />

third is set aside for National or State Park and the remaining is designated State Forest. The area<br />

covers a variety <strong>of</strong> forest types, ranging from Ash species (mountain ash and alpine ash) in the higher<br />

elevations. As the elevation decreases, these forest structures are replaced by Messmate, and in drier<br />

areas, Peppermints, Boxes and Candlebark.<br />

Figure 1. The location <strong>of</strong> the Central Forest Management Area, Victoria <strong>Australia</strong>. The grid<br />

area represents the distribution <strong>of</strong> laser scanning data collected, the red circular points<br />

represents the field plot data collected.<br />

Field plot data<br />

The field data collected comprise forest plot measurements for young ash-regrowth forests in the<br />

Central FMA. A total <strong>of</strong> 88 square 0.04 hectare sample plots were measured. The sampling design was<br />

based on a stratified random sampling methodology. The stratification was based on existing SFRI<br />

data to include stands which regenerated between 1950 to 1989, species dominated by one or more <strong>of</strong><br />

mountain ash (Eucalyptus regnans), alpine ash (Eucalyptus delegatensis) or shining gum (Eucalyptus<br />

nitens) and currently zoned for commercial timber harvesting. The location <strong>of</strong> each field plot was<br />

acquired using differential GPS. In each plot eucalypt top height, basal area and stems per hectare<br />

were measured. Table 1 summarises the plot data.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 122<br />

Table. 1. Summary <strong>of</strong> 88 field plot measurements.<br />

Minimum Maximum Mean Standard Deviation<br />

Top Height (m) 14.2 63.9 39.2 10.4<br />

Basal Area (m2/ha) 1.2 93.1 42.5 22.1<br />

Stems per hectare 25.6 1472.4 408.4 295.7<br />

Laser scanning data acquisition<br />

High-density laser scanning data were acquired over the study area with an Optech ALTM 3100 EA<br />

system during December 2007 and January 2008. The system settings and flight parameters are shown<br />

in Table 2. The vendor provided raw laser scanning data consisting <strong>of</strong> XYZ coordinates, <strong>of</strong>f-nadir<br />

angle, and intensity for all laser returns within the area. In addition, the vendor provided a filtered<br />

ground data set consisting <strong>of</strong> points presumed to be measurements <strong>of</strong> the terrain surface, identified via<br />

a proprietary filtering algorithm. These filtered ground returns were used to generate a 1 m digital<br />

terrain model (DTM). The airborne laser scanning system collected up to four returns from each laser<br />

pulse, and all returns were used in this paper.<br />

Table 2. Laser scanning data specifications.<br />

Flying height 1 300 m Flight Direction NE/SW<br />

Swath width 945 m Laser pulse density 0.96 pulses/m 2<br />

Swath overlap 25% Laser footprint size 0.26 m<br />

Other optical data available included digital 1:40 000 scale aerial photography flown in January and<br />

March 2007. The aerial photography was used to assist in validation <strong>of</strong> the laser scanning processing.<br />

METHODS<br />

The analysis <strong>of</strong> the data was divided into a four-stage process, with the main objective being to<br />

evaluate their potential to provide improved resource information for the Central FMA.<br />

• Calculation <strong>of</strong> laser scanning plot-level variables, such as height percentiles and coefficient<br />

<strong>of</strong> variation <strong>of</strong> above ground pulse responses. Laser scanning data were also used to<br />

determine ground height within the plots.<br />

• Exploratory analysis <strong>of</strong> the two datasets (field measurements and laser scanning data) to<br />

investigate their underlying structure.<br />

• Bivariate regression methods to generate relationships at plot level between field<br />

measurements (<strong>of</strong> top height, basal area and stems per hectare) and height percentile laser<br />

estimates.<br />

• Multiple regression analysis to assess whether using more than one height percentile laser<br />

variable and/or intensity and canopy structure variables improved the relationship.<br />

Laser scanning data extraction<br />

The following variables were calculated from the laser scanning data and extracted over co-located<br />

field plots for quantitative analysis: laser height percentiles; mean intensity percentiles; standard<br />

deviation <strong>of</strong> laser dispersals; percentage <strong>of</strong> ground returns; coefficient <strong>of</strong> variation; skewness and<br />

kurtosis.<br />

Laser height percentiles provide information on the structure <strong>of</strong> the forest canopy at different canopy<br />

height levels. Using the laser scanning data, the pulses above 0.5 m were divided into quantiles<br />

corresponding to every 10th percentile from the 10th to the 100th, as well as the 5th, 95th and 99th<br />

percentiles. The 0.5 m cut-<strong>of</strong>f was used as a threshold to account for undulations in terrain. Other<br />

researchers have used higher height thresholds (up to 3 m) to eliminate laser returns from the<br />

understorey vegetation layer (Naesset 2002; Riaño et al. 2004). Lower thresholds are appropriate in<br />

Victorian ash forests, because <strong>of</strong> the generally high numbers <strong>of</strong> trees regenerated which leads to early


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 123<br />

forest canopy closure and limited development <strong>of</strong> the understorey. This provided 13 variables <strong>of</strong> an<br />

average laser canopy height by percentile.<br />

Like most discrete return systems, the Optech system records intensity for each pulse in the near infrared<br />

(1064 nm) region at a very narrow spectral width <strong>of</strong> 10 nm. The intensity <strong>of</strong> each return pulse,<br />

sometimes referred to as laser amplitude, represents the reflected energy from a highly culminated<br />

beam <strong>of</strong> light. It provides a concentrated measurement <strong>of</strong> the object's reflectivity unaffected by<br />

shadows or occlusions. This reflectance may vary based on the reflectance properties and porosity <strong>of</strong><br />

the targeted material, path length and incidence angle <strong>of</strong> the pulse. Accordingly, for this study the data<br />

are regarded as uncalibrated, and are only used as a relative measure <strong>of</strong> intensity. The mean intensity<br />

for each <strong>of</strong> the 13 height percentiles was calculated for each plot.<br />

The percentage <strong>of</strong> vegetation returns provides a measure <strong>of</strong> canopy density, and is calculated by<br />

dividing the sum <strong>of</strong> all vegetation returns with height values above 0.5 m by the total number <strong>of</strong><br />

returns. All returns above this threshold are considered to be canopy hits. Areas with low numbers <strong>of</strong><br />

vegetation returns will be those with sparser, more open canopies.<br />

The standard deviation <strong>of</strong> laser dispersals provides a simple measurement <strong>of</strong> the variation or dispersal<br />

within the laser height distribution <strong>of</strong> each field measurement plot.<br />

The coefficient <strong>of</strong> variation (CV) summarises the relative variation, or dispersion, <strong>of</strong> the laser height<br />

distribution within each sample plot. It is the ratio <strong>of</strong> standard deviation and mean, and is expressed as<br />

a percentage. As a measure <strong>of</strong> crown density, higher CV values indicate sparse, open canopies and low<br />

CV values dense, closed canopies (e.g.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 124<br />

Multiple regression analysis was conducted to determine if further variation in the models could be<br />

explained by the inclusion <strong>of</strong> laser scanning-derived measures <strong>of</strong> intensity and canopy<br />

structure/density.<br />

RESULTS<br />

Exploratory data analysis<br />

Summary statistics for the field measurement data have been given in Table 1. The histograms in<br />

Figure 2 illustrate their distribution.<br />

Figure 2 shows the general distribution <strong>of</strong> the field measurement data. Top height and basal area are<br />

noticeable for their relatively normal distributions. Stems per hectare demonstrate a positive skew. A<br />

single plot with stems per hectare greater than 1400 trees/ha has a major influence on the distribution.<br />

Table 3 shows the correlations between field data variables and a selection <strong>of</strong> laser scanning derived<br />

variables. The data show that basal area is positively correlated with stems per hectare. Top height is<br />

negatively correlated with basal area due to biological growth. Top height is found to be negatively<br />

correlated with stems per hectare; this is basically due to mortality increasing as the plots get older<br />

(taller).<br />

The laser scanning data can be broken into three categories: height percentiles; intensity<br />

measurements; and canopy density/structure measurements. The height percentiles are generally<br />

positively correlated with each other. As expected the 50th height percentile is positively correlated<br />

with basal area and top height and negatively correlated with stems per hectare. The laser scanning<br />

canopy density/structure metrics have a stronger correlation with stems per hectare than the height<br />

percentiles.<br />

Table 3. Summary <strong>of</strong> correlations between field measured and LiDAR-derived variables.<br />

Stems per hectare<br />

(SPH)<br />

-0.30 SPH<br />

Basal Area (m2/ha) 0.54 0.20 BA<br />

Standard Deviation 0.90 -0.40 0.51 sd<br />

50 th Height percentile 0.73 -0.09 0.70 0.63 p50<br />

Mean intensity -0.50 -0.08 -0.17 -0.42 -0.31 p50i<br />

Coefficient Variation 0.26 -0.37 -0.14 0.52 -0.25 -0.25 CV<br />

Percentage <strong>of</strong><br />

Ground Returns<br />

0.08 -0.20 -0.24 0.11 -0.19 -0.19 0.24 pcveg<br />

Skewness -0.14 -0.25 -0.42 0.06 -0.61 -0.20 0.73 0.46 skew<br />

Eucalypt top height prediction using laser height percentiles<br />

The percentile height with highest R 2 and lowest RMS error was selected as the predictor for<br />

estimation <strong>of</strong> top height. In this case, laser height values corresponding to the 95th percentile were<br />

used (Table 4). When the height percentile is greater than 70 % the differences in R 2 values are small,<br />

which implies that if separate models for each percentile (> 70th) were generated then each model<br />

would explain a similar amount <strong>of</strong> variance. However, if RMS error <strong>of</strong> each top height model is<br />

considered then differences between the percentiles become apparent (Table 4).<br />

Figure 3 illustrates there is a strong linear relationship (R 2 = 0.87, RMSE error = 3.89 m) between the<br />

95th height percentile and field-measured top height. With an R 2 <strong>of</strong> 0.87 using a single variable it is<br />

clear that a simple model that uses a single height percentile is the most effective approach to a<br />

predication <strong>of</strong> top height. A small number <strong>of</strong> plots had much higher laser percentile heights than field<br />

measured heights. After examination <strong>of</strong> the digital aerial photography this was explained by large over<br />

hanging canopies from trees outside the field measurement plots, and as a result these plots were not<br />

removed as outliers. Canopy density variables did not add any additional value to the predictive model<br />

in terms <strong>of</strong> explaining the remaining variation.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 125<br />

Table. 4. Summary <strong>of</strong> laser height percentiles used to estimate eucalypt top height.<br />

Height<br />

percentile<br />

Frequency<br />

Frequency<br />

Basal Area (m2/ha)<br />

0 5 10 15 20<br />

0 5 10 15 20 25<br />

R2 RMSE<br />

(m)<br />

Height<br />

percentile<br />

R2 RMSE<br />

(m)<br />

Height<br />

percentile<br />

R2 RMSE<br />

(m)<br />

1 0.0047 10.61 40 0.2794 9.06 90 0.8725 3.81<br />

5 0.0139 10.59 50 0.5464 7.19 95 0.8736 3.79<br />

10 0.0065 10.63 60 0.7989 4.78 99 0.8444 4.21<br />

20 0.0134 10.59 70 0.8488 4.15 Max height 0.8399 4.27<br />

30 0.0985 10.12 80 0.8667 3.90<br />

10 20 30 40 50 60<br />

Top height (m )<br />

0 500 1000 1500<br />

Stocking (stems/ha)<br />

Frequency<br />

Frequency<br />

0 5 10 15<br />

0 5 10 15 20<br />

0 20 40 60 80 100<br />

Basal Area (m2/ha)<br />

1955 1965 1975<br />

Ageclass<br />

1985<br />

Top height (m)<br />

20 30 40 50 60<br />

20 30 40 50 60<br />

Lidar 95th Percentile Height (m)<br />

Figure 2. Histograms <strong>of</strong> field plot data. Figure 3. Eucalypt top height against laser<br />

scanning 95 th height percentile.<br />

80<br />

60<br />

40<br />

20<br />

0<br />

10<br />

20<br />

30 40 50<br />

50th Lidar Percentile Height(m)<br />

Basal Area (m2/ha)<br />

80<br />

60<br />

40<br />

20<br />

0<br />

20 30 40 50<br />

Mean intensity for 99th laser Percentile Height(m)<br />

Figure 4. Eucalypt basal area against 50 th laser height percentile and mean intensity from 95 th<br />

height percentile.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 126<br />

Eucalypt basal area prediction using laser scanning data<br />

Multiple regression showed that the 50th height percentile and the mean intensity value <strong>of</strong> the 95th<br />

percentile were important variables. The model containing these variables had an R 2 <strong>of</strong> 0.56 with an<br />

RMS error <strong>of</strong> 14.69 m 2 . Figure 4 shows the relationship between basal area and the two significant<br />

variables, both are positively correlated with basal area. The laser height percentile can be interpreted<br />

as representing stand development. The biological interpretation <strong>of</strong> the mean intensity value <strong>of</strong> the<br />

95 th height percentile was aided by plotting it against the proportion <strong>of</strong> eucalypt stems to total stems<br />

per hectare (Figure 5). It can be seen that the intensity variable is positively correlated with the<br />

proportion <strong>of</strong> eucalypts per hectare.<br />

Figure 6 shows that there are no strong outliers in the dataset causing undue influence on the<br />

regression. There are no major patterns or structure in the residuals, which indicates that the model<br />

predicts basal area reasonably well at both high and low basal area.<br />

Eucalypt stems per hectare prediction using laser scanning data<br />

Analysis <strong>of</strong> the field measurement data indicates that the stems per hectare (sph) range from 26 to<br />

1472 sph and that the distribution is skewed to the left. As a consequence a square root transformation<br />

was made to stabilise the residuals and variance.<br />

Multiple regression analysis was used to investigate whether height percentiles, intensity and canopy<br />

density measures were useful in predicting sqrt(stems per hectare). Beginning with an empty model,<br />

all variables were considered but a variable was only added if the p-value fell below 5%. (i.e. where<br />

inclusion <strong>of</strong> a variable was warranted because it significantly improved the model). In all four<br />

variables were added to the model including two laser height percentile (5th and 100th), one measure<br />

<strong>of</strong> intensity (60th height) and one canopy structural variable (skewness). The model containing these<br />

variables had an R 2 <strong>of</strong> 0.41 with an RMS error <strong>of</strong> 5.581 (sqrt(sph)).<br />

Given the strong relationships between the different height percentiles and top height observed in the<br />

scatterplots in Figure 3, collinearity between variables in the model may be a problem. The presence<br />

<strong>of</strong> collinearity was assessed using variance inflation factor. According to Rabe-Hesketh and Everitt<br />

(2000) multicollinearity exists if VIF values are larger than 10 and if the mean VIF is larger than the<br />

VIF for individual variables. This was not the case for the variables in the model.<br />

It can be seen in Figure 7 there a are no strong outliers in the dataset causing undue influence on the<br />

regression. However, the RMS is high (5.581 sqrt(trees/ha)) limiting its practical use for providing<br />

precise estimates <strong>of</strong> tree density.<br />

Mean Intensity for 95th Height Percentile<br />

50<br />

40<br />

30<br />

20<br />

0.2 0.4 0.6 0.8 1.0<br />

Proportion <strong>of</strong> Eucalypt stems/ha to total stems/ha<br />

Figure 5. Proportion <strong>of</strong> eucalypt stems total<br />

stems versus intensity <strong>of</strong> 95 th height<br />

percentile.<br />

35<br />

30<br />

sqrt(Stems per Hectare (sph/ha))<br />

25<br />

20<br />

15<br />

10<br />

5<br />

10 15 20 25 30<br />

Predicted sqrt(Stems per Hectare (sph)<br />

Figure. 6. Predicted versus actual sqrt<br />

(eucalypt stems per hectare).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 127<br />

Basal Area (m2/ha)<br />

80<br />

60<br />

40<br />

20<br />

0<br />

20 30 40 50 60 70 80<br />

Predicted Basal Area (m2/ha)<br />

Figure 7. Predicted versus actual eucalypt basal area.<br />

DISCUSSION<br />

This project sought to evaluate the potential benefit <strong>of</strong> using field and airborne laser scanning data to<br />

improve forest resource estimates <strong>of</strong> young regrowth stands. It is anticipated that some <strong>of</strong> the methods<br />

described in this paper will be operationalised within Victoria and that laser scanning data will be used<br />

to provide stand-based estimates <strong>of</strong> eucalypt top height and basal area, which are key growth and yield<br />

model parameters. At this stage, for the forest plots studied, it appears that laser scanning data is able<br />

to provide reasonable estimates <strong>of</strong> eucalypt mean top height (R 2 =0.87), basal area (R 2 =0.56). The<br />

following discussion compares this study against results in the literature and discusses some<br />

operational considerations.<br />

Comparisons with related research<br />

Eucalypt top height<br />

Since laser scanning data are capable <strong>of</strong> providing a very accurate measure <strong>of</strong> distance, it is well suited<br />

to providing estimates <strong>of</strong> canopy height. Over forests, previous research has shown that for a given<br />

plot size and canopy structure, a certain percentile in the height distribution exists that corresponds to<br />

the canopy height <strong>of</strong> interest (Magnussen & Boudewyn 1998; Næsset & Økland 2002; Næsset &<br />

Bjerknes 2001).<br />

In this project a strong relationship (R 2 = 0.87; RMS error = 3.89 m) between eucalypt top height and<br />

laser derived height was observed above the 70th percentile; within this range little difference was<br />

observed between height percentiles. This is likely due to factors such as the density and homogeneity<br />

<strong>of</strong> the forest canopies studied (Yu et al., 2005). The results from this study suggest that the linear<br />

model is relatively insensitive to laser height distribution percentiles above 70%. Combined, these<br />

results agree with the best available science which shows that the accuracy <strong>of</strong> laser-derived height,<br />

after calibration to field height measurements, is comparable to that <strong>of</strong> manual field survey methods<br />

(Donoghue & Watt, 2006; Holmgren, 2003; Hyyppä et al., 2000; Næsset, 1997b; Persson et al., 2002;<br />

Watt, 2005; Lim et al., 2003; Lim & Treitz, 2004; Nelson et al., 2004; Popescu et al., 2003, 2004).<br />

It should be noted that in order to achieve a good level <strong>of</strong> accuracy the density <strong>of</strong> laser returns must be<br />

sufficient to (a) define the underlying terrain and (b) capture variations in terms <strong>of</strong> stand height and<br />

spatial arrangement. Generally a survey that records at least 4 returns/m 2 at a scan angle <strong>of</strong>


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 128<br />

height measure can be interpreted as being a measure <strong>of</strong> the development phase <strong>of</strong> the plot which is<br />

related directly to changes in basal area. The intensity variable is most likely reflecting the eucalypt<br />

crown cover. The estimates <strong>of</strong> basal area observed in this study are similar to other relevant research.<br />

In European conifer-dominated forest Næsset (2002) reported R 2 values <strong>of</strong> 0.86 for basal area in<br />

southern Norway. Lim et al. (2003) reported basal area estimates <strong>of</strong> R 2 = 0.86 in a Canadian hardwood<br />

forest and Lucas et al. (2006) reported R 2 = 0.61 from the <strong>Australia</strong>n forests.<br />

Eucalypt stems per hectare<br />

Eucalypt stems per hectare were not as reliably predicted using laser scanning measurements. A<br />

model, based on height percentiles, intensity and canopy structure variables provided the best model<br />

(R 2 <strong>of</strong> 0.41). However, the RMS error is relatively high (5.581 sqrt(sph)) limiting its practical use for<br />

providing accurate estimates <strong>of</strong> eucalypt tree density.<br />

This result is in contrast to other studies where the inclusion <strong>of</strong> measures <strong>of</strong> canopy characteristics<br />

derived from the laser height distribution, in combination with selected laser height percentiles, have<br />

proven useful for estimating tree density (Hudak et al. 2006; Næsset 2002; 2004a). One explanation is<br />

that in young regrowth <strong>of</strong>ten contain non-eucalypt species which influences the laser distribution at<br />

certain points in time. This makes it difficult to model eucalypt stems per hectare in young regrowth<br />

stands because while the distribution <strong>of</strong> the returns may change the eucalypt stems per hectare may<br />

not.<br />

Operational considerations<br />

It is likely that in the near future that laser scanning data will be used operationally to provide resource<br />

information to aid the sustainable management <strong>of</strong> Victoria’s commercial native forests. An appropriate<br />

approach would be to fly laser scanning instruments over GPS-located field plots measured in<br />

commercial forest areas. A number <strong>of</strong> findings identified as part <strong>of</strong> this study are considered<br />

applicable to an operational roll out:<br />

Estimate volume directly from the laser scanning data: It may be prudent to do this as the analysis<br />

shows that there are errors associated with each <strong>of</strong> the forest variables predicted. If these are used as<br />

inputs to another model then the errors may become additive.<br />

Refinement and validation <strong>of</strong> relationships: Previous research has indicated that if the physical<br />

structure <strong>of</strong> the forest changes then relationships may change. Therefore, it is expected that more than<br />

one estimation model (per variable) may be needed to adequately capture variation between<br />

silviculture regimes and forest areas. Validation <strong>of</strong> model predictions should also be frequently<br />

reviewed to ensure estimates are reliable.<br />

For future surveys, use laser scanning data to map the variation prior to establishing field plots: Since<br />

laser scanning data are sensitive to changes in forest structure and canopy cover, it would be possible<br />

to use these data to map forest variation. This information could be used to stratify the area prior to the<br />

forest inventory. From a modelling perspective this is important, and especially if estimates are to be<br />

extrapolated to areas where no field plots are available.<br />

Stems per hectare estimates: The methods tested in this study have shown that stems per hectare<br />

estimates are generally difficult to obtain from laser scanning data. An alternative approach would be<br />

to calculate stems per hectare by running an automated tree detection routine over the laser scanning<br />

dataset. Although single tree detection algorithm generally work best when detecting the number <strong>of</strong><br />

stems on sparse stands, the error is likely to be high in dense eucalypt forests, limiting its practical use.<br />

It would probably be difficult to use this technique on very dense stands (greater than 1,000 sph) due<br />

to the coalescing <strong>of</strong> crowns. Increasing the laser scanning point density is one option to improve the<br />

accuracy <strong>of</strong> the detection, but this may not be economically viable.<br />

CONCLUSION<br />

This study has demonstrated that airborne scanning data provide an alternative operational approach to<br />

estimate, at the plot-level, resource information for Victorian ash forest. Eucalypt top height and basal<br />

area can be determined with acceptable accuracy.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 129<br />

REFERENCES<br />

DNRE, 2000. State-wide Forest Resource Inventory: Forest Stand Classification. Forest Services Technical<br />

Reports 00-1. Department <strong>of</strong> Natural Resources and Environment Victoria.<br />

Donoghue, D.N.M., Watt, P.J., Cox, N.J., Wilson, J., 2007. Remote sensing <strong>of</strong> species mixtures in conifer<br />

plantations using LiDAR height and intensity data. Remote Sensing <strong>of</strong> Environment 110, 509-522.<br />

Donoghue, D.N.M. and Watt, P.J., 2006. Using LiDAR to compare forest height estimates from IKONOS and<br />

Landsat ETM+ data in Sitka spruce plantations. International Journal <strong>of</strong> Remote Sensing. 27, 2161-217.<br />

Holmgren, J., Nilsson, M., Olsson, H., 2003. Estimation <strong>of</strong> tree height and stem volume on plots using airborne<br />

laser scanning. Forest Science 49, 419-428.<br />

Holmgren, J. and Wallerman, J., 2006. Estimation <strong>of</strong> tree size distribution by combining vertical and horizontal<br />

distribution <strong>of</strong> LIDAR measurements with extraction <strong>of</strong> individual trees. Proceedings from the EARSeL<br />

ISPRS Workshop, 3D Remote Sensing in Forestry, Vienna, 14-15 February 2006.<br />

Hyyppä, J., Hyyppä, H, Ruppert, G., 2000. Automatic Derivation <strong>of</strong> Features Related to Forest Stand Attributes<br />

Using Laser Scanner. IAPRS 2000 33, 421-428.<br />

Lim, K., Treitz, P., Wulder, M., St-Onge , B., Flood, M., 2003. LiDAR remote sensing <strong>of</strong> forest structure,<br />

Progress in Physical Geography 27 88–106.<br />

Lim, K., and P. Treitz, 2004. Estimation <strong>of</strong> aboveground forest biomass from airborne discrete return laser<br />

scanner data using canopy-based quantile estimators. Scandinavian Journal <strong>of</strong> Forest Research 19, 558-<br />

570.<br />

Lucas, R.M., Cronin, N., Lee, A., Moghaddam, M., Witte, C., and Tickle, P. 2006. Empirical relationships<br />

between AIRSAR backscatter and LiDAR-derived forest biomass, Queensland, <strong>Australia</strong>. Remote Sensing<br />

<strong>of</strong> Environment 100, 407–425.<br />

Næsset, E., 1997. Determination <strong>of</strong> mean tree height <strong>of</strong> forest stands using airborne laser scanner data. ISPRS<br />

Journal <strong>of</strong> Photogrammetry and Remote Sensing 52, 49-56.<br />

Næsset, E., 2002. Predicting forest stand characteristics with airborne scanning laser using a practical two-stage<br />

procedure and field data. Remote Sensing <strong>of</strong> Environment 80, 88-99.<br />

Næsset, E., 2004. Estimation <strong>of</strong> Above- and Below-Ground Biomass in Boreal Forest Ecosystems. In:<br />

Proceedings <strong>of</strong> the ISPRS working group VIII/2, Laser-Scanners for Forest and Landscape, Freiburg,<br />

Germany 03-06 October 2004, 145-148.<br />

Næsset, E. amd Bjerknes, K.O. 2001. Estimating tree heights and number <strong>of</strong> stems in young forest stands using<br />

airborne laser scanner data. - Remote Sensing <strong>of</strong> Environment 78, 328-340.<br />

Næsset, E. and Okland, T. 2002. Estimating tree height and tree crown properties using airborne scanning laser<br />

in a boreal nature reserve. Remote Sensing <strong>of</strong> Environment 79, 105–115.<br />

Nelson, R., Krabill, W., Maclean, G. 1984. Determining forest canopy characteristics using airborne laser data.<br />

Remote Sensing <strong>of</strong> Environment 15, 201-212.<br />

Nelson, R., Oderwald, R., Gregoire, T. G. 1997. Separating the ground and airborne laser sampling phases to<br />

estimate tropical forest basal area, volume, and biomass. Remote Sensing <strong>of</strong> Environment, 60, 311-326.<br />

Nilsson, M. 1996. Estimation <strong>of</strong> tree heights and stand volume using an airborne lidar system. Remote Sensing<br />

<strong>of</strong> Environment 56, 1-7.<br />

Persson, Å., Holmgren, J., Soderman, U., 2002. Detecting and measuring individual trees using an airborne laser<br />

scanner, Photogrammetric Engineering and Remote Sensing 68, 925-932.<br />

Popescu, S.C., Wynne, R.H., Nelson, R.F., .2002. Estimating plot-level tree heights with lidar: local filtering<br />

with a canopy-height based variable window size, Computers and Electronics in Agriculture 31, 71-95.<br />

Popescu, S.C., Wynne, R.H., Nelson, R.F., 2003. Measuring individual tree crown diameter with LiDAR and<br />

assessing its influence on estimating forest volume and biomass, Canadian Journal <strong>of</strong> Remote Sensing.<br />

Vol. 29(5): 564 - 577<br />

Riaño, D., Meier, E., Allgöwer, B., Chuvieco, E.,Ustin, S.L., 2003. Modelling airborne laser scanning data for<br />

the spatial generation <strong>of</strong> critical forest parameters in fire behaviour modelling. Remote Sensing<br />

Environment 86, 177-186.<br />

Stephens, P.R., Watt, P.J., Loubser, D., Kimberley, M.O. 2007. Estimation <strong>of</strong> carbon stocks in New Zealand<br />

planted forests using airborne scanning LiDAR. In proceedings <strong>of</strong> IAPRS Volume XXXVI, Part 3 / W52.<br />

Turner, R. 2007. An overview <strong>of</strong> Airborne LiDAR applications in New South Wales state forests. In proceedings<br />

<strong>of</strong> ANZIF 2007 Growing Forest Values conference, C<strong>of</strong>fs Harbour.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 130<br />

Turner, R. 2008 Applications for remote sensing tools in forestry operations. In proceedings <strong>of</strong> ForestTECH<br />

2008 Conference – Albury, NSW, <strong>Australia</strong> – April 2008.<br />

Watt, P.J., 2005. An evaluation <strong>of</strong> LiDAR and optical satellite data for the measurement <strong>of</strong> structural attributes<br />

in British upland conifer plantation forestry. Doctoral thesis. Department <strong>of</strong> Geography, University <strong>of</strong><br />

Durham, England.<br />

Watt, P.J., and Haywood, A. 2006. Evaluation <strong>of</strong> airborne scanning LiDAR generated data as input into<br />

biomass/carbon models. Pöyry Forest Industry, contract report 38A08068 to Ministry for the Environment,<br />

Wellington, New Zealand.<br />

Yu, X., Hyyppä, J., Kaartinen, H., Maltamo, M., 2004. Automatic Detection <strong>of</strong> Harvested Trees and<br />

Determination <strong>of</strong> Forest Growth Using Airborne Laser Scanning. Remote Sensing <strong>of</strong> Environment 90, 451-<br />

-462.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 131<br />

OBJECT-BASED ANALYSIS OF FOREST STAND DELINEATION<br />

ON HIGH SPATIAL RESOLUTION IMAGERY<br />

USING OPEN SOURCE SOFTWARE<br />

ABSTRACT<br />

Andrew Haywood 1 and Christine Stone 2<br />

For decades, aerial photo interpretation has been, and to a good extent is still, the<br />

method <strong>of</strong> choice for producing fine-scale native forest stand mapping. Recent<br />

computer techniques have eased the task <strong>of</strong> the interpreter, who is now able to delineate<br />

polygons through on-screen digitising in a Geographical Information System (GIS)<br />

environment. Even with these advances, a great deal <strong>of</strong> skill is required in the polygon<br />

delineation. In an effort to contribute to the automation <strong>of</strong> this process, we introduce an<br />

open-source object based solution to the mapping <strong>of</strong> forest stand boundaries using<br />

attributes derived from digital aerial photography and laser scanning data (lidar)<br />

acquired over a study area in the Victorian Central Highlands. This methodology<br />

transforms remotely sensed imagery (single or multi-channel) and laser scanning<br />

derived canopy raster layers into polygon vector layers. It is intended that the resultant<br />

polygon layer should resemble the product derived by an aerial interpreter, without any<br />

prior knowledge <strong>of</strong> the scene. The derived product aims to produce a layer comprising<br />

<strong>of</strong> relatively homogeneous polygons all exceeding a minimum size. This layer can be<br />

used as a template by the interpreter. The interpreter can then aggregate (and sometimes<br />

correct) pre-delineated regions by simple drag-and-click operations. The derived<br />

product is only meant to be an intermediate aid for the work <strong>of</strong> the interpreter. The<br />

relationship between spectral, texture and laser scanning derived features for forest<br />

stand boundary delineation, and human interpreted boundaries is not straight forward.<br />

However, this approach is relatively cheap and flexible, being a workable compromise<br />

between (expensive) s<strong>of</strong>tware automation and interpreter supervision for classifying<br />

native forest landscapes. Preliminary results are encouraging; both in regard to<br />

automating the process and the delivery <strong>of</strong> robust delineation <strong>of</strong> stand boundaries.<br />

Future research will focus on appropriate input resolution to reduce computation<br />

requirements and improved data fusion methods to obtain more accurate forest stand<br />

delineation.<br />

INTRODUCTION<br />

Approximately a third (8.3 million hectares) <strong>of</strong> Victoria’s land mass is covered by forest, <strong>of</strong> which 3.4<br />

million hectares is classified as State Forest and 3.7 million hectares as national parks and other<br />

reserves. Privately owned forest accounts for 1.2 million hectares, <strong>of</strong> largely native forest and 360 000<br />

hectares <strong>of</strong> plantations (predominantly Pinus radiata and Eucalyptus globulus) (DSE 2005).<br />

The State Forests are managed for wood production and the provision <strong>of</strong> non-wood production values<br />

including recreation, biological and landscape diversity. Since these forests provide many functions<br />

that are important to society, there is a necessity to monitor their sustainable management and to<br />

understand the causes <strong>of</strong> change. At the time <strong>of</strong> publication, the Victorian Department <strong>of</strong><br />

Sustainability and Environment (DSE) was responsible for the sustainable management <strong>of</strong> public land<br />

in Victoria, including the public forest estate. As a consequence, DSE engages in a number <strong>of</strong><br />

processes to monitor the sustainability <strong>of</strong> Victoria’s forests. These include Victoria’s State <strong>of</strong> the<br />

Forests Report (produced every five years), Sustainability Charter for Victoria’s State Forests (DSE<br />

2006) and associated Criteria and Indicators for Sustainable Forest Management in Victorian Forests<br />

(DSE 2007). These mechanisms are designed to enable Victoria to critically assess and evaluate<br />

progress towards achieving its sustainable forest management objectives and targets. The State <strong>of</strong> the<br />

1<br />

Victorian Department <strong>of</strong> Sustainability and Environment, PO Box 500, East Melbourne, Victoria, 3002, <strong>Australia</strong>. Email:<br />

Andrew.Haywood@dse.vic.gov.au.<br />

2<br />

Forest Science Centre, NSW Department <strong>of</strong> Industry and Investment, PO Box 100, Beecr<strong>of</strong>t, NSW, 2119, <strong>Australia</strong>.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 132<br />

Forests Report seeks to include all forest land tenures in Victoria. However, a focus on Victoria’s State<br />

Forests is evident due to the limited data available on other land tenures. This is an issue that needs to<br />

be addressed.<br />

Forest stand delineation is one <strong>of</strong> the key building blocks for both forest management and monitoring<br />

and reporting. The process <strong>of</strong> delineating forest stands should allow the 1) production <strong>of</strong> forest maps<br />

precise enough to derive wood volume estimates, 2) characterisation <strong>of</strong> ecosystem structures, 3)<br />

production <strong>of</strong> forest structure maps useful for habitat and biodiversity evaluation. Within Victoria’s<br />

State Forests, stands are currently mapped using aerial photography interpretation (API) in<br />

combination with field surveys and are manually mapped in accordance with the State-wide forest<br />

resource inventory (SFRI) Forest Stand Classification (DNRE 2000). This approach acknowledges<br />

three primary stand components: eucalypt species, eucalypt structure and disturbance. Eucalypt<br />

structure comprises <strong>of</strong> crown cover, cover form, stand height and crown size. All <strong>of</strong> these attributes<br />

are visually assessed and assimilated into a single code for each forest stand by the photo interpreters.<br />

The resulting map classes are called forest stand classes. Manual delineation and interpretation <strong>of</strong> air<br />

photos is a consuming process and accuracies are dependant <strong>of</strong> the skills and experience <strong>of</strong> the photo<br />

interpreters. The majority <strong>of</strong> this mapping has occurred within State Forests and is increasingly<br />

becoming outdated. In addition, there are major mapping gaps across other land tenures.<br />

Internationally it has been recognised that current forest mapping procedures are falling short <strong>of</strong> forest<br />

agency expectations for three reasons: 1) the decreasing availability <strong>of</strong> trained API analysts; 2) the<br />

laborious, time consuming nature <strong>of</strong> manual feature identification and 3) the high labour costs<br />

involved (e.g. Leckie et al. 2003; Opitz and Blundell 2008; Wulder et al. 2008a and b). As a<br />

consequence, cost effective, semi-automated stand classification mapping techniques to supplement<br />

traditional methods need to be developed (e.g. Leckie et al. 1998; 2003; Hay et al. 2005; Wulder et al.<br />

2008a and b).<br />

Over the last two decades here has been considerable research into semi-automated forest mapping<br />

based on remotely sensed data, with varying success (e.g. Kilpeläinen and Tokala 1999; Almeida-<br />

Filho and Shimabukuro 2002; Flanders et al. 2003; Leckie et al. 2003; Tiede et al. 2004; Zhang et al.<br />

2004; Hay et al. 2005; Chubey et al. 2006; Antonarakis et al. 2008; Castilla et al. 2008; Pascual et al.<br />

2008; Wulder et al. 2008a and b). For the purposes <strong>of</strong> this study it is hypothesised that spectral,<br />

textural and laser scanning attributes can be utilised in a practical open-source solution to semiautomate<br />

forest stand delineation in Victoria’s natural forests.<br />

SPECTRAL INDICES IN FOREST STAND DELINEATION<br />

Spectral reflectance patterns <strong>of</strong> forest vegetation in the visible spectrum is generally controlled by the<br />

absorption features related to chlorophyll content. In contrast, the spectrum in the near infrared region<br />

is generally influenced by water content and the contribution <strong>of</strong> other organic materials. Different<br />

vegetation types have characteristic reflectance spectra. Simple image transformations have been<br />

useful in classifying imagery into discrete categories. For instance, band ratios have been used to<br />

detect and map different vegetation types extracted from remotely acquired spectral imagery (e.g.<br />

Gamon et al. 1995). The Normalised Difference Vegetation Index (NDVI) is a commonly used index<br />

that provides a measure <strong>of</strong> the amount and vigour <strong>of</strong> vegetation at the surface (e.g. Gamon et al. 1995).<br />

Unfortunately in complex eucalypt forests the water content and contribution <strong>of</strong> chlorophyll are<br />

spatially and temporally dynamic even within the same species and this can cause considerable<br />

variation in reflectance values (Goodwin et al. 2005). In addition, the reflectance <strong>of</strong> two different<br />

species can be very similar (Coops et al. 2004). As a consequence it is likely that spectral indices<br />

alone will not be sufficient to successfully delineate forest stands and that additional information such<br />

as texture indices and Laser scanning derived indices will be required.<br />

TEXTURE INDICES IN FOREST STAND DELINEATION<br />

The structurally complex and dynamic nature <strong>of</strong> native forests limits the ability <strong>of</strong> traditional statistical<br />

classification procedures to create stand level polygons based solely on the spectral response pattern at<br />

individual pixel locations. Spectral indices are a cumulative measure <strong>of</strong> all components within a pixel<br />

that influence the spectral values. In addition to chlorophyll and water content, factors such as


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 133<br />

difference in crown closure, shadows, or stand density may result in markedly different stand<br />

structures, yet may still be represented by the same spectral value. The increased availability <strong>of</strong> higher<br />

spatial resolution sensors has provided imagery commonly having a pixel size smaller than the object<br />

<strong>of</strong> interest. This can result in visually discernable objects being populated with pixels having a high<br />

degree <strong>of</strong> spectral variation producing a ‘salt and pepper’ effect. Per-pixel image classifications are<br />

commonly applied at low or moderate spatial resolution but sometimes do not operate well at high<br />

spatial resolution. Classification accuracies, however, have increased through incorporating the ideas<br />

<strong>of</strong> image texture, shape and context. Texture, in the context <strong>of</strong> digital processing, is the spatial<br />

variability <strong>of</strong> image tones which describes the relationship between elements <strong>of</strong> surface cover (Wulder<br />

et al. 1998). The above-ground organisation <strong>of</strong> forest elements or forest structure is accordingly<br />

represented in texture. There are several approaches to texture processing. Commonly texture<br />

measures are derived from moving a fixed-size, odd-numbered window through the image and<br />

calculating a variety <strong>of</strong> different statistics such standard deviation (first-order) or second-order<br />

statistics derived from image spatial co-occurrence (Haralick et al. 1973) or spatial autocorrelation<br />

functions (e.g. semivariance). Numerous studies have added texture transformations with the<br />

spectrally derived indices and improved forest stand classification (e.g. Cohen et al. 1990; Gong et al.<br />

1992; Franklin and McDermid 1993; Ryherd and Woodcock 1996; Wulder et al. 1998; Franklin et al.<br />

2001; Coburn & Roberts 2004; Zhang et al. 2004).<br />

LASER SCANNING INDICES IN FOREST STAND DELINEATION<br />

Discrete-return lidar sensors operate by rapidly emitting a laser pulse toward a target (e.g. a forest<br />

stand) and recording the time, location and quantity <strong>of</strong> the reflected energy. Most attention has focused<br />

upon using lidar as a tool for characteristing vertical forest structure – primarily the estimation <strong>of</strong> tree<br />

and stand heights with volume and biomass also <strong>of</strong> interest (e.g. Rooker et al. 2006; Pascual et al.<br />

2008). A number studies have investigated fusing remotely sensed spectral imagery and laser scanning<br />

data and then statistically combining metrics derived from both sources <strong>of</strong> data for stand delineation<br />

(e.g. Hill and Thomson 2005; Hyde et al. 2006). There are two different sources <strong>of</strong> laser scanning<br />

information that can be used to stratify forest areas by canopy structure and/or composition.. Both can<br />

be applied at the tree or plot/stand level. The first, and most common, uses statitical measures based on<br />

the distribution <strong>of</strong> pulse laser returns from the forest canopy (Evans et al. 2006; Donoghue et al. 2007;<br />

Pascual et al. 2008). This assumes that the distribution <strong>of</strong> pulses will be different for forest class and /<br />

or density due to canopy structural characteristics such as canopy porosity, branching and form. The<br />

second alternative uses the near infrared intensity <strong>of</strong> the return from the forest canopy to help<br />

differentiate forest types. For discrete Laser scanning systems the intensity value (sometimes referred<br />

to as the amplitude) <strong>of</strong> each return pulse provides a measure <strong>of</strong> the amount <strong>of</strong> energy reflected from a<br />

target. Intensity values are influenced not only by sensor specifications and atmospheric conditions but<br />

also the reflectivity <strong>of</strong> the target in the near infrared range <strong>of</strong> wavelengths (Donoghue et al. 2007). It<br />

is assumed therefore that lidar intensity data contain information relating to forest type and condition<br />

(Kim et al. 2009).<br />

IMAGE SEGMENTATION IN FOREST STAND DELINEATION<br />

After spectral, texture and laser scanning pre-processing are conducted, further processing is required<br />

to achieve stand delineation, a task for which a range <strong>of</strong> different image segmentation methods can be<br />

employed (e.g. Wulder et al. 2008a). Image segmentation is the partitioning <strong>of</strong> a digital image into a<br />

set <strong>of</strong> jointly exhaustive and mutually disjointed regions than are more uniform within themselves than<br />

when compared to adjacent regions (Wulder et al. 2008a and b). In addition to the limitations already<br />

discussed with pixel-based methods applied to high-resolution imagery, classical edge-detection<br />

algorithms have had limited success, partly due to the fact that forest stand boundaries are <strong>of</strong>ten<br />

diffuse and difficult to define, even for an experienced aerial photo-interpreter (Leckie et al. 2003;<br />

Castilla et al. 2008; ). Attention recently, however, has been on object-orientated approaches, and in<br />

particular, automated segmentation that exploits the spatial information inherent in remotely sensed<br />

imagery in addition to the spectral information (Wulder et al. 2008a and b). Several <strong>of</strong> these methods<br />

have been applied to forest stand delineation, with the most common approach being based on region<br />

merging (e.g. Baatz and Schape 2000; Pekkarinen 2002; Hay et al. 2005; Castilla et al. 2008; Wulder<br />

et al. 2008a and b). This typically involves the sequential aggregation <strong>of</strong> adjacent regions according to


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 134<br />

their similarity until a stop criterion is reached, which can be base on either a similarity or a size<br />

threshold. Once the digital scene has been segmented into ‘objects’ or polygons, clustering methods<br />

are then applied to classify the objects based on features (i.e. spectral and spatial properties) and<br />

meaningful information extracted from the objects (Chubey et al. 2006).<br />

Although several region-based image analysis techniques have been used successfully for forest<br />

information extraction purposes, adoption <strong>of</strong> these techniques by <strong>Australia</strong>n forest agencies has been<br />

limited. This has been partly due the computational complexity <strong>of</strong> some <strong>of</strong> the procedures and the<br />

empirical nature <strong>of</strong> the customised requirements such as the trial and error basis for determining the<br />

optimal parameterization for the segmentation (e.g. Flanders et al. 2003; Hay et al. 2005).<br />

Nevertheless, with increasing availability <strong>of</strong> new commercial region-based image analysis s<strong>of</strong>tware<br />

such as Definiens Developer 7 ® (Definiens® , Germany, formerly eCognition) and Feature Analyst®<br />

(Visual Learning Systems Inc., Missoula, Montana), coupled with diminishing availability <strong>of</strong> skilled<br />

photo interpreters there is new interest in region-based image analysis <strong>of</strong> remotely sensed data to<br />

support manual delineation and interpretation (Hay et al., 2005; Chubey et al. 2006; Wulder et al.<br />

2008a and b) We advocate that robust open-source protocols for implementing semi-automated image<br />

segmentation for individual forestry applications are a cost-effective alternative to the current<br />

commercial s<strong>of</strong>tware packages.<br />

PRACTICAL AND OPEN-SOURCE APPROACH IN FOREST STAND DELINEATION<br />

Since it is likely that significant research remains until fully automated forest stand delineation can be<br />

achieved, the general approach should be as practical as possible. Consequently the interim goal rather<br />

than trying to replace photo interpreters, should be to support them in generating more timely,<br />

consistent and accurate products (Leckie et al., 1998; Castilla et al. 2008; Wulder et al. 2008a and b).<br />

Therefore, new and or better tools are required that produce incremental improvements in these areas.<br />

These tools need not provide final solutions or 100% correct results, they simply need to be tools that<br />

are useful and that can be easily corrected when things go awry. Specifically, they must be simple to<br />

apply, not require expensive equipment, not substantially alter the mapping workflow, nor involve<br />

inordinate fine-tuning by the interpreter (Leckie et al. 1998; Wulder 2008a and b). In order to facilitate<br />

these requirements, a semi-automated forest stand delineation methodology has been developed for<br />

application to moist native forests in south eastern <strong>Australia</strong>.<br />

Although there are a number <strong>of</strong> existing segmentation algorithms available for delineation <strong>of</strong> forests<br />

stands, they have all been developed overseas for non-eucalypt forests and significant effort is<br />

required to become familiar with these algorithms and associated s<strong>of</strong>tware modules (e.g. Definiens<br />

Developer 7; Castilla et al. 2008) . As a consequence, an alternative approach that developes an<br />

integrated workflow process utilising open-sourced s<strong>of</strong>tware was taken in this study. The<br />

Geographical Resources Analysis Support System platform (GRASS www.grass-itc.it; Neteler and<br />

Mitasoua 2004) was chosen due to its popularity within the open-source community and that it fully<br />

integrates with the open-source statistical s<strong>of</strong>tware R (www.r-project.org; R Development Core Team<br />

2005), along with the python scripting language (van Rossum and Drake 2001).<br />

In this paper we present a methodology that combines spectral and texture attributes, Laser scanning<br />

derived Canopy Height Model layers and a simple region growing algorithm to delineate and<br />

characterise moist sclerophyll forest stands. The study area is described in the next section followed by<br />

a description <strong>of</strong> the three datasets used in the analysis. A detailed presentation <strong>of</strong> the algorithms<br />

utilised is made, followed by an examination <strong>of</strong> the results. The paper is concluded with a discussion<br />

<strong>of</strong> the method and its extension.<br />

METHODS<br />

Study area and datasets<br />

The study area is located 70 km north east <strong>of</strong> Melbourne, in Victorian Central Highlands, southeastern<br />

<strong>Australia</strong>. A high proportion <strong>of</strong> this mountainous area supports Wet Forests, dominated by<br />

Eucalyptus regnans (Mountain Ash). The area was selected to represent a variety <strong>of</strong> ash-forest types<br />

and forest conditions.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 135<br />

The area experiences a cool temperate climate, with mild summers and cool winters. Average annual<br />

rainfall exceeds 1200 mm over most <strong>of</strong> the area. Soils tend to be free draining, friable, brown<br />

gradational, have high water holding capacities, and have developed on a variety <strong>of</strong> volcanic parent<br />

rock materials (DNRE 1998). Fire is the major natural disturbance associated with the study area.<br />

Several fires have occurred within the study area over the past 150 years, the most extensive being in<br />

1926; 1939 (McHugh 1991; Jeremiah and Roob, 1992) and the recent extreme fire event <strong>of</strong> January<br />

2009.<br />

The study area is subject to intensive hardwood timber harvesting. Large-scale timber cutting,<br />

generally selective harvesting and sawmilling occurred in these forests in the latter part <strong>of</strong> the<br />

nineteenth and early twentieth centuries. Massive salvage operations followed major wildfires,<br />

particularly the extensive 1939 fires. Since the 1960s, clearfellling has been the major silvicultural<br />

system practised (Squire et al. 1991). The study area is 4 km by 4km, for a total area <strong>of</strong> 1 600 ha.<br />

Stand delineation from aerial photographic interpretation<br />

In Victoria, API formed the basis <strong>of</strong> a State wide Forest Resource Inventory project, whereby the<br />

derived polygons describing the forest attributes were considered based upon homogenous attributes<br />

such as eucalypt species, eucalypt structure and disturbance. An interpreter delineated the inventory<br />

polygons using visual interpretation <strong>of</strong> stereo images. The forest inventory polygons are assumed to be<br />

positionally accurate to ± 20 m, following State guidelines (Black 1996). In this study, we utilised<br />

stand boundaries interpreted from aerial photography acquired in the year 1999 at a scale <strong>of</strong> 1:20 000<br />

(DSE 2007). These boundaries have been updated to 2008 using logging and fire history records.<br />

Digital aerial photography<br />

Digital aerial photography data at a scale <strong>of</strong> 1:40 000 were acquired over the study area during January<br />

and March 2007 using a ZI Digital Mapping Camera (AAM Hatch Ltd). The Land Victoria 20 m<br />

contours plus photogrammetric breaklines were used to orthorectify the image. Colour balancing,<br />

mosaicing and resampling to a pixel size <strong>of</strong> 0.5 m followed the orthorectification process. The images<br />

were saved as multiband Geotiff files having visible red (R), green (G) and blue (B) channels.<br />

According to the image supplier the average location error <strong>of</strong> the orthorectified images is less than 1<br />

m. Only the central parts <strong>of</strong> the photographs were used in the mosaic to decrease the effects <strong>of</strong><br />

bidirectional reflectance, exposure fall<strong>of</strong>f, and relief displacement on the image analysis. This photo<br />

mosaic was used as the dataset for the spectral and texture analysis.<br />

Laser scanning data<br />

High-density laser scanning data were acquired over the study area with an Optech ALTM 3100 EA<br />

system (AAM Hatch Ltd) during December 2007 and January 2008. The system settings and flight<br />

parameters are shown in Table 1. The vendor provided raw laser scanning data consisting <strong>of</strong> XYZ<br />

coordinates, <strong>of</strong>f-nadir angle, and intensity for all laser returns within the area in LAS file format. In<br />

addition, the vendor provided a filtered ground dataset consisting <strong>of</strong> points presumed to be<br />

measurements <strong>of</strong> the terrain surface, identified via a proprietary filtering algorithm. These filtered<br />

ground returns were used to generate a 1 m digital terrain model (DTM).<br />

Table 1: Laser scanning data specifications.<br />

Flying height 1 300 m Flight Direction NE/SW<br />

Swath width 945 m Laser pulse density 0.96 pulses/m 2<br />

Swath overlap 25% Laser footprint size 0.26 m<br />

Workflow structure<br />

The workflow presented here was set up on a LINUX platform using GRASS GIS as the environment<br />

for the implementation <strong>of</strong> new modules. GRASS GIS is designed for both interactive use with a<br />

Graphical User Interface (GUI) and command line use. The built in commands can be batched<br />

together with LINUX programs using shell scripting as an interface, which gives broad possibilities on<br />

adaptations <strong>of</strong> the existing GIS functionalities (Neteler and Mitsova 2004). Furthermore new modules<br />

are added as python scripts into the workflow (Python S<strong>of</strong>tware Foundation 2006).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 136<br />

The workflow included the following steps: Data processing, Imagery segmentation, Feature<br />

extraction, Unsupervised region classification, Merging and mapping <strong>of</strong> unsupervised classes to<br />

general forest classifications and Comparison between SFRI mapping and semi-automated delineation<br />

Processing the Laser scanning data; DCHM and colour aerial photography<br />

The elevations <strong>of</strong> the laser canning measurement were converted to a digital canopy height model<br />

(DCHM) by subtracting <strong>of</strong>f the elevation <strong>of</strong> the underlying terrain. The canopy height model was<br />

calculated at a resolution <strong>of</strong> 0.5 m and then re-sampled to 1.0, 2.0 and 5.0 m resolutions.<br />

Unfortunately the visible colour RGB aerial photography used in this study does not allow the<br />

calculation <strong>of</strong> the widely used Normalised Difference vegetation Index (NDVI) as reflectance in the<br />

NIR spectral region was not captured. Instead, a vegetation index (VI) (Equation 1) based on the green<br />

(G) and red (R) bands was used. The ratio <strong>of</strong> these two bands has been successfully used in Canada to<br />

map and monitor forest disturbance caused through stand mortality from mountain pine beetle attack<br />

(Wulder et al. (2009).<br />

Vegetation Index = R/G Equation 1<br />

The image attributes were obtained using a square shaped windows (size 20 by 20 m). Holopainen and<br />

Wang (1998) have stated this size appears generally to be the near-optimal window size for extracting<br />

aerial photograph attributes for forest classification. The image attributes were extracted using the<br />

original 0.5 m resolutions as well as images re-sampled to 1.0, 2.0 and 5.0 m resolutions.<br />

The extracted neighbourhood attributes derived using the R/G VI were average, standard deviation,<br />

number <strong>of</strong> unique pixel values and range <strong>of</strong> pixel values. Additionally, four texture features based on<br />

the image gray-level co-occurrence matrices (Haralick et al., 1973; Haralick 1979) were computed:<br />

angular second moment (ASM), contrast (CON), correlation (COR) and entropy (ENT).<br />

These texture features were calculated from the pixel window using horizontal (0), vertical and<br />

diagonal (45 and 135) directions resulting in four values per feature. Prior to conducting the texture<br />

analysis the band data were requantified (Equations 2, 3, 4) to a 6 bit image (64 levels). This process<br />

does not change the actual grey levels distributions but greatly reduces the computation time.<br />

The two ‘best” neighbourhood and texture features were selected for input into the segmentation<br />

algorithm. In this study ‘best’ was defined through a iterative process <strong>of</strong> re-running all pair<br />

combinations to visually identify the best segmentation results.<br />

Imagery Segmentation<br />

Within GRASS there are a number <strong>of</strong> segmentation modules, including pixel-based (maximum<br />

likelihood classifier- i.maxlik and the sequential maximum a posteriori algorithim-i.smap), edgedetection<br />

(v.lidar.edge detection) and region growing modules (v.lidar.growing).<br />

Studies presented in the literature and personal experience suggest that region-based segmentation<br />

methods are generally more appropriate for semi-automatic stand delineation. Unfortunately, the<br />

current region growing modules implemented in GRASS are tailored for pure laser scanning object<br />

identification and do not lend themselves to the inclusion <strong>of</strong> spectral and textural information. As a<br />

consequence part <strong>of</strong> this study was to develop a more generic region growing module for GRASS.<br />

Traditional region growing methodology can be viewed as an iterative process by which regions are<br />

merged starting from individual pixels (or any initial segmentation), and growing iteratively until<br />

every pixel is processed. In broad terms, it can be described by the following steps:<br />

1. Segment the entire image into pattern cells (1 or more pixels);<br />

2. each pattern cell is compared with its neighbouring cells to determine if they are similar, using<br />

a similarity measure. If they are similar, merge the cells to form a fragment and update the<br />

property used in the comparison;<br />

3. continue growing the fragment by examining all <strong>of</strong> its neighbours until no joinable regions<br />

remain. Label the fragment as a completed region; and<br />

4. move to the next uncompleted cell, and repeat these steps until all cells are labelled.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 137<br />

The drawback <strong>of</strong> this traditional scheme is that, at each iteration, one or several merges occur so that<br />

the resulting segmentation is dependent on the order <strong>of</strong> the merges (lack <strong>of</strong> global convergence). In<br />

order to solve this problem, a number <strong>of</strong> different approaches have been suggested; including iterative<br />

parallel (Tilton and Cox, 1983), Hierarchical Step-Wise (Beaulieu and Goldberg, 1989) and size<br />

constrained region merging (Castilla and Hay et al. 2008). The segmentation algorithm proposed here<br />

is simpler than these approaches but is has been providing satisfactory results in the segmentation <strong>of</strong><br />

some forest and agricultural images.<br />

New region growing segmentation algorithm<br />

The approach implemented was based on the traditional region growing technique, with some<br />

modifications which partially solve the problem <strong>of</strong> the dependence on the order <strong>of</strong> the merges. The<br />

algorithm can be applied to any number <strong>of</strong> layers but in this study was limited to the DCHM and the<br />

two best texture and neighbourhood metrics<br />

Feature extraction<br />

The aim <strong>of</strong> the feature extraction phase is to determine the statistical attributes <strong>of</strong> each candidate forest<br />

stand outlined by the polygons resulting from the segmentation process. The result <strong>of</strong> the feature<br />

extraction phase is a file that contains a list <strong>of</strong> candidate stands sorted in descending order according to<br />

their area. For each candidate stand on the list, the file contains several statistics from the images<br />

selected by the user. These statistical attributes are then used to conduct a supervised or unsupervised<br />

classification; in both cases, to determine the similarity measure between two candidate stands<br />

outlined by polygons.<br />

Unsupervised classification<br />

In this study an unsupervised classification based on a clustering algorithm was applied to the list <strong>of</strong><br />

candidate stands, characterized by their statistical attributes, defined in the feature extraction file. The<br />

clustering algorithm uses the covariance matrix and mean vector <strong>of</strong> the candidate stands to estimate<br />

the centres <strong>of</strong> the classes.<br />

Merging <strong>of</strong> classes<br />

This is an interactive phase where all resulting classes were reclassified into broad forest classes. The<br />

user defines the reclassification rules by comparing the aerial photography and the canopy height<br />

model with the classification results.<br />

Comparison <strong>of</strong> automated segmentation versus aerial photographic interpretation delineation<br />

Originally, two comparison assessment techniques were envisaged. The first accuracy assessment test<br />

was to use existing irregularly spaced Statewide Resource Forest Inventory plots. However, this<br />

dataset was found to be unsuitable. There were too few points and too large a gap between points,<br />

relative to the mapping scale and the resulting number <strong>of</strong> polygons delineated. The second accuracy<br />

assessment was a qualitative comparison with the manually mapped SFRI classes. A qualified photo<br />

interpreter assessed the semi-automated delineation and where possible attributed to the automatically<br />

delineated polygons. The comparison <strong>of</strong> the two delineations was largely qualitative, as per Leckie et<br />

al (2003) and Wulder et al (2008). It must be recognised that there are a number <strong>of</strong> limitations<br />

associated with this comparison approach:<br />

1. The original SFRI delineation does not necessarily represent the “truth”. This process is<br />

afflicted with some unknowns (subjectivity, abstraction, and different semantics).<br />

2. The two products are based on different imagery sources in terms <strong>of</strong> date <strong>of</strong> acquisition and<br />

resolution– the SFRI mapping based on benchmarked 1980’s imagery versus the semiautomated<br />

delineation based on 2007 imagery.<br />

3. Unfortunately, there is no universally accepted procedure for comparing thematic maps. In<br />

addition, there is increasing evidence that traditional accuracy assessment tools based on<br />

cross-tabulation spreadsheets referring to points or pixels may not be suitable for object-based<br />

classification (de Kok, 2001; Blaschke 2003, Radoux and Defourney, 2007). Because objects<br />

include information about shape, area, area-shape relationships, topology etc., this additional<br />

information may have to be evaluated in another manner.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 138<br />

RESULTS<br />

The automated polygon segmentation was generated using the following three layers:<br />

1. Digital Canopy Height Model (DCHM)<br />

2. Average Vi with a 5m × 5m moving window (VI average)<br />

3. Contrast Vi with a 5m × 5m moving window (VI contrast)<br />

The results <strong>of</strong> the semi-automated segmentation versus manual delineation applied to the study area<br />

are shown in Figure 1. Using a simple visual examination, several differences in the polygons are<br />

obvious. Specifically, the semi-automated segmented polygons are more detailed than those resulting<br />

from the manual delineation. While human interpretation intrinsically includes generalisation – which<br />

is <strong>of</strong>ten important and necessary for many applications – automated image segmentation can also<br />

result in very complex outlines following shapes <strong>of</strong> small stands <strong>of</strong> trees. These differences are<br />

reflected in the mean perimeter <strong>of</strong> the polygons generated from the two processes: the semi-automated<br />

output had a mean polygon perimeter that was approximately 100% greater than the perimeter <strong>of</strong> the<br />

SFRI-manually delineated polygons, while the mean polygon size for the semi-automated polygons is<br />

only 30% greater than the manual SFRI polygons (Table 2). The manual SFRI polygons have a much<br />

greater range in polygon sizes, with a standard deviation that is more than double the automated<br />

delineated polygons.<br />

Figure 1: Screenshots <strong>of</strong> the resulting maps benchmarked manual delineation (left) and<br />

semi-automated polygon delineation (right).<br />

In addition, the benchmarked SFRI manually interpreted polygons were provided using ortho aerial<br />

images and corporate fire and logging history GIS layers. The result is a combination <strong>of</strong> visual<br />

interpretation (pro<strong>of</strong>ed by field surveys) updated with fire and logging history post photo date.<br />

However, Figure 1 shows the benchmarked SFRI data contain errors.<br />

The red arrows mark two points, where the line management is not comprehensible. These errors are a<br />

result <strong>of</strong> the logging history data not being up to date.<br />

A qualified photo interpreter assessed the semi-automated delineation and attributed each polygon<br />

with a simplified SFRI code. The photo interpreter noted several attributes <strong>of</strong> the automated delineated<br />

polygons:<br />

1. As previously mentioned, the boundaries delineated by the automated process are not as<br />

straight as those produced manually by a human interpreter;<br />

2. Approximately 60% <strong>of</strong> the area contained sensible polygons that were easily attributed;<br />

3. The remaining 40% polygons would require minor modifications to be meaningful, or were<br />

not appropriate for reasons discussed below.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 139<br />

Table 2: Properties <strong>of</strong> delineated polygons.<br />

Attribute Manual Semi-automated<br />

Total # <strong>of</strong> polygons 261 189<br />

Mean polygon size (ha) 6.13 8.46<br />

Minimum polygon size (ha) 0.01 1.51<br />

Maximum polygon size (ha) 50.48 112.15<br />

Standard deviation polygon size (ha) 7.95 14.35<br />

Mean polygon perimeter (m) 1291.75 2458.87<br />

Minimum polygon perimeter (m) 9.09 701.45<br />

Maximum polygon perimeter (m) 9021.90 25225.73<br />

Standard deviation polygon perimeter (m) 1268.29 2829.54<br />

An example <strong>of</strong> the first issue <strong>of</strong> more ‘precise’ delineations compared to manual mapping is shown in<br />

Figure 2. This is caused by the fact that the automatic delineation depends on the boundaries <strong>of</strong> the<br />

individual tree crowns. This means the older the trees within a forest stand the more frayed<br />

(fragmented) the boundary will be.<br />

Figure 2:Geometric differences for object delineation <strong>of</strong> forest species classes (left: manually<br />

mapped by human interpreter; right: semi-automatically constructed through image<br />

segmentation).<br />

Examples <strong>of</strong> the second situation, where the semi-automated delineation produced polygons that were<br />

sensible and easy to type, are provided in Figure 3.<br />

Figure 3: Examples <strong>of</strong> where semi-automated delineation performed well.<br />

These polygons represent areas with homogenous texture and tone, which have relatively uniform<br />

species composition, age, and height –demonstrating the ability <strong>of</strong> semi-automated segmentation<br />

routines to perform well in uniform stand conditions.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 140<br />

In contrast, Figure6 illustrates situations where the semi-automated delineation did not perform well.<br />

In Figure 4 (left), some mature vegetation has been included within a larger re-generation polygon. In<br />

Figure 4 (right), the highlighted polygon seems no different than the area to the left. The main reason<br />

for these errors is how the algorithm deals with the weighting between the three input layers. In both<br />

cases shown below the digital canopy height model layer is driving the incorrect delineation. This<br />

coupled with the minimum mapping unit size tends to be the main issues.<br />

Figure 4: Examples <strong>of</strong> where semi-automated delineation performs poorly.<br />

The quantitative comparisons for the five major species groups identified within the study area results<br />

in an overall accuracy <strong>of</strong> 65% (Kappa Index: 44% see Table 3). A closer look at the classes shows<br />

partly very good results (for the mixed Eucalypt classes), but very bad results (user accuracy for the<br />

‘unknown’ class). One reason is the inexact and/or subjective reference dataset; another reason<br />

originates in the resemblances <strong>of</strong> the classes. Confusion occurred most between the classes ReM -<br />

DeM and ReP - DeP, while the most stable subdivision occurred between mixed (Mixed E. regnans<br />

and E. delegatensis) and the pure stands. If only the four stable classes (ReM, ReP, DeM, DeP) are<br />

taken into account, overall accuracy rises up to 80% (Kappa index: 62%).<br />

Table 3:Accuracy assessment for the five major forest species classes in the study area.<br />

Development phases<br />

Mixed<br />

E. regnans<br />

(ReM)<br />

Pure<br />

E. regnans<br />

(ReP)<br />

Mixed<br />

E.delegatensis<br />

(DeM)<br />

Pure<br />

E.delegatensis<br />

(DeP)<br />

Unknown<br />

User Accuracy* in % 80 42 69 54 12<br />

Producer Accuracy in % 54 48 89 49 30<br />

Overall Accuracy: 65% Kappa Index: 44%<br />

However, it should be noted that the reference benchmarked SFRI data are subjective (and in some<br />

places obviously incorrect). The SFRI class descriptions can be fuzzy and it is sometimes difficult to<br />

distinguish forest classes in the field, if they are proximate in time.<br />

CONCLUSION<br />

It is generally accepted that manual aerial photo interpretation provides better classification accuracy<br />

than automated delineation, because the interpreter can use additional clues, such as area, shape and<br />

contextual information, including topographic information, results <strong>of</strong> previous classification, etc. In<br />

this paper, a region-growing segmentation technique, in which the algorithm uses some contextual<br />

attributes besides spectral values <strong>of</strong> the pixels, is an attempt to improve digital classification. Even this<br />

alternative can result in an inadequate classification, so the procedure proposed here, similar to other<br />

techniques, will require editing after the digital analysis steps to eliminate misclassified set <strong>of</strong> pixels<br />

(polygons).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 141<br />

Overall, the segmentation accuracies are moderate from the semi-automated approach but it must be<br />

recognised that <strong>of</strong>ten these forest classes are spectrally similar and the forest texture and Laser<br />

scanning height data vary within each class. Notwithstanding, the results generally encourage further<br />

investments in data acquisition and methodological development. However, some critical points<br />

remain. In particular the following tasks need to be solved for supporting or partially replacing<br />

traditional aerial photography interpretation techniques:<br />

• Quantify to what extent the results simplify API using aerial photography combined with<br />

stand ground surveys (time, exactness, costs)<br />

• Supporting the API work with interim analysis results (e.g. stand height maps, differentiation<br />

vegetation, gap distribution etc) to get more objective and faster results in case <strong>of</strong> unclear<br />

classification.<br />

• Integrating statistical results in the classification which allow statements about how certain a<br />

forest vegetation classification is, or how uncertain the distinction from a particular other<br />

class is.<br />

For the foreseeable future, aerial photo interpretation coupled with ground based surveys will be<br />

necessary (in particular in the production forest) especially for planning purposes. But the proposed<br />

method in this paper is believed to potentially reduce the efforts required. API can then focus on<br />

uncertain situations and on the most important forest stands concerning forest planning and ecological<br />

questions.<br />

ACKNOWLEDGEMENTS<br />

This study was supported by the Victorian Government through the Resource Outlook project. We<br />

would like to thank Lee Miezis, Mike Sutton, Fiona Hamilton, Fred Cummings, Kristen Thrum,<br />

Andrew Mellor, Kristen Thrum and Kate Nolan for their support (Department <strong>of</strong> Sustainability and<br />

Environment).<br />

REFERENCES<br />

(Due to space limitations three pages <strong>of</strong> references are available upon request from the<br />

main author)


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 142<br />

THE GLOBAL FOREST RESOURCE ASSESSMENT<br />

REMOTE SENSING SURVEY<br />

Adam Gerrand 1 , Erik Lindquist 1 , Mette Wilkie 1 and Rod Keenan 2<br />

ABSTRACT<br />

Global concern is growing over deforestation because <strong>of</strong> its climate impacts as well as<br />

loss <strong>of</strong> biodiversity and other forest services. FAO has been reporting on the world's<br />

forests at 5 to 10 year intervals since 1946 through the Global Forest Resources<br />

Assessments (FRA). As part <strong>of</strong> FRA 2010, FAO, its member countries and partner<br />

organizations will undertake a new global remote sensing survey <strong>of</strong> forests between 2008<br />

and 2011. The survey will substantially improve knowledge on land use change including<br />

deforestation, reforestation and natural expansion <strong>of</strong> forests. The assessment will sample<br />

the whole land surface <strong>of</strong> the Earth with over 13 000 Landsat Global Land Survey image<br />

sections covering 10km by 10km at each <strong>of</strong> the intersections <strong>of</strong> the latitude and longitude<br />

degree lines.<br />

The main outcomes will be improved information at the global and ecozone level on<br />

changes in forest cover and land use including trends in the rate <strong>of</strong> deforestation,<br />

afforestation and natural expansion <strong>of</strong> forests from 1990 to 2005. The work also has<br />

potential as a global monitoring framework for a range <strong>of</strong> other purposes. A significant<br />

effort will also be made to provide training to build on the capacity <strong>of</strong> many countries to<br />

improve their forest monitoring and reporting. The paper will describe the FRA2010<br />

Remote Sensing Survey and encourage involvement <strong>of</strong> others who can provide useful<br />

additional data or analyses for validation.<br />

INTRODUCTION<br />

The world’s forests provide vital economic, social and environmental benefits. They mitigate climate<br />

change by storing carbon, provide wood and non-wood forest products, support human livelihoods, supply<br />

clean water and provide habitat for half the species on the planet (Wilson, 1988).<br />

The Food and Agriculture Organization (FAO) <strong>of</strong> the United Nations provides detailed information on<br />

global forest cover and forest land use at 5 to 10 year intervals from 1946 to 2005 in a series <strong>of</strong> reports<br />

called the Global Forest Resources Assessments (FRA). Each is based on data that countries provide<br />

to FAO in response to a questionnaire. FAO compiles and analyses the information and presents the<br />

current status <strong>of</strong> the world’s forest resources and their changes over time. Historically, FRA reporting<br />

has evolved to reflect the major issues <strong>of</strong> concern at the time. Early reports focused on timber stocks in<br />

response to post-war (WWII) needs for building materials while more recent emphasis has shifted to<br />

deforestation and conservation issues. The breadth and quality has also improved as individual<br />

countries gain reporting capacity and information availability increases. The most recent FRA report,<br />

in 2005, was the most comprehensive in scope ever and aimed at assessing progress towards<br />

sustainable forest management (FAO, 2006a).<br />

The need for reliable information on forests<br />

During the 2008 G-8 Summit, world leaders “encouraged actions for Reducing Emissions from<br />

Deforestation and Forest Degradation in Developing Countries (REDD) including the development <strong>of</strong><br />

an international forest monitoring network building on existing initiatives” (G8 Summit 2008).<br />

As part <strong>of</strong> the FRA 2010, FAO, its member countries and partners are undertaking a systematic remote<br />

sensing survey which will form the basis for a consistent long-term global forest monitoring system and<br />

1 UN Food and Agriculture Organisation, FAO Forestry Department, Viale delle Terme di Caracalla, Rome 00100, Italy<br />

Rome. Tel: +3906 5705 3063, Fax: +3906 5705 5137. adam.gerrand@fao.org<br />

2 School <strong>of</strong> Forest and Ecosystem Science, University <strong>of</strong> Melbourne, Water Street Creswick, Victoria <strong>Australia</strong> 3363.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 143<br />

help meet a number <strong>of</strong> international forest information needs, including the G-8 recommendations (FAO,<br />

2008). Forest loss has effects on local, regional, and global scales on a wide range <strong>of</strong> factors including<br />

impacts on climate, biodiversity and ecosystem services (Lambin and Geist, 2006). With respect to climate<br />

alone, the latest estimates state that “forestry” (including deforestation, forest biomass decomposition after<br />

logging, and release <strong>of</strong> CO2 from peat soils) is responsible for about 17 % <strong>of</strong> human-produced greenhouse<br />

gas emissions (IPCC, 2007). The Clean Development Mechanism under the UNFCCC, and other<br />

initiatives such as Reducing Emissions from Deforestation and Degradation (REDD) are being developed<br />

to help reduce the negative effects <strong>of</strong> forest loss on climate but issues <strong>of</strong> accounting for and monitoring<br />

forest resources remain difficult to resolve.<br />

According to the latest FRA assessment, approximately 13 million hectares <strong>of</strong> forest cover is cleared annually<br />

worldwide (FAO, 2006a). Few countries can adequately report on changes in forest area over time so there is<br />

a need to improve the quality <strong>of</strong> reporting and harmonisation <strong>of</strong> definitions <strong>of</strong> forests (FAO, 2006, Hansen et<br />

al., 2008b). FAO is working to improve the results through strengthening national capacity as part <strong>of</strong> FRA<br />

2010 and improved reporting guidelines (FAO, 2007b).<br />

Satellite remote sensing <strong>of</strong>fers the advantage <strong>of</strong> broad area coverage, repeatable and systematic<br />

observations <strong>of</strong> the Earth’s surface. Though remote sensing does not replace the need for field-collected<br />

data, it <strong>of</strong>fers distinct benefits when conducting large-area surveys for broad vegetation-type categories.<br />

One major benefit is the ability to provide maps rather than tabular numerical results summarised by<br />

country or region that better illustrate where changes in forest cover are occurring. Forest loss is variable<br />

across forest types and geographical locations, a fact increasingly well documented in regional remote<br />

sensing studies from the tropics to the boreal (Potapov et al., 2008; Hansen et al., 2008b, Achard et al.,<br />

2002).<br />

The FRA 2010 Remote Sensing Survey<br />

The FRA 2010 Remote Sensing Survey (RSS) brings together the comprehensive Landsat Global<br />

Land Survey satellite imagery database (Tucker et al., 2004) and is a partnership <strong>of</strong> many <strong>of</strong> the<br />

world’s leading land cover remote sensing scientists to analyse satellite data and engage with country<br />

experts in over 150 countries (see Acknowledgements).<br />

The goals <strong>of</strong> the RSS are to obtain systematic information on the distribution and changes in forest cover<br />

and forest land use from 1990 to 2000 to 2005 at regional, ecozone and global levels. The RSS also<br />

provides a consistent framework upon which future global assessments and more detailed regional<br />

assessments can be based. As presented in this paper, the RSS is a work in progress; the methods,<br />

discussions and conclusions are preliminary and may be amended as the study develops. Notably, the RSS<br />

is intended to provide information complementary to and in some cases strengthen existing national<br />

reporting systems - but not replace them.<br />

METHODS<br />

Sampling framework<br />

The FRA 2010 RSS global sampling grid consists <strong>of</strong> 13 689 sites and covers the globe between 75<br />

degrees North and South in latitude. A systematic sampling design based on each longitude and<br />

latitude intersection has been implemented, with a reduced intensity above 60 degrees North/South<br />

latitude due to the curvature <strong>of</strong> the Earth (every second intersection sampled in between 60 and 75<br />

degrees North/South).<br />

Each sample tile will cover a 10 by 10 kilometre box at every one-degree latitude and longitude<br />

junction (approximately 100km apart). This grid <strong>of</strong> sample plots is the same basic layout but a lower<br />

intensity (wider spacing) than the national forest assessments supported by FAO (FAO, 2008).<br />

Antarctica was excluded from the land mask. Sample locations within deserts and areas with<br />

permanent ice are also excluded from analysis leaving a total number <strong>of</strong> 9 078 tiles (Figure 1).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 144<br />

Figure 1: The 13 689 sample locations <strong>of</strong> the FRA RSS (light black dots) and the 9 078 sample sites<br />

where at least a portion <strong>of</strong> the 10x10 km sample tile has tree cover > 10% as derived from<br />

the circa 2000 MODIS VCF (dark black dots).<br />

Tasmania 10 sample tiles<br />

across different forest<br />

types<br />

Landsat samples available on-line: Draft viewing, analysis and labelling s<strong>of</strong>tware:<br />

Figure 2: The FRA RSS sampling grid with 700 samples across <strong>Australia</strong> and the 10 samples in<br />

Tasmania with one sample tile shown in detail with Landsat images from the FRA website<br />

(lower left) and draft analysis (lower right).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 145<br />

Remotely sensed inputs<br />

The United States Geological Survey’s Landsat Global Land Survey dataset (GLS) provided the<br />

imagery data for interpretation and classification. The GLS is a co-registered, multi-date dataset<br />

composed <strong>of</strong> the single best Landsat image acquisition covering most <strong>of</strong> the Earth’s land surface and<br />

centered on the years 1975, 1990, 2000, and 2005 (Gutman et al., 2008; Tucker et. al., 2004). The<br />

FRA 2010 RSS will focus initially on the GLS1990, GLS2000 and GLS2005. If time permits and<br />

techniques are developed to handle earlier Landsat MSS data, then we may include data from 1975 to<br />

make a 30 year time series.<br />

For each survey tile, Landsat bands 1-5 and 7 (also 8 in the case <strong>of</strong> ETM+) <strong>of</strong> the GLS acquisitions were<br />

compiled. These were clipped to a 20km by 20km box centered on each one-degree latitude and<br />

longitude intersection to create imagery ‘chips’. This produced 56 219 individual imagery chips for the<br />

three time periods. The central 10km by 10km box <strong>of</strong> each sampling tile will be used for area<br />

calculations and statistical analysis. This was adjusted to preserve an integer number <strong>of</strong> Landsat data<br />

pixels (30m spatial resolution). For 10km by 10km blocks, 334 by 334 pixels tiles were created for a tile<br />

extent <strong>of</strong> 10 020 by 10 020m.<br />

The images will be processed through a series <strong>of</strong> steps described in more detail in Gerrand et al. (2009);<br />

Achard et al (2009) and Bodart et al (2009). The aim is to develop polygons <strong>of</strong> a similar spectral signal<br />

and label these with a simple legend with 6 main land cover classes (Table 1).<br />

Labeling objects (polygons) in each segmented survey tile is a two-step process. The first step is a<br />

draft pre-labeling exercise followed by label correction and validation by in-country experts. Polygons<br />

are pre-labeled in one <strong>of</strong> two ways: (i) automatically using spectral training signatures and (ii) by<br />

visual interpretation. Automated labeling by the Joint Research Centre (JRC) will be applied to survey<br />

tiles in the humid tropics and Russian boreal forests (Bodart et al, 2009). For all other regions, FAO<br />

specialists will use visual interpretation and a FAO-developed s<strong>of</strong>tware program called GEOVIS<br />

Mapping Device – Change Analysis Tool (MADCAT) to label polygons.<br />

The FAO-developed Land Cover Classification System (LCCS) (FAO, 2005) has been adapted for<br />

labeling polygons by land cover. The legend has been simplified to include only 5 land cover classes<br />

(Table 1). Survey tiles processed by the JRC use a slightly different initial legend and will be re-coded to<br />

fit the FRA cover classes. A simple system <strong>of</strong> 9 Land-use codes has also been developed for use in the<br />

RSS based on FRA definitions (FAO, 2007b).<br />

Table 1: Land cover and land-use classes to be used in the FRA 2010 RSS.<br />

Land Cover Class Land-Use Class<br />

Tree Cover Forest<br />

Other wooded land<br />

Other land with tree cover<br />

Shrub Cover<br />

Herbaceous Grass and herbaceous cover<br />

Agricultural crops<br />

Other Land Built up habitation<br />

Bare land<br />

Water Water<br />

No data No data<br />

Validation and changes in forest cover and use<br />

Pre-labeled polygons in ESRI shapefile format and the remotely sensed imagery will be provided to all<br />

countries with sample tiles for in-country, expert validation. Polygon labels will be checked for<br />

accuracy against each time period <strong>of</strong> imagery. Ancillary, country-specific land cover data sets (such as<br />

forest inventory and vegetation type maps where available) and qualitative information obtained from<br />

the Degree Confluence Project (www.confluence.org) and Google Earth (www.earth.google.com) can<br />

also be used for validation. Important data is also held in many other organisation such as State and


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 146<br />

Territory and local organisations and these are encouraged to submit data to the on-line data portal being<br />

set up under the FRA website: http://www.fao.org/forestry/fra2010-remotesensing/en/.<br />

Forest cover changes both positive (afforestation or natural expansion) and negative (deforestation)<br />

between time periods will be calculated by summing polygon area for those polygons that changed<br />

labels from time one to time two or three. All polygons with tree cover and those identified as forest<br />

cover change will also be assigned a cover and land use code for each time period. Changes in land<br />

use will be tallied in the same manner as forest cover. Global ecological zones (FAO, 2001), and<br />

global political and administrative boundaries (FAO, 2006b) will be used as summary units.<br />

EXPECTED RESULTS<br />

Summaries <strong>of</strong> forest/tree cover and land use will be produced at regional, eco-zone and global levels<br />

for 1990, 2000, and 2005. Change in forest/tree cover and land use between periods will also be<br />

summarized. Country-level statistics will not be reported in the RSS as the global sampling design<br />

precludes statistically valid estimates at the national level in most cases due to the small number <strong>of</strong><br />

samples. The MODIS VCF product based on an updated version <strong>of</strong> Hansen (2003) will use the RSS<br />

sample tiles to validate the latest version (circa 2005) at 250-meter spatial resolution. This analysis will be<br />

used to generate a new global map <strong>of</strong> the world’s tree cover planned for release in 2011.<br />

A database <strong>of</strong> all imagery chips used in the RSS is available for free download via the internet<br />

(http://www.fao.org/forestry/fra2010-remotesensing/en/). This repository <strong>of</strong> global, multi-temporal<br />

imagery and ancillary information has the potential to become an important data source for researchers,<br />

land managers and policy makers and can serve as both a base for future studies and an archive <strong>of</strong> the<br />

RSS.<br />

A pilot study analysis <strong>of</strong> around 250 samples worldwide is currently underway at the time <strong>of</strong> writing this<br />

paper in July 2009. <strong>Australia</strong> is participating in the Pilot Study and has 13 sample tiles to process.<br />

DISCUSSION<br />

Statistical sampling versus wall-to-wall mapping<br />

The RSS is employing a systematic sampling approach covering most <strong>of</strong> the land surface <strong>of</strong> the Earth.<br />

Though global coverage remotely sensed datasets do exist, they exhibit a relatively coarse spatial resolution<br />

(250+ meters) and many forest cover changes take place at spatial scales smaller than can adequately be<br />

measured with ‘large’ pixels, especially in the tropics (Hansen, 2008b).<br />

Landsat data was chosen as the preferred data source for the FRA RSS because it has a suitable pixel size<br />

(30m) to detect small patches <strong>of</strong> forest change and because it has the best historical archive <strong>of</strong> global data<br />

(Williams, 2006). The commendable decision in 2008 by the USGS to open the entire Landsat archive for<br />

use overcomes one <strong>of</strong> the major historical limitations to use <strong>of</strong> these data. However, the large data volumes<br />

and difficulties <strong>of</strong> automating the processing to produce successful results across a wide variety <strong>of</strong><br />

ecosystems, currently limit the ability to develop a global, wall-to-wall map <strong>of</strong> the worlds forests at<br />

Landsat-scale spatial resolution. These constraints are being addressed and promising results are<br />

emerging on regional and continental scales in North America (Masek et al., 2006), <strong>Australia</strong> (Caccetta<br />

et al., 2007), central Africa (Hansen et al., 2008a), and Europe (Pekkarinen et al., 2009).<br />

A sampling approach using 30m spatial resolution imagery was chosen as a solution that achieves<br />

manageable data volumes, comprehensive coverage <strong>of</strong> sample plot locations, and returns statistically valid<br />

results at regional and global spatial scales. A thorough review <strong>of</strong> statistical sampling compared with wall-towall<br />

mapping methods can be found in Stehman et al., 2003. A detailed justification for the FRA RSS<br />

sampling strategy can be found in FAO, 2007a.<br />

Land cover and land use information are both needed<br />

The RSS will produce results on both forest cover and forest land use. Land cover is what can be detected<br />

from a remotely sensed instrument, like the Landsat satellite, and refers to the biophysical attributes <strong>of</strong> the<br />

Earth’s surface. Land use implies a human dimension or purpose characterizing a location. It can sometimes<br />

be inferred from remotely sensed data, however, is typically only verified with local, expert knowledge or


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 147<br />

data collected in situ for a particular location. Accurate information on land use is critical to understand the<br />

drivers <strong>of</strong> forest cover change (Lambin and Geist, 2005) and help develop effective policies and strategies to<br />

reverse forest loss that is one <strong>of</strong> the global objectives for forests (UNFF, 2007).<br />

Country involvement and capacity building<br />

FAO and its partners will work closely with national governments and institutions and with a wide range<br />

<strong>of</strong> other international and non-governmental organizations. Countries will be involved in the analysis to<br />

include national data and local knowledge to help ensure the results are validated and as accurate as<br />

possible.<br />

The FAO developed computer s<strong>of</strong>tware for viewing the imagery and labelling the changes will be made<br />

freely available to participating countries. A series <strong>of</strong> training workshops will be held around the world<br />

in regional centres to improve the technical capacity <strong>of</strong> many staff in analysing remote sensing imagery.<br />

The access to free remote sensing data and s<strong>of</strong>tware will particularly benefit developing countries with<br />

limited forest monitoring data or capacity. The long-term aim <strong>of</strong> this is to strengthen the abilities <strong>of</strong><br />

countries to regularly monitor and manage their forests and enable informed decisions as well as meet<br />

the reporting requirements <strong>of</strong> many national and international processes.<br />

<strong>Australia</strong>’s involvement<br />

<strong>Australia</strong>’s forests and ecosystems are significant natural assets and are <strong>of</strong> global importance. They are highly<br />

valued, have many uses and provide a wide range <strong>of</strong> valuable products and benefits for society. <strong>Australia</strong> has<br />

over 149 million hectares <strong>of</strong> forests (NFI, 2008), which equates to around 4% the world’s forests (FAO, 2006).<br />

There are increasing requirements to measure and monitor the extent and condition <strong>of</strong> <strong>Australia</strong>’s forests<br />

for management purposes and for domestic and international reporting requirements. The traditional focus<br />

on wood production values and estimates <strong>of</strong> sustainable yield has been long recognised as being too<br />

narrow; however, data collection systems for monitoring the broader range <strong>of</strong> forests and forest ecosystems<br />

are limited in both area and length <strong>of</strong> study. <strong>Australia</strong> has set up a sophisticated National Carbon<br />

Accounting System (Richards and Brack, 2004) to monitor the stock and change in carbon stocks. This<br />

system has been designed to meet the requirements for monitoring and reporting change in tree cover for<br />

the purposes <strong>of</strong> reporting to UNFCC and the Kyoto Protocol. There are opportunities for using the NCAS<br />

data to assist the validation <strong>of</strong> the FRA samples by being able to spatially compare the presence <strong>of</strong> trees.<br />

However, the NCAS on its own does not fulfil the data needs for FRA because methods and definitions are<br />

different (e.g. the NCAS is a pixel based system with one class <strong>of</strong> woody vegetation cover with a minimum<br />

mapping unit (MMU) <strong>of</strong> 1ha and the FRA RSS uses an object polygon based approach with several<br />

vegetation classes (trees, shrubs, herbaceous, and has a MMU <strong>of</strong> 5 ha. The <strong>Australia</strong>n National Forest<br />

Inventory data is based on State and Territory forest and vegetation mapping and includes forest type which<br />

is needed to translate into the global FRA classes. Neither NCAS nor the NFI datasets meet the FAO<br />

requirements alone as FRA forest definition includes a land-use component (FAO, 2006). Thus additional<br />

information such as that collected by the <strong>Australia</strong>n Collaborative Land Use Mapping Programme<br />

(ACLUMP Leslie et al. 2006) will be required to obtain satisfactory final results.<br />

There is still no adequate national forest monitoring system for a range <strong>of</strong> other values such as biodiversity<br />

and other values (Lindermayer et al. 2007, Brack, 2007). Gerrand and Clancy (2007) argued that <strong>Australia</strong><br />

should be setting up a national framework for a monitoring system because current approaches using<br />

periodic map compilations are not adequate to report on change in forest extent or condition. The<br />

proposed Continental Forest Monitoring Framework (Wood et al. 2007) developed by the National<br />

Forest Inventory and supported by States and Territory agencies as a concept would help address many<br />

<strong>of</strong> these concerns. The sampling system proposed under the CFMF was for a 20km grid spacing to<br />

provide many more samples to get better statistical results at a regional level.<br />

The FRA samples are approximately 100km apart and thus do not have enough samples to provide<br />

reliable results at a local scale. For example, in Tasmania there are only 10 samples - including one each<br />

on King and Flinders Islands (Figure 2). This is clearly not enough to capture the diversity or accurately<br />

estimate the area <strong>of</strong> forest or change in Tasmania and other methods are needed to report on forest extent<br />

and change at the local scale.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 148<br />

The CFMF has not yet been implemented in <strong>Australia</strong> due to a lack <strong>of</strong> funding. With the FRA sampling<br />

there now exists a starting platform <strong>of</strong> data that can help build the data inputs to get a CFMF style system<br />

up and running at reduced cost. The biggest costs <strong>of</strong> course are the field work but even the start with the<br />

image samples and processing is a useful step forward. <strong>Australia</strong> has not formally confirmed it will<br />

complete the analysis <strong>of</strong> the FRA RSS sample sites at the time <strong>of</strong> writing this paper (July 2009) and one<br />

aim <strong>of</strong> this paper is to encourage awareness <strong>of</strong> the work to build support and commitment to the work.<br />

CONCLUSIONS<br />

The FRA 2010 RSS will be a systematic, comprehensive, global study <strong>of</strong> tree cover and forest land-use<br />

changes from 1990 to 2000 to 2005. It presents a consistent methodology for monitoring forest change<br />

at a global level that can be expanded for more detailed studies. It is expected that the RSS will<br />

improve understanding <strong>of</strong> total forest area changed, the patterns resulting from this change, and the<br />

processes driving forest cover change globally. This is information that governments, land managers,<br />

researchers and civil society groups can use to make better-informed decisions regarding the world’s<br />

forest resources.<br />

FAO outreach and training activities will help build technical capacity to monitor forest resources in<br />

many countries. FAO will provide access to remote sensing imagery either through the internet or hard<br />

media. The image processing s<strong>of</strong>tware can be used for other studies and monitoring purposes.<br />

Additionally, a global network <strong>of</strong> FRA RSS specialists will be built representing a powerful human<br />

resource for improved technical capacity and pr<strong>of</strong>iciency in many countries.<br />

The FRA 2010 and the RSS combined will provide a basis for reporting on progress towards sections<br />

<strong>of</strong>: (i) the United Nations Convention on Biological Diversity’s target <strong>of</strong> reversing biodiversity loss by<br />

2010, (ii) the Millennium Development Goals, (iii) the Global Objectives <strong>of</strong> the UN Forum on Forests,<br />

(iv) the International Tropical Timber Organization’s Objective 2000. If countries choose and have the<br />

resources to do so, the methods have the potential to form a platform for developing more detailed<br />

reporting capabilities at a national level such as those required for the land use and land use change<br />

results for the UN Framework Convention on Climate Change and the Kyoto Protocol and the<br />

Reducing Emissions from Deforestation and Forest Degradation in Developing Countries (REDD).<br />

<strong>Australia</strong> is encouraged to participate in the FRA RSS and to complete the analysis <strong>of</strong> the samples as<br />

part <strong>of</strong> its contribution to this important global study and also as an important step to developing a<br />

better national forest monitoring system.<br />

ACKNOWLEDGEMENTS<br />

The FRA RSS is a partnership between FAO, the Joint Research Centre <strong>of</strong> the European Commission, US<br />

Geological Survey and NASA, South Dakota State University, and Friedrich-Schiller University and ,many<br />

national experts. We would like to thank: Ralph Ridder, Frederic Achard and colleagues at JRC (see Bodart et.<br />

al.), Thomas Loveland, John Latham, Antonio Di Gregorio, Antonio Martucci, Ilaria Rosati. Many others<br />

have helped — please accept our thanks. Funding was supported by initial grants from NASA, the<br />

Governments <strong>of</strong> Finland, <strong>Australia</strong>, and funding for 2009-11 from the European Commission. For more<br />

information please visit: http://www.fao.org/forestry/fra2010-remotesensing/en/<br />

LITERATURE CITED<br />

Achard, F., Eva, H.D, Stibig, H.J., Mayeaux, P., Gallego, J., Richards, T., Malingreau, J.P., 2002. Determination <strong>of</strong><br />

deforestation rates <strong>of</strong> the world’s humid tropical forests. Science, 297, pp. 999-1002.<br />

ACLUMP, 2006. Guidelines for land use mapping in <strong>Australia</strong>: principles, procedures and definitions. A technical<br />

handbook supporting the <strong>Australia</strong>n Collaborative Land Use Mapping Programme Edition 3. Eds R. Leslie et al.<br />

Bureau <strong>of</strong> Rural Sciences, Canberra, 83pp.<br />

Bodart, C., Beuchle, R., Simonetti, D., Eva, H., Raši, R., Carboni, S., Brink, A., Stibig, H. Achard, F., Mayaux, P.,2009.<br />

Global monitoring <strong>of</strong> tropical forest cover changes by means <strong>of</strong> a sample approach and object-based classification<br />

<strong>of</strong> multi-scene Landsat imagery: pre-processing and first results. (ISRSE – this meeting proceedings).<br />

Brack, C.L., 2007. National Forest Inventories and Biodiversity Monitoring in <strong>Australia</strong>. Plant Biosystem 141:<br />

104-112.<br />

Caccetta, P.A., Furby, S.L., O'Connell, J., Wallace, J.F. and Wu, X. 2007. Continental Monitoring: 34 Years <strong>of</strong><br />

Land Cover Change Using Landsat Imagery, 32nd International Symposium on Remote Sensing <strong>of</strong><br />

Environment, June 25-29, 2007, San José, Costa Rica.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 149<br />

Desclée, B, Bogaert, P., Defourny, P., 2006. Forest change detection by statistical object-based method. Remote<br />

Sensing <strong>of</strong> Environment, 102, pp. 1-11.<br />

Duveiller, G., Defourny, P., Desclée, B., Mayeaux, P., 2008. Deforestation in Central Africa: estimates at regional,<br />

national and landscape levels by advanced processing <strong>of</strong> systematically-distributed Landsat extracts. Remote<br />

Sensing <strong>of</strong> Environment, 112, pp. 1969-1981.<br />

FAO, 2001. Global Ecological Zoning for the Global Forest Resources Assessment, 2000. Forest Resources<br />

Assessment - Working Paper 56. Rome, Italy. http://www.fao.org/docrep/006/ad652e/ad652e00.htm.<br />

FAO, 2005. Land Cover Classification System (LCCS), Version 2: Classification Concepts and User Manual. Ed.<br />

Di Gregorio, A. http://www.fao.org/docrep/008/y7220e/y7220e00.htm<br />

FAO, 2006a. Global Forest Resources Assessment 2005, FAO Forestry Paper 147.<br />

ftp://ftp.fao.org/docrep/fao/008/a0400e/<br />

FAO, 2006b. The Global Administrative Unit Layers (GAUL).<br />

http://www.foodsec.org/docs/GAUL_Doc01_Ver15.pdf<br />

FAO 2007a. Global Forest Resources Assessment 2010, Options and recommendations for a global remote<br />

sensing survey <strong>of</strong> forests. Forest Resources Assessment programme Working Paper 141.<br />

ftp://ftp.fao.org/docrep/fao/010/ai074e/ai074e00.pdf.<br />

FAO, 2007b. Specification <strong>of</strong> National Reporting Tables for FRA 2010. Forest Resources Assessment<br />

Programme Working Paper 135. http://www.fao.org/forestry/media/14119/1/0/<br />

FAO, 2008. National Forest Monitoring and Assessment Manual for integrated field data collection. NFMA Working<br />

Paper 37/E.<br />

Gerrand, A. and Clancy, T. Why we need a national forest and ecosystem monitoring system: Combining the<br />

Continental Forest Monitoring Framework and Long Term Ecological Research network into a broader<br />

Terrestrial Ecosystem Research Network. Proc. Inst. <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong> Conference, C<strong>of</strong>fs Harbour, 2007.<br />

Gutman, G., Byrnes, R., Masek, J., Covington, S., Justice, C., Franks, S. and Kurtz, R., 2008. Towards monitoring<br />

land-cover and land-use changes at a global scale: The Global Land Survey 2005. Photogrammetric<br />

Engineering and Remote Sensing, 74, pp. 6–10.<br />

Hansen, M.C., Dimiceli, C., Sohlberg, R.A., 2003. Global percent tree cover at a spatial resolution <strong>of</strong> 500<br />

meters: First results <strong>of</strong> the MODIS vegetation continuous fields algorithm, Earth Interactions, 7, paper<br />

no. 10, 15pp.<br />

Hansen, M., Roy, D.P., Lindquist, E., Justice, C., Altstatt, A., 2008a. A method for integrating MODIS and<br />

Landsat data for systematic monitoring <strong>of</strong> forest cover and change in Central Africa. Remote Sensing<br />

<strong>of</strong> Environment, 112, pp. 2495–2513.<br />

Hansen, M.C., Stehman, S.V., Potapov, P.V., Loveland, T.R., Townshend, J.R.G., DeFries, R.S., Pittman, K.W.,<br />

Arunarwati, B., Stolle, F., Steininger, M.K., Carroll, M., Dimiceli, C., 2008b. Humid tropical forest clearing from<br />

2000 to 2005 quantified by using multitemporal and multi resolution remotely sensed data. Proceedings <strong>of</strong> the<br />

National Academy <strong>of</strong> Science: 105, pp. 9439-9444.<br />

Intergovernmental Panel on Climate Change, 2007. Climate change 2007—The Physical Science Basis:<br />

Contribution <strong>of</strong> Working Group I to the Fourth Assessment Report <strong>of</strong> the IPCC (Cambridge University<br />

Press, Cambridge, UK).<br />

Lambin, E., and Geist, H., (eds), 2005. Land Use and Land Cover Change: Local Processes, Global Impacts (Springer,<br />

New York).<br />

Lindenmayer, D. et al. 2007. On borrowed time: <strong>Australia</strong>'s environmental crisis and what we must do about it.<br />

Penguin books, <strong>Australia</strong>. 138 pp.<br />

Masek, J.G., Vermote, E.F., Saleous, N.E., Wolfe, R., Hall, F.G., Huemmrich, K.F., Gao, F., Kutler, J., Lim, T., 2006.<br />

A Landsat surface reflectance dataset for North America, 2990-2000. IEEE Geoscience and Remote Sensing<br />

Letters, 3, pp. 68-72.<br />

Pekkarinen, A., Reithmaier, L. Strobl, P., 2009. Pan-European forest/non-forest mapping with Landsat ETM+<br />

and Corine Land Cover 2000 data. ISPRS Journal <strong>of</strong> Photogrammetry and Remote Sensing 64, 171-183.<br />

Potapov, P., Hansen, M.C., Stehman, S.V., Loveland, T.R., Pittman, K., 2008. Combining MODIS and Landsat<br />

imagery to estimate and map boreal forest cover loss. Remote Sensing <strong>of</strong> Environment, 112, pp. 3708-<br />

3719.<br />

Richards, G. P. and C. Brack, 2004. "A continental biomass stock and stock change estimation approach for<br />

<strong>Australia</strong>." <strong>Australia</strong>n Forestry 67(4): 284-288.<br />

Stehman, S.V., 2003. Statistical sampling to characterize land cover change. Remote Sensing <strong>of</strong> Environment, 86, pp.<br />

517-529.<br />

The Group <strong>of</strong> Eight Summit, 2008. Toyako G8 Summit Leaders Declaration Hokkaido Toyako, 8 July 2008.<br />

http://www.g8summit.go.jp/eng/doc/doc080714__en.html<br />

Tucker, C.J., Grant, D.M., Dykstra, J.D., 2004. NASA’s global orthorectified Landsat data set. Photogrammetric<br />

Engineering and Remote Sensing, 70, pp. 313–322.<br />

UNFF, 2007. Non-legally Binding Instrument on all Types <strong>of</strong> Forests. UN General Assembly A/C.2/62/L.5.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 150<br />

Williams, D.L., Goward, S. and Arvidson, T., 2006. Landsat: yesterday, today, and tomorrow. Photogrammetric<br />

Engineering and Remote Sensing, 72, pp. 1171–1178.<br />

Wilson, Edward O. (ed.), 1988. Biodiversity. National Academy Press, Washington, DC, 521pp.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 151<br />

ABSTRACT<br />

SUSTAINABLE POLE SUPPLY PROJECT<br />

Hugh Stone 1 and David Wood 2<br />

Ergon Energy is facing a critical shortage in the future supply <strong>of</strong> hardwood power poles<br />

with demand expected to increase from 13,000 to 25,000 p.a. over the next 15 years.<br />

This is unlikely to be met by traditional suppliers. With over 900,000 poles in the<br />

network, <strong>of</strong> which approximately 94% are hardwood, Ergon Energy is facing a<br />

significant shortfall, commencing as early as 2009. Hardwood is the preferred pole type<br />

due to its economic advantages and its superior electrical insulation properties. It is safe<br />

and reliable.<br />

The Sustainable Pole Supply Project has been initiated to ensure that Ergon Energy has a<br />

reliable and economic future supply <strong>of</strong> poles. To achieve this, the corporation seeks to<br />

acquire native forests and to establish pole plantations. These forests will be located in<br />

diverse geographic areas to minimise risk and will be selectively harvested to provide a<br />

renewable supply <strong>of</strong> hardwood power poles while preserving ecosystems and<br />

maintaining habitats. Additionally, the benefits <strong>of</strong> carbon capture and biodiversity<br />

conservation can <strong>of</strong>fset other activities <strong>of</strong> the corporation in meeting its sustainability<br />

objectives.<br />

This paper discusses the risks and benefits to Ergon Energy and the community <strong>of</strong><br />

undertaking the project.<br />

INTRODUCTION<br />

Ergon Energy manages $5.6 billion worth <strong>of</strong> electricity infrastructure assets over one million square<br />

kilometres <strong>of</strong> regional Queensland. Ergon Energy’s service area effectively covers 97% <strong>of</strong> the state -<br />

equivalent to most <strong>of</strong> the eastern seaboard <strong>of</strong> the United States. This represents one <strong>of</strong> the largest and<br />

most diverse infrastructure networks in the western world (EECL 2007). A large majority <strong>of</strong> this<br />

network is held up, literally, by approximately 900 000 poles <strong>of</strong> which approximately 94% are<br />

hardwood. The map in Appendix 1 shows the area serviced by Ergon Energy’s network in Queensland<br />

and the area covered by the distribution network.<br />

DEMAND FOR POLES<br />

The demand for timber poles arises from network maintenance, network expansion to service<br />

consumer demand, and network upgrading to ensure feeders are not over loaded and enough capacity<br />

exists to meet peak demand.<br />

Maintenance <strong>of</strong> the Existing Network<br />

The maintenance <strong>of</strong> poles and other assets associated with the Ergon Energy network is driven by<br />

policies to meet customer expectations <strong>of</strong> reliability, employee, public and environmental safety; the<br />

present and future needs <strong>of</strong> the assets; legislative requirements; and internal and external benchmarks<br />

The vast geographical spread <strong>of</strong> Ergon Energy’s service area has a number <strong>of</strong> impacts on the<br />

performance <strong>of</strong> the distribution system. The geographic and related environmental features which<br />

reinforce the need for a robust maintenance program include:<br />

• High exposure to cyclone risk;<br />

• High storm and lightning activity;<br />

• Significant summer-winter and day-night temperature variations;<br />

• High rainfall areas;<br />

• Other weather impacts (e.g. the Channel Country flooded for hundreds <strong>of</strong> kilometres);<br />

• Active termite populations affecting power pole integrity;<br />

1<br />

Ergon Energy, Toowoomba, Qld. Email: Hugh.Stone@ergon.com.au.<br />

2<br />

Ergon Energy, Townsville, Qld.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 152<br />

• Unstable soil types (e.g. Darling Downs black cracking soils). and<br />

• Bushfire hazards.<br />

All electrical utilities have in place strategies and plans to inspect assets and ensure they are<br />

maintained and replaced on a cyclic basis. This usually entails physically inspecting each structure or<br />

piece <strong>of</strong> equipment, with cycles determined by legislation or engineering best practice (Sanders &<br />

Noonan 2008). Ergon Energy has a routine inspection and maintenance program for poles and<br />

associated equipment covering the entire network on a 4 yearly cycle, known internally as the Asset<br />

Inspection and Defect Management Program.<br />

This inspection program commenced in April 2003 (initially on a 3 year cycle) and from that time<br />

approximately 1.5 million pole inspections have been carried out. The first cycle <strong>of</strong> inspections<br />

resulted in 10,595 poles being identified as defective and consequently replaced, with a further 19,959<br />

poles treated with pole base reinforcement (installation <strong>of</strong> a large metal stake around the pole base as<br />

support, to extend the life <strong>of</strong> poles which have ground line or below ground defects) (EECL 1997).<br />

Network Expansion<br />

The growth in the population <strong>of</strong> Queensland has seen a commensurate growth in domestic demand for<br />

electricity supply, and has been accompanied by a significant industrial demand for electricity supply<br />

especially in the mining and allied sectors.<br />

For the last 5 years customer numbers have grown around 2% annually (EECL 2007). In 2007/08<br />

customer numbers increased by 2.6%. This increase has been driven largely by the expansion and<br />

development <strong>of</strong> the regional coastal centres <strong>of</strong> Hervey Bay, Mackay, Townsville and Cairns, and<br />

growth in the Toowoomba area.<br />

Network Augmentation<br />

As population increases and new industry is established there is an increase in demand from existing<br />

assets, with a need to expand the capacity <strong>of</strong> lines and substation assets to cope with increased loads..<br />

This can result in reconfiguration <strong>of</strong> distribution lines to transmission standards with larger poles and<br />

associated equipment to accommodate actual and projected increases in demand.<br />

Number <strong>of</strong> Poles<br />

26000<br />

24000<br />

22000<br />

20000<br />

18000<br />

16000<br />

14000<br />

12000<br />

10000<br />

8000<br />

6000<br />

4000<br />

2000<br />

0<br />

2004<br />

2006<br />

2008<br />

Future Hardwood Poles Requirement<br />

2010<br />

2012<br />

2014<br />

2016<br />

2018<br />

Year<br />

2020<br />

2022<br />

2024<br />

Defective Replacement New Extension Augmentation<br />

Figure 1 - Future hardwood pole requirement<br />

Combining the pole demand requirements from maintenance <strong>of</strong> the existing network, network<br />

expansion to meet demand from new customers and network augmentation, Ergon Energy’s current<br />

2026<br />

2028<br />

2030


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 153<br />

demand for hardwood poles is around 13,000 annually. Research conducted internally suggests that<br />

this demand is set to increase steadily to around 25,000 poles annually by 2031 (EECL, nda). Figure 1<br />

presents this projected pole demand.<br />

This extrapolation assumes the age pr<strong>of</strong>ile <strong>of</strong> existing poles created from historical data, will continue<br />

to apply. Analysis indicated a pole life <strong>of</strong> 51 years ± 11 years. The projected curve <strong>of</strong> future pole<br />

requirements flattens <strong>of</strong>f from about 2022 due largely to an expectation that a higher proportion <strong>of</strong><br />

new development will move cable underground.<br />

POLE SUPPLY<br />

The South East Queensland Forest Agreement (SEQFA) signed in 1999 was an agreement between the<br />

Queensland Government, the timber industry and conservation groups that sought to secure a viable<br />

timber industry while securing an adequate and representative conservation estate. The agreement<br />

involves the transition <strong>of</strong> publicly owned commercial native forests (State Forests) into the<br />

conservation estate over a 25 year period with essentially a final harvest <strong>of</strong> all merchantable wood<br />

products as each area is completed.<br />

In the area covered by the SEQFA the supply <strong>of</strong> poles from State land is governed by a policy<br />

stipulating that the current supply, which extends until 2014, will be maintained at historic levels<br />

existing prior to commencement <strong>of</strong> the SEQFA. Supply arrangements will be reviewed at the<br />

completion <strong>of</strong> this period and there is no in-principle reason why new contracts could not be entered<br />

into for pole supply up until the end <strong>of</strong> the SEQFA period in 2024. Present Government policy<br />

stipulates that there will be no net increase in the supply <strong>of</strong> pole products, and that substitute policies<br />

will promote a transition from products supplied from native forest to plantation products (Stephen<br />

Walker DNRW, pers. comm. 16/2/09) under the SEQFA.<br />

The phasing out <strong>of</strong> native forest harvesting is accompanied by a transition plan to move to plantation<br />

timber (both hardwood and s<strong>of</strong>twood) for all forest products. Despite comprehensive strategies to<br />

allow industry to make this transition, the use <strong>of</strong> plantation grown poles is largely untested and has not<br />

provided a degree <strong>of</strong> confidence sufficient for Ergon Energy to rely solely upon plantations to meet a<br />

shortfall <strong>of</strong> pole supplies as native forest logging is wound up.<br />

Traditionally, hardwood poles have been obtained 35% from public resources and 65% from private<br />

property. (Chris Bragg, DNRW pers comm. 20/02/09). However, while public land accounts for only<br />

35%, they are typically among the larger pole sizes (12.5m and 14m), while private resources provide<br />

poles <strong>of</strong> a lower length and strength range (EECL, n.d.).<br />

Given then that Ergon Energy faces a sharp increase in demand for hardwood poles, and that current<br />

suppliers are unable to increase their supply, and particularly for the common 12.5m and 14m pole<br />

lengths, either a new approach to sourcing hardwood timber poles is required, or a suitable alternative<br />

has to be found.<br />

ALTERNATIVES TO HARDWOOD POLES<br />

Plantation S<strong>of</strong>twood<br />

An in-service trial <strong>of</strong> slash pine poles commenced in the mid eighties. After 15 years, results were<br />

promising, suggesting physical deterioration <strong>of</strong> CCA treated slash pine poles is only marginally higher<br />

than that <strong>of</strong> CCA treated hardwoods (Powell , nda). However, the strength <strong>of</strong> s<strong>of</strong>twood poles has been<br />

questioned and consequently they are not currently considered suitable for all Ergon Energy<br />

applications.<br />

While the substitution <strong>of</strong> plantation grown s<strong>of</strong>twood poles is feasible, this option would require a<br />

relatively long planning horizon to realise, on the assumption that a) existing maturing plantations are<br />

likely to be committed already under wood supply agreements; and b) a s<strong>of</strong>twood pole clearly needs to<br />

be larger in diameter to be equivalent in strength to a hardwood pole, and would therefore need to be<br />

sourced from end <strong>of</strong> rotation stands. Thus while s<strong>of</strong>twood poles may provide a longer term option,<br />

they cannot be considered a realistic substitute for hardwood poles for at least the next 25 years, until<br />

young unallocated s<strong>of</strong>twood plantations reach the necessary pole size.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 154<br />

Concrete Poles<br />

Concrete poles can be readily supplied in specific sizes and they <strong>of</strong>fer an option that matches or<br />

exceeds the strength properties <strong>of</strong> hardwood poles. Additionally, they have greater structural<br />

reliability and reduced maintenance costs. As such they are preferred for use in transmission and sub<br />

transmission (66kV and above). Their use in the distribution network is not common due to the initial<br />

high capital cost <strong>of</strong> concrete poles (refer Table 1), higher field installation costs, considerations <strong>of</strong><br />

current leakage and the lack <strong>of</strong> versatility in accommodating additional future fittings (eg crossarms).<br />

While their use is warranted within the distribution network as transformer poles and on higher<br />

voltage sub transmission feeders, they are not as versatile as hardwood poles and their higher capital<br />

costs make it difficult to justify them for universal use.<br />

The environmental impact and sustainability <strong>of</strong> using concrete poles also needs to be considered.<br />

Manufacture <strong>of</strong> these poles requires use <strong>of</strong> non-renewable raw materials and large amounts <strong>of</strong> energy<br />

producing a relatively large carbon footprint in comparison to the use <strong>of</strong> timber which is both a<br />

renewable resource and a carbon sink.<br />

Ergon Energy continues to research alternatives to solid hardwood timber poles with long term field<br />

trials. Other alternatives include steel poles, fibre composites, laminated timber and hybrids using a<br />

combination <strong>of</strong> materials. Hardwood timber maintains its status as the preferred option for use in the<br />

distribution network mainly because <strong>of</strong> cost. It has desirable electrical properties, it is safe, strong and<br />

reliable and importantly, in a time when climate change is a real concern and must be considered in<br />

long term business planning, it is a renewable resource that sequestrates carbon.<br />

Table 1 A comparison <strong>of</strong> pole costs for timber and concrete<br />

Pole Size<br />

length/strength<br />

Average Pole cost ex<br />

depot ($ per pole)<br />

Con-<br />

Wood<br />

crete<br />

Typical use<br />

9.5/5 $140 $400 Urban Service pole<br />

11/5 $180 $490 SWER intermediate pole or Urban LV reticulation pole<br />

12.5/5 $220 $690 Urban HV/LV reticulation pole<br />

12.5/12 $490 $950 Urban Transformer pole<br />

14/8 $410 $800 Rural backbone feeder intermediate pole<br />

18.5/12 $870 $1750 66kV transmission intermediate pole<br />

20/12 $1360 $2000 66kV transmission intermediate pole<br />

SUSTAINABLE POLE SUPPLY PROJECT (SPSP)<br />

With no immediate substitute for hardwood timber poles, an immediate forecast shortage <strong>of</strong> poles over<br />

the next 20 years, a likely continuing demand for poles well beyond the 20 year projections demand<br />

cited above, it was logical step for Ergon Energy to embark on a project which will provide a<br />

sustainable supply <strong>of</strong> hardwood timber poles – the SPSP.<br />

The SPSP will ultimately provide a sustainable source <strong>of</strong> poles capable <strong>of</strong> satisfying 50% <strong>of</strong> the<br />

projected demand. The project scope allows for the purchase or lease <strong>of</strong> up to 10 000 hectares <strong>of</strong> land<br />

suitable for growing timber. The majority <strong>of</strong> the areas covered by the project will be native forests,<br />

where enrichment planting would be done in understocked or degraded forest. There is also provision<br />

for the conversion <strong>of</strong> previously cleared farm land to hardwood plantations. Exercising these options<br />

would optimise production <strong>of</strong> poles from areas with a mosaic <strong>of</strong> past land use, and add some benefits<br />

by means <strong>of</strong> more efficient land management and enhanced environmental values.<br />

Internally, the project is directed to the search for and acquisition <strong>of</strong> land considered suitable for the<br />

organisation’s needs, while an external forest manager is engaged to provide advice on the suitability<br />

<strong>of</strong> potential properties and to create and implement a forest management plan for each acquired<br />

property. It is intended that, once the SPSP is established, minimal internal resources will be required<br />

to manage the project with a heavy reliance on external resources to provide ongoing forest<br />

management expertise.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 155<br />

Ideally the properties will be 1,000 to 2,000 hectares in size to allow geographical separation <strong>of</strong> each<br />

property in an effort to minimise risks associated with pathogens, fire and climatic influences. The<br />

first property was purchased in 2008 under the project and selection <strong>of</strong> additional properties is well<br />

advanced.<br />

The SPSP contains other options than outright purchase or lease.. Agreement has been made with<br />

Wide Bay Water to purchase poles from plantations established to use sewage effluent in the Hervey<br />

Bay area. Discussions have also been held with representatives from <strong>Australia</strong>n Forest Growers on<br />

long term sales agreements for suitable species. No specific model <strong>of</strong> management and procurement is<br />

stipulated at this stage; rather, a range <strong>of</strong> options will be pursued to bring more flexibility into the<br />

supply chain.<br />

Property Selection<br />

To assist in the location <strong>of</strong> potential properties for purchase or lease, a combination is employed <strong>of</strong><br />

desktop GIS analysis, field investigation and inspection, followed by contact with stock and station<br />

agents. The first step is to define areas more likely to contain appropriate vegetation types with<br />

desirable pole species, soils and climate conditions. A GIS consultant was engaged to identify suitable<br />

land parcels consistent with a pre determined scope and set <strong>of</strong> criteria.<br />

Figure 2. Example <strong>of</strong> map output with Regional Ecosystems containing favoured species shaded<br />

in different colours. Orange–Corymbia citriodora, Light Blue-Eucalyptus acmenoides,<br />

Dark Blue-Eucalyptus sideroxylon. (Fitzpatrick 2009)<br />

The data sets used in the GIS analysis are:<br />

• Queensland Department <strong>of</strong> Natural Resources and Water (DNRW) FPC data for<br />

entire Queensland provided in MGA zones 54, 55, 56 (QLD DNRW 2007);<br />

• Queensland Department <strong>of</strong> Natural Resources and Water (DNRW) LANDSAT scene<br />

data for entire Queensland (QLD DNRW 2007);<br />

• Queensland Local Government Boundaries for entire Queensland (DNRW).<br />

• Queensland Environment Protection Agency’s Version 5.0 Regional Ecosystems and<br />

Remnant Vegetation (RE) data for entire Queensland (QLD EPA 2007); and


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 156<br />

• Cadastre with tenure attributes over the Stage One project area (QLD Lands<br />

Department).<br />

• Ergon list <strong>of</strong> acceptable and preferred pole species<br />

The criteria in use to identify potentially suitable properties that warrant further on-ground<br />

investigation are:<br />

• Searches are restricted to freehold land only.<br />

• A minimum land parcel size <strong>of</strong> 100 hectares will be used as a filter.<br />

• DNR&W Foliage Projection Cover (FPC) based on 12% FPC which equates to an<br />

approximate canopy cover <strong>of</strong> 20%.(to capture regrowth areas).<br />

• A refined list <strong>of</strong> Regional Ecosystems (RE’s) most likely to contain suitable forest<br />

structure (predominantly woodland and tall forest) and species with Durability Class 1<br />

& 2 species.<br />

• Only RE’s with a Vegetation Management Act status <strong>of</strong> ‘Not <strong>of</strong> Concern’ are<br />

desirable. It was considered filtering to RE’s listed as ‘Of Concern’ or ‘Endangered’<br />

carried risks for future availability for harvest. This list was created manually based on<br />

published descriptions <strong>of</strong> each Regional Ecosystem and local knowledge.<br />

This process creates a map output that can be used to target geographic locations with mapped suitable<br />

vegetation. It should be noted that non-remnant forested areas suitable for the project are also under<br />

consideration, and in some ways are more desirable in terms <strong>of</strong> <strong>of</strong>fset trading. These are difficult to<br />

refine through GIS desktop analysis as the vegetation type is not mapped; however, assuming the FPC<br />

criteria are met, and mapped ‘Not <strong>of</strong> Concern’ remnant vegetation is adjacent or in the vicinity, they<br />

are included in the final output.<br />

A comprehensive due diligence process has been developed to ensure that the growing stock and site<br />

quality <strong>of</strong> purchased properties will provide a reasonable return on investment. Legal due diligence<br />

has also included assessment <strong>of</strong> environmental risks. As a result, several properties have been rejected<br />

because their growth potential has not been able to match possible liabilities presented by major weed<br />

infestations.<br />

Investigation <strong>of</strong> suitable properties is ongoing and properties have been purchased at Ravensbourne<br />

(‘Eagle Rock’), North East <strong>of</strong> Toowoomba and at Tiaro, North <strong>of</strong> Gympie. A forest management plan<br />

for Eagle Rock has been created and its implementation commenced with trials <strong>of</strong> both commercial<br />

and non commercial thinning in progress, with a view to improving stands to focus site resources on to<br />

future pole crop trees. Some enrichment planting <strong>of</strong> old log dumps, rehabilitation <strong>of</strong> access tracks and<br />

weed control has also been carried out.<br />

Timber stands will be selectively harvested according to the Queensland Code <strong>of</strong> Practice<br />

administered by DNRW. Management will adopt the principles <strong>of</strong> Ecological Sustainable Forest<br />

Management (ESFM) with the intent <strong>of</strong> maximising all forest values including biodiversity and<br />

conservation values.<br />

BENEFITS ASSOCIATED WITH THE PROJECT<br />

The main benefit <strong>of</strong> the SPSP is a sustainable supply <strong>of</strong> hardwood poles to help meet internal current<br />

and projected demand for hardwood poles. While a proportion <strong>of</strong> Ergon’s forecast pole demand will<br />

still be sourced on the open market, having its own sustainable supply <strong>of</strong> poles <strong>of</strong>fers a degree <strong>of</strong><br />

protection from market forces and security for business planning.<br />

Other benefits associated with the project include carbon sequestration; biodiversity <strong>of</strong>fset credits, and<br />

conservation and ecosystem health.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 157<br />

Carbon Sequestration<br />

The introduction <strong>of</strong> an Emissions Trading Scheme (ETS) by the <strong>Australia</strong>n Government is unlikely to<br />

provide Ergon Energy with carbon credits available for trade, as sequestration <strong>of</strong> carbon from native<br />

forests is not included as a source <strong>of</strong> carbon credits within the current proposed ETS framework.<br />

While plantations established on previously cleared land acquired under the project could be certified<br />

for carbon trading benefits, the benefits associated with carbon sequestration in terms <strong>of</strong> corporate and<br />

community responsibility are desirable even without an economic gain in the form <strong>of</strong> emission<br />

credits.. The impacts on climate change from carbon emitted into the atmosphere through human<br />

activity are generally accepted as a threat to the health <strong>of</strong> the global environment, and individuals and<br />

corporations are encouraged to reduce their carbon footprint as best they can. The sustainable pole<br />

project works to capture and store atmospheric carbon and effectively reduce Ergon Energy’s<br />

footprint.<br />

Biodiversity Offset Credits<br />

The Queensland Government introduced the Queensland Government Environmental Offsets Policy<br />

(QGEOP) on the 1st July 2008 across the whole <strong>of</strong> government in an effort to replace environmental<br />

values lost through development. The policy is supported by <strong>of</strong>fset policies that provide detailed<br />

direction for specific environmental issues (vegetation, koala habitat, marine fish habitat). These<br />

Activity-Specific Offset Policies are developed and administered by different government agencies<br />

according to their environmental portfolio.<br />

The QGEOP defines an <strong>of</strong>fset as “an action taken to counterbalance unavoidable, negative<br />

environmental impacts that result from an activity or a development. An <strong>of</strong>fset may be located within<br />

or outside the geographic site <strong>of</strong> the impact. Offsets can be either direct or indirect. A direct <strong>of</strong>fset is<br />

aimed at on-ground maintenance and improvement <strong>of</strong> habitat or landscape values, such as long term<br />

protection <strong>of</strong> existing habitat, or legally securing regrowth vegetation for inclusion as remnant<br />

vegetation upon maturity. An indirect <strong>of</strong>fset aims to improve knowledge, understanding and<br />

management leading to improved conservation outcomes. Examples <strong>of</strong> indirect <strong>of</strong>fset include funding<br />

species-specific research, sponsoring implementation <strong>of</strong> species recovery plans, or providing<br />

infrastructure, such as nest boxes or fauna crossings, to minimise impacts on fauna. Generally<br />

speaking a direct <strong>of</strong>fset is preferable to an indirect <strong>of</strong>fset, because the impact is easier to measure.<br />

According to the QGEOP, where Ergon’s activities affect mapped remnant vegetation, essential<br />

habitat (under the Vegetation Management Act) or listed Rare, Vulnerable or Endangered species<br />

(under the Nature Conservation Regulations 2008) an <strong>of</strong>fset must be provided in line with the relevant<br />

Activity-Specific Offset Policy, potentially at a ratio greater than 1:1.<br />

While the implementation <strong>of</strong> the QGEOP with its associated activity-specific <strong>of</strong>fset policies is in its<br />

infancy, and remains largely untested, it currently requires any adverse impact on native vegetation to<br />

be <strong>of</strong>fset to ensure there is no overall loss <strong>of</strong> environmental value as a result <strong>of</strong> the activity.<br />

A general interpretation <strong>of</strong> the Activity-Specific Offset Policy for vegetation would be the securing <strong>of</strong><br />

like vegetation not currently mapped as “remnant” 3 . Preferably this vegetation would be located<br />

geographically close to the area being disturbed. For example, if 5 hectares <strong>of</strong> “Of Concern” remnant<br />

vegetation is cleared, the administering authority may issue a conditional permit requiring the holder<br />

to legally secure a floristically and ecologically “like” area <strong>of</strong> non-remnant vegetation for future<br />

inclusion as mapped remnant vegetation – essentially replace what will be removed from the mapped<br />

remnant vegetation map as a result <strong>of</strong> their development. Our experience to date has seen disturbed<br />

areas mandated to be <strong>of</strong>fset at ratios greater than 1:1 depending on the conservation status <strong>of</strong> the<br />

ecosystem disturbed.<br />

A general interpretation <strong>of</strong> the policy under an activity-specific <strong>of</strong>fset policy for biodiversity may<br />

require a net increase <strong>of</strong> individuals from pre disturbance levels for each individual species impacted.<br />

Given this, the impact to Ergon Energy is largely dependent on interpretation. As the NCA defines all<br />

native vegetation as protected, a strict interpretation <strong>of</strong> the policy would define any disturbance to a<br />

3<br />

“Remnant” status is technically achieved when the upper stratum <strong>of</strong> a stand reaches 70% <strong>of</strong> site height and crown cover <strong>of</strong><br />

the upper stratum is => 50%.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 158<br />

native species, regardless <strong>of</strong> its individual or ecosystem conservation status, as a conservation loss<br />

which therefore must be <strong>of</strong>fset. By definition <strong>of</strong> the policy, the <strong>of</strong>fset must result in a net increase in<br />

individual numbers for each species impacted.<br />

For land owners the introduction <strong>of</strong> this policy brings financial opportunity through the sale <strong>of</strong> <strong>of</strong>fset<br />

credits. In the vegetation example above, the permit holder could secure an <strong>of</strong>fset from a land owner<br />

by purchasing the vegetation community values <strong>of</strong> their property (or part there<strong>of</strong>) by way <strong>of</strong> a caveat<br />

over the property. Essentially this would result in the permit holder paying a nominated value per unit<br />

area to the land owner, who in return is legally bound to retain the native vegetation which in time<br />

would reach “remnant” status, be classified, and subject to legislative controls.<br />

Regardless <strong>of</strong> the interpretation <strong>of</strong> the policy, with Ergon’s continually expanding distribution<br />

network, the corporation will be regularly required to secure <strong>of</strong>fsets. Being a landowner will<br />

potentially make the process <strong>of</strong> sourcing and securing <strong>of</strong>fsets much more efficient. Furthermore, land<br />

acquired under the project will provide opportunities for sale <strong>of</strong> <strong>of</strong>fset credits to other entities seeking<br />

<strong>of</strong>fsets for their impacts. The sale <strong>of</strong> <strong>of</strong>fset credits would provide a positive financial injection back<br />

into the project, effectively reducing the price paid for the property and providing funds to manage<br />

biodiversity threats, such as fire, pests and weeds, to ensure remnant status is achieved.<br />

Current experience suggests the sale <strong>of</strong> <strong>of</strong>fset credits will provide a per hectare payment up-front to<br />

compensate the landowner for retaining the vegetation, and a further per hectare sum for management<br />

costs associated with restoring the vegetation to remnant status. Management costs are assigned for the<br />

management <strong>of</strong> fire, pests, weeds and any other long term threat to the health <strong>of</strong> the vegetation. An<br />

<strong>of</strong>fset agreement is drawn up detailing responsibilities <strong>of</strong> all parties in ensuring the vegetation reaches<br />

remnant status. Once the regulator assesses the area as remnant, the <strong>of</strong>fset area is included on the<br />

<strong>of</strong>ficial remnant vegetation mapping and the <strong>of</strong>fset agreement is deemed to be satisfied.<br />

In summary, the Sustainable Pole Supply Project will provide a residual benefit through biodiversity<br />

<strong>of</strong>fset credits which can be used internally to satisfy <strong>of</strong>fset obligations stemming from other arms <strong>of</strong><br />

the business, and/or sold externally to provide income to the project and funds for land management.<br />

Conservation and Ecosystem Health<br />

Timber poles can be produced without compromising biodiversity or conservation values. Through<br />

the application <strong>of</strong> the principles <strong>of</strong> Ecologically Sustainable Forest Management (ESFM) all forest<br />

values can be maintained. Sensitive areas, such as riparian zones, will be identified and then excluded<br />

from the harvest area. In the case <strong>of</strong> Eagle Rock, past land use has led to a degraded forest structure<br />

and health. Ergon Energy is confident that biodiversity and conservation values will increase over<br />

time as the benefits from sustainable silvicultural treatment and management are realised.<br />

Initial approaches have been made to universities for the conduct <strong>of</strong> surveys for threatened species and<br />

to monitor biodiversity programs. This will produce information on biodiversity changes over time,<br />

and provide an education resource for the universities.<br />

RISKS ASSOCIATED WITH THE PROJECT<br />

In managing the projected increase in demand for hardwood poles against a predicted reduction in<br />

market supply, the greatest risk to Ergon Energy would stem from doing nothing. The acquisition <strong>of</strong><br />

land provides Ergon with an asset which can be liquidated quickly, while the identified benefits justify<br />

the SPSP proceeding. Against these benefits, the risk associated with the project is considered<br />

acceptable. Other risks include:<br />

a) The objectives <strong>of</strong> the project may not be realised, or the percentage <strong>of</strong> poles produced may be less<br />

than anticipated. However if a tree does not achieve the form and quality standards to be utilised as a<br />

pole, it would be saleable as a quota sawlog or salvage grade round timber, at a lower price. In all<br />

stands there will be a proportion <strong>of</strong> trees that do not meet pole specifications.<br />

b) This brings with it an associated risk in the need to focus on non-core business, that is, managing<br />

timber sales. However the revenue returned from timber sales should at least neutralise the associated<br />

costs.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 159<br />

c) The major changes in government laws and regulations than occur over the lifetime <strong>of</strong> even a shortrotation<br />

pole plantation are recognised as a significant risk, and one that Ergon Energy cannot avoid<br />

through planning and good management. The greatest risk in this regard would be a change in<br />

government policy on the legality <strong>of</strong> harvesting privately owned native forests. Another area <strong>of</strong><br />

uncertainty are the permissible activities in mapped remnant areas. At the time <strong>of</strong> writing, the media<br />

has speculated on the impacts <strong>of</strong> a review <strong>of</strong> laws associated with native forest regrowth and its<br />

associated mapping. The conversion <strong>of</strong> mapped “regrowth” areas to a “remnant” vegetation category<br />

is one example <strong>of</strong> how a legislative change could dramatically change the pole supply resource.<br />

Should this happen, the concept <strong>of</strong> biodiversity <strong>of</strong>fset trading under the Vegetation Management Act<br />

could be compromised.<br />

d) Another significant risk, which has been evident from the commencement <strong>of</strong> the project, has been<br />

competition from other land uses, such as agriculture and horticulture for the better quality land sought<br />

for forestry.<br />

e) Another recognised risk is the false perception among some agricultural sectors that the demand for<br />

land for forest establishment will threaten their industry. This has been seen, for example, with some<br />

members <strong>of</strong> the sugar industry, and has been accompanied by public and political lobbying. However<br />

the Sustainable Supply Pole Project is not expected to meet this type <strong>of</strong> resistance because <strong>of</strong> its focus<br />

mainly on existing regrowth, the dispersed nature <strong>of</strong> property purchases, and the low key approach<br />

adopted in seeking suitable properties.<br />

f) The supply <strong>of</strong> poles from Ergon owned and managed forests could adversely impact private<br />

suppliers. This is an unlikely risk as pole delivery can be managed so that no suppliers are in fact<br />

adversely impacted. This could simply be effected by ensuring Ergon poles are distributed across<br />

Ergon regions and depots. As other electricity utilities face the same dilemma <strong>of</strong> future shortages <strong>of</strong><br />

hardwood poles, the formation <strong>of</strong> joint ventures with private growers, or other corporate structures, to<br />

ensure security <strong>of</strong> supply is anticipated.<br />

CONCLUSION<br />

The hardwood pole is a reliable, safe, renewable asset on which Ergon Energy relies for the<br />

distribution <strong>of</strong> electricity across Queensland. The likely unavailability <strong>of</strong> hardwood poles from current<br />

suppliers has been the catalyst for Ergon Energy’s proactive initiation <strong>of</strong> the Sustainable Pole Supply<br />

Project with the acquisition <strong>of</strong> native forests to help meet Ergon’s future pole needs, and the<br />

management <strong>of</strong> those forests sustainably for a wide range <strong>of</strong> forest values.<br />

ACKNOWLEDGEMENTS<br />

The authors wish to acknowledge the assistance <strong>of</strong> David Carleton, Manager <strong>of</strong> the Sustainable Pole<br />

Supply Project, and Bernard Fitzpatrick, CTG Consulting, who provided some <strong>of</strong> the background<br />

information presented in the paper.<br />

REFERENCES<br />

Ergon Energy Corporation Limited. n.d. Hardwood Pole Shortage Report. Internal Report prepared by J. Brooks<br />

and P. Le. Network Maintenance and Standards, Ergon Energy Corporation Ltd, Queensland.<br />

Ergon Energy Corporation Limited. 2008. Network Management Plan Part A - Electricity Supply for Regional<br />

Queensland 2008/09 to 2012/13. Ergon Energy Corporation Ltd, Queensland.<br />

Sanders, A & Noonan, K. 2008 Wood Pole Management , Overhead Lines Seminar 4 - 5 March 2008 Sydney.<br />

CIGRE <strong>Australia</strong>n Panel B2- Overhead Lines.<br />

Powell, M. n.d. Inspection and Assessment <strong>of</strong> Slash Pine Utility Poles after approximately 15 years in-service<br />

exposure in several areas <strong>of</strong> Queensland. Queensland Forestry Research <strong>Institute</strong>, Queensland.<br />

Wood Pole versus Concrete Pole Comparison. Internal Report, Network Maintenance and Standards, Ergon<br />

Energy Corporation Ltd, Queensland.<br />

Fitzpatrick, B. Identification <strong>of</strong> potential land parcels for poles, South East Queensland Biogeographic Region.<br />

Internal report produced for Ergon Energy by CTG Consulting February 2009.<br />

Queensland Government Environmental Offsets Policy Queensland Government Environmental Protection<br />

Agency, 2008, //www.epa.qld.gov.au/publications/p02501aa.pdf/


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 160<br />

APPENDIX 1 – ERGON ENERGY’S SERVICE AREA


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 161<br />

PROCESSING PERFORMANCE AND SAWN PRODUCT RECOVERY<br />

FROM THINNED NATIVE FOREST REGROWTH LOGS<br />

FROM SOUTHERN AUSTRALIA<br />

Russell Washusen 1 , Andrew Morrow 1 , Dung Ngo 1 , Graeme Siemon 2 ,<br />

Tim Wardlaw 3 , Mike Ryan 4 , Martin Linehan 5 and Daniel Tuan 5<br />

ABSTRACT<br />

Logs <strong>of</strong> conventional sawlog quality, diameter and length were obtained from thinned<br />

E. fastigata (New South Wales), E. sieberi (Victoria), E. regnans (Victoria), E. regnans<br />

(Tasmania) and E. diversicolor (Western <strong>Australia</strong>). They were sawn and dried using<br />

normal industry processing methods for the respective species. The wood processing<br />

performance and product quality were compared with samples <strong>of</strong> logs <strong>of</strong> similar grade,<br />

diameter and length, from normal log supplies from much older unthinned regrowth.<br />

Losses due to board end-splitting, board deflection and drying performance <strong>of</strong> the<br />

wood from the thinned regrowth was comparable to, and in some cases, better than the<br />

older regrowth; and the general wood quality tended to be better in the younger thinned<br />

logs due to less frequent kino and insect damage. The results give support to the<br />

thinning programs applied in even-aged forests by forest management agencies across<br />

southern <strong>Australia</strong>.<br />

INTRODUCTION<br />

In eucalypt species that tend to produce even-aged stands, thinning has been conducted experimentally<br />

by forest management agencies across southern <strong>Australia</strong> since about the 1950’s. This early work led<br />

to the first serious attempts to develop commercially viable and silviculturally beneficial mid-rotation<br />

thinning operations during the “Young Eucalypt Program” in the late 1980’s (Kerruish and Rawlins<br />

1991). Commercial thinning <strong>of</strong> small areas <strong>of</strong> eucalypt production forests is now a common practice.<br />

The majority <strong>of</strong> these thinned forests are yet to contribute to a significant supply <strong>of</strong> sawlogs. However,<br />

some <strong>of</strong> the earlier experiments have produced trees <strong>of</strong> a harvestable size and the thinned stands are<br />

generally approaching the time when they will produce a significant supply <strong>of</strong> logs for conventional<br />

processing.<br />

Thinning treatments, where they have been monitored, have <strong>of</strong>ten produced growth responses in the<br />

retained trees (Webb 1966, Goodwin 1990, West 1991, O’Shaughnessy and Jayasuriya 1993, Fagg<br />

2006). The most important effect is accelerated diameter growth rate and as a consequence increased<br />

sawlog yields.<br />

Despite the work <strong>of</strong> Wardlaw et al (2004) who examined differences between large thinned regrowth<br />

logs and smaller unthinned regrowth, and Innes et al. (2005) who examined differences in wood<br />

properties <strong>of</strong> trees <strong>of</strong> different ages and sizes, there is little information available about the comparable<br />

performance <strong>of</strong> logs from thinned regrowth and older conventional sources. This is because direct<br />

comparisons have not been made between logs <strong>of</strong> comparable diameter and grade obtained from older<br />

conventional sawlog supplies and thinned forests. In the latter case, the logs will tend to produce a<br />

sheath <strong>of</strong> wood after thinning (Figure 1) that has different wood properties than much slower grown<br />

unthinned regrowth.<br />

Some processors are concerned that this growth response may alter wood properties and consequently<br />

processing performance. The magnitude <strong>of</strong> longitudinal tensile growth stresses may be different,<br />

1<br />

CSIRO Materials Science and Engineering and CRC for Forestry, Contact author. Email: Russell.Washusen@csiro.au<br />

2<br />

Forest Products Commission<br />

3<br />

Forestry Tasmania<br />

4<br />

VicForests<br />

5<br />

Forests NSW


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 162<br />

impacting on board end-split severity, board deflection during sawing and product undersizing and, as<br />

a consequence, impact on product recovery and value. During drying the wood may produce different<br />

drying performance and influence the commercially important defects <strong>of</strong> surface and internal<br />

checking.<br />

Figure 1. Radial strips from the base <strong>of</strong> the stem from trees <strong>of</strong> similar diameter from 1939<br />

regrowth E. regnans (bottom) and 1972 regrowth E. regnans thinned in 1990 (top)<br />

showing differences in growth ring width in the outer heartwood and sapwood.<br />

This project was initiated at the request <strong>of</strong> the processing industry in <strong>Australia</strong> which wanted to<br />

quantify differences in processing performance and general wood quality in thinned regrowth forests<br />

where there had been a growth response. To do this, log samples were selected from thinned and<br />

unthinned forests so that log diameter, quality and height in the stem were similar.<br />

The species examined were suggested by industry because <strong>of</strong> their importance to processors and their<br />

availability from thinning experiments. They were E. fastigata from New South Wales, E. sieberi from<br />

Victoria, E. regnans from Tasmania and Victoria, and E. diversicolor from southwest Western<br />

<strong>Australia</strong>. It was the primary objective <strong>of</strong> this project to quantify differences in wood processing<br />

performance and product quality in logs <strong>of</strong> a size and quality suitable for processing in existing<br />

sawmills.<br />

This work produced five reports for Forest and Wood Products <strong>Australia</strong> (FWPA), one report for each<br />

species by region (Washusen et al. 2007 a,b,c,d,e). This paper summarises these sub-projects and<br />

presents the main findings.<br />

METHODS<br />

The evaluation <strong>of</strong> each species by region was conducted in sawmills located close to the forest <strong>of</strong><br />

interest, and currently processing the respective species from unthinned regrowth sources. This meant<br />

that each evaluation had unique attributes, partly because <strong>of</strong> differences in processing methods, and<br />

partly due to differences in target markets and consequential customer product requirements. While<br />

these differences were inevitable, each sub-project aimed to understand the effect <strong>of</strong> increased growth<br />

rates following thinning on key indicators <strong>of</strong> wood quality and value. In order to do this each subproject<br />

adhered as close as possible to the structure that follows.<br />

Initially, trees from thinned stands were selected and harvested to produce either a single butt log or<br />

two logs from the lower stem. These were then matched with logs from the same height so that the<br />

diameter and log grade (for the respective forest management agencies) were as similar as possible.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 163<br />

Coupe selection was also structured to minimise variation in elevation, slope, rainfall, soil types and<br />

genetic origins. Examples <strong>of</strong> coupe maps and thinned forests are shown in Figures 2 and 3. These are<br />

coupe maps for the E. regnans from the Styx Forest Block in Tasmania (Fig. 2) and thinned E.<br />

regnans from the Central Highlands <strong>of</strong> Victoria (Figure 3).<br />

Figure 2. Coupe maps for unthinned E. regnans (left) and thinned E. regnans from the Styx<br />

Forest Block, Tasmania.<br />

Figure 3. Thinned E. regnans from the Central Highlands, Victoria.<br />

Where possible, harvesting <strong>of</strong> the respective thinned and unthinned coupes for each species by region,<br />

was conducted at about the same time and by the same methods. Logs were transported to the mills<br />

either in sawlog length for a single log per tree, or bushlog length where two logs were produced. At<br />

the mill the logs were measured and graded according to the requirements <strong>of</strong> the respective State forest<br />

management agencies. The logs were colour coded so each board produced could be tracked<br />

throughout processing and individual log statistics produced.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 164<br />

The randomization <strong>of</strong> the logs for sawing ensured that the following conditions were imposed on the<br />

processing:<br />

• Identity <strong>of</strong> the logs was unknown throughout processing except by supervising research<br />

staff.<br />

• Logs were processed by identical strategies in each sawmill.<br />

• Boards from each sample were randomised within a single drying batch and so were dried<br />

with the same schedule, and the position within the kiln was reduced as much as possible<br />

as a variable in drying performance.<br />

• During the grading by industry graders (Figure 4), resource identity was unknown.<br />

• During processing staff were employed on one task, and normal staff rotations avoided to<br />

ensure consistency in processing strategies.<br />

Many <strong>of</strong> the conditions imposed on the research restricted the size <strong>of</strong> the project because <strong>of</strong> the<br />

expense in handling and tracking material. Sample size was restricted to a maximum <strong>of</strong> 30 m 3 <strong>of</strong><br />

sawlogs from each thinned and unthinned forest. This was considered imperative in order to eliminate<br />

as many as possible <strong>of</strong> the variables that may impact on processing performance. These variables<br />

included sawing techniques, resulting in differences in product orientation and sizing, differences in<br />

drying schedules and drying conditions, and differences in interpretation <strong>of</strong> board grade and final<br />

product size.<br />

A general summary <strong>of</strong> the log sample locations, age at establishment and thinning, sawing and drying<br />

methods and final products produced are given in Tables 1 and 2. The year <strong>of</strong> thinning for most<br />

species ranged from 1990 to 1994 so that at harvest at least 12 to 16 years had elapsed since thinning.<br />

One sample <strong>of</strong> E. regnans from the Styx Forest Block in southern Tasmania was thinned in 1970.<br />

The coupes selected from the unthinned forests were scheduled for normal harvesting operations. They<br />

ranged from mature stands or stands dominated by mature trees <strong>of</strong> unknown age (E. fastigata, E.<br />

diversicolor and E. sieberi) with 1939 and 1934 regrowth E. regnans from Victoria and Tasmania<br />

respectively. One stand <strong>of</strong> unthinned 1974 regrowth E. diversicolor was also selected in Western<br />

<strong>Australia</strong>.<br />

Processing methods (sawing and drying) and final products for the respective species were normal for<br />

the sawmills where the processing was conducted. Back-sawing strategies and quarter-sawing<br />

strategies to produce sized boards for drying were applied for E. diversicolor and E. sieberi and E.<br />

regnans (from Tasmania) and E. fastigata respectively. Quarter-sawing strategies to produce slabs<br />

dimensioned only in thickness were applied to E. regnans in Victoria. The slabs were ripped to final<br />

board sizes after drying. With exception <strong>of</strong> E. regnans from Victoria, the production was conventional<br />

appearance grade products suitable for flooring or furniture, and tile battens and pallet grade boards. A<br />

dual strategy where both F17 structural products and appearance grade products was employed for the<br />

Victorian E. regnans.<br />

Figure 4. Industry graders grading dried and skip dressed boards from E. sieberi.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 165<br />

Table 1. Log sample coupe locations, year <strong>of</strong> establishment and thinning and log diameter, length and grade<br />

Species Location Year<br />

Established<br />

Year<br />

Thinned<br />

Log length<br />

(m)<br />

Log diameter<br />

(cm)<br />

Log grade<br />

E. fastigata Nalbaugh SF, NSW 1956 1994 4.9 52.2 SFNSW Quota grade<br />

E. fastigata Cathcart SF, NSW Unknown Unthinned 4.9 54.0 SFNSW Quota grade<br />

E. diversicolor Pemberton, WA 1972 1994 6.0 40.4 WA FPC Grade 1<br />

E. diversicolor Pemberton, WA 1974 Unthinned 6.0 36.9 WA FPC Grade 1<br />

E. diversicolor Pemberton, WA Unknown Unthinned 6.0 35.7 WA FPC Grade 1<br />

E. regnans Central Highlands, Vic 1971 1990 5.4 52.3 VicForests B, B/C and C<br />

E. regnans Central Highlands, Vic 1939 Unthinned 5.4 50.0 VicForests B, B/C and C<br />

E. regnans Styx Forest Block, Tas 1946 1970 5.4 / 4.9 45.7 FT Category 3<br />

E. regnans Styx Forest Block, Tas 1934 Unthinned 5.4 / 4.9 49.3 FT Category 3<br />

E. sieberi Jirrah Block, Orbost, Vic 1965 1991 4.5 49.1 VicForests B and C<br />

E. sieberi Jirrah Block, Orbost, Vic Mature & 1957 Unthinned 4.5 59.7 VicForests B and C<br />

Table 2. Log sample locations, year <strong>of</strong> thinning, processing methods and final products<br />

Species Location Established Thinned Sawing Drying Final product thickness /grade<br />

E. fastigata Nalbaugh SF, NSW 1956 1994 Quarter-sawn Predrier / kiln 19 mm / flooring and tile battens<br />

E. fastigata Cathcart SF, NSW Unknown Unthinned Quarter-sawn Predrier / kiln 19 mm / flooring and tile battens<br />

E. diversicolor Pemberton, WA 1972 1994 Back-sawn Air dry / kiln 21 and 25 mm / appearance and pallet<br />

E. diversicolor Pemberton, WA 1974 Unthinned Back-sawn Air dry / kiln 21 and 25 mm / appearance and pallet<br />

E. diversicolor Pemberton, WA Unknown Unthinned Back-sawn Air dry / kiln 21 and 25 mm / appearance and pallet<br />

E. regnans Central Highlands, Vic 1971 1990 Quarter-saw Predrier / kiln 38 mm / appearance and F17<br />

E. regnans Central Highlands, Vic 1939 Unthinned Quarter-saw Predrier / kiln 38 mm / appearance and F17<br />

E. regnans Styx Forest Block, Tas 1946 1970 Quarter-sawn Air dry / kiln 24 mm / appearance grades<br />

E. regnans Styx Forest Block, Tas 1934 Unthinned Quarter-sawn Air dry / kiln 24 mm / appearance grades<br />

E. sieberi Jirrah Block, Orbost, Vic 1965 1991 Back-sawn Predrier / kiln 19 mm / flooring and tile battens<br />

E. sieberi Jirrah Block, Orbost, Vic Mature & 1957 Unthinned Back-sawn Predrier / kiln 19 mm / flooring and tile battens


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 166<br />

From data collected during processing a number <strong>of</strong> wood quality indicators were calculated. These<br />

included the recovery <strong>of</strong> all boards after docking end-splits, the grade recovery, the percentage <strong>of</strong> the<br />

volume <strong>of</strong> standard grade and better boards


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 167<br />

Table 4. Log sample means for the calculated processing, feature and recovery and product value variates (bold = significant differences between<br />

thinned and unthinned samples at p≤0.05).<br />

Species<br />

Thinned<br />

Surface check (mm m -2 )<br />

Processing characteristics Feature characteristics Recovery and product value<br />

# internal checks : # boards<br />

Spring (mm m -1 )<br />

End split docked: board volume<br />

Knot (m m -1 )<br />

E. fastigata 1994 82 0.078 0.58 0.064 0.14 0.63 0.10 21.4* 8.8* 15.3 268<br />

E. fastigata Unthinned 73 0.021 0.63 0.053 0.17 0.80 0.30 20.8* 3.8* 16.7 208<br />

E. diversicolor 1994 15 0.000 3.34 0.056 0.02 0.35 0.35 27.6** 15.5** 23.8 170<br />

E. diversicolor Unthinned 43 0.000 2.86 0.038 0.05 0.40 0.44 27.6** 11.2** 23.4 141<br />

E. diversicolor Unthinned 51 0.000 2.19 0.025 0.03 0.57 0.59 28.2** 7.9** 32.0 126<br />

E. regnans 1990 16 0.431 0.17 0.050 0.32 0.79 0.04 27.7** 8.8** 7.1** 0.9 279<br />

E. regnans Unthinned 41 0.844 0.18 0.080 0.36 0.87 0.05 25.8** 3.7** 11.4** 2.3 234<br />

E. regnans 1970 12 0.000 0.37 0.027 0.24 0.46 0.00 28.7** 13.9** 4.2 220<br />

E. regnans Unthinned 13 0.000 0.19 0.023 0.21 0.32 0.00 30.8** 19.3** 4.0 263<br />

E. sieberi 1991 347 0.151 0.77 0.059 0.54 0.87 0.86 29.1** 1.0** 9.4 278<br />

E. sieberi Unthinned 738 0.113 0.72 0.089 0.53 0.99 0.95 31.1** 0.0** 5.1 291<br />

* Recovery based on final product size<br />

** Recovery based on nominal dry dimensions after skip dressing<br />

Kino (m m -1 )<br />

Insect damage (m m -1 )<br />

Recovery all grades (% log<br />

volume)<br />

Recovery <strong>of</strong> select grade (% log<br />

vol)<br />

Recovery <strong>of</strong> F17 (% log vol)<br />

Standard grade +


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 168<br />

Spring<br />

There were significant differences in spring for E. regnans from Tasmania and E. diversicolor. In both<br />

cases it was worse in the thinned regrowth. However, spring was generally minor and not value<br />

limiting. For example, in the case <strong>of</strong> the thinned E. regnans from Tasmania the calculated mean spring<br />

<strong>of</strong> 0.37 mm m -1 equates to approximately 1.6 mm <strong>of</strong> spring for a 5.0 m long board. Spring <strong>of</strong> this<br />

magnitude was not grade limiting. Spring was much worse in E. diversicolor than any other species,<br />

because the long log length, and relatively small diameter <strong>of</strong> the logs, accentuated the effect <strong>of</strong> growth<br />

stress release. In this case spring was reduced to acceptable levels for strip flooring during processing.<br />

End-splits<br />

The only significant difference in the volume docked to eliminate end-splits : original board volume<br />

was for E. diversicolor. The thinned sample had significantly more end-splits. The ratio can be<br />

expressed as a percentage <strong>of</strong> log volume by multiplying by 100 and then by the recovery <strong>of</strong> all grades<br />

(Table 4). In this case the ratio <strong>of</strong> 0.056:1 indicates that less than 1.6% <strong>of</strong> log volume was lost due to<br />

docked end-splits indicating that it was a minor value limiting defect.<br />

Knots, kino and insect damage<br />

There were no significant differences for knots. There were significant differences in kino in all<br />

species by region and there were significant differences in insect damage for E. fastigata and E.<br />

diversicolor. Generally kino and insect damage were less common in the thinned samples. The only<br />

case where this did not occur was for kino from the thinned E. regnans from Tasmania.<br />

Kino and insect damage were the major value limiting defects observed. These defects limited board<br />

grade to standard grade, utility grade, pallet or F17, depending on the grading strategy applied. As a<br />

consequence while overall recoveries tended to be similar for each species by region the product value<br />

per cubic metre <strong>of</strong> log input was generally higher for the thinned samples.<br />

The only case where this did not occur was with the E. regnans from Tasmania, where many boards<br />

were limited to standard grade because <strong>of</strong> the presence <strong>of</strong> kino. However, it is important to note that <strong>of</strong><br />

the 11 log samples processed in this series <strong>of</strong> trials, the incidence <strong>of</strong> kino in the thinned Tasmanian E.<br />

regnans was the 4 th lowest recorded, and in comparison to both the thinned and unthinned E. regnans<br />

samples from the comparable Victorian trials, was a much scarcer defect (0.460 m m -1 versus 0.792 m<br />

m -1 and 0.875 m m -1 for the thinned and unthinned samples respectively). This indicates that kino vein<br />

occurrence may have been low in the unthinned Tasmanian sample by the overall E. regnans forest<br />

standards.<br />

Despite the lower product values from the thinned Tasmanian E. regnans logs, there appears to be a<br />

considerable advantage from thinning in terms <strong>of</strong> sawn timber value MAI ($ ha -1 year -1 ) and pulpwood<br />

yields that more than compensates the lower sawn product values and extra harvesting costs. Estimates<br />

from harvest records in the two coupes selected for this work found that the sawn timber value MAI<br />

was 191 ($ ha -1 year -1 ) and 880 ($ ha -1 year -1 ) for the unthinned coupes and thinned coupes<br />

respectively. At the same time MAI <strong>of</strong> pulpwood production was 5.0 (m 3 ha -1 year -1 ) and 6.1 (m 3 ha -<br />

1 -1<br />

year ) from the unthinned and thinned coupes respectively.<br />

Differences in tree dominance and kino and insect damage<br />

By ensuring that the trees were selected so that the logs were as close as possible to the same diameter<br />

for the thinned and unthinned regrowth, there remains a question about the differences in tree<br />

dominance and its possible association with defect occurrence. This may account for some <strong>of</strong> the<br />

differences in kino and insect damage. However, for this evaluation, which was interested only in the<br />

impact <strong>of</strong> differences in growth rates on processing performance and product quality in conventional<br />

sawmills, this was always a compromise that had to be accepted. It would therefore be useful to obtain<br />

further information on the occurrence <strong>of</strong> these defects in both thinned and unthinned regrowth.<br />

CONCLUSION<br />

While more work is needed to understand differences in kino and insect damage, this series <strong>of</strong> trials<br />

gave a clear indication that the existing processing industry would not be disadvantaged by processing<br />

logs from thinned regrowth <strong>of</strong> conventional sawlog size and quality.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 169<br />

Faster growth after thinning did not contribute to any serious decline in wood processing performance<br />

and there is some evidence <strong>of</strong> improvements in wood drying performance, possibly due to more<br />

uniform wood properties.<br />

Given increases in sawlog yields that may be expected from thinning in even age stands these results<br />

support these thinning strategies in production forests.<br />

ACKNOWLEDGEMENTS<br />

This research was funded by the FWPA, CSIRO, VicForests, Forests NSW, Forestry Tasmania and the<br />

WA Forest Products Commission as part <strong>of</strong> the FWPA PN06.3015. Processing was conducted by Blue<br />

Ridge Hardwoods, Eden, NSW; McCormack Demby Timbers, Morwell, Victoria; ITC Southwood and<br />

Launceston, Tasmania and Auswest Timbers, Pemberton, Western <strong>Australia</strong>.<br />

REFERENCES<br />

Fagg, P. (2006). Native Forest Silviculture Guideline No. 13. Thinning <strong>of</strong> Ash Eucalypt Regrowth. Department<br />

<strong>of</strong> Sustainability and Environment Victoria.<br />

Goodwin, A. (1990) Thinning response in eucalypt regrowth. Tasforests, 2(1): 27-35.<br />

Innes, T., Armstrong, M. and Siemon, G. (2005). The impact <strong>of</strong> harvesting age/tree size on sawing, drying and<br />

solid wood properties <strong>of</strong> key regrowth eucalypt species. Project No. PN03.1316, Forest and Wood<br />

Products Research and Development Corporation, Melbourne. pp 99.<br />

Kerruish, C.M. and Rawlins, W.H.M. (Eds.) (1991) The Young Eucalypt Report: Some Management Options for<br />

<strong>Australia</strong>’s Regrowth Forests. CSIRO, East Melbourne.<br />

O’Shaughnessy, P. J. and Jayasuriya, M. D. A. (1993). (Eds) Third progress report, North Maroondah,<br />

Melbourne Water, Report Number MMBW-0017.<br />

Wardlaw T.J., Plumpton B.S., Walsh A.M. and Hickey J.E. (2004) Tasforests Volume 15 June 2004.<br />

Washusen, R., Morrow, A., Linehan, M,. Bojadzic, M. Ngo Dung and Tuan, D. (2007a). The effect <strong>of</strong> thinning<br />

on wood quality and solid wood product recovery <strong>of</strong> regrowth forests: 1. E. fastigata from southern<br />

New South Wales. Ensis Client report no. 1770 for Forests and Wood Products <strong>Australia</strong> PN06.3015.<br />

May 2007. 35 pp.<br />

Washusen, R., Morrow, A., Siemon, G. and Ngo Dung (2007b). The effect <strong>of</strong> thinning on wood quality and solid<br />

wood product recovery <strong>of</strong> regrowth forests: 2. E. diversicolor from southwest Western <strong>Australia</strong>. Ensis<br />

Client report no. 1821 for Forests and Wood Products <strong>Australia</strong> PN06.3015.<br />

Washusen, R., Morrow, A., Ryan, M. and D. Ngo (2007c). The effect <strong>of</strong> thinning on wood quality and solid<br />

wood product recovery <strong>of</strong> regrowth forests: 3. E. regnans from Central Highlands Victoria. Ensis Client<br />

report no. 1828 for Forests and Wood Products <strong>Australia</strong> PN06.3015.<br />

Washusen, R., Morrow, A., Wardlaw, T. and D. Ngo (2007d). The effect <strong>of</strong> thinning on wood quality and solid<br />

wood product recovery <strong>of</strong> regrowth forests: 4. E. regnans from Southern Tasmania. Ensis Client report<br />

no. 1827 for Forests and Wood Products <strong>Australia</strong> PN06.3015.<br />

Washusen, R., Morrow, A., Ryan, M. and D. Ngo (2007e). The effect <strong>of</strong> thinning on wood quality and solid<br />

wood product recovery <strong>of</strong> regrowth forests: 5. E. sieberi from Eastern Victoria. Ensis Client report no.<br />

1826 for Forests and Wood Products <strong>Australia</strong> PN06.3015.<br />

Webb, A. W. F. (1966). The effect <strong>of</strong> thinning on the growth <strong>of</strong> Mountain Ash in Victoria. MSc thesis, Forest<br />

Commission, Victoria.<br />

West, P.W. (1991). Thinning response and growth modelling. In Kerruish C. M. and Rawlins W. H. M.editors<br />

(1991) The young eucalypt report - some management options for <strong>Australia</strong>'s regrowth forests. CSIRO<br />

<strong>Australia</strong>, pp 28-49.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 170<br />

DIESEL-ELECTRIC HYBRID DRIVE TECHNOLOGY FOR<br />

REDUCED FUEL CONSUMPTION AND CARBON EMISSIONS<br />

IN FOREST OPERATIONS<br />

Mark Brown 1<br />

ABSTRACT<br />

The forest industry is well placed to play a positive role in climate change through the<br />

management <strong>of</strong> forests as carbon sinks. However, forest operations generate significant<br />

carbon emissions. Therefore forest companies have the opportunity to play a positive role<br />

in climate change by reducing their carbon emissions by implementing environmentally<br />

friendly technology and work methods.<br />

The CRC for Forestry has partnered with a developer <strong>of</strong> retr<strong>of</strong>it hybrid-drive technology,<br />

Innovative Transport Solutions Pty Ltd (ITS Electric), to develop and test diesel-electric<br />

hybrid conversions for forestry applications. The overall aim is a 10-25% reduction in<br />

fuel consumption for a reduction in operating costs and carbon emissions. The first trial<br />

vehicle, a 6X6 truck used to move chip trailers around forest harvesting sites, will be<br />

tested under field conditions in Western <strong>Australia</strong> during the last half <strong>of</strong> 2009.<br />

This paper discusses why hybrid technology is believed to have significant potential for<br />

reducing fuel use and carbon emissions for some areas <strong>of</strong> forest operations; provides an<br />

overview <strong>of</strong> the unique conversion approach to create hybrid drive vehicles; and<br />

describes the field trial that will be conducted over the next two years.<br />

INTRODUCTION<br />

Forest operations are energy intensive with harvest and haulage methods being responsible for the<br />

bulk <strong>of</strong> the energy consumed. Commonly used forest harvesting equipment in <strong>Australia</strong> consumes up<br />

to 40 L <strong>of</strong> fuel per hour and the typical haulage trucks use up to 80 L/100 km. The use <strong>of</strong> such a large<br />

amount <strong>of</strong> fuel in forest operations is <strong>of</strong> concern with its increasing cost, but it also has important<br />

carbon emission implications. In Sweden, where harvesting systems are among the most fuel efficient<br />

in the world with fuel consumption between 10 and 25 L/hr, studies have shown CO2 emissions <strong>of</strong><br />

about 5.9 t per m 3 <strong>of</strong> wood harvested and over 6.7 t per m 3 for haulage (Berg & Lindholm, 2005).<br />

In an effort to further advance the positive role <strong>of</strong> forestry in the climate change issue the industry is<br />

actively exploring options to improve the fuel efficiency <strong>of</strong> their harvesting and haulage operations.<br />

This interest is highlighted by handbooks <strong>of</strong> best practice, such as ‘In Forest Operations Fuel<br />

Economy Counts’ from Canada, which focus entirely on what can be done both by equipment<br />

selection and by adjusting work techniques to reduce the use <strong>of</strong> fuel. While these handbooks are<br />

popular and successful in achieving outcomes in Canada, it is a fact that up to 80% <strong>of</strong> the fuel use is<br />

influenced by machine and engine design. This highlights the need for a continued drive for more<br />

efficient machine and engine designs.<br />

HYBRID TECHNOLOGY<br />

Hybrid technology is the use <strong>of</strong> more than one power source in a vehicle’s drive or power system that<br />

allows for the optimisation <strong>of</strong> the two power sources over the work cycle for minimum energy use.<br />

Many examples <strong>of</strong> hybrid technologies have the added advantage that one or more <strong>of</strong> the power source<br />

technologies have the capability <strong>of</strong> recovering and storing energy when not providing power, either<br />

through the normal function <strong>of</strong> the operations (moving down hill, braking, etc.) or by recovering<br />

excess energy output. Without the energy recovery aspect <strong>of</strong> hybrid technology, optimisation between<br />

power sources can only <strong>of</strong>fer very limited efficiency gains and in many cases none at all. This is due<br />

1 University <strong>of</strong> Melbourne, CRC Forestry, Harvesting and Operations. Email: mwbrown@unimelb.edu.au


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 171<br />

to the difficulty <strong>of</strong> effectively optimising the power sources over complex work cycles (Lin et.al.,<br />

2003). As a result the application <strong>of</strong> hybrid technologies which <strong>of</strong>fer the greatest efficiency gains are<br />

those used in work cycles that have frequent and large variations in power load. In the transport<br />

industry these are operations with frequent starts and stops, as highlighted by Langer (2004) where<br />

medium duty trucks are projected to be able to increase fuel economy by 71% due to the start and stop<br />

nature <strong>of</strong> the work.<br />

The two most common hybrid technology combinations are an internal combustion engine (ICE) with<br />

a battery powered electric motor, and an ICE with an accumulator powered by a hydraulic motor. The<br />

characteristics <strong>of</strong> the hydraulic hybrid system are a limited capacity to store energy, and best operation<br />

with a quick release <strong>of</strong> the stored energy. Due to these characteristics, the hydraulic hybrid system is<br />

only practical in a limited number <strong>of</strong> applications where the power input from the hydraulic power<br />

system is relatively short and regeneration <strong>of</strong> the hydraulic power is sufficient between power input<br />

demands. Electric hybrid technology tends to be more versatile, with an increased capacity to respond<br />

to more frequent and/or sustained power demands depending on the size and type <strong>of</strong> battery used. In<br />

part, this increased flexibility has led to this project’s exploration <strong>of</strong> electric hybrid technology.<br />

The development <strong>of</strong> electric hybrid technology has largely focussed on integrated drive systems where<br />

the electric drive and regenerative system is integrated with the transmission system. This integrated<br />

approach <strong>of</strong>fers the highest level <strong>of</strong> control to return the highest possible gains in efficiency from the<br />

hybridisation. However, this integrated approach also means that retr<strong>of</strong>itting the technology to existing<br />

ICE is either impossible due to a lack <strong>of</strong> space and/or cost prohibitive. Therefore the introduction <strong>of</strong><br />

the electric hybrid technology would be limited to purchasing new equipment and this would involve a<br />

relatively high cost. With forestry vehicles having a potential working life up to 15 years and<br />

companies trying to minimise purchase costs, the realisation <strong>of</strong> the potential efficiency gains would<br />

therefore be slow.<br />

The <strong>Australia</strong>n company, Innovative Transport Solutions Pty. Ltd. (ITS Electric), has developed an<br />

electric hybrid approach which can be applied to many ICE applications as a conversion. Like other<br />

electric hybrid technology, the ITS approach adds an electric motor to the ICE to maximise the<br />

efficiency <strong>of</strong> both and uses the electrical system and battery to recover energy for further efficiency<br />

gains. The main difference in this case is that the electric motor is introduced directly at the axle<br />

without interfering with the traditional drive system and the hybridization is controlled through an<br />

independent external controller.<br />

This approach means that while there are limitations in the level <strong>of</strong> integration which can be achieved<br />

with the ICE drive system, it also means most ICE drive systems can be retr<strong>of</strong>itted in a cost effective<br />

way, with projected payment periods for many truck applications <strong>of</strong> less than three years.<br />

POTENTIAL APPLICATION OF THE HYBRID CONVERSION TECHNOLOGY<br />

IN <strong>FORESTRY</strong> OPERATIONS<br />

As stated earlier, hybrid technology is most efficient in operations with frequent and large variations in<br />

power demand. In transport operations, some <strong>of</strong> the best results with hybrid technology have occurred<br />

where there are a large number <strong>of</strong> starts and stops in a normal work cycle, such as urban garbage<br />

collection, local delivery trucks and city busses.<br />

Blohm and Anderson 2004 looked at the potential <strong>of</strong> hybridisation in garbage trucks in the United<br />

Kingdom. Through GPS tracking they identified that, in an average day’s work, the trucks stopped 459<br />

times and spent over 40% <strong>of</strong> the time idling. They identified fuel economy benefits <strong>of</strong> 30-40%<br />

depending on the type <strong>of</strong> hybridization used. A US company, United Parcel Service (UPS), which<br />

operates a large fleet <strong>of</strong> delivery vehicles, announced in a 2007 press release that they would continue<br />

their the expansion <strong>of</strong> hybrid vehicles in their fleet with an additional 50 hybrid trucks, and expected<br />

to save about 166,500 litres <strong>of</strong> fuel per year, based on their experience since 2001 with hybrid<br />

vehicles.<br />

Much <strong>of</strong> the equipment used in forest operations has frequent and large variations in power demand,<br />

including skidders, forwarders, shunt trucks, chippers, haul trucks, feller-bunchers and harvesters. All<br />

<strong>of</strong> this equipment may benefit from an adapted hybrid system.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 172<br />

When the work cycle <strong>of</strong> forestry primary transport equipment is examined most tend to cover short<br />

distances, frequently start and stop, and <strong>of</strong>ten have large amounts <strong>of</strong> idle time or low power demands. .<br />

A series <strong>of</strong> short operational trials carried out by the CRC for Forestry in Western <strong>Australia</strong> examined<br />

the basic productivity <strong>of</strong> harvesting equipment. Skidders working in plantation clearfelling operations<br />

have a minimum <strong>of</strong> 100 to 140 start/stops in a ten hour shift and spend only about 40% <strong>of</strong> the time<br />

under heavy power demand, travelling loaded. Similarly, forwarders have a minimum <strong>of</strong> 120 to 200<br />

start/stops in a ten hour shift and spend less than 25% <strong>of</strong> the time under heavy power demand<br />

travelling loaded.<br />

A desktop analysis <strong>of</strong> the work cycle <strong>of</strong> shunt trucks as it relates to chipper operations indicates a<br />

minimum <strong>of</strong> 90 to 140 start/stops in a ten hour shift, over 50% <strong>of</strong> the time idling, and only about 20%<br />

<strong>of</strong> the time pulling a full load that has a large power demand.<br />

Fuel savings <strong>of</strong> 10-25% are achievable for these three primary transport functions. For a skidder this<br />

represents up to 15,000 litres <strong>of</strong> fuel saved per year, and about half as much for the forwarder and for<br />

the shunt truck, since a forwarder is considerably more efficient than a skidder, while a shunt truck<br />

typically has a much smaller work load.<br />

In the longer term, or with more aggressive hybrid conversions, even greater savings are possible by<br />

using a smaller ICE in combination with the electric motor, as the electric system could support the<br />

limited amount <strong>of</strong> heavy power demand.<br />

Beyond these primary transport applications, a more detailed analysis <strong>of</strong> work cycles is required to<br />

explore the potential for other forestry machines and such equipment is not addressed in this report.<br />

With such applications, while there would be a limited start/stop type function for energy recovery,<br />

there would be an opportunity to recover energy at times <strong>of</strong> low demand in order to power an electric<br />

motor to support times <strong>of</strong> high demand.<br />

THE CRC <strong>FORESTRY</strong> INDUSTRY TRIAL<br />

After extensive discussion with ITS Electric and the forestry industry, an initial trial has been planned<br />

with shunt trucks. Significant effort had already been put into designs to fit class 8 trucks with hybrid<br />

technology that would directly apply to shunt trucks, since shunt trucks are a version <strong>of</strong> a class 8 truck.<br />

Being able to directly apply this existing knowledge and design to the shunt truck application will<br />

minimize risk and cost.<br />

From the forest industry perspective, while shunt trucks have limited application, the 6X6 trucks<br />

required are difficult to get affordably on the used truck market, where they are typically sourced, due<br />

to competition from the mining industry. As a result the approach <strong>of</strong> purchasing a more readily<br />

available used class 8-truck and converting it to 6X6 with the electric hybrid conversion technology<br />

can be done for about the same cost as the traditionally used 6X6 trucks, thereby reducing any<br />

associated risks and costs.<br />

From an operational trial perspective, the use <strong>of</strong> the shunt truck as the “pro<strong>of</strong> <strong>of</strong> concept” for this fuel<br />

saving technology is very attractive since the work cycle <strong>of</strong> the shunt truck is highly repetitive, the<br />

operating conditions are relatively easy to control and there is frequent operator change over.<br />

Considering:<br />

o the high level <strong>of</strong> repetition,<br />

o the opportunity to control certain operating conditions, and<br />

o the ability to eliminate or at least identify operator influences,<br />

the project will be able to evaluate the performance <strong>of</strong> the shunt truck effectively and isolate, with<br />

reasonable confidence, the influence <strong>of</strong> the hybrid technology without the high cost <strong>of</strong> highly<br />

controlled bench tests that are <strong>of</strong>ten required to measure fuel impacts.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 173<br />

As shunt trucks have been identified as a good approach for all three collaborators, the trial design has<br />

four main objectives:<br />

1. To determine if the hybrid conversion concept is practical and functional.<br />

2. To test the functional capabilities <strong>of</strong> the converted hybrid 6X6 shunt truck against<br />

currently acceptable 6X6 diesel trucks to ensure it can perform the task.<br />

3. To quantify the fuel efficiency gains and emission reduction potential <strong>of</strong> the technology in<br />

a 6X6 shunt truck application.<br />

4. To evaluate the short to medium term reliability and costs <strong>of</strong> the hybrid conversion<br />

technology.<br />

The evaluation <strong>of</strong> the hybrid conversion technology will be done in four distinct phases over a 24 to 30<br />

month period <strong>of</strong> time.<br />

o Phase one will look at the costs <strong>of</strong> the vehicle and its setup as the basis for economic<br />

comparisons.<br />

o Phase two will include some specific comparative functional tests on turning and pulling<br />

capabilities to ensure the hybrid truck is able to complete the required operational tasks<br />

o Phase three will be a detailed evaluation <strong>of</strong> operating cost, performance and reliability on a<br />

short to medium term to complete the economic evaluation as well as environmental impacts.<br />

o Phase four will be a medium to long term evaluation <strong>of</strong> operating costs and reliability to<br />

better understand maintenance costs.<br />

To evaluate the comparative initial costs <strong>of</strong> the hybrid conversion against the traditional 6X6 truck in<br />

Phase one, all components and their cost for the conversion to hybrid will be recorded, including<br />

labour and customization by the forest company operator for safety and ease <strong>of</strong> operation.<br />

Considering that prototypes tend to cost as much as three times that <strong>of</strong> production models and, if the<br />

technology is successful, on a large scale there will be cost benefits to large volume production <strong>of</strong><br />

certain components, estimates <strong>of</strong> future costs with sensitivity analysis will be done. All these initial<br />

and preparation costs will be compared against the cost <strong>of</strong> the last three to six 6x6 trucks acquired,<br />

based on availability <strong>of</strong> information from the cooperating forest company, including all preparation<br />

costs. This comparison will look at both simple purchase/setup costs and annual distribution <strong>of</strong> these<br />

costs based on projected lives for the two different trucks.<br />

With Phase two, functional tests to ensure the truck will be able to perform the required operational<br />

tasks without significant risk <strong>of</strong> failure or decreased safety will be carried out. The turning radius will<br />

be checked with controlled test methods in good (flat, dry asphalt), moderate (flat, loose gravel) and<br />

poor (sloped, very loose gravel) conditions. Pulling strength as compared to the traditional 6X6 truck<br />

will be examined with a controlled acceleration test where the distance and time to attain a defined<br />

operating speed from a dead stop will be compared under good (flat, dry asphalt), moderate (flat, loose<br />

gravel) and poor (sloped, very loose gravel) conditions with a loaded trailer. For both sets <strong>of</strong><br />

controlled tests in phase two, each test condition will be repeated to ensure three results are obtained<br />

that are within +/- 5% <strong>of</strong> each other (to ensure the results are true and repeatable).<br />

During Phase three a detailed, side-by-side operational tracking <strong>of</strong> the hybrid truck as compared to the<br />

traditional 6X6 truck will be performed for six months. With the support <strong>of</strong> onboard tracking<br />

technology, shift level manual data records and detailed observation studies data will be collected on:<br />

• equipment availability,<br />

• utilization and effective time<br />

• ICE use and performance (idle time RPM pr<strong>of</strong>ile, travel speeds, fuel consumption, etc.)<br />

• hybrid technology use and performance (use <strong>of</strong> electric motor, battery<br />

levels/charge time, etc.)<br />

• work or duty cycle <strong>of</strong> the truck


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 174<br />

• short and long operational delays and root cause <strong>of</strong> the delays<br />

• fuel use<br />

• scheduled and unscheduled repair and maintenance<br />

• harvesting system results that influenced shunt truck use/performance<br />

• general operating conditions and operational productivity<br />

The data will be analysed to quantify the performance <strong>of</strong> the hybrid shunt truck relative to the<br />

traditional 6X6 trucks and clearly identify any impacts on operating cost, operating performance and<br />

fuel use. Particular emphasis will be given to identifying conditions where the hybrid demonstrates<br />

significant strengths or weaknesses.<br />

Phase four <strong>of</strong> the project will see the continued collection <strong>of</strong> similar data as for phase three, but at a<br />

lower data collection intensity for an additional 18 months, and will focus on longer term issues with<br />

performance and operating costs, such as maintenance and delays.<br />

CONCLUSION<br />

While the forestry industry has a positive role to play in climate change through carbon sequestration<br />

in the growth and sustainable management <strong>of</strong> forests, forest operations represent an opportunity for<br />

further improvement by reducing fuel use and emissions through the appropriate application <strong>of</strong> new<br />

technology.<br />

One such opportunity is the use <strong>of</strong> hybrid drive technology, as a large amount <strong>of</strong> the equipment used<br />

in forest operations have work cycles that match the conditions where hybrid technology has been<br />

most successful — a large amount <strong>of</strong> start and stop activity, with large periods <strong>of</strong> work time at idle or<br />

low power demand levels.<br />

With the more common integrated hybrid technology, its introduction would be limited to new<br />

equipment and even then, within the price sensitive forest industry, this would be slow, based on the<br />

current mechanical life <strong>of</strong> forest equipment and replacement cycles. As a result, the new “Made in<br />

<strong>Australia</strong>” conversion approach is an attractive alternative which would allow the introduction <strong>of</strong> the<br />

technology without equipment replacement and with relatively short payback periods.<br />

The CRC for Forestry will work with the technology developer and the <strong>Australia</strong>n forest industry to<br />

adapt the hybrid conversion technology for shunt trucks in the forest industry. The project will seek to<br />

verify the base performance and any impact the technology has on the costs, efficiency and<br />

performance <strong>of</strong> the shunt trucks. If successful, the hybrid shunt truck will provide a cost effective,<br />

environmentally friendly technology option for the industry. In addition, the hybrid shunt truck will<br />

provide a good base for investment into hybrid conversion technology, and pave the way for its<br />

introduction to other forestry-related machinery, or the conversion <strong>of</strong> existing equipment.<br />

REFERENCES<br />

Berg S. and Lindholm E-L., 2005, ‘Energy use and environmental impacts <strong>of</strong> forest operations in Sweden’,<br />

Journal <strong>of</strong> Cleaner Production vol. 13, pp. 33-42.<br />

Blohm T. and Anderson S., 2004, ‘Hybrid Refuse Truck Study’ paper presented to Product Development<br />

Conference, Huntington Beach, California.<br />

Forest Innovation Partnership, In Forest Operations Fuel Economy Counts, Sainte-Foy, Quebec, Canada.<br />

Langer T., 2004, ‘Energy Savings Through Increased Fuel Economy For Heavy-Duty Trucks’, prepared for the<br />

National Commission on Energy Policy.<br />

Lin C-C., Peng H., Grizzle J.W. & Kang J-M., 2003, ‘Power Management Strategy for a Parallel Hybrid Electric<br />

Truck’, IEEE Transactions on Control Systems Technology, vol. 11 no. 6, pp. 839-849.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 175<br />

MODELLING MILL DOOR LOG PRICES FOR<br />

PLANTATION EUCALYPTS WITH CSIROMILL<br />

Russell Washusen 1<br />

ABSTRACT<br />

CSIROMILL incorporates financial models <strong>of</strong> sawmills, developed from processing<br />

experiments with the assistance <strong>of</strong> industry, to assist establishment <strong>of</strong> new plantation<br />

eucalypt resources. During a recent international project supported by the <strong>Australia</strong>n<br />

Centre for International Agricultural Research (ACIAR), one module <strong>of</strong> CSIROMILL<br />

was used to model mill door prices <strong>of</strong> unpruned 9-year-old E. pilularis, processed with a<br />

HewSaw R200 PLUS linear sawmill. The results indicated that it would not be viable to<br />

grow the unpruned E. pilularis, despite the efficiencies <strong>of</strong> the linear flow sawmill. Further<br />

work was conducted with two additional modules <strong>of</strong> CSIROMILL using pruned 17-22<br />

year-old Corymbia spp. (spotted gum), processed with a HewSaw SL250 Trio linear<br />

sawmill; and, pruned 17-22 year-old spotted gum, processed with a conventional<br />

reciprocating flow eucalypt sawmill. The mill door prices generated demonstrate the<br />

importance <strong>of</strong> producing high quality logs. The modelling also gave an advantage in mill<br />

door log values <strong>of</strong> $90-120 per cubic metre with the linear system compared to the<br />

reciprocating system. The results demonstrate how CSIROMILL can be applied, and the<br />

importance <strong>of</strong> both log quality and sawmill efficiency in developing viable plantations for<br />

the production <strong>of</strong> sawlogs.<br />

INTRODUCTION<br />

Recent experiments with plantation-grown E. nitens, E. pilularis and Corymbia citriodora subsp.<br />

variegata indicated that the HewSaw R200 and R250 sawmills can be applied to saw small-diameter<br />

eucalypt logs (Washusen et al. 2007, 2008, 2009). The sawing technology incorporated into these<br />

sawmills is usually associated with the s<strong>of</strong>twood industry where, unlike in eucalypts, longitudinal<br />

peripheral growth stresses pose few problems during processing.<br />

In each <strong>of</strong> the experiments it was found that, with the application <strong>of</strong> appropriate cutting patterns, the<br />

technology appears to be capable <strong>of</strong> processing small diameter eucalypts with a large range <strong>of</strong> growth<br />

stress levels. The main characteristic associated with growth stress release that may hinder processing<br />

efficiency was bow in boards cut from near the log periphery. However, the magnitude <strong>of</strong> bow was<br />

considered small enough to allow efficient transport <strong>of</strong> boards in the green mill if minor modifications<br />

were made to board conveyors.<br />

In comparison to conventional eucalypt processing, the HewSaw incorporates technology that<br />

produces fundamental differences in processing performance for eucalypt saw logs. These differences<br />

are:<br />

(i) Chippers are applied with small diameter circular saws during sawing in such a way that<br />

wood near the log periphery (assumed to be the most highly stressed wood) is removed<br />

and boards are pr<strong>of</strong>iled before or simultaneously with sawing. The removal <strong>of</strong> chip from<br />

near the log periphery eliminates the need to handle distorted ‘roundbacks’ in the mill, it<br />

tends to reduce board deflection (bow), and board and log end-splitting appears to be<br />

controlled if sufficient material is chipped from the log periphery (Washusen et al 2007,<br />

2008, 2009).<br />

(ii) End dogging <strong>of</strong> logs is eliminated which can reduce log end-splitting induced during the<br />

sawing process (Washusen and Innes 2007).<br />

1 Dr Russell Washusen, Principal Research Scientist, CSIRO Materials Science and Engineering, Project Leader 2.3, CRC<br />

for Forestry, Bayview Av, Clayton South, email: russell.washusen@csiro.au


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 176<br />

(iii) The adoption <strong>of</strong> multi-saw (and chipper) technology allows linear flow <strong>of</strong> material. In<br />

contrast, conventional eucalypt sawmills, even those where twin saw systems are<br />

employed, require the saws to be moved repeatedly through the logs and flitches, or the<br />

logs and flitches to be passed back and forwards through a stationary saw (reciprocating<br />

flow). Either <strong>of</strong> these methods produces a relatively slow throughput <strong>of</strong> logs and flitches,<br />

and bottle necks in the processing line <strong>of</strong>ten occur.<br />

The potential for improved material flow, while maintaining or improving board behaviour, allows<br />

high log throughput rates to be achieved. For example, the most modern dedicated conventional<br />

reciprocating flow sawmills built in <strong>Australia</strong> have log throughput rates <strong>of</strong> around 25,000-45,000 m 3<br />

<strong>of</strong> log annum -1 in a single shift (pers. comm. Ge<strong>of</strong>f Bertolini, Whittakers Timber Products, Western<br />

<strong>Australia</strong>; Ian Whiteroad, Integrated Tree Cropping, Tasmania).<br />

In comparison, single shift annual log throughput rates for sawing s<strong>of</strong>twood with the HewSaw R250<br />

and HewSaw SL250 Trio PLUS are around 130,000 m 3 <strong>of</strong> log annum -1 and 200,000 m 3 <strong>of</strong> log annum -1<br />

respectively (pers. comm. John Marshall, Carter Holt Harvey, Victoria; Kennett Westermark, Veisto<br />

Oy, Finland). In comparison to conventional reciprocating systems, this improved material flow has<br />

the potential to reduce processing costs for eucalypts, with flow-on effects <strong>of</strong> improved returns to both<br />

processors and growers.<br />

In order to quantify the impact <strong>of</strong> improved material flow a HewSaw R200 PLUS system was<br />

modelled to predict mill door prices for 9-year-old unpruned E. pilularis with the CSIROMILL H 200<br />

PLUS 12000 AD BB module (Washusen et al. 2008). CSIROMILL is a modelling tool developed by<br />

CSIRO in partnership with industry to assist planning <strong>of</strong> new plantation-grown eucalypt resources.<br />

The primary function <strong>of</strong> CSIROMILL is to simultaneously model sawmill pr<strong>of</strong>itability and mill door<br />

prices using information collected from ‘real life’ experimental processing trials.<br />

The modelling <strong>of</strong> the unpruned E. pilularis indicated that processing could be pr<strong>of</strong>itable, although mill<br />

door prices were too low for pr<strong>of</strong>itable growing because <strong>of</strong> poor log quality. As a result <strong>of</strong> this work<br />

further modelling was undertaken with high quality sawlogs to further assess the potential mill door<br />

prices for eucalypts processed with linear sawing systems. For the purpose <strong>of</strong> comparing outcomes<br />

with conventional processing, comparisons were made with a modern reciprocating flow sawing<br />

system currently used to process small diameter native forest eucalypts.<br />

For this work, two additional modules <strong>of</strong> CSIROMILL were employed. CSIROMILL also has a data<br />

base <strong>of</strong> information obtained from numerous trials in sawmills across <strong>Australia</strong>. The data base has<br />

recorded log quality information for pruned eucalypts that can be readily produced by growers, and<br />

measured product recoveries and values matched to logs <strong>of</strong> a specified quality. This information was<br />

used to model mill door prices for pruned logs so that comparisons could be made between the<br />

unpruned E. pilularis and linear and reciprocating sawing systems processing larger diameter, pruned<br />

eucalypts.<br />

LOG QUALITY, RECOVERIES AND PRODUCT VALUES USED FOR MODELLING<br />

For modelling the following information was used. The pruned Corymbia spp (spotted gum) log Grade<br />

2 adopted by Washusen (2006) was used. The log specifications are given in Table 1.<br />

Table 2 gives the recoveries <strong>of</strong> wood chip (percent <strong>of</strong> log volume) and sawn product based on nominal<br />

dry dimensions (percent <strong>of</strong> log volume) for the pruned log Grade 2 for spotted gum.<br />

These values were obtained from trials conducted at the Auswest Timbers Pemberton sawmill<br />

(Washusen 2006). Grade recoveries met the requirements <strong>of</strong> the <strong>Australia</strong>n Standard AS 2796.1<br />

(Timber-Hardwood-Sawn and milled products, Part 1: Product specification). The synonym for<br />

standard grade is medium feature grade in AS 2796.1. The logs were sourced from Boyup Brook and<br />

Vasse in southwest Western <strong>Australia</strong> and ranged in age from 17-22 years.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 177<br />

Table 1. Specifications for pruned spotted gum log Grade 2.<br />

Log characteristic Specifications<br />

Diameter <strong>of</strong> defect core < 15.0 cm<br />

Log mean diameter 30.1-35.0 cm<br />

Sweep in 2.4 m length*


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 178<br />

Figure 1. Example sawing pattern for a 35 cm small end diameter log showing the sequence <strong>of</strong><br />

cuts on the twin saw and resaw cuts.<br />

All wood prices used were AUD mill-door-wholesale prices free <strong>of</strong> GST.<br />

Prices for sawn boards were adapted from Washusen (2006) using the values <strong>of</strong> the actual board sizes<br />

recovered. Victorian ash prices were used to value sawn wood with a 50% discount for select and<br />

standard grade boards shorter than 1.8 m and a 10% discount for all other boards


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 179<br />

THE CSIROMILL METHOD AND MODULES USED<br />

The CSIROMILL system and the methodology used for modelling are described by Washusen et al.<br />

(2008) and readers are directed to that report for background information.<br />

Two CSIROMILL modules were used for modelling (Table 4). One was a WH module and the other a<br />

HSL 250 module. These are described below.<br />

Table 4. CSIROMILL Modules and the processing systems they represent.<br />

Module System Annual log intake range<br />

(m 3 )<br />

WH 30000 AD SG Reciprocating system with twin and<br />

multi-saws<br />

40,000-60,000<br />

SL 250 150000 AD SG Linear system with three<br />

chipping/sawing units in line<br />

250,000-350,000<br />

The WH Modules <strong>of</strong> CSIROMILL<br />

The WH modules <strong>of</strong> CSIROMILL are reciprocating flow systems representing modern twin-saw<br />

technology coupled with a downstream multisaw resaw and trim-saw systems. The reciprocation <strong>of</strong> a<br />

35 cm SED log during part <strong>of</strong> the sawing process through the actual twin saw represented in the<br />

module is shown in Figure 2.<br />

The WH modules are based on the Whittakers Timber Products small log line in Western <strong>Australia</strong><br />

with some modifications. This is the newest hardwood mill in <strong>Australia</strong> having commenced<br />

production in 2006. The modules reflect actual construction costs <strong>of</strong> the mill (for most components).<br />

The operating costs have been sourced from a number <strong>of</strong> hardwood mills located in Western <strong>Australia</strong>,<br />

Victoria and Tasmania.<br />

Capital, operating costs, staffing levels and summary sheets <strong>of</strong> the CSIROMILL WH 30000 AD SG<br />

module are given in the output files in Washusen (2008).<br />

For the WH 30000 module, logs are cut to length either in the forests or the log yard. They are<br />

debarked on the sawmill in-feed, sawn with the twin bandsaw and multisaw resaw coupled with a<br />

trim-saw system. The twin saw is controlled by a computer. The logs are scanned and the computer<br />

determines the best sawing pattern and the sawing is then automatic. The overhead end-dogging<br />

system is equipped with a ‘turn-down’ device eliminating the need to release the log. The debarking<br />

and sawing system has capacity to process logs from 20 cm SED to 40 cm SED.<br />

Figure 2. Pruned eucalypt log being processed on the twin saw at the Whitakers Timber Product<br />

sawmill in Western <strong>Australia</strong>, showing the reciprocation <strong>of</strong> the log and log rotation to<br />

produce the log break-down pattern shown in Figure 1.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 180<br />

The log throughput range is 20,000-30,000 m 3 annum -1 for a single shift. Based on trials conducted in<br />

the Whittakers Timber Products mill, the actual capacity is approximately 25,000 m 3 annum -1 in a<br />

single shift (Washusen unpublished data). The sawing and material handling systems appear to have<br />

greater capacity than this, but log throughput is highly dependant on staff capability. The throughput<br />

rates tested in the analysis are 45,000-55,000 m 3 annum -1 for a two shift operation.<br />

Sawn residues are chipped and transported by conveyors to a stock pile or for loading directly into<br />

trucks. Bark and 50% <strong>of</strong> the sawdust is transported to bins and sold to local markets. The remaining<br />

sawdust is used green as fuel to heat kilns.<br />

Immediately after sawing, boards containing sapwood are segregated and treated with a standard<br />

boron diffusion treatment.<br />

Drying incorporates air-drying in sheds with drop down sides to control air flow. Following air-drying<br />

the wood is reconditioned in steam then kiln dried to a final moisture content <strong>of</strong> 10-12%. Heating for<br />

kilns is via pressurised hot water, heated from a furnace with sawdust and dry shavings being used as<br />

fuel.<br />

The analysis is applicable to the production <strong>of</strong> oversized skip dressed boards that are sold for further<br />

manufacturing into flooring. Each analysis assumes that the entire intake is <strong>of</strong> the logs specified. In<br />

reality there would be a range <strong>of</strong> log sizes and possibly more than one species.<br />

The SL 250 Modules <strong>of</strong> CSIROMILL<br />

The SL 250 modules are based on the HewSaw SL250 Trio sawing line. Figure 3 shows the linear<br />

configuration <strong>of</strong> the separate machines that make up the sawing line. Figure 4 shows an example<br />

sawing pattern produced during the sawing process. The sawing strategies are different to those<br />

produced by the HewSaw R250 and R200 in that the sawing is not completed within the one machine.<br />

However, the rip saw produces a similar cutting pattern to the R250, and the chipping prior to<br />

application <strong>of</strong> these saws produces the same stress release.<br />

The separation <strong>of</strong> the machines in the HewSaw SL250 sawing line has several advantages over the<br />

HewSaw R250 and R200. It allows the cant to be turned so that the sawing produces back-sawn<br />

boards. The cant saw recovers either 1+1 or 2+2 side boards that are not recovered by the HewSaw<br />

R250 and R200. This results in higher recovery in larger diameter logs and in pruned logs these boards<br />

would be produced from the clearwood zone. The minimum log length that can be sawn on the SL250<br />

Trio is also 2.4 m compared to 3.0 m for the HewSaw R250. This gives greater capacity to reduce<br />

board deflection by reducing log length, and to reduce the impact <strong>of</strong> log taper on recovery. Because <strong>of</strong><br />

these features, in comparison to the HewSaw R250 and R200, the SL 250 Trio is better suited to<br />

processing pruned eucalypts.<br />

The SL 250 modules are based on a number <strong>of</strong> mills in Finland, as no mills <strong>of</strong> this type exist in<br />

<strong>Australia</strong>. Two examples <strong>of</strong> Finnish mills are shown in Figure 5. One <strong>of</strong> the features <strong>of</strong> the modules is<br />

the incorporation <strong>of</strong> tray sorters on the outfeed from the saw for the four or eight outer boards (Figure<br />

5) and bins for the centre boards. Tray sorters can be set up so they pick up boards from specific parts<br />

<strong>of</strong> the sawing pattern. As these boards are the same dimension for logs <strong>of</strong> similar diameter, sorting <strong>of</strong><br />

boards is not required. Transport distances can also be reduced and some elevators eliminated. All <strong>of</strong><br />

these features can be potentially beneficial to transporting and sorting hardwood boards that may be<br />

unstable to transport on conventional s<strong>of</strong>twood conveying systems (pers. comm. Kennett Westermark,<br />

Viesto Oy, Finland).<br />

As with the WH modules, actual capital and construction costs applicable to <strong>Australia</strong> have been used<br />

using a conversion rate <strong>of</strong> € 1 = AUD 1.69, where capital costs were supplied to CSIRO in Euros. The<br />

operating costs have been sourced from a number <strong>of</strong> mills located in Victoria and Tasmania, with the<br />

maximum cost used for calculations.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 181<br />

Figure 3. The configuration <strong>of</strong> the HewSaw SL250 Trio line (adapted from www.hewsaw.com).<br />

Capital, operating costs, staffing levels and summary sheets <strong>of</strong> the CSIROMILL SL 250 150000 AD<br />

SG module used for modelling are given in the output files in Washusen (2008).<br />

The module represents a mill where logs are scanned and merchandised in the log yard and debarked<br />

on the sawmill in-feed. The sawing system incorporates a HewSaw SL 250 PLUS trio sawing line<br />

with scan and set capability and a ring reducer immediately after the debarker. The debarking and<br />

sawing system can process logs from 2.4 to 6.0 m in length and from 10 cm SED to 40 cm SED and<br />

50 cm LED. Line speed is 60 to 150 m min -1 . Side boards are handled with a tray sorter and stacker,<br />

and there is a separate stacker for centre boards.<br />

Figure 4. Pattern produced by the HewSaw SL250 Trio line (adapted from www.hewsaw.com).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 182<br />

Figure 5. The HewSaw SL250 sawing line at United Sawmills Ltd., Pori, Finland (left) and a<br />

tray sorter on the outfeed <strong>of</strong> the HewSaw R250 at Havesa Timber, Hamina, Finland<br />

(right).<br />

In normal processing <strong>of</strong> s<strong>of</strong>twood in Finland (Norway spruce and Scots pine), the HewSaw SL 250<br />

Trio has a maximum log throughput rate <strong>of</strong> 220,000 m 3 annum -1 for a single shift. In the CSIROMILL<br />

module the capacity has been reduced for spotted gum to a range <strong>of</strong> 125,000 to 175,000 m 3 annum -1<br />

for a single shift operation. This is to take into account log volume, higher wood density, increased<br />

board deflection and shorter logs (shorter boards), all <strong>of</strong> which will slow down material flow rates.<br />

The throughput rates modelled in the analysis ranged from 250,000 to 350,000 m 3 annum -1 for a two<br />

shift operation (approximately 60-80% <strong>of</strong> the maximum log throughput rate for s<strong>of</strong>twood).<br />

Sawn residues are chipped and transported by conveyors to a stock pile or for loading directly into<br />

trucks. Bark and 50% <strong>of</strong> the sawdust is transported to bins and sold to local markets. The remaining<br />

sawdust is used green as fuel to heat kilns.<br />

Immediately after sawing, boards containing sapwood are treated with a standard boron diffusion<br />

treatment.<br />

Drying incorporates air-drying in sheds with drop down sides to control air flow. Following air drying<br />

the wood is reconditioned in steam then kiln dried to a final moisture content <strong>of</strong> 10-12%. Heating for<br />

kilns is via pressurised hot water, heated from a furnace with sawdust and dry shavings used as fuel.<br />

The modelling is applicable to the production <strong>of</strong> oversized skip dressed boards that are sold for further<br />

manufacturing into flooring. Each model assumes that only the entire intake is <strong>of</strong> the logs specified. In<br />

reality there would be a range <strong>of</strong> log sizes and possibly more than one species.<br />

RESULTS<br />

Mill door prices for three annual log intake rates for both mills and IRR’s for the mills <strong>of</strong> 10, 15 and<br />

20% are given in Table 5 and Figure 6. Taxation rate on pr<strong>of</strong>it was 30%. Detailed output files for an<br />

example analysis for the SL 250 150000 and WH 40000 modules are given in Washusen (2008). Table<br />

5 also gives the Mill door prices for E. pilularis from the H200 PLUS 120000 module from Washusen<br />

et al. (2008).<br />

For 50,000 and 300,000 m 3 throughput, for the WH 30000 and SL 250 150000 modules respectively,<br />

the mill door prices were approximately AUD 90-110 per cubic metre higher for the SL 250 150000.<br />

The difference in log prices was due to reduced costs in all sectors <strong>of</strong> the mill, with greatest reductions<br />

in the sawmill. Costs attributed to sawing were AUD 14 and AUD 52 per cubic metre <strong>of</strong> log input for<br />

the linear and reciprocating system respectively.<br />

The results indicate that, on a comparable basis, there is an improvement in mill door prices with<br />

improved log quality and with the application <strong>of</strong> linear sawing systems for a species that can produce<br />

back-sawn boards that can be dried readily. Therefore, for such species there is considerable potential<br />

to improve returns to growers with the adoption <strong>of</strong> pruning to produce high quality sawlogs, and<br />

through the application <strong>of</strong> linear sawing systems. For this potential to be realised, resources <strong>of</strong> the


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 183<br />

appropriate size need to be developed within an economically viable supply zone. For plantations<br />

producing 10 m 3 ha -1 annum -1 <strong>of</strong> sawlogs this would require a total plantation area <strong>of</strong> 25,000 – 30,000<br />

ha and 4,500 – 5,500 ha for the linear and reciprocating system respectively. For the linear system, it<br />

may be difficult to achieve this plantation area in a single species. However, the mill design has<br />

capacity to process s<strong>of</strong>twood and eucalypt logs in batches. This was the approach taken by FEA,<br />

processing Pinus radiata and E. nitens with the HewSaw R200. This would require a modest<br />

plantation area that could be expanded given good economic returns to the mill and growers. In<br />

southern <strong>Australia</strong>, there are already in existence a number <strong>of</strong> s<strong>of</strong>twood mills processing below their<br />

capacity that could take advantage <strong>of</strong> appropriate resources. However, the plantation area available is<br />

very small and it may be several years before there are serious attempts to process large volumes <strong>of</strong><br />

high quality pruned eucalypts in linear sawing systems.<br />

Table 5. Mill door prices (in bold AUD per cubic metre <strong>of</strong> log intake) for pruned spotted gum<br />

estimated with CSIROMILL modules WH 30000 AD SG and SL 150000 AD SG and<br />

unpruned E. pilularis estimated with CSIROMILL module H200 PLUS 120000 AD<br />

BB.<br />

WH 30000 log intake (m 3 annum -1 )<br />

Logs processed IRR (%) 55,000 50,000 45,000<br />

Pruned spotted gum 10 123 111 95<br />

Pruned spotted gum 15 102 88 71<br />

Pruned spotted gum 20 81 65 46<br />

SL 250 150000 log intake (m 3 annum -1 )<br />

Logs processed IRR (%) 350,000 300,000 250,000<br />

Pruned spotted gum 10 208 201 191<br />

Pruned spotted gum 15 193 185 173<br />

Pruned spotted gum 20 178 168 154<br />

H 200 PLUS 120000 log intake (m 3 annum -1 )<br />

Logs processed IRR (%) 260,000 220,000 180,000<br />

Unpruned E. pilularis 10 42 35 26<br />

Unpruned E. pilularis 15 32 24 13<br />

Unpruned E. pilularis 20 22 13 1<br />

CONCLUSIONS<br />

This paper estimates mill door prices for pruned 30-35 cm small end diameter (SED) plantation-grown<br />

spotted gum logs modelled using the WH 30000 ADSG and HSL250 150000 ADSG modules <strong>of</strong><br />

CSIROMILL. These modules represent the state <strong>of</strong> the art reciprocating and linear sawing<br />

technologies that are capable <strong>of</strong> processing eucalypt logs within the SED range <strong>of</strong> 20-40 cm.<br />

Variables tested were a 30% taxation rate, IRR’s <strong>of</strong> 10, 15 and 20%, product values <strong>of</strong> $295 per cubic<br />

metre <strong>of</strong> log input and log input ranges <strong>of</strong> 45,000-55,000 m 3 annum -1 and 250,000-350,000 m 3 annum -1<br />

for the WH 30000 ADSG and SL 250 150000 AD SG modules respectively.<br />

Mill door prices were approximately AUD 90-100 higher for the SL 250 150000 ADSG module<br />

(linear system) for the most likely log throughput rates for both systems. This difference in log prices<br />

was due to reduced processing costs in all sections <strong>of</strong> the mill, with the largest reductions attributed to<br />

the sawmill (AUD 52 to AUD 14 per cubic metre <strong>of</strong> log intake).<br />

The Mill Door prices were also much higher from both systems than those reported for unpruned and<br />

unthinned E. pilularis mill door prices modelled with the H 200 PLUS 150000 module <strong>of</strong><br />

CSIROMILL in earlier work.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 184<br />

The results indicate considerable potential to improve returns to growers with the adoption <strong>of</strong> both<br />

pruning to produce high quality sawlogs, and the application <strong>of</strong> linear sawing systems. For this to be<br />

realised resources <strong>of</strong> the appropriate size would need to be developed within an economically viable<br />

supply zone.<br />

REFERENCES<br />

Washusen, R. (2006). Evaluation <strong>of</strong> product value and sawing and drying efficiencies for low rainfall hardwood<br />

thinnings. A report for the RIRDC/Land and Water <strong>Australia</strong>/FWPRDC/MDBC Joint Venture<br />

Agr<strong>of</strong>orestry Program. Project No CSF-65A. pp 52.<br />

Washusen, R. and Innes, T. C. (2007). Processing plantation eucalypts for high value timber. In proceedings <strong>of</strong><br />

the Plantation eucalypts for high value timber conference, 9-12 October 2007, Moorabin, Melbourne.<br />

pp 91-108.<br />

Washusen, R., Morrow, A., Ngo, D,. Bojadzic, M., Henson, M., Porada, H., Northway, R., Boynton, S.,<br />

Chen Shaoxiong, Peng Yan, Nguyen Quang Trung and Bui Chi Kien. (2007). Genetic variation in<br />

growth stress related wood behaviour <strong>of</strong> small diameter Eucalyptus nitens logs processed with a<br />

Hewsaw R250 sawmill. Report for ACIAR project FST/2001/021: Improving the value chain for<br />

plantation-grown eucalypts in China, Vietnam and <strong>Australia</strong>: sawing and drying. Client Report No<br />

1799. pp 39.<br />

Washusen, R., Morrow, A., Dung Ngo, Northway, R., Boynton, S. and Henson, M. (2008). Processing air and<br />

kiln dried sawn wood from unthinned and unpruned 9-year old Eucalyptus pilularis from northern New<br />

South Wales with a HewSaw R200. Report for ACIAR project FST/2001/021: Improving the value<br />

chain for plantation-grown eucalypts in China, Vietnam and <strong>Australia</strong>: sawing and drying. CSIRO<br />

Material Science & Engineering Client Report No CMSE(C)-2008-300. pp 50.<br />

Washusen, R. (2008). A comparison <strong>of</strong> mill door log prices for pruned Corymbia spp processed with<br />

conventional reciprocating and linear flow sawing systems. Report for ACIAR project FST 2001/021.<br />

Improving the value chain for plantation grown Eucalypts in China, Vietnam and <strong>Australia</strong>: sawing and<br />

drying. CSIRO Materials Science & Engineering Client Report No: CMSE (C)-2008-301. pp 42.<br />

Washusen, R., Morrow, A., Dung Ngo, Harding, K. and Innes, T. (2009). Processing air and kiln dried sawn<br />

wood from unthinned and unpruned 7.5 year-old Corymbia citriodora subsp. variegata from<br />

Queensland with a HewSaw R200. Report for ACIAR Project FST 2001/021. Improving the value<br />

chain for plantation grown Eucalypts in China, Vietnam and <strong>Australia</strong>: sawing and drying. CSIRO<br />

Materials Science & Engineering Client Report No: CMSE-2009-046. pp 43.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 185<br />

REDUCED FUEL USE IN <strong>FORESTRY</strong> TRANSPORTATION<br />

THROUGH THE USE OF HIGHER PRODUCTIVITY VEHICLES (HPV)<br />

Mark Brown 1<br />

ABSTRACT<br />

In transportation, payload is critical to vehicle efficiency and economics. The focus <strong>of</strong><br />

improvements in transportation has mainly been on the economics <strong>of</strong> improved payload,<br />

but fuel usage for a given transport task is equally important.<br />

In the load-constrained context <strong>of</strong> transport on public road networks, potential gains in<br />

payload must be achieved by reduced tare weight or the introduction <strong>of</strong> new<br />

configurations that have higher load limits. With the introduction <strong>of</strong> performance based<br />

standards (PBS) in <strong>Australia</strong>, the forest industry is exploring the opportunity to introduce<br />

new high productivity vehicles (HPV).<br />

From an analysis <strong>of</strong> weighbridge data from six forestry companies, a potential 13%<br />

reduction in fuel use has been identified through lower tare weight specifications. A<br />

further 71% reduction may be possible through new HPV designs.<br />

INTRODUCTION<br />

Transportation in forestry has the draw back that a significant cost is involved in the delivery <strong>of</strong> wood<br />

and it users a large amount <strong>of</strong> energy. Studies in Sweden have shown that transportation represents 53<br />

to 56% <strong>of</strong> the total energy used in the production and delivery <strong>of</strong> timber (Berg and Lindholm, 2005).<br />

Similarly, in western Canada the Forest Engineering Research <strong>Institute</strong> <strong>of</strong> Canada (FERIC) found<br />

haulage accounts for 51% <strong>of</strong> the total energy used in forest operations (Sambo, 2002). While<br />

operations in <strong>Australia</strong> certainly differ from those in Sweden and Canada, the fact that similar results<br />

are reported across a broad geographical area, and that the equipment used is similar to that used in<br />

<strong>Australia</strong>, suggests that transportation is likely to be a significant proportion <strong>of</strong> total energy use in<br />

<strong>Australia</strong>n forest operations.<br />

As fuel use in truck transport is <strong>of</strong> interest outside <strong>of</strong> forest transportation, there is no shortage <strong>of</strong><br />

technology and training available to help reduce fuel use. While many solutions <strong>of</strong>fer significant<br />

opportunity for reducing fuel use, they <strong>of</strong>ten come at a high cost and/or are difficult to implement<br />

effectively. Solutions typically involve improving technology so that the same level <strong>of</strong> work is<br />

achieved with lower energy use, or by adapting work methods to reduce “wasted energy”.<br />

An alternative approach is for the technology to do more work with the same or a nominal increase in<br />

energy usage in order to achieve an overall reduction <strong>of</strong> energy to complete a freight task. An<br />

example <strong>of</strong> this approach was demonstrated by FERIC when examining the potential to reduce fuel<br />

consumption by using large <strong>of</strong>f-highway trucks. In the study FERIC found that an <strong>of</strong>f-highway truck<br />

that consumed almost twice as much fuel as the normal haulage truck on a l/100km basis actually used<br />

36% less fuel on a l/tonne transported basis to complete the same freight task (Michaelsen, 2007).<br />

While the introduction <strong>of</strong> very large <strong>of</strong>f-highway trucks is not feasible across many forest operations<br />

in <strong>Australia</strong>, the concept <strong>of</strong> doing more work with nominally the same or even more energy is<br />

applicable here. As most wood in <strong>Australia</strong> is transported via public roads for part <strong>of</strong> the journey from<br />

forest to mill, the total load or gross vehicle weight (GVW) for a given vehicle configuration is limited<br />

by public road restrictions. In some cases the actual vehicle configuration that is accepted on the route<br />

will also be limited.<br />

Until recently these vehicle limits have been very restrictive with most <strong>Australia</strong>n states allowing only<br />

two or three configurations which fall within the GVW restriction <strong>of</strong> 42 to 62.5 tonnes. Consequently,<br />

1 University <strong>of</strong> Melbourne, Program Leader, CRC Forestry, Harvesting and Operations


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 186<br />

to increase the vehicle work output under these restrictions, the empty or tare weight needs to be<br />

reduced, to allow an increase in payload without exceeding the GVW. In Case Study 1 (below) data<br />

from six <strong>Australia</strong>n forest companies are examined to see how well tare weights are managed, and<br />

what opportunities exist for improvement.<br />

There has recently been a change in the <strong>Australia</strong>n road rules for vehicle configurations, with a shift<br />

from prescriptive standards (predefined configurations like B-doubles) to performance based standards<br />

(PBS). As a result, industry is no longer confined to a limited set <strong>of</strong> predefined configurations.<br />

Industry is now able to design a configuration ideally suited to a given application, provided it can<br />

demonstrate that it meets the performance standards which ensure it is safe and infrastructure-friendly<br />

(Prem, Ramsay, and McLean, 1999). Having a greater variety <strong>of</strong> vehicle configurations is <strong>of</strong><br />

particular interest to the forest industry, as the energy efficiency <strong>of</strong> its vehicles can be increased, and<br />

the transportation costs and associated pollution can be reduced.<br />

Through discussion with the forest industry and trailer manufacturers, the CRC for Forestry has<br />

identified two new vehicle configurations which <strong>of</strong>fer attractive potential gains in efficiency. These<br />

gains will be evaluated in operational trials.<br />

Case Study 2 examines the potential impact <strong>of</strong> a PBS design which uses <strong>of</strong>f-the-shelf technology to<br />

produce a quad axle B-double. Case Study 3 looks at the application <strong>of</strong> new stearable wheel<br />

technology to produce a high productivity semi-trailer.<br />

CASE STUDY 1: Improving the current configuration with a focus on tare weight<br />

Before exploring the use <strong>of</strong> new technology and trailer designs, the first step is to see what potential<br />

exists for improvements in vehicle efficiency through improvements with currently used<br />

configurations. This approach is attractive, as potential improvements might be made at relatively low<br />

cost and low risk simply by promoting the application <strong>of</strong> current best practice without venturing into<br />

the unknown.<br />

This approach is based on the theory that each configuration is limited by the GVW it can transport —<br />

a combination <strong>of</strong> the vehicle tare weight and the payload. As forest products are transported as a bulk<br />

commodity, load management is highly important and care must be taken to ensure each load is as<br />

close as possible to the maximum load, without exceeding load limits.<br />

When this is done, every kilogram <strong>of</strong> vehicle tare weight represents a kilogram <strong>of</strong> lost payload each<br />

trip. This is based on the core assumption that those vehicles with a lower tare weight will be able to<br />

transport more wood per trip at the same GVW than those vehicles with higher tare weights, but still<br />

use only the same amount <strong>of</strong> fuel.<br />

To understand what the potential savings might be, weighbridge data (by load delivered) providing<br />

details <strong>of</strong> truck identification, date, time, GVW and tare weight were collected from six forest<br />

companies from Western <strong>Australia</strong>, South <strong>Australia</strong> and Victoria over a period from 6 to 12 months.<br />

After eliminating corrupted data points, partial loads, and trucks making only few deliveries (less than<br />

10), the combined fleet for analysis comprised 427 trucks <strong>of</strong> four configurations which delivered a<br />

total <strong>of</strong> 52,306 loads within the analysis period.<br />

Table 1 presents a summary <strong>of</strong> the data, including the average tare weight for the lightest 20%, the<br />

average tare, and the average tare weight for the heaviest 20% <strong>of</strong> the vehicles in a configuration class.<br />

The table indicates there is a large variation in vehicle tare weights within the same configuration<br />

class, with a difference <strong>of</strong> up to 26% between the lightest 20% and heaviest 20% in a group.<br />

In the case <strong>of</strong> the 82t road trains, the tare weight variation within the configuration is greater than the<br />

gain achieved by moving up to the 82t road train configuration from the 79t road train.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 187<br />

Table 1. Summary <strong>of</strong> weigh bridge data<br />

Trucks Loads<br />

Allowable<br />

GVW<br />

Lightest<br />

20 %<br />

Average tare<br />

Whole<br />

group<br />

Heaviest<br />

20 %<br />

# # Kg Kg Kg Kg<br />

Semi-Trailers 97 9962 42500 16627 18275 20056<br />

B-doubles 87 17212 62500 21107 22651 24679<br />

79t Road trains 209 20071 79000 26008 28585 32867<br />

82t Road trains 34 5061 82500 27194 29450 31924<br />

For simplicity, the analysis in Table 2 assumes that vehicle fuel consumption is unchanged with a<br />

change in tare weight. Therefore if more wood can be moved per trip there is an opportunity to<br />

decrease fuel use for a given freight task with no change in vehicle configuration. Since vehicle fuel<br />

consumption increases with weight (USEPA,2004) further savings are likely as the time traveling<br />

empty would experience lower fuel consumption under the same operating conditions. Further<br />

research however is required to quantify fuel savings when the vehicle is empty.<br />

Table 2. Fuel savings relative to the heaviest 20% through increased payload<br />

Potential payload<br />

Fuel savings as compared to<br />

heaviest 20 % in class<br />

Lightest<br />

20 %<br />

Average<br />

Heaviest<br />

20 %<br />

Lightest<br />

20 %<br />

Average<br />

Kg Kg Kg<br />

Semi-Trailers 25873 24225 22444 13 % 7 %<br />

B-doubles 41393 39849 37821 9 % 5 %<br />

79t Road trains 52992 50415 46133 13 % 8 %<br />

82t Road trains 55306 53050 50576 9 % 5 %<br />

To put this into perspective, many forest companies have indicated they would invest in upgrading<br />

from an average B-double to an average 79t road train were that configuration allowed elsewhere apart<br />

from Western <strong>Australia</strong>. Such an investment would yield just over 20% improvement in efficiency, but<br />

requires significantly more effort to introduce and a higher initial investment.<br />

CASE STUDY 2: HPV Quad-B-double using current self-steering axle technology<br />

With the opportunity to develop alternative high productivity configurations through PBS, many<br />

transportation industries have explored options to increase payload to improve efficiency. One concept<br />

which has shown promise for several bulk transport applications, including wood chip transport, is the<br />

Quad-B-double configuration shown in Figure 1.<br />

Figure 1. Quad-B-double


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 188<br />

By including two additional self-steering axles over a traditional B-double, the GVW is increased by<br />

14.5 tonnes. This approach is favoured by the transport industry as it stays with design and technology<br />

with which they are familiar, and the rig is only 1 metre longer than a B-double. Furthermore, the<br />

industry believes public acceptance for this configuration will be possible with a reasonable level <strong>of</strong><br />

effort.<br />

Initial evaluation <strong>of</strong> the new Quad-B-double configuration for wood chip transport has led to the<br />

conclusion that the configuration will have a sufficient storage volume to reach full loads based on<br />

normal wood chip density. As the Quad-B-double is not currently hauling woodchips, determining the<br />

tare weights for evaluation <strong>of</strong> the impact on efficiency requires the use <strong>of</strong> certain assumptions. Firstly,<br />

as the Quad-B-double design is the same as the standard B-double, it is assumed the tare weights<br />

would be about the same with an additional increase in weight for the two additional axles and the<br />

extra one metre <strong>of</strong> length. Each axle weighs between 800 kg and 1200 kg depending on the axle<br />

design and suspension, with self-steering axles being at the heavier end <strong>of</strong> the range. The weight <strong>of</strong> the<br />

extra metre in length would vary between 200 Kg and 400 Kg depending on the frame and wall<br />

materials used. Based on these ranges, it is assumed that the Quad-B-double would weigh, on<br />

average, about 2400kg more than a standard B-double. Finally, since the heavier trucks in the fleet<br />

tend to be older, it is assumed the new Quad-B-doubles would cover a range <strong>of</strong> tare weights<br />

comparable to the lower end <strong>of</strong> the range reported in Case Study 1, giving a probable tare weight<br />

range <strong>of</strong> 23 500 Kg to 25 100 Kg, and a potential payload range <strong>of</strong> 51 900 Kg to 53 500 Kg.<br />

Table 3 shows the impact on fuel use with the shift to the Quad-B-double from the standard B-double<br />

for a given freight task. With this shift in configuration there are potential fuel savings <strong>of</strong> over 40%.<br />

Table 3. Difference in fuel use between Quad-B-double and Standard B-double.<br />

B-double<br />

Difference in fuel use Lightest<br />

Heaviest<br />

Average<br />

20 %<br />

20 %<br />

Quad-B- Light 29 % 34 % 41 %<br />

double Heavy 25 % 30 % 37 %<br />

CASE STUDY 3: HPV semi-trailers using new steerable wheel technology<br />

In addition to new and creative uses <strong>of</strong> existing technology, the introduction <strong>of</strong> PBS has provided an<br />

opportunity to incorporate new technology into the design <strong>of</strong> completely new HPV configurations.<br />

One example <strong>of</strong> this is the steerable wheel concept developed by Steerable Wheel Systems Pty Ltd<br />

(SWS). SWS has developed an innovative modular steerable wheel system that comprises<br />

electronically controlled twin wheels with integrated suspension (Prem, May, & Davey, 2008). The<br />

advantage <strong>of</strong> this new technology, <strong>of</strong> particular interest for woodchip transport, is the lack <strong>of</strong> an axle<br />

allowing for greater load volumes with a lower centre <strong>of</strong> gravity. In addition, the actively steered<br />

wheels allow for longer single trailer units which can be unloaded more quickly than trains.<br />

Furthermore, the wheel spacing options allow for a higher GVW within current semi-trailer and Bdouble<br />

overall dimensions.<br />

To examine the potential <strong>of</strong> this new concept for the forest industry, two steerable wheel<br />

configurations are being evaluated with the industry, including a 19 metre configuration (Figure 2)<br />

which would be an alternative to the current semi-trailer configuration, and a 26 metre configuration<br />

(Figure 3) which would be an alternative for the current B-double configuration.<br />

Figure 2. A 19 m steerable wheel<br />

configuration — 55.7 t GVW and 41.3 t<br />

payload


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 189<br />

Figure 3. A 26 m steerable wheel configuration — 68 t GVW and 51 t payload<br />

Since steerable wheel trailers are a new concept based on a new technology, with new trailer design<br />

tailored to the technology, there are no examples on the road from which to draw actual weight<br />

information. Therefore tare weights and payload need to be derived from theoretical weights provided<br />

by design models. In the case <strong>of</strong> the 19 m configuration, the allowable GVW based on PBS will be<br />

55.7 t and the current design predicts a tare weight <strong>of</strong> 14.4 t for a payload <strong>of</strong> just over 41 t.<br />

For the 26 metre configuration, the PBS design would allow for a 68 t GVW and the design model<br />

predicts a tare weight <strong>of</strong> 17 t for a payload <strong>of</strong> 51 t. Assuming the predicted tare weights may be<br />

optimistic, the analysis in Table 4 summarises the impact on fuel use <strong>of</strong> moving from the current semitrailer<br />

configuration to the 19 m steerable wheel configuration:<br />

o at the predicted tare weight, and<br />

o at a tare weight that is 10% higher than predicted.<br />

Table 4. Difference in fuel use between semi-trailers and 19 m steerable wheel configuration<br />

Difference in fuel use<br />

Semi-trailer<br />

Lightest 20 % Average Heaviest 20 %<br />

19 m Predicted tare 60 % 70 % 84 %<br />

Steerable +10 % 54 % 65 % 78 %<br />

Table 5 summarises the impact <strong>of</strong> moving from the current B-double configurations to the 26 m<br />

steerable wheel configuration on the same basis.<br />

Table 5. Difference in fuel use between B-doubles and 26 m steerable wheel configuration<br />

Difference in fuel use<br />

Lightest 20 %<br />

B-double<br />

Average Heaviest 20 %<br />

26 m Predicted Tare 23 % 28 % 35 %<br />

Steerable +10 % 19 % 24 % 30 %<br />

Even at a tare weight 10% higher than predicted for the steerable wheel configurations, fuel use<br />

savings <strong>of</strong>:<br />

o up to 78 % are possible for the semi-trailer alternative, and<br />

o up to 30 % for the B-double alternative.<br />

These results are quite comparable to the savings achieved with the Quad-B-double in Case Study 2,<br />

but in this case the steerable wheel configuration has fewer wheels, a lower center <strong>of</strong> gravity and can<br />

be easily unloaded as a single unit.<br />

CONCLUSION<br />

Fuel consumption is an important factor in both the cost and environmental impact <strong>of</strong> transportation<br />

operations. One major focus <strong>of</strong> the transportation industry is the control these impacts by reducing<br />

fuel use while doing the same work.<br />

The three case studies presented, however, show more work can be achieved by using the same<br />

amount <strong>of</strong> fuel and, similarly to the FERIC study in Canada, by consuming a more nominal amount <strong>of</strong><br />

fuel.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 190<br />

Each <strong>of</strong> the reported case studies examines progressively higher risk approaches to improving payload<br />

and thus the efficiency <strong>of</strong> forest transportation on public roads in <strong>Australia</strong>. In Case Study 1, fuel<br />

savings <strong>of</strong> up to 13% can be achieved by specifying vehicles that are as light as the current best 20%<br />

in <strong>Australia</strong>. As this involves no change in vehicle configuration nor the application <strong>of</strong> new<br />

technology, the effort to implement the change is relatively small, and returns potentially high.<br />

In Case Study 2, by taking advantage <strong>of</strong> new PBS rules and using <strong>of</strong>f-the-shelf technology, even<br />

greater improvements are possible, with potential fuel savings <strong>of</strong> up to 41%. This approach requires<br />

the introduction <strong>of</strong> a new configuration built from relatively proven technology. Therefore the<br />

technical risk is low, but effort may be high to win public acceptance and ultimately route access for<br />

the use this new configuration. But the reward is high.<br />

Finally, Case Study 3 examines the use <strong>of</strong> a new configuration based on new design and technology to<br />

achieve savings as high as 84%. This approach is accompanied by a significant cost and risk in<br />

dealing with new technology and, as with the Quad-B-double, will require significant effort to gain<br />

route access.<br />

However this vehicle configuration <strong>of</strong>fers significant potential fuel savings, and <strong>of</strong>fers some design<br />

features that will benefit other aspects <strong>of</strong> the operations.<br />

This paper has focused more on the fuel savings that can be achieved. However, changes in vehicle<br />

configuration may also reduce green house gas emissions. Based on the fact that each litre <strong>of</strong> diesel,<br />

when burned, produces 2.7 kg <strong>of</strong> C02, green house gas emissions would be reduced by the same<br />

percentages, with increased payloads. Thus the industry would demonstrate its social responsibility in<br />

protecting the environment. The outcome <strong>of</strong> the case studies is important, and will become<br />

increasingly so with the introduction <strong>of</strong> carbon emission trading or carbon tax schemes in <strong>Australia</strong>.<br />

REFERENCES<br />

Berg S. and Lindholm E-L., 2005. ‘Energy use and environmental impacts <strong>of</strong> forest operations in<br />

Sweden’, Journal <strong>of</strong> Cleaner Production vol. 13, pp. 33-42.<br />

Michaelsen J., 2007. Reducing fuel consumption, emissions <strong>of</strong> GHG and transportation costs by the<br />

implementation <strong>of</strong> partial <strong>of</strong>f-highway hauls in forest operations, Contract Report- CR-0293-1,<br />

Forest Engineering Research <strong>Institute</strong> <strong>of</strong> Canada, Montreal, Quebec.<br />

Prem H., Mai L., and Davey G., 2008. ‘A new steerable wheel system for road transport applications’,<br />

Proceedings <strong>of</strong> the International Conference on Heavy Vehicle, Paris 2008, paper 31.<br />

Prem H., Ramsay E., and McLean J., 1999. Performance based standards for heavy vehicles:<br />

Assembly <strong>of</strong> case studies, Report to National Transport Commission <strong>Australia</strong>.<br />

Sambo S.M., 2002. Fuel consumption for ground-based harvesting systems in western Canada,<br />

Advantage Report v.3 no.29, Forest Engineering Research <strong>Institute</strong> <strong>of</strong> Canada, Montreal,<br />

Quebec.<br />

Smartway Transport Partnership, 2004. A glance at clean freight strategies weight reduction, U.S.<br />

Environmental Protection Agency, viewed 8 April 2009,<br />


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 191<br />

ONBOARD SYSTEMS FOR<br />

AUSTRALIAN FOREST OPERATIONS<br />

Martin Strandgard 1<br />

ABSTRACT<br />

Management <strong>of</strong> harvesting operations using onboard technology has been highly<br />

developed overseas, particularly in Scandinavia, but is relatively still in its infancy in<br />

<strong>Australia</strong>. Overseas experience has identified productivity gains <strong>of</strong> over 30% in some<br />

areas and expectations <strong>of</strong> further significant gains. On the basis <strong>of</strong> this overseas<br />

experience, the CRC for Forestry Program 3 has commenced to examine the range <strong>of</strong><br />

onboard technology available, and to identify and trial equipment suitable for <strong>Australia</strong>n<br />

forest harvesting conditions. The paper presents the preliminary findings <strong>of</strong> the project<br />

and the expected costs and gains from the adoption <strong>of</strong> the technology.<br />

INTRODUCTION<br />

Harvesting and transport account for over 50% <strong>of</strong> the costs for a rotation. Mechanical harvesting<br />

operations in <strong>Australia</strong> typically cost between $20 - $40 per tonne (averaging about $25 per tonne) and<br />

are currently used on about 80% <strong>of</strong> the 25 million tonnes harvested annually, a percentage that is<br />

increasing. International experience has shown that effective use <strong>of</strong> onboard computing technology<br />

can increase machine utilisation and productivity by up to 30% through identifying operational<br />

bottlenecks, delays and improving operator performance. Targeted operator training and damage<br />

prevention techniques have produced a reduction <strong>of</strong> up to 40% in particular maintenance costs.<br />

Limited detailed time-studies in <strong>Australia</strong> have found machine utilisation rates are typically between<br />

70 and 85%, with some under 60% (Acuna et al., 2009).<br />

OBJECTIVES<br />

This paper is based on a project funded by Forests and Wood Products <strong>Australia</strong> (FWPA) and three<br />

industry partners: Timbercorp Ltd, ForestrySA and VicForests. As the industry partners cover a broad<br />

cross-section <strong>of</strong> the <strong>Australia</strong>n forest industry it is expected that the results will be widely applicable to<br />

<strong>Australia</strong>n forestry.<br />

The project objectives are:<br />

• Improvement in utilisation and/or productivity <strong>of</strong> forestry equipment <strong>of</strong> at least 10%<br />

• Improvement <strong>of</strong> machine availability <strong>of</strong> at least 5%<br />

• Measurable improvements in fuel efficiency <strong>of</strong> the operations<br />

• Measurable improvement in operation safety and operator job satisfaction<br />

The project will consider only <strong>Australia</strong>n harvesting operations and only those activities associated<br />

with harvesting, processing, transport to the coupe edge and onsite chipping.<br />

Utilisation is the proportion <strong>of</strong> scheduled time that a machine actually works. For most machines it<br />

can be further broken down into travel and work times. Utilisation is improved mainly through<br />

identifying and reducing delays. Reducing travel time can also be important.<br />

Productivity is the amount <strong>of</strong> wood cut or transported in a given time (typically m 3 or tonnes/hr). It<br />

can be expressed in terms <strong>of</strong> either productive (actual work hours) or scheduled machine hours.<br />

Productivity is strongly related to piece size (Hartsough and Cooper, 1998) and utilisation and, for<br />

forwarding, load capacity and hauling distance (Jiroušek et al. 2007). Operator skill is also a factor<br />

(Ovaskainen et al., 2004). Both utilisation and productivity can be affected by machine interactions in<br />

1 Department <strong>of</strong> Forest and Ecosystem Science, University <strong>of</strong> Melbourne. Email: mnstra@unimelb.edu.au.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 192<br />

a “hot deck” operation (Spinelli and Visser, 2008), for instance if a forwarder has to wait for<br />

harvesters to produce wood or a mobile chipper has to wait for wood to be delivered.<br />

In sawlog cut to length operations, value recovery is <strong>of</strong> more interest to forest owners than<br />

productivity. For example in ForestrySA’s operations, cross-cutting decisions can result in large<br />

differences in value recovery without changing volume output. Value recovery is being investigated<br />

in another project but potential synergies between the projects will be identified.<br />

Machine availability is the proportion <strong>of</strong> scheduled time that the machine is mechanically available for<br />

work. As such, it reflects machine characteristics (e.g. make/model, age, work and maintenance<br />

history), site factors (e.g. rockiness, slope) and also, to an extent, operator skills and attitude.<br />

Fuel efficiency is generally expressed in terms <strong>of</strong> fuel consumed per unit <strong>of</strong> wood cut or transported.<br />

Fuel use per unit <strong>of</strong> wood extracted in Sweden has been reduced from 5.4 l/m 3 to 3.7 l/m 3 over the 20<br />

years to 2006, mainly through the introduction <strong>of</strong> lighter, more fuel efficient machines and operator<br />

training. For a given site and machine, fuel efficiency is strongly related to utilisation and productivity.<br />

Operation safety is measured in terms <strong>of</strong> number <strong>of</strong> lost time injuries and amount <strong>of</strong> lost time per unit<br />

time (<strong>of</strong>ten expressed as Lost Time Injury Frequency Rate (LTIFR) which is the lost time per million<br />

hours worked). Job satisfaction is difficult to measure accurately and is dependent on both work<br />

related and external factors.<br />

DELIVERABLES<br />

The primary project deliverable is a selection guide to allow users to match their circumstances (e.g.<br />

plantation, native forest, grower, contractor, machine types, hot/cold decking, etc) and the issues they<br />

would like to resolve or understand (e.g. poor utilisation, productivity, etc) with an appropriate<br />

onboard system (or systems). This will be accompanied by an implementation guide and basic<br />

s<strong>of</strong>tware data analysis tools.<br />

It is expected that the primary delivery mechanism will be the web, due to the ease <strong>of</strong> implementation,<br />

distribution, and management <strong>of</strong> the selection guide and other material via this medium.<br />

AVAILABLE ONBOARD COMPUTING EQUIPMENT<br />

A review <strong>of</strong> available equipment identified three major categories <strong>of</strong> onboard computing technology<br />

which have the potential to meet the project requirements:<br />

1. Purpose – built forestry systems<br />

Purpose-built systems have the advantages <strong>of</strong> associated s<strong>of</strong>tware to manage, analyse and<br />

display forestry specific data and the ability to obtain user inputs, particularly delay causes.<br />

• Manufacturer supplied systems. All recently manufactured harvesting equipment<br />

has some level <strong>of</strong> onboard computing technology. Often a number <strong>of</strong> computers are<br />

available with different capabilities (and prices). Data are transferred by USB<br />

memory device, or over the mobile phone network, or in some cases by floppy disk.<br />

The data complies with the StanForD standard (Skogforsk, 2007), allowing data from<br />

different manufacturers to be processed using any manufacturer’s s<strong>of</strong>tware.<br />

• Multidats. Multidats were developed by FPInnovations (formerly Feric) in Canada to<br />

collect long-term utilisation data from forestry machines. The basic configuration<br />

records machine activity using a vibration sensor. There are 4 channels for<br />

connection <strong>of</strong> analogue sensors to switches, current/pressure sensors, etc and an<br />

optional GPS input. Operators can also enter delay causes in categories. Operator<br />

feedback can be provided with the optional Multipad display. Data are transferred<br />

with a PDA or laptop connected to the Multidat by cable.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 193<br />

2. Telematics<br />

Telematics is the remote monitoring <strong>of</strong> machine function and location. It originated with the<br />

trucking industry and has recently been applied to <strong>of</strong>f-road vehicles. Several major<br />

manufacturers <strong>of</strong> forestry equipment have produced proprietary telematic products for their<br />

construction equipment (e.g. Product Link from Caterpillar and Komtrax from Komatsu).<br />

There are also a number <strong>of</strong> generic telematic solutions e.g. Mix Telematics and Topcon Tierra.<br />

Telematic equipment will soon be standard on new construction equipment and can be<br />

retr<strong>of</strong>itted to older machines or to machines from other manufacturers (with limited<br />

functionality).<br />

The key characteristic <strong>of</strong> telematic equipment is that data collected by the onboard technology<br />

are sent periodically to a central computer via satellite or mobile phone, and accessed via a<br />

web site. The service is provided by a one-<strong>of</strong>f payment or ongoing fee-for-service basis for<br />

each machine.<br />

The basic function <strong>of</strong> telematics is tracking <strong>of</strong> vehicle movements via GPS. Most telematic<br />

equipment can also collect data from the machine’s Canbus (see below). Some can collect<br />

additional data using digital or analogue sensors. There are currently no standards as to what<br />

data are collected and how they are presented to end-users, though this is being addressed by<br />

the Association <strong>of</strong> Equipment Management Pr<strong>of</strong>essionals.<br />

Telematics is being largely driven by the requirements <strong>of</strong> the trucking and construction<br />

industries. That said, there is a great deal <strong>of</strong> potential overlap in requirements, such as vehicle<br />

location, hours worked and fuel consumption. However, some degree <strong>of</strong> customisation will be<br />

required to meet the requirements <strong>of</strong> the forest industry.<br />

There are a small number <strong>of</strong> telematic systems that encompass both heavy vehicles and<br />

forestry applications, such as RouteHawk (Strong Engineering), FCGIS-H (Bracke Systems)<br />

and TAGASIS (Aldata).<br />

3. Generic Dataloggers<br />

A wide range <strong>of</strong> generic dataloggers is available which could be used to collect data from<br />

forest harvesting machines (e.g. Datataker, BusDAQ and ASLH309). Some have been used<br />

for this purpose, mainly in research trials. Typically they have multiple channels that can<br />

collect digital, analogue, or Canbus (see below) data, or a combination <strong>of</strong> these. GPS input is<br />

an option on some dataloggers, and some can be interfaced with displays to provide feedback<br />

to, or input from, machine operators.<br />

Depending on the make and model, generic dataloggers can transfer data by a variety <strong>of</strong><br />

means including USB memory devices, wifi (Bluetooth or wlan), radio modems and mobile<br />

phone modems.<br />

Equipment installation<br />

With the exception <strong>of</strong> manufacturer supplied equipment, data collection hardware needs to be wired<br />

into each machine to collect the required data using either stand-alone sensors, such as a vibration<br />

sensor, connections to existing switches or installation <strong>of</strong> new switches to detect on/<strong>of</strong>f actions (e.g.<br />

pedal movements), or electric current or hydraulic pressure changes. The alternative to this approach<br />

is to monitor the Canbus. Canbus is a two wire communications bus originally developed by Bosch to<br />

simplify wiring looms in luxury cars. It has been widely adopted in many areas, including heavy<br />

vehicles such as trucks and forest machines. Canbus promises to deliver a great deal <strong>of</strong> useful<br />

information about machine performance and activities with a single connection (eg. total operating<br />

time, work time, travel time, fuel use, engine rpm/load). Forest harvesting machines control the<br />

engine and transmission using a Canbus compliant with the heavy equipment standard J1939 and the<br />

remainder <strong>of</strong> the machine functions using proprietary manufacturer codes which differ between<br />

manufacturers and machine models. The data available from the J1939 compliant component is


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 194<br />

unknown at this stage as there is no compulsion for a manufacturer to fully implement the standard or<br />

to publicly document their implementation.<br />

The Canbus is accessed either via an existing diagnostic port or by wiring an additional port into the<br />

system (if the system has this capability) which then must be enabled programatically.<br />

As the use <strong>of</strong> Canbus for data collection on forestry machines is a relatively unknown factor,<br />

equipment selected with Canbus capability will need to be able to collect data using analogue and/or<br />

digital channels to minimise the risk <strong>of</strong> not being able to collect useful data via the Canbus.<br />

INDUSTRY PARTNER REQUIREMENTS<br />

Each industry partner will be installing equipment on 3 to 4 forest machines which work together on<br />

an operation. To test a range <strong>of</strong> onboard computing equipment and to avoid data integration problems,<br />

it is anticipated that each partner will test a different type <strong>of</strong> equipment and will use one brand <strong>of</strong><br />

equipment (where possible) within their organisation.<br />

ForestrySA manage radiata pine plantations primarily for sawlog production. As mentioned above,<br />

ForestrySA have a particular interest in value recovery while their contractors are focussed on<br />

utilisation and productivity. Manufacturer supplied equipment is the only means to achieve both these<br />

requirements. Equipment for this study has to be capable <strong>of</strong> optimising cross-cutting decisions and<br />

allowing operators to enter delay causes. Contractors are unlikely to allow ForestrySA staff to view<br />

data related to utilisation and productivity, so segregation <strong>of</strong> data needs to be factored into the data<br />

collection and transfer process. To meet these requirements Dasa computers will be installed. Dasa<br />

equipment is installed by manufacturers and as aftermarket equipment which makes it ideally suited<br />

for this project.<br />

Timbercorp are a major blue gum MIS company producing woodchips for export that owns and<br />

operates their own harvesting equipment. Their typical harvesting operation is to fell and debark<br />

stems at the stump, and use forwarders to transport whole stems to a roadside chipper. This type <strong>of</strong><br />

operation is primarily concerned with efficient throughput <strong>of</strong> timber. The fundamental drivers are the<br />

number <strong>of</strong> stems each harvester processes, the number <strong>of</strong> forwarder loads and the tonnes <strong>of</strong> wood<br />

processed by the chipper. These are in turn dependent on the utilisation percentage. Machine<br />

interactions can also affect individual machine performance.<br />

Timbercorp are interested in all aspects <strong>of</strong> utilisation, productivity, fuel efficiency, etc. Timbercorp<br />

currently have Multidats installed on a range <strong>of</strong> harvesting equipment as part <strong>of</strong> another trial.<br />

RouteHawks will be installed on the Timbercorp equipment to test the telematics approach including<br />

data collection from the Canbus and potentially remote data collection via mobile or satellite phones.<br />

Although Timbercorp are currently under administration, it is anticipated that this will be resolved in<br />

time for the project to continue. If this proves not to be the case, the project will proceed with the<br />

other two industry partners.<br />

VicForests extract timber from Victoria’s native forest estate using a mixture <strong>of</strong> hand-felling and<br />

mechanised operations depending on tree size. In clearfell operations, stems are generally cut using a<br />

felling head on an excavator and transported to a landing by a grapple skidder. Excavator-mounted<br />

grapples are then used to debark, cut to length and load logs onto trucks for transport to the mill.<br />

Although price premiums apply to high quality log products, it is generally not possible to identify log<br />

quality due to internal defect until the stem has been cut. Regrowth thinning operations generally use<br />

harvesters and/or feller-bunchers and forwarders to harvest and process trees.<br />

As with ForestrySA, VicForests’ interests differ from those <strong>of</strong> their contractors. The recommended<br />

solution is to install Multidats on a range <strong>of</strong> equipment, to obtain basic productivity and utilisation<br />

information, and then examine the breaking down <strong>of</strong> utilisation to travel and work times, and the<br />

breaking down <strong>of</strong> causes <strong>of</strong> delay into categories, such as mechanical, personal and operational delays.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 195<br />

Each mobile machine will also have a Garmin GPS/GIS to display coupe boundaries, site hazards and<br />

the machine’s location.<br />

MACHINE DATA COLLECTION<br />

This section describes one or more alternative means <strong>of</strong> collecting the data required to achieve the<br />

project objectives.<br />

Utilisation and Productivity<br />

• Harvester utilisation can be measured using a vibration sensor or engine rpm/load to<br />

determine when the machine is active. Travel and work components <strong>of</strong> utilisation can be<br />

distinguished in a number <strong>of</strong> ways, including sensors to detect when the boom or harvester<br />

head is active, the travel alarm, speed (GPS), or connections to the travel pedal or brake<br />

circuit,. The only practical source <strong>of</strong> detailed information on log dimensions is the<br />

manufacturer supplied system. Simpler production measures include stem counts based on the<br />

harvester head tilt up switch and log counts based on number <strong>of</strong> main saw cuts or time spent<br />

by the harvester head travelling along the stem as an indication <strong>of</strong> stem length.<br />

• Forwarder productivity is measured by the amount <strong>of</strong> wood transported in a given time. To<br />

a large extent this is determined by the load capacity <strong>of</strong> the forwarder, piece and stack size and<br />

travel distance. Other factors include the operator’s loading and unloading efficiency and<br />

delays. Productivity can be estimated using a GPS to estimate travel distance and number <strong>of</strong><br />

loads or a load cell to estimate number <strong>of</strong> loads and total weight transported. The operator can<br />

also manually enter into the onboard computer when the machine is being loaded or unloaded.<br />

Forwarder utilisation can be determined with a vibration sensor or engine load. Utilisation<br />

can be broken down to work and travel, using a GPS or by monitoring engine load.<br />

• Skidder productivity (as with the forwarder) is measured by the amount <strong>of</strong> wood transported<br />

in a given time. However, it is more difficult to estimate the amount <strong>of</strong> wood a skidder<br />

transports because a skidder drags logs rather than carrying them. Estimates <strong>of</strong> productivity<br />

can be made using a GPS to estimate travel distance and number <strong>of</strong> loads or a strain gauge to<br />

estimate load weight. Utilisation can be determined with a vibration sensor or engine load.<br />

Utilisation can be broken down to work and travel with a GPS or by monitoring engine load.<br />

• Mobile chipper productivity is directly related to its utilisation, all else being equal. Spinelli<br />

and Visser (2009) reported that organisational delays, such as waiting for wood delivery and<br />

chip trucks, were the most significant factors affecting chipper utilisation. Operator and<br />

mechanical delays were also important. Travel time and operator skill are relatively minor<br />

issues for chipper utilisation and productivity.<br />

Utilisation can be determined using a vibration sensor or Canbus monitoring <strong>of</strong> engine load.<br />

Given the significance <strong>of</strong> delays to chipper utilisation and productivity, recording <strong>of</strong> delay<br />

causes is an essential requirement for a chipper. Entry <strong>of</strong> delay causes is currently only<br />

possible using manufacturer supplied equipment or a Multidat or RouteHawk.<br />

Fuel efficiency<br />

Fuel use can be measured using a flow meter installed in machine fuel lines or by recording refill<br />

quantities. The Canbus also reports fuel consumption but it is potentially inaccurate as the value is<br />

calculated rather than being measured directly.<br />

Fuel efficiency is strongly related to machine utilisation and productivity hence, unless a specific need<br />

is identified, it will be monitored in the project by recording machine refill quantities rather than by<br />

fitting sensors to machines.<br />

Machine availability<br />

Machine availability is monitored by recording machine utilisation and delay causes, in order to<br />

identify delays that are mechanical in nature and hence result in the machine being unavailable for


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 196<br />

work. Currently only manufacturer supplied equipment, Multidats and RouteHawks are capable <strong>of</strong><br />

recording delay causes to calculate and manage machine availability.<br />

Operator health, safety and job satisfaction<br />

Lost time injuries in forestry where harvesting is mainly mechanical are typically around 10-20 per<br />

million work hours. At this frequency it is unlikely that safety improvements could be monitored with<br />

the small sample in the trial. Therefore the trial will focus on means to reduce identified injury causes.<br />

Most injuries occur outside the cabin. Machine maintenance was identified by Nieuwenhuis and<br />

Lyons (2002) as a major source <strong>of</strong> injuries in Irish forest operations, so steps to identify and reduce<br />

delays resulting from machine breakdowns will have the additional benefit <strong>of</strong> reducing time lost<br />

through injury.<br />

Operating machines has created another range <strong>of</strong> health and safety issues. Stress and fatigue have<br />

been identified as important issues for machine operators, which can be alleviated by frequent short<br />

breaks (Kirk et al., 1997) that can be monitored by onboard equipment. Anecdotal evidence suggests<br />

that fitting onboard optimising computers has reduced stress in harvester operators as the computers<br />

now make most <strong>of</strong> the decisions as to what products to cut from each stem. Onboard GPS/GIS<br />

systems can improve safety by alerting operators to potential site hazards and by keeping machines<br />

within site boundaries.<br />

A major means <strong>of</strong> improving operator job satisfaction is to provide feedback on operator performance<br />

and rewards for improved performance. Several onboard computing options can provide operator<br />

feedback (manufacturer systems and Multidats with the Multipad option and RouteHawks). It may<br />

also be possible to achieve by adding screens to generic dataloggers.<br />

PRELIMINARY COSTS AND RETURNS<br />

Indicative purchase costs (<strong>Australia</strong>n dollars) for each unit used in the study are as follows:<br />

• DASA computer. A price was not available at the time <strong>of</strong> writing<br />

• Multidat. A new price list is being compiled with reduced prices. Each unit is expected to<br />

cost under $5,000 including optional GPS. The Garmin GIS display units are ~$900 each.<br />

• RouteHawk. Basic unit cost is $5,000. Use <strong>of</strong> the central database to store and distribute data<br />

will incur a small ongoing cost per machine. Mobile or satellite phone modem will add costs<br />

(up to ~$2,000 for a satellite phone) as well as call costs which will depend on the amount and<br />

frequency <strong>of</strong> data transferred.<br />

As all the units are imported, exchange rate movements will impact these prices.<br />

Installation and operating costs will depend on the options installed and the machine type. The most<br />

basic installation is to wire the unit to the machine’s power supply. Other options will include<br />

installation <strong>of</strong> GPS aerials, and switches or connections to existing switches. Operating costs will<br />

include data transfer and analysis and will depend on the degree to which this is automated. Manual<br />

data transfer may require purchase <strong>of</strong> a suitable PDA.<br />

At the time <strong>of</strong> writing, the equipment for the project had not been installed. However, the CRC for<br />

Forestry Program 3 has several other projects using Multidat onboard computers. In one project,<br />

~10% <strong>of</strong> potential productivity gains were identified through the use <strong>of</strong> the onboard technology. This<br />

represents an estimated annual saving ($/tonne) <strong>of</strong> over $40,000 per harvester.<br />

Details <strong>of</strong> the changes required to achieve the productivity gains will be made available initially to<br />

CRC for Forestry Program 3 members.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 197<br />

CONCLUSION<br />

Onboard computing technology has been shown in overseas studies to be able to significantly increase<br />

forest harvesting machine utilisation and productivity, and reduce maintenance and fuel costs. Studies<br />

performed in <strong>Australia</strong>n forest harvesting operations have indicated that there is the potential for these<br />

gains to be achieved in <strong>Australia</strong>.<br />

The initial stage <strong>of</strong> the project has been to identify available onboard computing technology and match<br />

it to industry partner requirements.<br />

• ForestrySA manage radiata pine plantations primarily for sawlog production and would be<br />

best suited with Dasa onboard computers as they can both meet ForestrySA’s requirement to<br />

increase value recovery as well as improving the harvesting contractor’s performance.<br />

• Timbercorp own their harvesting equipment and so are interested in the full gamut <strong>of</strong> project<br />

objectives. As they have a Multidat installation on a related project, their needs will best be<br />

met by trialling the RouteHawk telematics solution potentially including testing <strong>of</strong> Canbus<br />

connections and remote data collection techniques.<br />

• VicForests use contractors to harvest timber from Victoria’s native forests. The nature <strong>of</strong><br />

their harvesting operations is best suited to installing Multidats and Garmin GPS/GIS units on<br />

a range <strong>of</strong> equipment to obtain basic utilisation and productivity data and to examine delay<br />

causes and improve operator safety.<br />

REFERENCES<br />

Acuna, M., Wiedemann, J. and Strandgard, M. (2009). Evaluation <strong>of</strong> an in-field chipping operation in<br />

Western <strong>Australia</strong>. CRC for Forestry Program 3, Bulletin 4.<br />

Hartsough, B.R. and Cooper D.J. (1998) Cut-To-Length Harvesting <strong>of</strong> Short Rotation Eucalyptus at<br />

Simpson Tehama Fiber Farm. Proceedings <strong>of</strong> the Short-Rotation Woody Crops Operations<br />

Working Group, Second Conference, 25-27 August 1998, Vancouver, Washington, USA<br />

Jiroušek, R. Klvač, R. and Skoupý, A. (2007) Productivity and costs <strong>of</strong> the mechanised cut-to-length<br />

wood harvesting system in clear-felling operations. J. For. Sci. 53(10): 476-482<br />

Kirk, P.M., Byers, J.S., Parker, R.J. and Sullman, M.J. (1997) Mechanisation Developments Within<br />

the New Zealand Forest Industry: The Human Factors. Int. J. For Eng. Vol. 8(1)<br />

Nieuwenhuis, M. and Lyons, M. (2002). Health and Safety Issues and Perceptions <strong>of</strong> Forest<br />

Harvesting Contractors in Ireland Int. J. For Eng.Vol. 13(2)<br />

Ovaskainen, H., Uusitalo, J. and Väätäinen, K. (2004) Characteristics and Significance <strong>of</strong> a Harvester<br />

Operators' Working Technique in Thinnings Int. J. For Eng. Vol. 15(2)<br />

Ovaskainen, H. (2009). Timber harvester operators’ working technique in first thinning and the<br />

importance <strong>of</strong> cognitive abilities on work productivity. Dissertationes Forestales 79 University<br />

<strong>of</strong> Joensuu, Faculty <strong>of</strong> Forest Sciences<br />

Skogforsk (2007). Standard for Forest Data and Communications. Skogforsk.<br />

Spinelli, R and Visser R.J.M. (2008). Analyzing and estimating delays in harvesting operations. Int. J.<br />

For. Eng. 19(1)<br />

Spinelli, R and Visser R.J.M. (2009). Analyzing and estimating delays in wood chipping operations.<br />

Biomass and Bioenergy 33(3): 429-433


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 198<br />

TIMBER FROM NATIVE FOREST AND PLANTATION EUCALYPTS -<br />

USERS WILL QUICKLY FIND THAT THEY ARE<br />

NOT THE SAME THING<br />

Gregory Nolan 1<br />

ABSTRACT<br />

Solid eucalypt hardwood timber has largely shifted in <strong>Australia</strong> from a general<br />

construction material to primarily an architectural selection in building. Yet, this section<br />

<strong>of</strong> <strong>Australia</strong>’s construction industry has established preferences for relatively clear, stable,<br />

larger section timber recovered from older native forest eucalypt logs, traditionally the<br />

base resource for the Tasmanian solid hardwood production industry. Over recent<br />

decades, the transition from this resource to younger regrowth forced significant industry<br />

changes, but these are comparatively minor to the changes that are likely to occur as the<br />

industry’s base resource becomes smaller eucalypt plantation logs selected for their<br />

growth and not necessarily wood quality. This paper reports on recent work on the<br />

comparative properties <strong>of</strong> the mature, regrowth and plantation resource available for the<br />

Tasmanian solid hardwood production industry and discusses the likely implications their<br />

different properties for pr<strong>of</strong>itable processing, value adding and use.<br />

INTRODUCTION<br />

Forests provide society with a range <strong>of</strong> environment and economic services and products, and an array<br />

<strong>of</strong> human experiences and associations. This paper deals with just a few, focusing on the solid wood<br />

products drawn from native and plantation eucalypt forests, the associations these products have when<br />

used in artefacts such as buildings, and the potential for plantation eucalypts to maintain those<br />

associations and support (or maintain) a viable processing industry.<br />

Wood products can be as varied as the forests from which they are drawn. A large part <strong>of</strong> this variety<br />

results from the quality <strong>of</strong> the log and the log’s transformation during different production processes<br />

into increasingly reduced primary wood elements: first, solid wood products, then short grained<br />

elements such as chips or flakes; fiberal elements, and finally chemical elements (Marra 1972, p.202).<br />

These primary wood elements can be viewed as products ready for artefact-making or as the<br />

ingredients for second generation products such as plywood, laminated veneer lumber (LVL) and fibre<br />

boards, again widely used in artefact-making. Another large part <strong>of</strong> the variety <strong>of</strong> wood products<br />

results from the temporal, genetic, and site characteristics <strong>of</strong> the forests that yield the logs for<br />

processing. If Tasmania’s eucalypt log supply is taken as an example, logs for sawn boards can<br />

originate from one <strong>of</strong> a dozen tree species from 80 to hundreds <strong>of</strong> years old drawn from native forests<br />

or from 15 to 25 year old monocultural native plantation forests <strong>of</strong> Southern blue gum, E. globulus,<br />

and Shining gum, E. nitens. Each forest provides logs, and these logs can be converted into at least<br />

some part <strong>of</strong> the range <strong>of</strong> potential wood products. The best (largest, straightest, and cleanest) logs<br />

provide a resource for the widest range <strong>of</strong> products, including high quality solid timber boards. The<br />

worst (smallest, youngest or most crooked) logs provide a resource for a much more limited product<br />

range, usually only chip or fibre production.<br />

The vast bulk <strong>of</strong> wood products generated through this matrix <strong>of</strong> resource and process and used in<br />

artefact construction are general commodity products: structural and industrial grade timbers and sheet<br />

materials used where considerations <strong>of</strong> economy, strength or durability dominate their use. The unseen<br />

structural elements <strong>of</strong> buildings are an example <strong>of</strong> this. As pine products are consistently cheaper to<br />

produce than hardwood ones, they have come to dominate this sector over recent decades, largely<br />

replacing hardwood products. However, there is a group <strong>of</strong> applications where the importance <strong>of</strong> the<br />

1 Centre for Sustainable Architecture with Wood, School <strong>of</strong> Architecture and Design, University <strong>of</strong> Tasmania, Locked Bag<br />

1-324, Launceston, Tasmania 7250 +61 3 6324 4478, Email: Gregory.Nolan@utas.edu.au.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 199<br />

aesthetic character <strong>of</strong> the timber and wood’s special association with nature and human culture are<br />

relatively high and more expensive differentiated products are used. These are primarily appearance<br />

applications in buildings and furniture.<br />

Unable to compete against cheaper pine products, solid eucalypt hardwood timber has largely shifted<br />

from being a general construction material to become primarily an architectural selection in building.<br />

In <strong>Australia</strong>, most appearance wood products are hardwoods and those produced locally are drawn<br />

almost exclusively from native forests. In Tasmania in particular, the most valuable products are<br />

drawn from mature native forest logs. Over the last two decades, industry’s transition from processing<br />

this resource to younger native forest regrowth, coupled with the shift in market from structural to<br />

appearance products has forced significant industry changes but these are likely to be relatively minor<br />

compared to the changes that are and will occur as native forest logs become relatively rare and the<br />

base resource for industry moves from regrowth to smaller eucalypt plantation logs <strong>of</strong> species selected<br />

their growth characteristics and not necessarily wood quality.<br />

There are several factors to explore in this transition: the nature <strong>of</strong> the desire for particular<br />

characteristics in wood and the demand this generates; the characteristics found in the timber from<br />

native forest logs necessary for the appearance applications that will support a differentiate product;<br />

the combination <strong>of</strong> age, species, and growing conditions necessary to produce trees with those<br />

characteristics; the particular characteristics <strong>of</strong> the plantation resources due to come on stream in<br />

Tasmania from about 2020, and the effects this difference will have on the timber processing industry<br />

and designers.<br />

DESIRE AND DEMAND<br />

Wood products possess a culturally-embedded attraction or appeal far beyond their simple form or<br />

functionality and designers draw on this to enhance the artifacts they design or make. This attraction is<br />

not just based on an aesthetic reaction to the wood’s colour, grain and touch but includes wood’s<br />

important associations with nature, its approachability <strong>of</strong> form and assembly, and senses <strong>of</strong> continuing<br />

culture.<br />

Designers recognise wood as a medium for creating more friendly and approachable environments for<br />

living (Kock 1972, p.15). For renowned Finish architect, Alvar Aalto, wood was a primary means <strong>of</strong><br />

expressing nature’s impact upon the built environment and its capacity to maintain and even enrich<br />

life. Wood’s ‘biological characteristics, its limited heat conductivity, its kinship with man and living<br />

nature, the pleasant sensation to the touch it gives (Aalto 1956, p.142)’ made it a suitable material<br />

through which to design a sympathetic world. Menin and Flora, (2003, p.79) believe Aalto’s use <strong>of</strong><br />

wood ‘became a trope for the individuality <strong>of</strong> his buildings and indeed his mind, as well as for their<br />

yearning for something natural amidst the new forms and material <strong>of</strong> the modern epoch’. Earle (1972,<br />

p. 223) reflected on this kinship between man, nature and wood, similarly contrasting it to the<br />

characteristics <strong>of</strong> man-made materials.<br />

The infinite shape plasticity <strong>of</strong> modern technology's production, made <strong>of</strong> materials and<br />

forms that deny any self identification in their total commitment to their function, is not<br />

easily related to man's experience with living. Instead the knotty, splintery, irregularly<br />

dimensioned, warping, shrinking, swelling, directionally grained, uneven strengthened<br />

piece <strong>of</strong> wood expresses better man's grasp on the imperfect realities <strong>of</strong> life.<br />

Wood’s democracy <strong>of</strong> use (Nolan 1994) is also part <strong>of</strong> making more livable environments. Kock (1972<br />

p.16) believed that wood’s workability provides us with an opportunity to create for ourselves the<br />

objects <strong>of</strong> beauty and usefulness that we desire. That is, many people can make (and enjoy making)<br />

useful wooden things for themselves. By inference, those living in wood-rich environments are likely<br />

to have a better intuitive understanding <strong>of</strong> the origins and character <strong>of</strong> their surrounds and be more<br />

comfortable in them.<br />

Banhan (1972, p. 8) noted a more ingrained cultural association as ‘our hands become immemorially<br />

conditioned to the grasp <strong>of</strong> wood and it has acquired for us an ancient familiarity… The grasp <strong>of</strong> wood<br />

is probably as basic to human culture as the making <strong>of</strong> fire and the writing <strong>of</strong> language.’


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 200<br />

Figure 1. Unique high feature woods (Left to Right): figured burl, wavy grain, gum cluster,<br />

and blackheart figure<br />

Designers <strong>of</strong> buildings and other artefacts intuitively recognise these desirable associations and<br />

attempt to borrow and blend them with the more obvious aesthetic appeal <strong>of</strong> grain and texture in the<br />

items that they design and make. For this, they regularly exploit high-quality appearance woods, <strong>of</strong>ten<br />

with limited amounts <strong>of</strong> natural feature, or unique woods, where the extent <strong>of</strong> the natural feature<br />

establishes the material’s appeal. Shown in Figure 1, these unique woods can contribute to the most<br />

special applications.<br />

Wood included in any visible surface or structural element carries these associations but it is strongest<br />

in elements close to the eye, the hand and the touch. These close-at-hand objects, such as the pews<br />

shown in Figure 2, also place the most demanding requirements on the appeal <strong>of</strong> the wood. Elements<br />

in view but further away, such as architectural structures, have lesser demands on quality.<br />

Figure 2. Tasmanian oak cathedral pews Figure 3. Cathedral nave and wood panelling


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 201<br />

The key wood properties for appearance applications are colour, grain, feature, workability and critical<br />

aspects <strong>of</strong> functionality (such as board size, stability, hardness and durability). Ranking the importance<br />

<strong>of</strong> each <strong>of</strong> these properties for particular applications is generally subjective (due to the number <strong>of</strong><br />

variables included) but examining price in the market place can provide a reasonable objective ranking<br />

<strong>of</strong> the desirability <strong>of</strong> grade, dimension and colour for more general building design. Tables 1 to 3<br />

(from Nolan et al. 2005, p.7) show key relativities. Material graded Select under AS 2796 (Standards<br />

<strong>Australia</strong> 1999) was worth about 75% more per m 3 than structural material <strong>of</strong> the same size (Nolan et<br />

al. 2005, p.8). Select appearance material was worth significantly more than Standard grade and about<br />

twice as much as high feature material. A premium is paid for wide material. The additional price paid<br />

for increased thickness can probably be attributed to both increased desirability and production costs.<br />

Table 1. Relative prices <strong>of</strong> flooring milled form nominal 100 x 25 mm material<br />

Grade Select Standard High Feature<br />

Relative price /m 3 119% 100% 61%<br />

Table 2. Relative prices for Select boards nominally 25 mm thick<br />

Width (nom) 75 100 125 150<br />

Relative price /m 3 93% 100% 100% 117%<br />

Table 3. Relative prices for Select boards nominally 100 mm wide<br />

Thickness 25 38 50<br />

Relative price /m 3 80% 100% 114%<br />

In summary, there is a general pattern that appearance products are valued significantly more than<br />

structural ones, and wide, select (low feature) boards are valued far more than other appearance<br />

products. Designers wish to use these products more and their clients are willing to pay more for them.<br />

Like all general patterns however, there are notable exceptions. When a wood is particularly unique or<br />

the level <strong>of</strong> feature in the wood reaches a certain level <strong>of</strong> consistency and intensity (Nolan 1998), their<br />

value and appeal can also increase. They become unique woods, used in special applications.<br />

The <strong>Australia</strong>n, and especially the Tasmanian, hardwood production industry naturally recognized<br />

these realities. Largely unable to compete with cheaper sawn and engineered pine product in the<br />

structural markets, it generally focuses on maximizing recovery <strong>of</strong> appearance product, and <strong>of</strong> select<br />

material in particular. Logs from mature and large regrowth native forest tree have shown that they<br />

can supply these products in an economically sustainable way. They can also provide a supply <strong>of</strong><br />

special timbers.<br />

AGE AND SPECIES EFFECTS<br />

To successfully substitute for older material, a replacement resource will need to provide logs <strong>of</strong> a<br />

suitable wood quality and <strong>of</strong> a size that can be economically processed. In this, age and species are<br />

important. The age <strong>of</strong> the resources generally influences log size, the uniformity <strong>of</strong> the wood and the<br />

generation <strong>of</strong> special features. Species characteristics also significantly influence wood quality.<br />

As shown in Figure 4, Washusen and Clark (reported in Nolan et al. 2005) established that the<br />

recovery <strong>of</strong> select grade board is strongly related to log size. This is due to a range <strong>of</strong> factors including<br />

geometric characteristics, the percentage <strong>of</strong> lyctus-susceptible sapwood in the log and growth stress.<br />

However, as shown in Figure 5, the plantation logs to be provided to the Tasmanian industry are<br />

forecast to be considerably smaller on average than those from native forest. As a result, the<br />

proportion <strong>of</strong> Select material recovered is likely to decrease.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 202<br />

Recovery <strong>of</strong> select grade<br />

(% <strong>of</strong> log volume)<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

r = 0.68<br />

0<br />

15 20 25 30 35 40 45 50 55 60 65<br />

Log small end diameter (cm)<br />

Figure 4. Plot <strong>of</strong> log diameter and recovery for<br />

pruned eucalypts milled with std. industry<br />

practice (Washusen and Clark, in prep)<br />

Figure 5. Log size distribution <strong>of</strong> “high<br />

quality” sawlogs from hardwood plantation<br />

and native forest (reproduced from FT 2007).<br />

Age also affects the uniformity <strong>of</strong> the wood. A tree grows by depositing a new layer <strong>of</strong> wood, under<br />

the bark, on the outside <strong>of</strong> the previous year's deposited wood. Freshly deposited wood differs from<br />

the wood deposited in the previous year’s growth, up until the tree reaches an age nominally between<br />

20 and 40 years old, after which the characteristics <strong>of</strong> the deposited wood are relatively consistent.<br />

Figure 6 is a general diagram <strong>of</strong> the relationship <strong>of</strong> wood density to age. As a result, as shown in<br />

Figure 7, a quarter sawn board cut from a 25- year-old plantation-grown log, cut from the zone <strong>of</strong><br />

wood property change (in yellow), is likely to have much greater variability in wood properties across<br />

the board than a board cut from the zone <strong>of</strong> unchanging wood properties (brown) from a mature log<br />

from a natural stand. The material currently supplied to Tasmanian hardwood producers probably has<br />

an average age <strong>of</strong> 90 years, while the plantation material to be supplied has largely been planted since<br />

1995 and will be about 25 years old when harvested.<br />

Figure 6. Generalised pith-to-bark variation in<br />

average basic density (Nolan et al. 2005, p.72)<br />

Figure 7. Boards cut position from 110 year<br />

and 25 year logs (Nolan et al. 2005, p.72)<br />

Age affects generate many <strong>of</strong> the most desirable characteristics <strong>of</strong> special timbers. Fiddle back and<br />

wavy grain form as the mass <strong>of</strong> the standing tree compresses and buckles the grain at the tree’s base..<br />

Burl is formed through the repeated healing <strong>of</strong> a wound. Blackheart is from long fungal infestations.<br />

Few if any <strong>of</strong> these occur in younger regrowth trees and it is unlikely any will occur in plantations.<br />

Species characteristics affect wood quality. In southern <strong>Australia</strong>, plantation eucalypt are generally<br />

one <strong>of</strong> two species: Southern (or Tasmanian) blue gum, E. globulus, planted in areas not susceptible to<br />

frost and Shining gum, E. nitens, planted in areas that are frost susceptible. In 2004, about 60% <strong>of</strong><br />

<strong>Australia</strong>’s plantation resource was E. globulus. As most <strong>of</strong> Tasmania’s plantations are occasionally<br />

susceptible to frost, E. nitens makes up the majority <strong>of</strong> material planted in that state to supply an<br />

ongoing sawn hardwood industry with logs. However, these species were generally selected for their<br />

growth characteristics and general pulping properties, not their properties as solid wood products.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 203<br />

TASMANIA’S YOUNG EUCALYPT AND PLANTATION RESOURCE<br />

By 2020, about 50% <strong>of</strong> Tasmania’s high quality sawlog resource will be drawn from a plantation<br />

eucalypt resource. Young regrowth will make up some proportion <strong>of</strong> the remainder. However, recent<br />

studies indicate that recovery <strong>of</strong> appearance grade products from young regrowth and plantation<br />

material in Tasmania is likely to be significantly less than from mature and older regrowth material.<br />

Reduced log size is only one reason for this. Increased drying degrade such as distortion and internal<br />

check is likely to lead to loss <strong>of</strong> grade and pr<strong>of</strong>itability and for the main Tasmanian plantation sawlog<br />

species, thinned and pruned E. nitens, potentially even complete loss <strong>of</strong> economic viability.<br />

After milling regrowth messmate, E. obliqua, from 1901, 1934, 1949 and 1967, Innes et al. (2005)<br />

reported a range <strong>of</strong> age and site effects, including:<br />

• Younger messmate logs had a lower heartwood proportion than older logs.<br />

• The youngest messmate (1967 regrowth) underwent significant internal checking and its<br />

properties were generally more variable.<br />

• Boards cut from younger material shrank less than older, but yield <strong>of</strong> select grade was<br />

substantially lower due to gum vein.<br />

• It appears that the oldest logs had the highest growth stress, as those logs had the most end<br />

splitting and boards cut from them the most spring.<br />

• There was no trend with age <strong>of</strong> basic density, initial moisture content, drying rate, strength or<br />

hardness.<br />

Recovery <strong>of</strong> boards graded Select to AS 1684 on both sides and <strong>of</strong> boards subject to internal check is<br />

included in Table 4.<br />

Table 4. E. obliqua - % <strong>of</strong> boards Select both sides and % <strong>of</strong> boards internally checked<br />

Year % Select both sides % Internally checked<br />

1901 92 0<br />

1934 78 4<br />

1949 44 0<br />

1967 20 15<br />

In 2005, a major study sought to assess the economics <strong>of</strong> processing the major southern <strong>Australia</strong>n<br />

species <strong>of</strong> plantation eucalypts <strong>of</strong> known origin using current industry-standard equipment and<br />

procedures (Innes et al. 2008). It also sought to identify the factors most directly affecting the value <strong>of</strong><br />

dry output given current market conditions.<br />

Table 5. Log classes<br />

species E.globulus E.globulus E nitens E. nitens E.globulus<br />

age 47 13 26 19<br />

silviculture fibre wide spaced, pruned thinned, pruned fibre thinned, pruned<br />

location Gippsland Gippsland Ridgley Ulverstone<br />

forest owner GRP Frank Hirst Gunns Robyn May<br />

log description run-<strong>of</strong>-bush pruned butt pruned<br />

butt<br />

unpruned<br />

upper<br />

run-<strong>of</strong>bush<br />

pruned butt<br />

As detailed in Table 5, two species, E. nitens and E. globulus, harvested from four sites in two States<br />

were milled across three operational hardwood mills, ITC Timber’s small regrowth log mill at<br />

Newood in Tasmania and two conventional mills, ITC Timber Heyfield in Victoria and Gunns<br />

Lindsay Street in Tasmania. A total <strong>of</strong> 12,778 sawn boards were produced. This material was dried in<br />

a combination <strong>of</strong> conventional drying yards (ITC Heyfield and ITC Mowbray) or recently<br />

commissioned predryers (Gunns Lindsay Street). Some boards were milled into final products and all


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 204<br />

boards were graded and tallied. Value reducing defects were assessed. Market values were estimated<br />

and applied to the recovered material.<br />

Of the logs milled in Tasmania, sawn recovery was dependent on the milling equipment available<br />

while total high-grade recovery was low. Total dry recovery from the newer ITC Newood mill was<br />

significantly higher from the same resource than the conventional mills. ITC Newood achieved a dry<br />

recovery <strong>of</strong> 40% from 26-year-old thinned-pruned E. nitens butt logs. Gunns Lindsay Street recovered<br />

34%. For E. globulus butt logs, recoveries were 37% and 29% respectively.<br />

dried board recovery (nominal m3<br />

board/m3 log)<br />

45%<br />

40%<br />

35%<br />

30%<br />

25%<br />

20%<br />

15%<br />

10%<br />

5%<br />

0%<br />

13% 10% 11% 4%<br />

4%<br />

4%<br />

5%<br />

21% 22% 20%<br />

4%<br />

25%<br />

0% 0% 0%<br />

0%<br />

1 2 3 4 5 6<br />

log grade<br />

HF/merch standard/MF select<br />

Figure 8. 26-year-old E. nitens thinned-pruned butt recovery- ITC<br />

dried board recovery (nominal m3<br />

board/m3 log)<br />

45%<br />

40%<br />

35%<br />

30%<br />

25%<br />

20%<br />

15%<br />

10%<br />

5%<br />

0%<br />

6%<br />

3%<br />

6%<br />

18% 17%<br />

0%<br />

2%<br />

3% 4%<br />

26%<br />

3%<br />

17% 17%<br />

3%<br />

2% 5%<br />

1 2 3 4 5 6<br />

log grade<br />

HF/merch standard/MF select<br />

Figure 9. 26-year-old E. nitens thinned-pruned butt recovery – Gunns<br />

Assessment <strong>of</strong> recovered boards milled into final products showed that:<br />

• Boards from E. nitens butt logs (both thinned-pruned and fibre managed) showed higher levels<br />

<strong>of</strong> surface check as the primary reason for downgrade (25-30%).<br />

• There was an unexpected high loss in recovery when skim-dressed boards were milled into<br />

final products, especially <strong>of</strong> E. globulus. 12% <strong>of</strong> sawn volume from the thinned-pruned E.<br />

nitens butt logs and 30% <strong>of</strong> sawn volume from the thinned-pruned E. globulus butt logs was<br />

rejected. This was mainly due to distortion related effects.<br />

• Boards cut from E. globulus thinned-pruned butt logs show higher level <strong>of</strong> “gum features” as<br />

the primary reason for downgrade.<br />

0%<br />

13%


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 205<br />

100%<br />

90%<br />

80%<br />

70%<br />

60%<br />

50%<br />

40%<br />

30%<br />

20%<br />

10%<br />

0%<br />

Gunns ENIT TP butt - 100mm - all grades<br />

Gunns ENIT UTUP - 125mm - all grades<br />

Gunns ENIT TUP tops - 125mm - all grades<br />

Gunns EGLOB TP butt - 125mm - all grades<br />

NST EGLOB TP butt - all sizes, all grades<br />

NST ENIT TP butt - all sizes, all grades<br />

no check minimal check moderate check heavy check<br />

Figure 10. Levels <strong>of</strong> internal check by log type.<br />

A high level <strong>of</strong> internal check was also found in the Tasmanian E. nitens butt logs harvested from both<br />

the fibre-managed and thinned and pruned stands. As shown in Figure 10, 25-31% displayed minimal<br />

check while between 6-17% showed moderate/heavy check. The check classed as minimal was still<br />

sufficient to be a serious defect and be grade limiting. The rate <strong>of</strong> checking was consistent for air-dried<br />

and pre-dried material, the two major hardwood drying technologies. Internal check in the E. nitens<br />

top logs and the E. globulus was significantly less.<br />

The high level <strong>of</strong> internal check observed was (and remains) <strong>of</strong> serious concern as the studied logs<br />

were from an expensive silvicultural regime, similar to the regime adopted by major growers<br />

managing E. nitens plantations for clear wood log production. The results indicate that between 15 and<br />

40% <strong>of</strong> boards from similar thinned and pruned full-rotation E. nitens can be expected to contain<br />

significant levels <strong>of</strong> internal check if dried with current industry technology. The situation could be<br />

even worse in full-scale production if 5.4 m logs were milled into long length boards. As stated above,<br />

there was an unexpected high loss in recovery when skim-dressed boards were milled into final<br />

products with dimensional affects making the boards unacceptable at any grade under the relevant<br />

<strong>Australia</strong>n standard. Further, quartersawn boards <strong>of</strong> both species cut at normal full length (5.4 – 6.0 m)<br />

can be expected to incur significant downgrade from spring distortion. These combined results cast<br />

doubt on assumptions <strong>of</strong> board and value recovery from other studies based on the assessment <strong>of</strong><br />

skim-dressed material only.<br />

product value (/m3 log)<br />

$400<br />

$350<br />

$300<br />

$250<br />

$200<br />

$150<br />

$100<br />

$50<br />

$0<br />

log grade 1 log grade 2 log grade 3 log grade 4 log grade 5 log grade 6<br />

sawlog grade<br />

E. nitens 26yo - thinned, pruned<br />

butt log - ITC Newood<br />

E. nitens 26yo - thinned, pruned<br />

upper log (unpruned) - ITC<br />

Newood<br />

E. nitens 26yo - fibre - ITC<br />

Newood<br />

,E. globulus 19yo - thinned<br />

pruned butt log - ITC Newood<br />

Figure 11. Total product value per cubic metre <strong>of</strong> sawlog at ITC Newood


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 206<br />

The economics analysis for the study found that the value <strong>of</strong> solid products recovered from the thinned<br />

and pruned E. globulus logs and probably logs salvaged from the fibre-managed E. nitens stands were<br />

likely to be sufficient for a suitable and sustainable return. See Figure 11. However, growing and<br />

processing the thinned and pruned E. nitens was marginal or uneconomic, mainly due to the loss <strong>of</strong><br />

value from internal checking, especially in timber from the pruned butt log.<br />

In a further study, 21-year-old trees, pruned at age six years to a height <strong>of</strong> 6.4 m, were selected from a<br />

silvicultural trial in north-east Tasmania which tested five stockings (100, 200, 300, 400 stems per<br />

hectare (SPH) and un-thinned control with ~700 SPH (Washusen et al. 2007). The logs were milled<br />

and the recovered boards were dried in a research predryer before final drying in a commercial kiln.<br />

On assessment, internal checking was found to be a common defect in the boards and was<br />

significantly more prevalent in boards from the lower logs than from the upper log. Frequency <strong>of</strong><br />

occurrence did not differ significantly between the two sawing methods (see Table 6). However, the<br />

mean numbers <strong>of</strong> internal checks per board was significantly greater for back-sawn boards than<br />

quarter-sawn boards, as well as there being significantly more checks in the boards from the lower<br />

logs than the upper logs for both sawing categories.<br />

Table 6. Percentage <strong>of</strong> boards displaying internal checking<br />

Lower log Upper log<br />

Back-sawn 73.3 34.3<br />

Quarter-sawn 73.1 23.1<br />

CONCLUSIONS<br />

This paper focuses on the solid wood products drawn from native and plantation eucalypt forests, the<br />

associations these products have when used in artefacts such as buildings, and the potential for<br />

plantation eucalypts to maintain those associations and support (or maintain) a viable processing<br />

industry.<br />

It links together the supply chain for these solid hardwood products in <strong>Australia</strong>: from the user seeking<br />

to enjoy an artefact; the producer suppling the material for the artefact, and the grower providing the<br />

raw material. Because <strong>of</strong> its visual qualities and its inherent inability to compete in more competitive<br />

markets, solid hardwood is used primarily as an appearance product in this country. It demands a high<br />

price in the market because <strong>of</strong> its innate attractions in building and especially in applications close to<br />

the eye, the hand and the touch. The key wood properties for these applications, established during<br />

centuries <strong>of</strong> practice, are colour, grain, feature, workability and critical aspects <strong>of</strong> functionality.<br />

The industry milling and drying <strong>of</strong> these products depends on the premium they demand over other<br />

wood products, as the cost <strong>of</strong> producing hardwood is systemically more expensive than milling<br />

s<strong>of</strong>twood. To mill these products, the industry requires a resource, and in Tasmania, 50% <strong>of</strong> the high<br />

quality log supply to be provided to industry from state forests from 2020 is predominantly plantation<br />

E. nitens.<br />

Unfortunately, given current technology, this resource is not likely to be suitable for the applications<br />

on which the current hardwood production industry depends. The boards either distort too much to be<br />

acceptable for high quality appearance products, or suffer from an unacceptable and unrecoverable<br />

drying degrade — that <strong>of</strong> internal check. These large hurdles overwhelm consideration <strong>of</strong> more subtle<br />

price and recovery influences, such as colour, stability, hardness and log size.<br />

This mismatch <strong>of</strong> resource and market is likely to have several effects. <strong>Australia</strong>’s designers (or more<br />

correctly, suppliers acting on their behalf) will import hardwood from overseas to overcome the lack<br />

<strong>of</strong> a suitable local product. The supply <strong>of</strong> unique wood from Tasmania’s forest, already significantly<br />

constrained, will decrease further. Lacking a resource, a proportion <strong>of</strong> hardwood producers will cease<br />

to trade. Others will continue to operate with the remaining native resource and some proportion <strong>of</strong><br />

recovered plantation wood. Some will attempt to compete in the structural market but they will need to<br />

pursue very selective production strategies to overcome the added complications and associated costs


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 207<br />

<strong>of</strong> hardwood production. The grower will be left to try to recover some value from the standing<br />

resource. However, the economics <strong>of</strong> this will be interesting, given the expensive silvicultural regimes<br />

pursued to produce the logs in the first instance.<br />

There is potentially a technological solution to processing the resource currently in the ground.<br />

However, current indications are that this solution will require a radical rather than incremental change<br />

to current production techniques. Recent studies have generally failed to reveal any statistically<br />

relevant differences to exploit in results from variations on current approaches. Unfortunately, it<br />

appears that E. nitens wants to check, especially in timber from the butt log, and E. globulus wants to<br />

distort.<br />

REFERENCES<br />

Aalto, A 1956, ‘Wood as a building material’, Arkkitehti-Arkkitekten in Schildt, Sketches.<br />

Banhan, R 1972, ‘Is there a substitute for wood grain plastic’ in Anderson EA & and Earle GF (ed.), Design and<br />

aesthetics in wood, State University <strong>of</strong> New York Press, Albany.<br />

Earle, GF 1972, ‘Is there a substitute for wood grain plastic’ in Anderson EA & and Earle GF (ed.), Design and<br />

aesthetics in wood, State University <strong>of</strong> New York Press, Albany.<br />

Forestry Tasmania 2007, Sustainable High Quality Eucalypt Sawlog Supply from Tasmanian State Forest:<br />

Review No. 3, Forestry Tasmania Planning Branch, August.<br />

Innes, T and Greaves, BL and Nolan, G & Washusen, R 2008, Determining the economics <strong>of</strong> processing<br />

plantation eucalypts for solid timber products, Forest & Wood Products Research & Development<br />

Corporation, PN04.3007<br />

Innes, T Armstrong, M & Siemon G 2005, The impact <strong>of</strong> harvesting age on sawing, drying and solid wood<br />

properties <strong>of</strong> key regrowth eucalypt species, Forest & Wood Products Research & Development<br />

Corporation, PN03.1316<br />

Kock 1972, ‘Design and the Product in the Future Economy’ in Anderson EA & and Earle GF (ed.), Design and<br />

aesthetics in wood, State University <strong>of</strong> New York Press, Albany.<br />

Marra 1972, ‘Wood Products in the Future – A Technological Extrapolation’ in Anderson EA & and Earle GF<br />

(ed.), Design and aesthetics in wood, State University <strong>of</strong> New York Press, Albany.<br />

Menin, S., Flora, S 2003, Nature and Space: Aalto and Corbusier, Routledge<br />

Nolan, G Greaves, BL Washusen, R & Parson M 2005, Plantations for Solid Wood Products in <strong>Australia</strong> – A<br />

Review, Forest & Wood Products Research & Development Corporation, PN04.3002<br />

Nolan, GB 1994, The Culture <strong>of</strong> using Timber as a Building Material in <strong>Australia</strong>, Proceedings, the Pacific<br />

Timber Engineering Conference, Surfer's Paradise, July<br />

Nolan, GB 1998, ‘Specifying natural feature timber’, 1st International Furniture Technology Conference<br />

Proceedings, 1998, Sydney, pp. 9.<br />

Standards <strong>Australia</strong> 1999, <strong>Australia</strong>n Standard AS2796: Timber – Hardwood – Sawn and Milled Products,<br />

Standards <strong>Australia</strong>, Sydney.<br />

Washusen, R Harwood C, Morrow A, Valencia JC, Volker P, Wood M, Innes T, Ngo D, Northway R &<br />

Bojadzic M 2007 Technical report 168: Gould’s Country Eucalyptus nitens thinning trial: solid wood<br />

quality and processing performance using conventional processing strategies, CRC Forestry, Hobart.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 208<br />

WOOD PROPERTY VARIATION IN MATURE QUEENSLAND<br />

EXOTIC PINE SILVICULTURE EXPERIMENTS<br />

Kevin Harding 1 , Terry Copley 1 , Marks Nester 2 , Kerrie Catchpoole 2 & Anton Zbonak 1<br />

ABSTRACT<br />

Three mature silvicultural experiments that produced large differences in growth rate<br />

responses were sampled to investigate the impact <strong>of</strong> these treatment responses on wood<br />

properties among and within trees. These experiments were a Caribbean pine<br />

cultivation and weed control experiment (26-years-old), an exotic pine mounding by<br />

taxa experiment (21-years-old) and a Caribbean pine spacing by thinning experiment<br />

(28-years-old). Results for wood density varied in consistency but overall displayed a<br />

small negative trend with growth rate. Micr<strong>of</strong>ibril angle tended to increase with faster<br />

growth rate but across the trials increases were small and produced inconsistent trends.<br />

Overall the responses in these wood properties were small compared to the large<br />

DBHOB differences evident in the experiments. Taxa differences studied in one<br />

experiment displayed a negative relationship between average DBHOB and average<br />

wood density ranking from highest to lowest: slash pine seedlings, F2 hybrid seedlings,<br />

F1 hybrid cuttings and Caribbean pine seedlings.<br />

INTRODUCTION<br />

In 2005, Forestry Plantations Queensland (FPQ) and the Wood Quality Improvement group within the<br />

Product Quality section <strong>of</strong> Horticulture and Forestry Science (H&FS), DPI&F, commenced a long<br />

term project to obtain samples from three mature silviculture/genetics trials to collect data on tree<br />

growth and wood quality. The main goal was to collect robust data from the three trials that would<br />

underpin wood property and growth modeling as part <strong>of</strong> an overall decision support system being<br />

developed by FPQ (Catchpoole et al. 2007). By modeling growth rate and wood property interactions<br />

produced by the silvicultural treatments present in these trials it should be possible to identify an<br />

optimised silvicultural regime to deliver the best structural grade recovery and value returns able to be<br />

produced from exotic pine plantings in Queensland.<br />

The data collection focused on within and between-tree variation to consider the impacts <strong>of</strong> different<br />

silvicultural factors on wood density, micr<strong>of</strong>ibril angle and juvenile wood proportion, both at breast<br />

height, and at multiple heights within the merchantable stem. The sampled experiments (132BOW,<br />

195MBR and 532NC) provided opportunities to examine the effect <strong>of</strong> site preparation and weed<br />

control, mounding and taxon , and initial spacing and on wood density, micr<strong>of</strong>ibril angle and juvenile<br />

wood proportion. The experiments were targeted due to their maturity (20 to 28 years old), their<br />

design and their observed large differences in growth rates in response to the treatments applied. This<br />

variation provided excellent scope to evaluate and model the relationship between large productivity<br />

responses and wood traits.<br />

The first experiment, Expt 132BOW, is a Caribbean pine cultivation and weed control trial planted in<br />

1980 and wood sampled at age 26 years. Expt 195MBR is an exotic pine mounding by taxa trial,<br />

comparing slash pine, Caribbean pine, F1 hybrid cuttings and F2 hybrid seedlings, planted in 1985 and<br />

wood sampled at age 21 years. Expt 532NC is a Caribbean pine spacing and thinning trial planted in<br />

1977, wood sampled at age 28-29 years, and used in a sawing study at age 30 years.<br />

The use <strong>of</strong> these designed and controlled experiments to study silviculture and growth rate effects on<br />

wood quality is much more cost effective than attempting to model these responses in routine<br />

1<br />

Horticulture and Forestry Science, Department <strong>of</strong> Primary Industries and Fisheries, Queensland<br />

2<br />

Forestry Plantations Queensland


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 209<br />

plantation stands. The confounding influence <strong>of</strong> site (both within and between site differences) age<br />

and spacing/thinning/fertiliser history means much larger samples <strong>of</strong> trees on many more sites would<br />

need to be studied to attempt to define the significance <strong>of</strong> responses. The latter approach very quickly<br />

becomes prohibitively costly and lacks the scientific rigour needed to produce repeatable results.<br />

MATERIAL<br />

Silviculture experiments<br />

Table 1. Relevant experiment details for the selected treatments sampled in three experiments<br />

grown in coastal southern and central Queensland.<br />

Experiment 132BOW 195MBR<br />

PEE seedlings, PEE ×<br />

532NC<br />

Species†<br />

and stock type<br />

PCH seedlings<br />

PCH F1 cuttings, PCH<br />

seedlings and PEE × PCH<br />

F2 seedlings<br />

PCH seedlings. Kennedy<br />

seed orchard batch T251<br />

Location<br />

Byfield<br />

22˚ 48’ S, 150˚ 39’ E<br />

Tuan<br />

25˚ 35’ S, 152˚ 45’ E<br />

Toorbul<br />

27°05'S 152°58'E<br />

Planted 1/1980 at 3.0 × 3.0 m 6/1985 at 5.0m × 2.1m 2/1977<br />

Date sampled -<br />

wood properties<br />

5/2006 3/2006 6/2005<br />

Spacing/stocking;<br />

Treatments Ploughing; Weed Control Mounding<br />

Thinning (at age 10<br />

years)<br />

Ploughing … Nil;<br />

Ploughed<br />

Weed control … Nil,<br />

Treatment levels woody weeds on inter- Nil, 0.5 m<br />

rows only, all woody<br />

weeds, and total weed<br />

control<br />

2<br />

Spacing (stocking)<br />

…1.8m 2 ( 3088spha),<br />

2.4m 2 (1737spha), 3.0m 2<br />

(1111spha) and 3.6m 2<br />

(772 spha)<br />

Thin … Nil, 200, 400<br />

and 600 spha<br />

† PCH = Pinus caribaea var. hondurensis; PEE = Pinus elliottii var. elliottii; PEE × PCH = interspecific hybrid cross<br />

METHODS<br />

Standing tree assessments<br />

In each experiment sample trees were selected to represent the diameter at breast height over bark<br />

(DBHOB) distribution in the treatment combinations studied. Each tree was assessed for standing tree<br />

acoustic velocity, height, and DBHOB and a 12mm core was extracted. Two acoustic velocity<br />

measurements (ST300) were obtained on opposite sides <strong>of</strong> each tree in approximately the same<br />

alignment as the increment core sample was taken; i.e. along the shortest diameter to minimise<br />

inclusion <strong>of</strong> compression wood. The wood cores were removed from the closest mid-whorl below<br />

breast height. All wood cores were resin extracted in methanol and hot water in a soxhlet apparatus<br />

using a boiling cycle <strong>of</strong> 8 hours in methanol, followed by 10 hours in hot water.<br />

In Expt 532NC approximately 105 trees per thinning treatment were assessed providing 403 trees in<br />

total due to incomplete survival or unacceptable tree form, size or damage. Numbers sampled varied<br />

across initial spacing × thinning treatment combinations as surviving tree numbers were small in<br />

widely spaced and heavily thinned plots and were large in unthinned plots. In Expt 195MBR 20 trees<br />

were sampled for each taxon × mounding combination (160 trees in total). Fifteen trees per weed<br />

control treatment plus the nil plough × weed control treatment used as a control were sampled in two<br />

replications <strong>of</strong> Expt 132BOW (150 trees in total).<br />

SilviScan<br />

Based on the DBHOB and gravimetric density results, 80 trees (5 trees per treatment) from 532NC<br />

and 40 trees (10 trees per taxon) from 195NC were systematically selected for felling and analysis <strong>of</strong><br />

wood properties with height within the stem. For this study only results for the breast height samples<br />

have been reported. The best defect-free radius <strong>of</strong> each core was used for SilviScan analysis. X-ray


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 210<br />

scanning was performed to capture density readings at every 50 microns from pith to cambium along<br />

each radial sample. The positioning <strong>of</strong> growth ring boundaries were assigned by comparing SilviScan<br />

pith to bark density patterns to standard pr<strong>of</strong>iles developed using a selection <strong>of</strong> samples with<br />

prominent and readily discernable growth ring patterns, as well as reference to actual samples under<br />

magnification. This combined approach was used to minimise the erroneous inclusion <strong>of</strong> false growth<br />

rings in the individual growth ring data. Radial increments to each growth ring boundary were used to<br />

estimate the basal area <strong>of</strong> each growth ring and to provide basal area weighted averages <strong>of</strong> density<br />

results. The controlled environment used for x-ray scanning equates to approximately 8% moisture<br />

content in the scanned samples so SilviScan density refers to density assessed at this moisture content.<br />

Analysis<br />

Analysis <strong>of</strong> variance was undertaken using GenStat (2002).<br />

RESULTS<br />

Expt 132BOW<br />

Ploughing was found by Debuse and Nester (2004) to have had a significant effect on the impact <strong>of</strong><br />

weed control on mean adjusted DBHOB at age 18.1 years (weed control x ploughing: F 3,20 = 3.2, P =<br />

0.045). At age 18.1 years, complete weed control in combination with ploughing yielded significantly<br />

greater adjusted DBHOB measures than all other treatments, with the exception <strong>of</strong> ‘nil plough +<br />

complete weed control’, and ‘plough + complete woody weed control’ treatments. These DBHOB<br />

results for the full set <strong>of</strong> experiment treatments indicated large differences in growth rate between the<br />

control treatment (nil plough and nil weed control) and the four ploughed weed control treatments that<br />

were sampled for this wood study.<br />

Analysis <strong>of</strong> variance revealed no significant difference among sampled ploughing × weed control<br />

treatments for area-weighted SilviScan means <strong>of</strong> wood density (F4,145=1.05, P= 0.385) and micr<strong>of</strong>ibril<br />

angle (F4,145=1.57,P=0.186) estimated on breast height radial core samples. Wood density displayed a<br />

consistent negative trend with more intensive weed control treatments which reflect growth rate<br />

improvements. Micr<strong>of</strong>ibril angle increased with better weed control from around 13 degrees in the<br />

control to about 14.6 degrees in the treatments with more intensive weed control. Trends for these<br />

traits with relative distance from the pith are plotted in Figure 1.<br />

Wood density (kg m -3 )<br />

750<br />

700<br />

650<br />

600<br />

550<br />

500<br />

450<br />

400<br />

350<br />

Nil plough-Nil weed Plough-1 nil weed Plough-2 Partial woody weed<br />

Plough-3 Complete woody weed Plough-4 Complete weed control<br />

MfA (deg)<br />

35<br />

30<br />

25<br />

20<br />

15<br />

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

Wood density MfA<br />

Relative distance from pith (0%) to bark (100%)<br />

Figure 1. Wood density and micr<strong>of</strong>ibril angle variation among Expt 132BOW plough × weed<br />

control treatments relative to proportional distance from pith. Weed control<br />

treatments: 1=nil; 2=woody inter-row weeds; 3= all woody weeds and 4 = complete.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 211<br />

Treatment (Number) Nil ploughing Ploughing<br />

Nil weeding<br />

(1)<br />

Partial woody weed<br />

removal and inter-row<br />

cultivation<br />

(2)<br />

Complete woody weed<br />

removal, inter-row<br />

cultivation, and basal<br />

spray<br />

(3)<br />

Complete weed control -<br />

grass and woody weeds<br />

(4)<br />

Wood density zones:<br />

400 – 500 kg/m 3<br />

500 – 600 kg/m 3 600 – 700 kg/m 3<br />

Figure 2. Graphic illustration <strong>of</strong> the average relative proportion <strong>of</strong> wood density zones within<br />

Expt 132BOW treatments. Tree cross-sections have been scaled to indicate relative<br />

mean size <strong>of</strong> trees in the treatments.<br />

Using the actual growth ring increment values from SilviScan results, Figure 2 was constructed to<br />

relative scale to illustrate the differences between treatments in both average tree size and proportions<br />

<strong>of</strong> wood density zones within each.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 212<br />

Although arbitrarily assigned, the three wood density zones defined approximate juvenile,<br />

intermediate and mature wood values. Although the incremental increase in tree size is very small<br />

from weed control treatment 2 to treatment 3, the proportion <strong>of</strong> wood in the lowest density zone varies<br />

markedly. Treatment 3 has resulted in a lower density inner core than other treatments, which is likely<br />

to have negative implications for the grade recovery <strong>of</strong> sawn framing from such a regime.<br />

As significant amounts <strong>of</strong> the higher density mature wood zone are lost in squaring up the log during<br />

sawing, the strength and stiffness properties <strong>of</strong> timber framing sawn from trees displaying these<br />

differences are likely to vary considerably. Ongoing analysis and data modelling <strong>of</strong> these relationships<br />

will be undertaken for the decision support system development that these studies support. The latter<br />

will need to consider the potential to gain economic returns from optimising silviculture regimes to<br />

balance average plantation stem volume production and the quality and pr<strong>of</strong>ile <strong>of</strong> the wood produced.<br />

The density pr<strong>of</strong>iles observed within this experiment suggest that there may be considerable scope for<br />

optimisation.<br />

Expt 195MBR<br />

This experiment compared three mound sizes (0.5m 2 , 1.0m 2 and 1.5m 2 ), which Debuse and Nester<br />

(2002) found had a significant effect on DBHOB growth for all taxa combined at age 16 years (F 3,5 =<br />

9.2, P = 0.018). Trees planted on all sizes <strong>of</strong> mound showed significantly greater DBHOB than those<br />

planted on no mounds. However, this mounding effect did not vary significantly among taxa (F 12,31 =<br />

0.5, P = 0.898) and therefore only the 0.5m 2 mounds were sampled in the wood study as these were<br />

closest in size to the current commercial practice. DBHOB measures differed significantly among taxa<br />

for all mound sizes combined at age 16 years (F 4,31 = 86.2, P < 0.001). PCH exhibited significantly<br />

larger DBHOB measures than all other taxa, while PEE had significantly smaller diameters than the<br />

other taxa (Debuse and Nester, 2002).<br />

Analysis <strong>of</strong> variance revealed a significant difference among taxa for breast height area-weighted<br />

SilviScan mean wood density (F3,152=45.70,P=0.001) but no significant difference due to mounding<br />

(F1,152=1.21,P=0.274) or for the interaction between these treatments (F3,152=0.56,P=0.645). There was<br />

also no significant difference in breast height area-weighted mean micr<strong>of</strong>ibril angle due to taxa<br />

(F3,152,P= 0.541), mounds (F1,152=2.85,P=0.094) or treatment interaction (F3,152=1.27,P=0.287). The<br />

area-weighted averages obtained from SilviScan analysis are presented in Figures 3 and 4.<br />

Wood density (kg m -3 )<br />

750<br />

700<br />

650<br />

600<br />

550<br />

500<br />

F1 CUT F2 PCH PEE<br />

Taxon<br />

Nil mound<br />

0.5m^2 mound<br />

Figure 3. Comparison <strong>of</strong> average area weighted wood density for taxa planted on nil mounds<br />

(treatment 1) and 0.5m 2 mounds (treatment 2) in Expt 195MBR.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 213<br />

Micr<strong>of</strong>ibril angle (deg)<br />

18<br />

17<br />

16<br />

15<br />

14<br />

F1 CUT F2 PCH PEE<br />

Taxon<br />

Nil mound<br />

0.5m^2 mound<br />

Figure 4. Comparison <strong>of</strong> average area weighted micr<strong>of</strong>ibril angle for taxon planted on nil<br />

mounds (Treatment 1) and 0.5m 2 mounds (treatment 2) in Expt 195MBR.<br />

The wood density ranking among taxa is in agreement with the findings <strong>of</strong> Kain (2003) and Harding<br />

and Copley (2000). Although it is not statistically significant, the parental taxa appear to show a<br />

response to mounding for micr<strong>of</strong>ibril angle that is much less pronounced in the hybrid taxa. The latter<br />

is likely to reflect better growth rate achieved on mounds.<br />

Wood density trends for the taxa are plotted in Figure 5. Although the relative trends among the taxa<br />

are similar both with mounds and without there are larger differences among the taxa on the<br />

unmounded site, and these differences are more accentuated in the outer wood.<br />

Wood density (kg m -3 )<br />

900<br />

800<br />

700<br />

600<br />

500<br />

400<br />

300<br />

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

Nil mound 0.5m^2 mound<br />

Relative distance from pith to bark<br />

F1 CUT<br />

F2<br />

PCH<br />

PEE<br />

Figure 5. Air dry density variation relative to proportional distance from pith for taxa in Expt<br />

195MBR planted with and without mounds.<br />

Expt 532NC<br />

The DBHOB trends with age plotted in Figure 6 indicate the large differences in growth rate observed<br />

among treatments that were studied. The magnitude <strong>of</strong> these DBHOB differences rendered this


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 214<br />

experiment an ideal candidate for wood property sampling to provide the opportunity to investigate<br />

wood property response to very large differences in growth rate.<br />

DBHOB (cm)<br />

50<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 5 10 15 20 25 30<br />

Age (years)<br />

S 772 - UT<br />

S 1111 - UT<br />

S 1736 - UT<br />

S 3086 - UT<br />

S 772 - Thin to 600<br />

S 1111 - Thin to 600<br />

S 1736 - Thin to 600<br />

S 3086 - Thin to 600<br />

S 772 - Thin to 200<br />

S 1111 - Thin to 200<br />

S 1736 - Thin to 200<br />

S 3086 - Thin to 200<br />

S 772 - Thin to 400<br />

S 1111 - Thin to 400<br />

S 1736 - Thin to 400<br />

S 3086 - Thin to 400<br />

Figure 6. DBHOB (cm) trends with age for unthinned initial spacing × thinning treatments<br />

sampled for wood studies in Expt 532NC.<br />

Wood density responses to the pre- and post-thinning stocking combinations illustrated in Figure 7 are<br />

not consistent across thinning treatments within initial spacing treatments. There is a tendency for<br />

area-weighted density to be higher in the more heavily thinned treatments. However, analysis <strong>of</strong><br />

variance produced no significant effects for initial stocking (F3,49=1.49, P=0.229), for postthinning<br />

stocking (F3,49=1.76, P=0.168) or for the interaction <strong>of</strong> these treatments (F9,49=0.49,<br />

P=0.871).<br />

Wood density (kg m -3 )<br />

650<br />

620<br />

590<br />

560<br />

530<br />

500<br />

772 spha 1111 spha 1737 spha 3088 spha<br />

Initial stocking<br />

200 spha<br />

400 spha<br />

600 spha<br />

Unthinned<br />

Figure 7. SilviScan area-weighted wood density variation in Expt 532NC with pre- and postthinning<br />

stocking levels.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 215<br />

The underlying trend with age plotted in Figure 8 is strong and clearly illustrates juvenile and mature<br />

wood density zones while also indicating the variability among the treatments.<br />

Basic density (kg m 3 )<br />

600<br />

550<br />

500<br />

450<br />

400<br />

350<br />

300<br />

Pith-5 Rings 06-10 Rings 11-15 Rings 16-20 Rings 21-25 Rings 26-bark<br />

Growth ring segment<br />

772 - 200spha<br />

772 - 400spha<br />

772 - 600spha<br />

772 - Unthinned<br />

1111 - 200spha<br />

1111 - 400spha<br />

1111 - 600spha<br />

1111 - Unthinned<br />

1737 - 200spha<br />

1737 - 400spha<br />

1737 - 600spha<br />

1737 - Unthinned<br />

3088 - 200spha<br />

3088 - 400spha<br />

3088 - 600spha<br />

3088 - Unthinned<br />

Figure 8. Basic density variation from pith to bark for initial stocking × thinning treatment<br />

combinations in Expt 532NC<br />

There is an indication that micr<strong>of</strong>ibril angle is influenced by growth rate and analysis <strong>of</strong> variance<br />

found a significant response to initial stocking (F3,49=9.15, P=0.001) which is reflected in the negative<br />

trend in Figure 9 for area-weighted micr<strong>of</strong>ibril angle with initial stocking level. No significant<br />

difference was found for post-thinning stocking level (F3,49=0.39, P=0.760) or for the initial stocking ×<br />

post-thinning stocking interaction (F9,49=1.12, P=0.366).<br />

Micr<strong>of</strong>ibril angle (deg)<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

772 spha 1111 spha 1737 spha 3088 spha<br />

Initial stocking<br />

200 spha<br />

400 spha<br />

600 spha<br />

Unthinned<br />

Figure 9. Area-weighted micr<strong>of</strong>ibril angle variation in Expt 532NC with pre- and post-thinning<br />

stocking levels.<br />

The results <strong>of</strong> a sawing study on 56 stems from this experiment have been reported separately<br />

(Harding et al. 2009) and draw on wood property results from this study to consider relationships<br />

between standing tree wood properties and sawn product quality and value.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 216<br />

CONCLUSION<br />

Wood property variation in the three mature silviculture experiments studied has revealed some<br />

significant variation with growth rate. However, given the size <strong>of</strong> the growth rate responses observed<br />

in these experiments (particularly between control treatments and the silviculture and/or taxon<br />

treatments that have provided the largest differences in DBHOB), the average changes in wood<br />

density and micr<strong>of</strong>ibril angles observed are relatively small. The data collected provide a very detailed<br />

data set for ongoing wood property modelling, which will underpin decision support systems for<br />

Queensland exotic pine plantations. The practical significance <strong>of</strong> the results observed requires detailed<br />

analysis and conversion modelling to refine predicted impacts on product recovery, grade and value.<br />

However, it is clear that there is considerable potential to gain economic returns from optimising<br />

silvicultural regimes to balance average plantation productivity and wood quality. The latter<br />

opportunities require the pith to bark wood density and wood properties pr<strong>of</strong>iles <strong>of</strong> the trees to be<br />

monitored and managed. Links are needed to key environmental drivers and their interaction with<br />

silvicultural regimes to optimise the quality and value returns from these plantations.<br />

REFERENCES<br />

Catchpoole, K.J., Nester, M.R. and Harding , K.J. (2007) Predicting wood value in Queensland Caribbean pine<br />

plantations using a decision support system. Aust. J. Forestry 70(2): 120-124.<br />

Debuse, V.J. and Nester, M.R. (2002) The effect <strong>of</strong> mound size on the growth <strong>of</strong> Pinus elliottii var elliottii,<br />

Pinus caribaea var hondurensis and their hybrids on poorly drained sites.– Experiment 195MBR.<br />

Unpublished confidential QFRI project report for DPI Forestry, Queensland. 27pp.<br />

Debuse, V.J. and Nester, M.R. (2004) The effect <strong>of</strong> weed competition and cultivation on growth and early<br />

survival <strong>of</strong> Pinus caribaea var. hondurensis – Experiment 132BOW. Unpublished confidential QFRI<br />

project report for DPI Forestry, Queensland. 22pp.<br />

GenStat (2002) GenStat for Windows, version 6.1 Lawes Agricultural Trust.<br />

Harding, K.J. and Copley, T.R. (2000) Wood property variation in Queensland-grown slash ×Caribbean pine<br />

hybrids. In “Hybrid Breeding and Genetics <strong>of</strong> Forest Trees” Proceedings <strong>of</strong> QFRI/CRC-SPF Symposium,<br />

9-14 April 2000, Noosa, Queensland, <strong>Australia</strong>. (Compiled by Dungey, H. S., Dieters, M. J. and Nikles, D.<br />

G.), Department <strong>of</strong> Primary Industries, Brisbane.<br />

Harding, K.J., Copley, T.R., Nester, M.J., Catchpoole, K.J. and Zbonak, A.(2009) Caribbean pine graded sawn<br />

recovery variation in a mature spacing by thinning experiment grown in Queensland. Paper submitted for<br />

presentation to the <strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong> National Conference, 6-10 September 2009,<br />

Caloundra, QLD<br />

Kain, D.P. (2003) Genetic parameters and improvement strategies for the Pinus elliottii × Pinus caribaea var.<br />

hondurensis hybrid in Queensland <strong>Australia</strong>. PhD Thesis, <strong>Australia</strong>n National University. 456pp.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 217<br />

ABSTRACT<br />

MODELLING THE INFLUENCE OF CLIMATE CHANGE<br />

ON PLANTATION WOOD PROPERTIES<br />

Robert Evans 1 , Josh Bowden 1 , Kevin Harding 2 , Terry Copley 2 ,<br />

Kerrie Catchpoole 3 , and Marks Nester 3<br />

As world-wide climate change is accelerating, the need for understanding forest<br />

responses to these changes is <strong>of</strong> increasing importance for commercial and environmental<br />

risk management. Development <strong>of</strong> models to predict wood quality changes for a range <strong>of</strong><br />

species over a range <strong>of</strong> predicted climatic conditions is a step towards optimisation <strong>of</strong> the<br />

value <strong>of</strong> future forest resources. The modelling is based on statistical analysis <strong>of</strong> growth<br />

and wood properties in many trees over a range <strong>of</strong> known climatic conditions. Wood<br />

quality data were obtained from a range <strong>of</strong> Pinus caribaea var. hondurensis samples<br />

using the SilviScan system. The samples were from a Forestry Plantations Queensland<br />

exotic pine thinning and spacing trial. This study is a work in progress and begins by<br />

showing the qualitative connections between environmental variation and wood property<br />

variation (density, tracheid diameter, micr<strong>of</strong>ibril angle, wall thickness, coarseness and<br />

modulus <strong>of</strong> elasticity). The plantation growth model 3-PG has been used to give<br />

predictions <strong>of</strong> annual growth for use in mapping wood properties onto a time axis. All<br />

properties are shown to be strongly affected by short-term variations in weather,<br />

especially rainfall. False rings in earlywood, produced in dry conditions, result in<br />

increased density and modulus <strong>of</strong> elasticity, and decreased micr<strong>of</strong>ibril angle and tracheid<br />

diameter.<br />

INTRODUCTION<br />

1.2 million cubic metres <strong>of</strong> s<strong>of</strong>twood sawlog are produced commercially in Queensland every year<br />

from trees that have experienced variations in climate over decades. Once the seeds (or cuttings) and<br />

sites have been chosen, the quantity and quality <strong>of</strong> the wood are controlled by weather and<br />

management practices such as thinning and application <strong>of</strong> fertiliser. Commercial forests are valued on<br />

the quantity and quality <strong>of</strong> the wood produced, therefore climate variations have a direct impact on the<br />

value, and sometimes the viability, <strong>of</strong> these resources (Broadmeadow et al., 2005).<br />

By combining SilviScan analyses with climate model predictions (for example, the <strong>Australia</strong>n<br />

Community Climate and Earth-System Simulator (ACCESS) (Kowalczyk et al. 2006)) we aim to<br />

estimate the effects <strong>of</strong> climate change on properties such as timber stiffness, strength, pulp yield and<br />

paper density and strength. The premise is that the current (and recent) variation in climate and<br />

corresponding wood properties over the range <strong>of</strong> commercial wood resources is sufficient to generate<br />

relationships that will allow the prediction <strong>of</strong> the effects <strong>of</strong> climate change using existing climate<br />

models.<br />

In the last decade, very large numbers <strong>of</strong> wood samples from across <strong>Australia</strong> and around the world<br />

have been measured by SilviScan. Encoded in these results are climate signals that could be used to<br />

find climate/wood property relationships. In this context, ‘climate’ initially involves temperature and<br />

rainfall, which are two <strong>of</strong> the most important extrinsic variables affecting tree growth and wood<br />

structure. The goal is to find climate signals in the noise arising from the many uncontrolled and<br />

unmeasured variables (e.g. variations in site, microclimate and genetics), and eventually to generate<br />

1<br />

CSIRO Materials Science and Engineering. Bayview Ave Clayton, Victoria.<br />

2<br />

Queensland Primary Industries and Fisheries, Department <strong>of</strong> Employment, Economic Development and Innovation,<br />

Townsville and Indooroopilly, Qld.<br />

3<br />

Forestry Plantations Queensland, Brisbane and Gympie, Qld.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 218<br />

predictive models that can be applied to different species over a range <strong>of</strong> sites in <strong>Australia</strong> and around<br />

the world.<br />

We will here describe the data, methods and initial observations for a set <strong>of</strong> trees collected from a<br />

South East Queensland Pinus caribaea var. hondurensis trial.<br />

MATERIAL AND METHODS<br />

A set <strong>of</strong> 30-year-old Pinus caribaea var. hondurensis samples from Beerburrum, in south-east<br />

Queensland (about 60km north <strong>of</strong> Brisbane), analysed in a previous project, was found to have a<br />

complex ‘false ring’ structure resulting from environmental variation over many years. These rings are<br />

regions <strong>of</strong> high wood density which occur within the normal annual ring growth <strong>of</strong> a tree, either before<br />

or during the normal rise in density at the end <strong>of</strong> a season, and is thought to occur at the cessation <strong>of</strong><br />

shoot and needle elongation (Chudn<strong>of</strong>f and Geary, 1973). Twelve radial samples from an original 520<br />

(65 trees, 8 heights per tree) were selected for high resolution analysis on SilviScan-3 at the 4m<br />

sampling level in 12 trees. The samples were in the form <strong>of</strong> 2mm wide (tangential dimension) and<br />

7mm high (longitudinal dimension) pith-to-bark (radial) strips.<br />

They were scanned in three stages. First, automated microscopy and image analysis <strong>of</strong> the polished top<br />

(radial-tangential) plane gave growth-ring boundary orientation, ray orientation, radial tracheid<br />

diameter and tangential tracheid diameter at 25 micron intervals. Second, automated x-ray scanning<br />

densitometry gave density variation (10 micron intervals), normalised to the actual average sample<br />

density obtained by independent gravimetric analysis. Third, automated scanning x-ray diffractometry,<br />

using 100 micron step size, gave micr<strong>of</strong>ibril angle (MFA) and together with the density information<br />

was used to estimate longitudinal modulus <strong>of</strong> elasticity (MOE). Tracheid wall thickness and<br />

coarseness were calculated from tracheid diameter and density on the assumption that dry cell wall<br />

density is constant at 1500 kg/m 3 . All properties were placed on a 25 micron step scale. Annual ring<br />

boundary identification was performed interactively to avoid the false rings. An annual ring boundary<br />

was generally recognisable by the large, rapid drop in density going from latewood formed at the end<br />

<strong>of</strong> one growing season to the earlywood <strong>of</strong> the next growing season..<br />

Weather data for the Beerburrum site (27.05S, 153.00E) were interpolated using 0.05 degree steps<br />

from nearby weather station information by the <strong>Australia</strong>n Bureau <strong>of</strong> Meteorology Silo Data Drill, and<br />

consisted <strong>of</strong> maximum temperature, minimum temperature, rainfall, evaporation, radiation, vapour<br />

pressure, maximum relative humidity and minimum relative humidity.<br />

The Keetch-Byram Drought Index (KBDI, http://www.kruckysweather.org/kbdi.htm) is an indicator<br />

<strong>of</strong> water deficiency commonly used in bush fire research. KBDI integrates weather conditions such as<br />

temperature and rainfall. During periods <strong>of</strong> high temperature and inadequate rainfall, this index<br />

increases rapidly as the soil dries out. The index returns to a low value after sufficient rain. Such<br />

intermittent conditions are responsible for the false rings seen in these trees.<br />

The wood property pr<strong>of</strong>iles were then remapped on to an annual scale by linear interpolation between<br />

ring boundaries. Intra-annual alignment was carried out using the false rings to ensure that subsequent<br />

averaging <strong>of</strong> the sample pr<strong>of</strong>iles did not eliminate the false ring structure. Prominent false ring<br />

positions were marked using the density pr<strong>of</strong>iles, identifying rising and falling edges as well as peak<br />

positions. The medians <strong>of</strong> all wood property pr<strong>of</strong>iles were then produced. As the ring boundaries are<br />

associated with the beginning <strong>of</strong> spring (July-August, or nominally 60% <strong>of</strong> the way through the<br />

calendar year – Shepherd et al., 2003) the time axes are annotated accordingly in the following Figures<br />

below. The range <strong>of</strong> years displayed was chosen to avoid the relatively featureless juvenile wood<br />

closest to the pith, the thinning operation in 1987, and crown closure.<br />

RESULTS AND DISCUSSION<br />

Figure 1 shows the radial pr<strong>of</strong>iles for MOE (Evans, 2006), which combines information from<br />

diffractometry and densitometry. Density can also be considered a composite property <strong>of</strong> wood,<br />

combining cell wall thickness, cell diameter and cell wall density. The pr<strong>of</strong>iles in Figure 1, which also<br />

show the annual rings and false ring structures, may therefore be considered to represent a summary <strong>of</strong>


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 219<br />

many <strong>of</strong> the properties measured on SilviScan. The lack <strong>of</strong> structural variation in the very juvenile<br />

wood is evident in these pr<strong>of</strong>iles.<br />

Under ideal conditions, Southern pines such as P. caribaea grow rapidly, displaying a distinctive<br />

“square-wave” pattern <strong>of</strong> alternating low density early wood and high density latewood. In these<br />

samples, the local weather patterns have caused intermittent wood growth to produce very complicated<br />

radial pr<strong>of</strong>iles <strong>of</strong> wood properties, and a relatively wide range <strong>of</strong> average growth rates.<br />

longitudinal modulus <strong>of</strong> elasticity<br />

0 50 100<br />

distance from pith / mm<br />

150 200<br />

Figure 1. Longitudinal modulus <strong>of</strong> elasticity pr<strong>of</strong>iles for the 12 selected radial samples.<br />

Radius / mm<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

thinned<br />

1987<br />

1980 1985 1990 1995 2000 2005<br />

Year<br />

Figure 2. Individual radial pr<strong>of</strong>iles for the 12 samples. Thinned to 200 stems/ha (squares), 400<br />

stems/ha (diamonds), 600 stems/ha (triangles). Unthinned - 1100 stems/ha (solid<br />

circles).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 220<br />

The ring boundary positions were used to generate radial growth curves for the samples, as shown in<br />

Figure 2. The plots from which the samples were taken had been thinned to 200, 400 and 600 stems/ha<br />

(plus unthinned control) in 1987 at age 10. The change in rate <strong>of</strong> growth after thinning increased with<br />

thinning intensity. It should be noted that the false ring patterns were not affected by the level <strong>of</strong><br />

thinning. As ring boundary positions were difficult to pinpoint within the juvenile wood (Fig 1), only<br />

data after 4 years is shown in Figure 2.<br />

An initial attempt to connect the annual growth <strong>of</strong> the samples with the weather was made using the<br />

modelling s<strong>of</strong>tware 3-PG (Landsberg et al., 2001, Coops and Waring 2001, Coops et al., 2005) and<br />

results are shown in Figure 3. Manual parameterisation <strong>of</strong> 3-PG was carried out by modifying the<br />

parameters for P. radiata as shown in Table 1. Note that Tmax, Tmin and Topt are dramatically<br />

different from those <strong>of</strong> the temperate climate pine. Figure 3, which also leaves out juvenile growth<br />

prior to 1982, shows a reasonable prediction <strong>of</strong> the 3-PG model <strong>of</strong> annual yearly increment, although<br />

some features, such as the spike in growth during 1999-2000, are inadequately modelled. The adjusted<br />

parameters reasonably predict yearly increments, but the prediction <strong>of</strong> monthly growth is far from<br />

optimal, indicating further work is needed before 3-PG can be used to map wood properties onto a<br />

time scale. Optimisation <strong>of</strong> the parameters Tmin, Topt and Tmax for this species over a range <strong>of</strong> sites<br />

with varying climate along the Queensland coast would improve the 3-PG model and is now in the<br />

planning stages.<br />

ring width / mm<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

1980 1985 1990 1995 2000 2005 2010<br />

year <strong>of</strong> growth<br />

Figure 3. Average ring widths for the 12 samples predicted by 3-PG (diamonds) using the<br />

available weather conditions and 3-PG parameters modified from radiata pine as<br />

seen in Table 1. Actual ring widths are also shown from SilviScan data (circles).<br />

Table 1. 3-PG parameters for P. radiata and modified for P. caribaea.<br />

3-PG parameters P. radiata P. caribaea<br />

(this study)<br />

Foliage: stem partitioning ratio @ D=20 cm (pFS20) 0.4 0.5<br />

Constant in the stem mass v. diam. relationship (aS) 0.0243 0.11<br />

Power in the stem mass v. diam. relationship (nS) 2.72 2.7<br />

Maximum fraction <strong>of</strong> NPP to roots (pRx) 0.6 0.95<br />

Minimum fraction <strong>of</strong> NPP to roots (pRn) 0.25 0.45<br />

Minimum temperature for growth (Tmin) / °C 0 8<br />

Optimum temperature for growth (Topt) / °C 20 32<br />

Maximum temperature for growth (Tmax) / °C 32 38<br />

Relative age to give fAge = 0.5 0.5 0.55


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 221<br />

Minimum basic density - for young trees (rhoMin) / t/m 3 0.480 0.500<br />

Maximum basic density - for older trees (rhoMax) / t/m 3 0.480 0.600<br />

Age at canopy cover 0 24<br />

3-PG Initialisation data:<br />

Fertility rating 0.6<br />

Initial stocking / stems/Ha 1100<br />

Soil class Sandy loam<br />

Max avail. soil water/ mm 250<br />

Initial avail. soil water 200<br />

Silvicultural event Age = 10 Stocking = 600<br />

Although many properties were measured by SilviScan, interpretation <strong>of</strong> their variation is greatly<br />

complicated by strong local correlation between properties. An example is shown in Figure 4. MFA<br />

and density are inversely correlated in general. The quantitative nature <strong>of</strong> the correlation varies over<br />

time and between samples, so that one property cannot be estimated from the other. Such correlations<br />

affect the interpretation <strong>of</strong> variation in all wood properties.<br />

density / kg/m3<br />

1200<br />

700<br />

200<br />

-300<br />

wood density (upper line), MFA (lower line)<br />

-800<br />

0 20 40 60 80 100 120 140<br />

distance from pith / mm<br />

Figure 4. Micr<strong>of</strong>ibril angle and density radial pr<strong>of</strong>iles for a typical radial sample.<br />

Mapping on to a time scale<br />

For comparison with weather / climate information, the wood property data should be presented on a<br />

time scale. No detailed intra-annual growth information was available for these trees, therefore<br />

converting the distance-based property pr<strong>of</strong>iles into time-based pr<strong>of</strong>iles could only be done using<br />

annual ring boundaries.<br />

density kg/m3<br />

1200<br />

700<br />

200<br />

-300<br />

-800<br />

5<br />

wood density (upper line), MFA (lower line)<br />

1989.6 1990.6 1991.6 1992.6 1993.6 1994.6 1995.6 1996.6 1997.6 1998.6 1999.6<br />

Figure 5. Median pr<strong>of</strong>iles <strong>of</strong> wood density and MFA showing their complementary local<br />

variation, including that caused by environmental influences. The pr<strong>of</strong>iles have been<br />

linearly remapped on to annual intervals.<br />

year<br />

After the mapping as described in the Methods section, the intra-annual position <strong>of</strong> wood property<br />

features is not strictly time-based because radial growth rate is not constant within a year. Faster<br />

30<br />

25<br />

20<br />

15<br />

10<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

MFA / deg<br />

MFA / deg.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 222<br />

spring growth, for example shifts the features to the right between the annual ring boundaries. This<br />

should be noted for all Figures from Figure 5 onwards.<br />

density / kg/m3<br />

1200<br />

700<br />

200<br />

-300<br />

-800<br />

wood density (upper line), tracheid population density (lower line)<br />

1989.6 1990.6 1991.6 1992.6 1993.6 1994.6 1995.6 1996.6 1997.6 1998.6 1999.6<br />

year<br />

Figure 6. Environmental influences on median wood density and the median population density<br />

<strong>of</strong> tracheids. The pr<strong>of</strong>iles have been linearly remapped on to annual intervals.<br />

Figure 5 shows approximate time-based pr<strong>of</strong>iles for density and MFA. The inverse relationship<br />

between these properties is very clear. In Figure 6, wood density variation is compared with that <strong>of</strong><br />

tracheid population (tracheids per mm 2 ). Note the similarity <strong>of</strong> the patterns <strong>of</strong> intra-annual features in<br />

Figures 5 and 6. The three properties (density, tracheid population and MFA) are measured<br />

independently. Environmental conditions have had a very large effect on the wood properties. Peaks in<br />

false ring density and tracheid population are <strong>of</strong>ten double the levels in the surrounding wood. MFA<br />

fell by up to a third in the same regions.<br />

KBDI is shown in Figure 7, together with wood density for a sequence <strong>of</strong> 9 years. There are many<br />

spikes in KBDI that could be associated with the density spikes (false rings), bearing in mind that the<br />

intra-annual positions <strong>of</strong> the false rings and the strictly time-based KBDI peaks are not expected to<br />

coincide, but lag by varying amounts. Wood radial growth is much more rapid in the first half <strong>of</strong> the<br />

year than in the second half, therefore it is expected that climate features in spring (true time scale)<br />

should occur before those in the property pr<strong>of</strong>iles.<br />

density kg/m3<br />

1200<br />

700<br />

200<br />

-300<br />

-800<br />

wood density (upper line), KBDI (lower line)<br />

1989.6 1990.6 1991.6 1992.6 1993.6 1994.6 1995.6 1996.6 1997.6 1998.6 1999.6<br />

Figure 7. Median wood density variation and drought index. The density pr<strong>of</strong>ile has been<br />

remapped on to annual intervals for closer comparison with the time-based drought<br />

index.<br />

year<br />

Rainfall appears to be the limiting factor in the earlywood growth <strong>of</strong> fast-growing tropical pines (Slee<br />

1972). Figures 8 and 9 show rainfall accumulated over 30 days (moving mean) with pr<strong>of</strong>iles <strong>of</strong> density<br />

and radial tracheid diameter. Note that over any 30 day period, the accumulated rainfall rarely exceeds<br />

200mm, and never in early spring. The peaks in rainfall are not associated with peaks in density and<br />

decreases in tracheid diameter, but with the restoration <strong>of</strong> ‘normal’ wood properties. Rain is expected<br />

to bring an end to the false rings and increase tracheid diameter.<br />

Therefore, the peaks in accumulated rainfall should be associated with rapid decreases in density and<br />

rapid increases in tracheid diameter. As with KBDI, many <strong>of</strong> the features seen in the rainfall patterns<br />

are reflected in the false ring structures.<br />

2200<br />

2000<br />

1800<br />

1600<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

KB drought index<br />

tracheids /mm2


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 223<br />

density kg/m3<br />

1200<br />

700<br />

200<br />

-300<br />

-800<br />

wood density (upper line), rainfall (lower line)<br />

1989.6 1990.6 1991.6 1992.6 1993.6 1994.6 1995.6 1996.6 1997.6 1998.6 1999.6<br />

Figure 8. Median wood density pr<strong>of</strong>ile remapped on to annual intervals, and rainfall<br />

accumulated over the previous 30 days.<br />

year<br />

Current efforts are directed towards the quantification <strong>of</strong> the effects <strong>of</strong> environmental variation on the<br />

wood properties at different scales. False ring structures that can be reliably associated with weather<br />

events will be used to estimate the magnitude <strong>of</strong> the local excursions in wood properties. Average<br />

wood properties within rings (and perhaps within earlywood and latewood) will be associated with<br />

climate, as well as climate variability, to generate relationships more suitable for the prediction <strong>of</strong><br />

future macroscopic wood properties.<br />

tracheid radial<br />

diameter /um<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

tracheid radial diameter (upper line), rainfall (lower line)<br />

1989.6 1990.6 1991.6 1992.6 1993.6 1994.6 1995.6 1996.6 1997.6 1998.6 1999.6<br />

Figure 9. Median tracheid radial diameter remapped on to annual intervals, and rainfall<br />

accumulated over the previous 30 days (moving mean).<br />

CONCLUSIONS<br />

year<br />

• Growth modelling s<strong>of</strong>tware 3-PG gave a qualitatively similar ring width pattern to that<br />

measured. Further efforts on 3-PG parameterisation to predict growth at higher<br />

(monthly) time resolution could improve the mapping <strong>of</strong> the distance scale on to the<br />

time scale. Accurate time scale mapping will improve statistical relationships between<br />

climate and wood properties, allowing more accurate predictions <strong>of</strong> future wood<br />

properties.<br />

• False ring structures associated with environmental variation are evident in all<br />

measured wood properties.<br />

• Initial indications are that the false rings result from an alternation between insufficient<br />

and sufficient rainfall for efficient growth.<br />

• Reduced rainfall during earlywood formation would be expected to increase density<br />

and MOE, decrease MFA and tracheid diameter.<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

0<br />

30 day rainfall<br />

/mm<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

30 day rainfall /mm


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 224<br />

REFERENCES<br />

Broadmeadow, M. S. J., D. Ray and. Samuel C.J.A. 2005, Climate change and the future for broadleaved tree<br />

species in Britain. Forestry, 78 (2), 145-161.<br />

Chudn<strong>of</strong>f, M. and Geary, T.F. 1973, Terminal shoot elongation and cambial growth rhythms in Pinus caribaea.<br />

Commonwealth Forestry Review, 52(4), 317-324.<br />

Coops, N.C. and Waring, R.H. 2001, Assessing forest growth across South-Western Oregon under a range <strong>of</strong><br />

current and future global change scenarios using a process model, 3-PG. Global Change Biology, 7(1), 15-<br />

29.<br />

Coops, N.C., Waring, R.H. and Law, B.E. 2005, Assessing the past and future distribution and productivity <strong>of</strong><br />

ponderosa pine in the Pacific Northwest using a process model, 3-PG, Ecological. Modelling, 183, 107–<br />

124.<br />

Evans, R. 2006, Wood stiffness by x-ray diffractometry. In: Characterisation <strong>of</strong> the cellulosic cell wall, Chapter<br />

11. Proceedings <strong>of</strong> the workshop 25-27 August 2003, Grand Lake, Colorado, USA. Southern Research<br />

Station, University <strong>of</strong> Iowa and the Society <strong>of</strong> Wood Science and Technology. D. Stokke and L. Groom,<br />

eds. Blackwell Publishing.<br />

Kowalczyk, E.A., Wang, Y. P., Law, R. M., Davies, H. L., McGregor, J.L. and Abramowitz, G. 2006, The<br />

CSIRO Atmosphere Biosphere Land Exchange (CABLE) model for use in climate models and as an <strong>of</strong>fline<br />

model. CSIRO Marine and Atmospheric Research paper 013.<br />

Landsberg, J. J., Johnsen, K. H., Albaugh, T. J., Allen, H. L. and McKeand, S.E. 2001, Applying 3-PG, a Simple<br />

Process-Based Model Designed to Produce Practical Results' to Data from Loblolly Pine Experiments.<br />

Forest Science, 47(1), 43-51.<br />

Shepherd, M., Cross M., Dieters M. J., Harding K., Kain D. and Henry R. 2003, Genetics <strong>of</strong> physical wood<br />

properties and early growth in a tropical pine hybrid. Canadian Journal <strong>of</strong> Forest Research, 33(10), 1923–<br />

1932.<br />

Slee, M.U. 1972, Growth patterns <strong>of</strong> slash and Caribbean pine and their hybrids in Queensland. Euphytica,<br />

21(1), 129-142.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 225<br />

COMPETITION AND COLLABORATION<br />

IN THE AUSTRALIAN TIMBER FURNITURE INDUSTRY –<br />

A VALUE CHAIN APPROACH TOWARD HIGHER VALUE<br />

CREATION FOR LOCAL AND GLOBAL MARKETS AND PATHWAYS<br />

FOR SUSTAINABLE FOREST RESOURCE USE<br />

ABSTRACT<br />

Sasha Alexander 1<br />

This paper explores the development <strong>of</strong> an innovative value creation methodology which<br />

extend value chain analysis techniques that highlight opportunities for performance,<br />

pr<strong>of</strong>itability and inter-firm relationship improvements to a complete timber value chain<br />

for a furniture product, from forest to end-consumer. The research design employs a<br />

mixed method approach comprising qualitative industry survey interviews, and<br />

quantitative and qualitative industry and end-consumer surveys, guided by a value chain<br />

analysis approach.<br />

The findings support more direct engagement with end-consumer needs, providing value<br />

chain insights toward those timber furniture attributes which require the greatest<br />

attention in the creation <strong>of</strong> value in the timber value chain, including timber material<br />

sourcing and production phases.<br />

A key theme to emerge is the positive correlation <strong>of</strong> organisational commitment toward<br />

creating further value for timber end-consumers through environmental stewardship, and<br />

capitalising on timber’s unique end-consumer-valued attributes, with the end-consumer<br />

as central to the development <strong>of</strong> the timber value proposition.<br />

INTRODUCTION<br />

The need to increase the competitiveness <strong>of</strong> <strong>Australia</strong>n firms participating in the timber furniture<br />

industry has become more critical in recent years with the marked increase in the volume <strong>of</strong> imported<br />

furniture products representing about AUD2 billion <strong>of</strong> a AUD5 billion furnishings industry (FIAA<br />

2002). The global furniture marketplace is considerable at about AUD80 billion <strong>of</strong> which <strong>Australia</strong>n<br />

production contributes a small fraction, especially to global export and import replacement (United<br />

Nations, 2005). <strong>Australia</strong> is grouped within a series <strong>of</strong> countries which possess high costs in<br />

production and therefore is seen at a competitive disadvantage in comparison to countries representing<br />

low cost production. This poses a considerable direct threat to competitiveness <strong>of</strong> so-called high cost<br />

countries (HCC).<br />

Coupled with a perception <strong>of</strong> reduction in access to particular <strong>Australia</strong>n forest resources, the need to<br />

focus on higher value-added products has become an imperative in planning for sustainable forest<br />

resource use and achieving returns on investment in the <strong>Australia</strong>n furniture sector (AEGIS, 1999).<br />

<strong>Australia</strong>n furniture products have surged in domestic and export popularity but the industry faces a<br />

strong threat from imports. The United Nations International Trade Database has valued furniture<br />

imports into <strong>Australia</strong> in excess <strong>of</strong> $470m in 1998 (United Nations 1999), $583m in 2001, moving<br />

progressively to $1.5bn in 2005 (United Nations 2005). <strong>Australia</strong>’s furniture sector exports in a trade<br />

balance comparison were $55.3m in 1998 (United Nations 1999), $76m in 2001, then climbing to<br />

$100m in 2005.<br />

The nature <strong>of</strong> international competition has become more complex during this period and <strong>Australia</strong>n<br />

industries have experienced reduced import tariff protection in a bid to encourage competitiveness, a<br />

1<br />

University <strong>of</strong> Western Sydney, Locked Bag 1797, Penrith South DC, NSW 1797. Ph: 61 2 9852 5222.<br />

Email: S.Alexander@uws.edu.au.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 226<br />

rise in labour costs in comparison to countries with relatively lower labour structures, cost increases in<br />

local materials and other commodities, and consolidation <strong>of</strong> providers in several global trading sectors.<br />

This trade gap imbalance creates an increasing concern for the competitiveness <strong>of</strong> the <strong>Australia</strong>n<br />

timber furniture industry. If sustained, this trend could render traditional timber furniture<br />

manufacturing based employment unsustainable, and erode domestic growth opportunities.<br />

Furthermore, a state <strong>of</strong> reduced competitiveness in adding value to processed materials could equate to<br />

further losses in wealth generation capability, potentially relegating <strong>Australia</strong> to a raw materials<br />

supplier to the global timber furniture products sector and a reduced economic value generator.<br />

The negative impact on the competitiveness <strong>of</strong> high cost countries (HCC) such as <strong>Australia</strong>, the USA,<br />

and the members <strong>of</strong> the European community and others, from low cost countries (LCC) <strong>of</strong><br />

manufacture, namely China, has transformed global competition and challenged competitors to adapt<br />

to new consumer demands (Navarro et al 2008).<br />

A firm’s competitive ability to succeed in meeting new consumer demands within an industry sector<br />

and the marketplace is based on maintaining consumer interest in the products and services it <strong>of</strong>fers<br />

(Woodruff 1997). The creation <strong>of</strong> unique value for the end-consumer is pivotal in building competitive<br />

advantage over rival firms (Barney 2002).<br />

THEORETICAL BACKGROUND: A VALUE CHAIN APPROACH<br />

In the analysis and building <strong>of</strong> competitive advantage, value chain methods have gained popularity as<br />

the efficiency <strong>of</strong> each process step is assessed in detail from raw materials to the eventual user,<br />

whereas the preceding supply chain models focused on activities that take raw materials and subassemblies<br />

into a manufacturing operation smoothly and economically (Porter 1985, Barney 2002).<br />

The importance an end-consumer places on the purchase <strong>of</strong> a forest product not only underwrites the<br />

viability <strong>of</strong> the timber furniture retailer but the suppliers <strong>of</strong> each preceding process. Firms can be seen<br />

to merely move money around by paying each other for services but it is the end consumer that is the<br />

ultimate source placing money back into the entire system or chain <strong>of</strong> activities known as the value<br />

chain.<br />

The value chain perspective highlights interdependence between firms aligned along a common value<br />

chain embracing raw material supply, logistics, product development, manufacturing and<br />

merchandising instead <strong>of</strong> primarily being about purchasing and selling. In recent years increases in the<br />

dominance <strong>of</strong> some value chains global competition has tended to be between value chains and<br />

precluded individual firms from re-entering the marketplace. This methodology is <strong>of</strong> value to a wide<br />

range <strong>of</strong> forest product industries and the timber furniture industry serves as an example <strong>of</strong> the<br />

potential.<br />

Morris (2000) suggests that a micro-level understanding <strong>of</strong> value creation <strong>of</strong> each firm within an<br />

interdependent and inter-firm context can be effective in focussing attention toward process efficiency<br />

and increases in productivity and economic rewards. This level <strong>of</strong> understanding if incorporated in<br />

cooperative intra-firm, inter-firm, regional and an industry-wide basis suggests a move to greater<br />

understanding <strong>of</strong> higher added value performance based on collaboration between firms and a<br />

heightened awareness <strong>of</strong> the cost in creating value for all suppliers and value for money for endconsumers.<br />

The increase in global competition has driven many organisations to search for new ways to achieve<br />

and retain competitive advantage by pursuing new value creation.<br />

In describing the attributes in new value creation, Bitici (2004) states that the more successful<br />

companies have a very clear and unambiguous value proposition, with the value proposition based on<br />

focussed and refined developments <strong>of</strong> key competencies, enabling value proposition delivery to<br />

consumers and suppliers. The key competencies include product leadership, operational excellence,<br />

and customer intimacy. Woodruff (1997) contends that the next major source <strong>of</strong> competitiveness will<br />

come from more outward orientation toward consumers and call on organisations to compete on<br />

superior consumer value delivery.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 227<br />

The extant literature does not directly link the concept <strong>of</strong> end-consumer value to the overall efficiency<br />

<strong>of</strong> the value chain, though there is discussion that value is reliant on value attributed by the endconsumer<br />

at the timber product retail level or retail showroom, as is the case with domestic or<br />

commercial timber furniture.<br />

Traditionally, each firm level decision is relatively independent <strong>of</strong> the end-consumer as value chain<br />

members are mostly isolated from having an overall viewpoint in how value creation is prioritised,<br />

information shared across value chain levels, nor the effectiveness and efficiency <strong>of</strong> activities and<br />

pr<strong>of</strong>itability in their value chain. The focus <strong>of</strong> value chain research to date has been on production<br />

efficiency and inter-firm communication and collaboration, rather than an extended model <strong>of</strong><br />

collaboration integrating end-consumer requirements with overall value chain ambitions which are<br />

<strong>of</strong>ten fragmented, at best.<br />

The sought-after linkage between value chain effectiveness and end-consumer value may be<br />

addressed, in part, by the concept <strong>of</strong> the co-creation <strong>of</strong> value (Prahalad and Ramaswamy 2004, p.8).<br />

The co-creation <strong>of</strong> value between firms and consumers provides the linkage between the needs <strong>of</strong> the<br />

consumer and the needs <strong>of</strong> the firm, without isolating the firm or the consumer from the context <strong>of</strong> use<br />

and the creation <strong>of</strong> the consumer experience.<br />

A co-created forum suggests a market where there is an uninterrupted flow <strong>of</strong> information from<br />

individual consumer experience to the value chain and back “rather than passive pockets <strong>of</strong> demand<br />

for the firm’s <strong>of</strong>ferings” (Prahalad and Ramaswamy 2004, p.16). The relationships are summarised in<br />

an adapted industry map placing the furniture consumer as the central focus <strong>of</strong> value creation activities<br />

and that all firm-level activities have a co-dependent effect on the success <strong>of</strong> the whole (Figure 1).<br />

Source: Adapted from Prahalad and Ramaswamy 2004, p.9<br />

Figure 1. Co-creation experience model<br />

The challenge for the future growth <strong>of</strong> value creation opportunities in the forest products sector or<br />

furniture sector may be reliant on the success <strong>of</strong> moving from an informal understanding <strong>of</strong> a value<br />

chain member’s role to a more formally recognised interdependent value creation model. Value chain<br />

members <strong>of</strong>ten conduct isolated efforts toward value chain outcomes communicating as a supplier to<br />

only one buyer. Firms may be unaware that their efforts are contained within a chain in which all<br />

inter-firm actions are interdependent.<br />

An alternate industry model where each value chain member is acknowledged by all value chain<br />

members in the value contributed to the value proposition delivered to the end-consumer has merit in<br />

the mounting <strong>of</strong> strategies to compete against rival chains with similar motivations. The prospect <strong>of</strong><br />

receiving and providing assistance to enhance the efficiency and pr<strong>of</strong>itability for collaborators within<br />

one’s own value chain also adds impetus to building competitive advantage and creating further value<br />

for end-consumers as a common objective.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 228<br />

METHODOLOGY<br />

This paper identifies specific areas for improving the competitiveness <strong>of</strong> the <strong>Australia</strong>n timber<br />

furniture industry, thereby influencing the forest products sector. The outcomes may have resonance in<br />

forest products industries and other sectors as the concept <strong>of</strong> competitiveness and a user-centred<br />

approach is not exclusive to the forest products sector alone. The most promising mechanisms for<br />

building value through industry and end-consumer collaboration using a value chain approach are<br />

explored. A mixed method research design was implemented in three phases and included the<br />

participation <strong>of</strong> several levels <strong>of</strong> the value chain from the forest source to the end-consumer.<br />

The confluence <strong>of</strong> decision making in the value chain was observed during interviews at multiple<br />

levels <strong>of</strong> the value chain. The levels investigated included the forest harvest, timber mill, timber<br />

merchant, the furniture manufacturer, furniture sub-contractor, and furniture retail levels <strong>of</strong> the value<br />

chain. As the research was focused primarily on the <strong>Australia</strong>n timber furniture industry, the timber<br />

furniture manufacturing level was selected as the starting point in the identification <strong>of</strong> further players<br />

in the timber furniture value chain.<br />

The analysis <strong>of</strong> data from surveys gathered in the production areas <strong>of</strong> the value chain and with endconsumers<br />

in the furniture retail environment was interpreted and conclusions were drawn to ascertain<br />

if the intentions and ambitions <strong>of</strong> the value chain were complementary to the needs <strong>of</strong> the endconsumer.<br />

Participation from the industry levels <strong>of</strong> the value chain was a key component to the study. This<br />

included firms associated with forest supply, timber milling and distribution, furniture manufacture,<br />

transport logistics and furniture retail. Respondents from timber furniture manufacturing level <strong>of</strong> the<br />

value chain overwhelmingly cited the threat posed by low cost entrants into the <strong>Australia</strong>n timber<br />

furniture market as a continuing competitive barrier to gaining market share. This emphasis and preoccupation<br />

in gaining local market share may be adversely affecting the psychology <strong>of</strong> firms in the<br />

creating <strong>of</strong> export opportunities at the furniture manufacturer level. Furniture manufacturer level<br />

resources are primarily focused on survival rather than growth as most firms in the furniture industry<br />

possess less than ten employees and retain few resources to generate innovations through new product<br />

development.<br />

The overall research strategy focused on the timber furniture value chain and the value creation<br />

process from forest plantation to furniture retail environment and the end-consumer. Barnes and<br />

Morris (1999) suggest that a focus on an individual product rather than only firm specific issues<br />

provides the ability to assess a production pipeline and therefore its value chain in its capacity to meet<br />

particular market demands (Womack and Jones 1996 cited in Barnes and Morris 1999). The<br />

production pipeline signifies the actions <strong>of</strong> value chain member firms in the production <strong>of</strong> a product.<br />

The selection <strong>of</strong> furniture type in this study was based on the suitability for timber value chain study<br />

within available time and resources. The selection suitability required the timber product to be<br />

relatively simple in construction including range <strong>of</strong> furniture construction materials used, as very<br />

complex products <strong>of</strong>ten constitute complex and extended value chain study, and resources. Access to a<br />

reliable study sample through industry referrals and the researcher’s familiarity with single seat<br />

furniture design promoted timber dining chairs as the favoured timber product to be traced through the<br />

timber value chain.<br />

A mixed method approach employing a triangulated research design (Creswell and Plano Clark 2003<br />

cited in Creswell and Plano Clark 2007 p.62) was implemented with the view to interpret the results <strong>of</strong><br />

both industry and end-consumer surveys in establishing the extent <strong>of</strong> the value chain and factors<br />

affecting value chain competitiveness (Figure 2).<br />

The research design was formulated in two industry data collection phases: Phase 1 - exploratory and<br />

primarily qualitative; and Phase 2, a more targeted survey informed by the industry interviews <strong>of</strong><br />

Phase 1, with both Phases containing a mixed approach seeking quantitative and qualitative data. A<br />

third, and final value chain level survey phase seeking end-consumer data provides the emphasis<br />

placed on successful timber furniture attributes and end-consumers’ perception <strong>of</strong> value (Phase 3).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 229<br />

Source: Adapted from Creswell & Plano Clark (2007 p.63).<br />

Figure 2. Mixed Method – Triangulation Design: Convergence Model<br />

The gathering <strong>of</strong> separate industry and consumer data was proposed as the most effective data<br />

gathering starting point until data from industry and consumer survey could be analysed for overlap<br />

and actors <strong>of</strong> significance (AEGIS 1999). This is unique to this study and not yet identified in the<br />

extant literature in the <strong>Australia</strong>n furniture sector.<br />

OPERATIONALISATION: THE VALUE CHAIN SAMPLING TECHNIQUE<br />

Initially nine value chains were administered. This later was reduced to three value chains based on<br />

the timeframe, level <strong>of</strong> cooperation, and geographic differentiation and complexity <strong>of</strong> the value chain.<br />

Limitations experienced included the securing and timing <strong>of</strong> interviews in several <strong>Australia</strong>n states;<br />

intellectual property issues; or issues undisclosed. Using this sample frame, over twenty-five<br />

respondents were engaged equating up to eight respondents per value chain on average.<br />

The measurement technique used in data gathering Phases 1, 2, and 3 <strong>of</strong> the study was a selfadministered<br />

survey technique containing a semi-structured interview, industry and end-consumer<br />

survey questionnaires respectively. The semi-structured interview technique is particularly effective in<br />

generating descriptive, explanatory, and exploratory perspectives. This method was used to test a<br />

proposed model based on the constructs <strong>of</strong> trust and collaboration, communication, and value creation.<br />

The standardisation <strong>of</strong> the survey for very different industry levels gives the study the potential to<br />

provide a balanced approach well suited to a value chain approach. A potential drawback is that the<br />

researcher may miss what is most important to the responder in their core competence area. Questions<br />

for this reason were as general and as open-ended as possible.<br />

In addition, the consideration <strong>of</strong> data from a wide sample <strong>of</strong> industry respondents also enhances the<br />

potential generality <strong>of</strong> findings. The semi-structured interview represented eight constructs <strong>of</strong> interest<br />

providing a number <strong>of</strong> variables <strong>of</strong> consequence to value chain performance and contained over thirtytwo<br />

open-ended questions within eight categories. The categories included firm level strategy; interfirm<br />

relations; resources and core competencies; local and global competitive threats; competitiveness<br />

from quality, policy and market data perspective, firm level performance; end-consumer value; and<br />

pr<strong>of</strong>itability trends.<br />

The research design Phase 2 incorporated a structured industry survey questionnaire targeted to the<br />

identified primary decision-making level <strong>of</strong> the value chain representing leadership in the particular<br />

forest product area: the timber furniture manufacturer. The structured industry survey instrument <strong>of</strong><br />

Phase 2 contained an extended and refined version <strong>of</strong> the semi-structured interview process


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 230<br />

represented in Phases 1 and more targeted survey questions. These questions included new product<br />

development (NPD), innovation processes, intellectual property including registered designs or<br />

patents, quality assurance, performance measurement, market intelligence, competitive advantage<br />

strategies, brand recognition, collaboration with suppliers and buyers, communication with suppliers<br />

and buyers, and timber certification. A series <strong>of</strong> questions were also directed to the level <strong>of</strong><br />

understanding <strong>of</strong> the end-consumer needs, industry perceptions <strong>of</strong> consumer value, and the successful<br />

timber attributes industry believed the end-consumer held including environmental issues and the<br />

influence on consumer purchasing decisions.<br />

The final survey Phase 3 was represented by an end-consumer survey questionnaire administered at<br />

the timber furniture retail showroom environment. The questions were directed at members <strong>of</strong> the<br />

general public whose intention was to purchase timber furniture. The end-consumer questionnaire<br />

sought to establish an end-consumer’s opinion <strong>of</strong> the success <strong>of</strong> timber as a furniture material and the<br />

timber attributes that contributed to this consumer perspective. The survey contained five sections<br />

including the level <strong>of</strong> consumer approval <strong>of</strong> timber as a furniture material, successful end-consumer<br />

timber attributes, end-consumer buyer preferences, brand character responses, and end-consumer<br />

perceptions <strong>of</strong> timber certification and related environmental issues affecting timber furniture<br />

ownership.<br />

RESULTS: INDUSTRY SURVEY<br />

From the industry survey results the following points can be summarised as aspects concerning the<br />

timber furniture industry by firms contributing to the furniture value chain (FVC).<br />

A) Strategy: Respondents were confident that their firms more likely held sound new product<br />

development strategies; that they possessed unique product or services over competitors; that<br />

competitors were assessed for competitiveness periodically through mostly informal methods; and<br />

that internal performance measurement applying benchmarking techniques were prevalent.<br />

B) Inter-firm relations: Respondents were more likely to seek better relationships with suppliers,<br />

buyers, and industry bodies and less likely to seek further collaboration with government in their<br />

keen need to derive more reliable market intelligence. Higher collaboration was linked to better<br />

quality outcomes with suppliers and buyers, and that firms, mostly manufacturers, were more<br />

likely to be the initiators <strong>of</strong> communication with suppliers and buyers. Business strategies <strong>of</strong><br />

interacting firms were not widely shared, nor were pr<strong>of</strong>it margins.<br />

C) Resources / core competencies: Respondents were positive in high and appropriate technology<br />

access and utilisation; reliable access to skilled human resources, reliable materials inventory and<br />

handling competencies, strong order processing efficiency and procurement effectiveness. The<br />

marketing <strong>of</strong> services was seen as a concern but effective selling was not.<br />

D) Competitive threats: Respondents were more likely to encounter difficulties in maintaining<br />

reliable suppliers and customers, and that international trade policy would adversely affect pr<strong>of</strong>its<br />

and with competition based on cost alone.<br />

E) Competitiveness: An increase in wider industry compliance to quality standards was linked to<br />

competitiveness. The design <strong>of</strong> imported products was not seen as a key competitive determinant,<br />

nor did timber certification give the imported products the competitive edge.<br />

F) Assessing efficiency – firm level performance: Strong support for understanding efficiency in<br />

firm’s and with collaborative firms, and preferences to meeting suppliers and buyers face-to-face.<br />

A very high level <strong>of</strong> satisfaction was present with current internal performance systems.<br />

G) Consumer value: The needs <strong>of</strong> end-consumers were held in very high regard by respondent firms<br />

and the furniture buyers according to the survey <strong>of</strong> firms, but the suppliers were thought to have<br />

less interest. Respondents perceived that end-consumers were more likely to have an interest in<br />

environmental issues and the belief that timber held superior material qualities over comparable<br />

materials in the end-consumer’s mind and that origin <strong>of</strong> timber both location and type had some<br />

influence on end-consumer choice. Respondents did link higher consumer knowledge <strong>of</strong> the<br />

process to higher pr<strong>of</strong>its by better informing the end-consumer’s purchase decision.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 231<br />

RESULTS: END-CONSUMER SURVEY<br />

From the end-consumer survey results the following points can be summarised as aspects concerning<br />

end-consumer attitudes to timber.<br />

A) Timber as a furniture material: End-consumers supported timber as a recyclable material but did<br />

not convincingly support timber as being environmentally safe. Timber was seen to be attractive to<br />

almost all respondents, and furniture construction material that was practical with inherent<br />

timeless qualities. Timber was seen as highly durable and possessing inheritable qualities with<br />

intergenerational appeal. There was some disconnection in terms <strong>of</strong> timber as fashionable with<br />

timber seen as youthful and a material <strong>of</strong> the future and also as an old-fashioned material with<br />

positive and negative associations possibly related to timber furniture design.<br />

B) Successful end-consumer timber attributes: A positive end-consumer attitude was expressed in<br />

timber possessing superior qualities as a furniture material in comparison to alternate materials.<br />

Emphasis was placed on attributes in five key areas including ecology, craftsmanship, touch<br />

characteristics, timber in use, and visual response characteristics. Successful visual characteristics<br />

account for the most favoured timber attributes with warmth, ageless, natural and good looking.<br />

The single most referred to timber attribute was the durability <strong>of</strong> timber.<br />

C) End-consumer buyer preferences: Respondents were highly to value the timber name and variety<br />

though the forest source location was not <strong>of</strong> considerable interest to end-consumers. Timber<br />

characteristics associated with timber quality were highly regarded by end-consumers as were<br />

levels <strong>of</strong> quality detail in timber exposing the qualities <strong>of</strong> the natural timber material with a high<br />

proportion in strong agreement.<br />

D) Brand character: Tradition linked to timber craftsmanship in production was respected by endconsumers<br />

though not overwhelmingly. Having knowledgeable retail sales staff was in demand<br />

and post-purchase furniture timber care was also sought-after. More than one visit to the retail<br />

showroom was essential for end-consumer prior to a furniture buying decision with cost and<br />

timber quality factoring highly in end-consumer decision-making.<br />

E) Timber certification: Selection <strong>of</strong> timber from approved forest sources was supported by half <strong>of</strong><br />

respondents with a quarter strongly agreeing that timber certification would influence their buying<br />

decision. Over seventy percent <strong>of</strong> end-consumers inaccurately described the time required for a<br />

hardwood timber to reach maturity prior to harvest, and were only moderately more successful in<br />

estimating s<strong>of</strong>twoods.<br />

RELATIONSHIP BETWEEN THE INDUSTRY AND END-CONSUMER SURVEY RESULTS<br />

There was encouragingly both interest from the industry level <strong>of</strong> the value chain toward end-consumer<br />

needs, and interest form the end-consumer level <strong>of</strong> the value chain towards the forestry and furniture<br />

manufacturing levels <strong>of</strong> the value chain.<br />

The end-consumers appear to possess a genuine interest in the variety <strong>of</strong> and qualities in timber but are<br />

less drawn to environmental issues than the industry level responses to the survey had expected. The<br />

industry has appeared to over-estimate the level <strong>of</strong> interest <strong>of</strong> the end-consumer in timber certification.<br />

In the industry position defence associated with timber certification, end-consumer responses in the<br />

survey also appear ambiguous toward timber certification where many respondents were undecided if<br />

timber certification would or would not affect their purchasing decisions. This may suggest that endconsumers<br />

may be experiencing a dilemma <strong>of</strong> environmental conscience where the end-consumer may<br />

need furniture for example but may be reluctant make a decision that may ‘adversely’ impact on a<br />

forest resource that appears to have a direct connection and resonance to overall planetary<br />

environmental health. This may be more a result <strong>of</strong> interpretations <strong>of</strong> general media than strategic<br />

long-range forestry management.<br />

The industry survey responses also linked more highly informed end-consumer purchasing decisions<br />

including forest source, timber variety, and qualities in craftsmanship with higher industry<br />

pr<strong>of</strong>itability. This was supported comprehensively in the end-consumer survey in the strong interest<br />

from end-consumers to understand the forest product intended for purchase from a timber variety and<br />

timber quality perspective. This is very encouraging and signifies precedence toward a level <strong>of</strong>


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 232<br />

cooperation on the key determinants in the process <strong>of</strong> development <strong>of</strong> new products. It also supports<br />

the value chain technique advancing as a legitimate evolutionary step <strong>of</strong> a well-planned and connected<br />

industry focussed on end-consumer value, and industry sustainability.<br />

DISCUSSION<br />

It is apparent that there is a strong emotional and empathy-to-nature-based connection to timber in<br />

general, shared at all levels <strong>of</strong> the value chain including the end-consumer. This is a positive<br />

implication for the future <strong>of</strong> targeted timber use for particular industries, as all parties in the value<br />

chain converge on the same level <strong>of</strong> agreement.<br />

This level <strong>of</strong> consensus may be represented by the follows points:<br />

A) Timber holds unique qualities over alternate materials in particular applications, including timber<br />

furniture products; and<br />

B) The use <strong>of</strong> timber in high pr<strong>of</strong>ile applications in close visual and textural contact with endconsumers<br />

holds long-term, endearing qualities which impact on longer term social,<br />

environmental and economic aspects that few alternate consumer products can provide over time.<br />

The barriers to maintaining industry competitiveness focused on the heightened future impairment <strong>of</strong><br />

finding buyers, but not suppliers. Further obstacles included a consistent, formal or informal, internal<br />

performance measurement system, despite cited confidence in efficient internal procedures which was<br />

inconsistent in the analysis <strong>of</strong> survey responses. Furthermore, a lack <strong>of</strong> access to market intelligence<br />

hindered firm-level planning and targeted investment.<br />

End-consumer survey phase returns in the furniture retail environment established consumer attitudes<br />

to timber as a furniture material. Both the industry survey and consumer surveys concurred that timber<br />

was seen a material that held superior qualities in comparison to other materials. Future timber<br />

furniture purchasing was also positively affected by timber certification but not comprehensively.<br />

The industry survey returned the view that consumers may be interested in the process <strong>of</strong> production<br />

from forest to retail, and timber related environmental issues. This was consistent with end-consumers<br />

though also focused on timber quality, craftsmanship, and how to care for timber. Most consumers had<br />

little idea how long a tree may take to grow to maturity to be <strong>of</strong> use in timber furniture manufacture,<br />

especially hardwood species. Only about half <strong>of</strong> end-consumer respondents preferred timber from an<br />

approved forest source leaving room for industry discussion.<br />

CONCLUSION<br />

This paper explored value creation in the <strong>Australia</strong>n timber furniture value chain and examined the<br />

level <strong>of</strong> synchronicity between the needs <strong>of</strong> end-consumers and value chain suppliers in what<br />

constitutes overall value for the end-consumer. The study establishes that in a fragmented industry<br />

despite high levels <strong>of</strong> trust and communication there is a search for leadership. A method by which the<br />

collective knowledge and expertise <strong>of</strong> the value chain is brought together to address competitive<br />

shortfalls is sought. It is not evident that the players in the timber furniture industry understand that<br />

they reside and participate within a value chain. As a consequence the industry may be unaware <strong>of</strong> the<br />

benefits that may be within arm’s reach in building competitive advantage.<br />

The gap in knowledge <strong>of</strong> the industry players namely the timber furniture manufacturers appears<br />

substantial. Their consistent request for access to more reliable buyers and access through buyers to<br />

the needs <strong>of</strong> the end-consumers is urgently required. Access to this vital information through focus on<br />

the entire value chain from forest to consumer may produce positive outcomes for all stakeholders.<br />

A mixed method research design approach to value chain thinking suggests that multiple level analysis<br />

<strong>of</strong> the industry and its consumer base is essential in understanding the attributes <strong>of</strong> the industry. This<br />

may assist in further understanding competition in local and global markets and how best to prepare to<br />

compete in those markets. By focusing on new value creation the paper indicates how the industry<br />

may synchronise the value needs <strong>of</strong> the industry and the end-consumer within the timber value chain.<br />

The targeting and prioritising the efficient use <strong>of</strong> forest resources toward the highest value-added<br />

yields per unit <strong>of</strong> material may support economic and environmental imperatives.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 233<br />

The focus on a high value added timber commodity such as timber furniture provides an opportunity<br />

to explore the application <strong>of</strong> value chain methodology. This methodology is much attributed in the<br />

building <strong>of</strong> competitive advantage in sectors outside wood and forest products industry and <strong>Australia</strong><br />

in general and can act as a point <strong>of</strong> industry reflection and change. Within the understanding <strong>of</strong> the<br />

consequence <strong>of</strong> all inputs and outputs in the creation <strong>of</strong> value throughout production and the endconsumer-led-decision-making<br />

process lie the answers to building sustainable competitive advantage<br />

and challenges for forest products industries.<br />

LIMITATIONS<br />

This study is exploratory and does not claim any larger population or industry generalisations other<br />

than supporting a theoretical model for end-consumer centred value creation within the timber<br />

furniture value chain. The study did not seek the participation <strong>of</strong> entire value chain based on time and<br />

resources precluding a wider sampling <strong>of</strong> all value chain level inputs and therefore a sample<br />

representing the timber furniture manufacturing level has been emphasised in its leadership role in<br />

decision making in the selected timber value chain.<br />

REFERENCES<br />

Barnes,J. and Morris,M., 2000. Value Chains, Production Pipelines and Systemic Efficiency. Authors’<br />

manuscript, South Africa.<br />

Barney,J.B., 2002. Gaining and Sustaining competitive advantage. New Jersey,USA: Prentice Hall.<br />

Bititci,U., et al 2004. Creating and managing value in collaborative networks. International Journal <strong>of</strong><br />

Physical Distribution and Logistics Management; 2004, 34,3/4; 251-268.<br />

Commonwealth <strong>of</strong> <strong>Australia</strong> and AEGIS, 1999. The <strong>Australia</strong>n Furnishings Industry: An Analysis.<br />

Commonwealth <strong>of</strong> <strong>Australia</strong>: Canberra.<br />

Creswell,J.W., and Plano Clark,V.L., 2007. Designing and conducting mixed methods research .<br />

Thousand Oaks, California. Sage Publications.<br />

FIAA, 2002. Furniture Trend 2002. Furnishing Industry Association <strong>of</strong> <strong>Australia</strong>. Sydney <strong>Australia</strong><br />

Morris, M., and Barnes,J. 1999. Using Production Pipelines as a research methodology for<br />

understanding competitiveness : a case study <strong>of</strong> an automotive plastics component. School <strong>of</strong><br />

Development Studies (incorporating CSDS) University <strong>of</strong> Natal, South Africa.<br />

Navarro,J., Hayward,P., and Voros,J. 2008. How to solve a wicked problem? Furniture foresight case<br />

study. Foresight, 2008; 10, 2; 11-29.<br />

Pakarinen, T.,1999 Success factors as wood as a furniture material. Forest Products Journal: 1999, 49,<br />

9; 79-85.<br />

Porter, M., (1985). Competitive advantage: creating and sustaining superior performance. New York<br />

:Free Press.<br />

Prahalad,C.K. and Ramaswamy.V. 2004. The future <strong>of</strong> competition: co-creating unique value with<br />

customers. Boston:Harvard University Press.<br />

United Nations, 2005. United Nations International Trade Database. http://comtrade.un.org/<br />

Woodruff,R.B. (1997). Customer Value: The next source for competitive advantage. Journal Academy<br />

<strong>of</strong> Marketing Science: Spring, Vol. 25, Issue 2;p139,15pgs.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 234<br />

CARIBBEAN PINE GRADED SAWN RECOVERY VARIATION<br />

IN A MATURE SPACING BY THINNING EXPERIMENT<br />

GROWN IN QUEENSLAND<br />

Kevin Harding 1 , Terry Copley 1 , Marks Nester 2 ,<br />

Kerrie Catchpoole 2 , Anton Zbonak 1<br />

ABSTRACT<br />

A mature Caribbean pine silviculture experiment provided initial square spacing<br />

treatments <strong>of</strong> 1.8 m 2 , 2.4 m 2 , 3.0 m 2 and 3.6 m 2 (equal to 3088, 1737, 1111 and 772<br />

stems/ha) that were thinned at age 10 years to 600, 400 and 200 stems/ha and an<br />

unthinned control. The trial was assessed for standing tree wood properties at 28 years<br />

and some <strong>of</strong> these trees were felled for a sawing study at 30 years.<br />

Dried dressed recovery (DDR) was strongly related to tree size and the volume ‘lost’ to<br />

defects tended to be lower in unthinned treatments. Ingrade recovery decreased with<br />

higher post-thinning stocking level for plots originally planted at 1111 and 772 spha<br />

and value per tree was highest in the lowest stocking treatment (772 spha). Log value<br />

decreased across all treatments consistently from butt to top log. The value per hectare<br />

was highest in unthinned plots due to high stem numbers per hectare. A complete<br />

economic analysis considering all cost structures is required to investigate the optimal<br />

silviculture required to maximise economic returns.<br />

INTRODUCTION<br />

In 2005, Forestry Plantations Queensland (FPQ) and the Wood Quality Improvement group within the<br />

Product Quality section <strong>of</strong> Horticulture and Forestry Science (H&FS), DPI&F, commenced a project<br />

to obtain samples from three mature silviculture/genetics trials to collect data on tree growth and wood<br />

quality. The main goal was to collect robust data from the three trials that would underpin wood<br />

property and growth modeling as part <strong>of</strong> an overall decision support system being developed by FPQ<br />

(Catchpoole et al. 2007). The data collection focused on within and between-tree variation to consider<br />

the impacts <strong>of</strong> different silvicultural factors on wood density, micr<strong>of</strong>ibril angle and juvenile wood<br />

proportion, both at breast height and at multiple heights within the merchantable stem. The<br />

experiments were targeted due to their maturity (20 to 28 years old), their design and their observed<br />

large differences in growth rates in response to the treatments applied. This variation provided<br />

excellent scope to evaluate and model the relationship between large productivity responses and wood<br />

traits.<br />

This paper focuses on the sawn recovery variables assessed in Pinus caribaea var. hondurensis<br />

experiment 532NC which was the only trial for which sawing was undertaken. This 28-year-old<br />

clearfall age experiment was intensively studied in a staged sampling program that provided:<br />

1. Standing tree non-destructive evaluations <strong>of</strong> wood properties in breast height increment cores and<br />

using standing tree acoustic assessment.<br />

2. Destructive sampling to examine up the stem variation with height including individual growth<br />

ring patterns <strong>of</strong> variation from pith to bark captured for key heights by SilviScan x-ray<br />

densitometric scans complemented by average 5-ring segment data at other tree heights.<br />

3. Sawn board graded recovery for 70 × 35mm structural product.<br />

Each <strong>of</strong> these assessment stages was based on fewer trees as the resources required and the expense <strong>of</strong><br />

the assessments increased. However, the selection <strong>of</strong> each subsample was informed by the results <strong>of</strong><br />

previous characterisations so as to select trees representative <strong>of</strong> the diameter distribution and acoustic<br />

properties observed in the experimental treatments.<br />

1<br />

Horticulture and Forestry Science, Department <strong>of</strong> Primary Industries and Fisheries, Queensland<br />

2<br />

Forestry Plantations Queensland


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 235<br />

MATERIAL<br />

Experiment 532NC was planted in 2/1977 at Toorbul (27°05'S 152°58'E) and was 28-years-old when<br />

wood sampled in 6/2005 and 30-years-old when sampled for sawing in 08/2007. The experiment was<br />

planted at six initial stockings: 772, 1111, 1310, 1370, 1737 and 3088 spha resulting from five square<br />

spacings (1.8 m 2 , 2.4 m 2 , 2.7 m 2 , 3.0 m 2 and 3.6 m 2 ) plus one rectangular (3.5 m x 2.1 m). It was<br />

thinned at age 10 years to 600 spha, 400 spha and 200 spha with a control treatment left unthinned.<br />

For the purposes <strong>of</strong> this study four initial stocking levels (772, 1111, 1737 and 3088 spha from 1.8 m 2 ,<br />

2.4 m 2 , 3.0 m 2 and 3.6 m 2 square spacings) were sampled as they provided extreme treatment levels<br />

and spacing comparisons with regular incremental increases across the treatments.<br />

For the sawing study, sampling was restricted to ten <strong>of</strong> the 16 spacing × thinning treatment<br />

combinations to meet logistical and resourcing requirements. The selection strategy used is detailed<br />

below.<br />

METHODS<br />

In each thinning treatment, approximately 105 trees were assessed for standing tree acoustic velocity,<br />

height, diameter at breast height and 12mm cores extracted giving a sample <strong>of</strong> 403 trees due to<br />

incomplete survival or unacceptable tree form, size (


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 236<br />

Sawing strategy<br />

All stems were cut to 4.8m log lengths to allow direct comparison <strong>of</strong> recovery with height across<br />

treatments. Top logs were also recovered and sawn to a small end diameter merchantable limit <strong>of</strong><br />

120mm DUB in a commercial length <strong>of</strong> 2.4, 3.0, 3.6 or 4.2m when the top log was shorter than 4.8m.<br />

Logs were sawn to maximise recovery <strong>of</strong> 70 × 35 mm framing dimension boards from centre cant/s<br />

and wings. All logs were sawn to only 70 × 35 mm framing dimension boards so non-structural<br />

recovery was not captured. The aim was to maximise the number <strong>of</strong> structural dimension boards<br />

recovered from each log and stem to provide a basis for comparison among and between treatments<br />

that would not be biased by product size differences among trees and logs. Standardising log length<br />

and product size should provide a sensitive basis for comparisons among trees and treatments.<br />

Analysis<br />

Correlation analysis at both individual tree level and treatment mean level was undertaken using the<br />

data analysis tools <strong>of</strong> GenStat (2002).<br />

RESULTS<br />

At the tree level, dried dressed recovery (DDR) ranged from 16% to 39%, with an average recovery <strong>of</strong><br />

33%. DDR for individual treatment levels are presented in Figure 1. Results indicate that the highest<br />

recovery is achieved from the treatment with the lowest initial stocking level (772 stems/ha). It is also<br />

clear that recovery increases with lower numbers <strong>of</strong> stems retained after thinning.<br />

Dried dressed recovery (%)<br />

38<br />

36<br />

34<br />

32<br />

30<br />

28<br />

26<br />

Unthin 400spha 600spha Unthin 400spha 600spha Unthin 200spha 400spha Unthin<br />

1 (3088) 2 (1737) 4 (1111) 5 (772)<br />

Initial and thinning stocking (stems/ha)<br />

Figure 1. Average dried dressed recovery by initial stocking and thinning treatment in<br />

Experiment 532NC<br />

Tree size/volume significantly affects DDR. In general terms the more intense thinning regimes tend<br />

to produce larger trees and is reflected in higher DDR values. The differences in DDR were more<br />

pronounced for the smaller trees (< 1m 3 volume) covering a range <strong>of</strong> nearly 20% DDR with lower<br />

differences observed in DDR among the larger trees (>1m 3 volume) with a range <strong>of</strong> around 10%.<br />

Lost volume due to defects<br />

During visual grading wood defects such as knots, splits, resin pockets, resin shakes, etc. were<br />

recorded. If the measured value <strong>of</strong> the defects exceeded the visual grading threshold values defined in<br />

the machine stress grading standard (AS/NZS 1748, 2006), the stick was docked to remove these<br />

defects. The initial low stocking treatment (772 spha) exhibited the highest percentage <strong>of</strong> lost volume,


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 237<br />

which is most likely associated with heavier branching and larger knot size in these trees. With the<br />

exception <strong>of</strong> treatment 772-unthinned, the unthinned plots showed lower values <strong>of</strong> PLV.<br />

Graded recovery<br />

All structural size sticks (70 × 35mm) were assessed for acoustic velocity using the CIRAD Forêt<br />

BING measurement system (Brancheriau and Baillères, 2003). Using the acoustic velocity properties<br />

with the stick average air-dry density and the maximum knot area ratio, an estimate <strong>of</strong> average<br />

modulus <strong>of</strong> elasticity (MoE) was calculated for each stick. Using these MoE estimates assignment <strong>of</strong><br />

sticks to MGP (Machine graded pine) classes was undertaken. Grade assignment starts with the<br />

highest values <strong>of</strong> MoE and assigning these sticks to an MGP grade population until its average reaches<br />

the desired threshold value for each grade (MGP15 = 15200MPa, MGP12 = 12700MPa, MGP10 =<br />

10000MPa, Utility < 10000MPa). The upper and lower limits <strong>of</strong> each grade were recorded and used to<br />

assign the grade for each stick in that population.<br />

Figure 2 shows an overall distribution <strong>of</strong> grades across all investigated treatments. The level <strong>of</strong> out-<strong>of</strong>grade<br />

(reject and utility) is approximately 47% when results for all treatments are combined. MGP 15<br />

grade makes up only 2.4% <strong>of</strong> the total recovery <strong>of</strong> all sticks. In reality, it is not practical to have a<br />

single grade with this low proportion so the recovery <strong>of</strong> MGP15 was also included in a separate<br />

analysis as part <strong>of</strong> the MGP12 population to examine impacts on total product value. This latter<br />

process also impacted the distribution <strong>of</strong> MGP12, MGP 10 and utility grades (Figure 2). All<br />

subsequent analysis undertaken is reported only for Option B (i.e. without an MGP 15 population).<br />

Proportion <strong>of</strong> grades (%)<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Incl MGP15 Excl MGP 15<br />

MGP15 MGP12 MGP10 Utility Reject<br />

Grade<br />

Figure 2. Proportion <strong>of</strong> machine graded pine classes for all stocking and thinning treatments<br />

combined with options <strong>of</strong> inclusion and exclusion <strong>of</strong> MGP15 grade.<br />

The overall distributions <strong>of</strong> product grades in Figure 3 differ considerably across the treatments. In<br />

general, the grade proportions suggest that initial high stocking (3088 and 1737 stems/ha) produced a<br />

higher quantity <strong>of</strong> MGP 12 grade. Additionally, the proportion <strong>of</strong> MGP12 grade recovery generally<br />

increases with stronger intensity <strong>of</strong> thinning as might be expected due to improved recovery <strong>of</strong> mature<br />

wood from the larger trees produced. However, although grade proportions are an important indicator<br />

<strong>of</strong> quality they must be considered with the total volume recovered to quantify the total economic<br />

potential <strong>of</strong> a treatment response.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 238<br />

Actual value per log (AUD)<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Unthin 400spha 600spha Unthin 400spha 600spha Unthin 200spha 400spha Unthin<br />

1 (3088) 2 (1737) 4 (1111) 5 (772)<br />

Initial and thinning stocking (stems/ha)<br />

Figure 3. Proportion <strong>of</strong> grade classes across different stocking and thinning treatments in<br />

Experiment 532NC<br />

Economic value recovery<br />

To consider the economic importance <strong>of</strong> the grade recovery distribution across treatments the value <strong>of</strong><br />

recovery by product was estimated. Product value was assigned to each stick based on its grade using<br />

the wholesale price estimates per m 3 provided in Table 2.<br />

Table 2. Approximate December 2008 wholesale s<strong>of</strong>twood product prices (AUD/m 3 ) for 70mm<br />

structural product grades and length<br />

MGP15 MGP12 MGP10<br />

Grade Length Length Length Length Length Length Utility<br />

>4.6m 2.4-4.6m >4.6m 2.4-4.6m >4.6m 2.4-4.6m<br />

Price<br />

AUD/m 3 $729 $663 $605 $582 $501 $456 $247<br />

Value recovery at log level<br />

Figure 4 presents the distribution <strong>of</strong> actual value <strong>of</strong> log along the longitudinal tree axis and across<br />

treatments. In all treatments the butt log is the most valuable part <strong>of</strong> the tree and total log value<br />

declines up the stem. The gradient <strong>of</strong> value decline is not uniform across treatments. Most notable is<br />

the steep difference observed in the fastest grown treatment, 772-200 stems/ha. On the other hand the<br />

unthinned compartment with the highest initial stocking level (3088 stems/ha) exhibited a moderate<br />

change in log value with log position in tree. It can be expected that these trends in part reflect branch<br />

size and therefore knot size impacts on grade.<br />

Value recovery at tree level<br />

The full distribution <strong>of</strong> DBHOB classes within each individual plot was used to calculate a weighted<br />

average value recovery at tree level across different spacing and thinning treatments. Using this<br />

approach the effect <strong>of</strong> influential individuals with low representation within the sample distribution is<br />

reduced. Figure 5 shows differences in value recovery for simple average and weighted average<br />

values. In general, unthinned plots with a higher initial stocking level provide less valuable trees than<br />

compartments with less initial stocking and more intense thinning regime. This is to be expected with<br />

average tree size increasing with more intensive thinning and providing improved recovery <strong>of</strong> better<br />

quality mature wood timber from the larger average log size.<br />

butt<br />

2<br />

3<br />

4<br />

5


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 239<br />

Proportion <strong>of</strong> grades<br />

100%<br />

80%<br />

60%<br />

40%<br />

20%<br />

0%<br />

Unthin<br />

1<br />

(3088)<br />

400spha<br />

600spha<br />

Unthin<br />

400spha<br />

600spha<br />

Unthin<br />

200spha<br />

400spha<br />

2 (1737) 4 (1111) 5 (772)<br />

Initial and thinning stocking (stems/ha)<br />

Unthin<br />

Reject<br />

Utility<br />

MGP10<br />

MGP12<br />

Figure 4. Value recovery at log level across different stocking and thinning treatments in<br />

Experiment 532NC.<br />

Value recovery/tree (AUD)<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

Averaged value/tree Weighted average value/tree<br />

Unthin 400spha 600spha Unthin 400spha 600spha Unthin 200spha 400spha Unthin<br />

1 (3088) 2 (1737) 4 (1111) 5 (772)<br />

Initial and thinning stocking (stems/ha)<br />

Figure 5. Comparison between average value recovery and weighted average value recovery at<br />

tree level across different stocking and thinning treatments in Experiment 532NC.<br />

Value recovery prediction for per hectare comparisons<br />

The experiment results were used to extrapolate to per hectare value comparisons by multiplying<br />

weighted average tree value with utilizable number <strong>of</strong> trees per hectare taking into account survival<br />

and tree size. Only trees meeting a minimum DBHOB <strong>of</strong> ≥ 22cm were considered merchantable.<br />

These comparisons are presented in Figure 6. Unthinned plots, despite lower average tree value,<br />

provide higher actual value on a per hectare basis than plots with high thinning intensity. The latter is<br />

unsurprising as it reflects numbers <strong>of</strong> trees and biomass on a site and is divorced from the commercial<br />

realities <strong>of</strong> establishment costs including costs per tree planted, and market requirements for both


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 240<br />

piece size in log and sawn form that impact harvesting, haulage and processing cost structures and<br />

product size requirements. The 600 spha post-thinning treatments in the higher initial stocking<br />

treatments (1737 and 1111 spha) and the 400 spha treatment in the low initial stocking (772 spha)<br />

treatment display intermediate returns. These results suggest that economic modelling <strong>of</strong> returns<br />

taking into account all establishment, harvesting and processing costs should provide important<br />

insights in to the optimal establishment and final stocking regimes needed to maximise value returns.<br />

Value recovery per hectare (AUD)<br />

100000<br />

80000<br />

60000<br />

40000<br />

20000<br />

0<br />

Unthin 400spha 600spha Unthin 400spha 600spha Unthin 200spha 400spha Unthin<br />

1 (3088) 2 (1737) 4 (1111) 5 (772)<br />

Initial and thinning stocking (stems/ha)<br />

Figure 6. Value recovery per hectare comparison across stocking and thinning treatments in<br />

Experiment 532NC for stems ≥ 22 cm DBHOB.<br />

Correlation analysis<br />

The matrices in Tables 3 and 4 (Appended) show the directions and strength <strong>of</strong> the correlations among<br />

investigated properties at individual tree level (all 56 trees sawn, ignoring treatment) and at treatment<br />

mean level combining 5-6 trees from each thinning treatment into single average values (10 data<br />

points). The results revealed that the tree growth properties such as DBHOB, tree height and tree<br />

volume were significantly correlated with almost all sawing properties. The most notable correlation<br />

(r=0.98) was found between tree volume and actual dollar value <strong>of</strong> graded products per tree. This is<br />

not surprising since increased dried dressed recovery would be expected as tree size increases. Larger<br />

trees should also produce a higher proportion <strong>of</strong> good quality grades from improved recovery <strong>of</strong><br />

mature wood as a proportion <strong>of</strong> overall recovery which is then reflected in overall higher mean value<br />

<strong>of</strong> structural graded products.<br />

There were no significant correlations between tree growth characteristics and basic density at<br />

individual tree level but they emerged at treatment level showing that density is positively related with<br />

DBHOB and tree volume. This reflects the balance <strong>of</strong> juvenile wood versus mature wood increment<br />

with lower proportions <strong>of</strong> higher density mature wood in the unthinned plots with slower growth rates<br />

compared to heavily thinned stands which produce bigger trees from better growth rates throughout<br />

the life <strong>of</strong> the stand.<br />

Acoustic velocity as assessed using an ST300 device was poorly correlated with all other investigated<br />

properties. The only significant correlation found for acoustic velocity was with MOE measured on<br />

framing sticks (r=0.49) and with the proportion <strong>of</strong> MGP12 and MGP10 grades combined (r=0.42) at<br />

individual tree level. However, it should be noted that ST300 velocity is used to predict wood stiffness<br />

in the outer wood <strong>of</strong> the tree which is then mostly lost during sawing in waney-edge boards.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 241<br />

Basic density <strong>of</strong> core samples was weakly correlated with dried dressed recovery (r=0.32 at individual<br />

tree level; r=0.80 at treatment level), and the proportion <strong>of</strong> MGP12 and MGP10 grades recovered.<br />

There was a strong relationship between average core density and average framing stick density<br />

(r=0.77) as might be expected.<br />

CONCLUSIONS<br />

This sawing study has produced very encouraging results which suggest that returns may be optimised<br />

by silvicultural regimes that are within the spectrum <strong>of</strong> current commercial practice. Initial planting at<br />

772 to1111 spha with thinning to between 400 and 600 spha produces relatively good dried dressed<br />

recovery percentages, structural grade distribution, log value and predicted product value per hectare..<br />

Although the results suggest that higher stocking regimes can produce higher average log and tree<br />

values and returns per hectare, these are unlikely to be economically viable once all establishment,<br />

harvesting and processing costs are considered, as well as the product requirements (piece size and<br />

grade distribution) <strong>of</strong> processors.<br />

ACKNOWLEDGEMENTS<br />

This work was funded by Forestry Plantations Queensland and represents the planning and<br />

collaborative efforts <strong>of</strong> a large team <strong>of</strong> colleagues from both FPQ and DPI&F Innovative Forest<br />

Products including: Ian Last and Paul Keay (FPQ) and Megan Prance, Martin Davies, Bill Atyeo,<br />

Adam Redman, Rapheal Tschupp, Dr Henri Bailleres, Eric Littee and Rob McGavin (Innovative<br />

Forest Products group, DPI&F).<br />

REFERENCES<br />

Brancheriau, L. and H. Baillères (2003) Vibration Spectra as Grading Technique for Structural<br />

Timber. Holzforschung 57: 644–652.<br />

Catchpoole, K.J., Nester, M.R. and Harding , K.J. (2007) Predicting wood value in Queensland<br />

Caribbean pine plantations using a decision support system. Aust. J. Forestry 70(2): 120-124.<br />

GenStat (2002) GenStat for Windows, version 6.1 Lawes Agricultural Trust.<br />

Standards Association <strong>of</strong> <strong>Australia</strong> and Standards Association <strong>of</strong> New Zealand. (2006) Timber -<br />

Mechanically stress-graded for structural purposes. AS/NZS 1748, Standards <strong>Australia</strong>/Standards<br />

New Zealand, Sydney and Wellington 12 pp.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 242<br />

Table 3. Correlation coefficients with p-level value at individual tree level. Trees from all sampled treatments combined (only statistically<br />

significant coefficients with p-level


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 243<br />

Table 4. Correlation coefficients with p-level value at treatment mean level (only statistically significant coefficients with p-level


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 244<br />

REHABILITATION OF FORESTS IN DECLINE:<br />

MT. LINDESAY STATE FOREST<br />

Peter St.Clair 1<br />

ABSTRACT<br />

Forests suffering decline were rehabilitated using a combination <strong>of</strong> harvesting, burning,<br />

planting and herbicide treatments at Mt. Lindesay State Forest in Northern NSW. The<br />

severity <strong>of</strong> the decline ranged from slight through to severe, where, in the latter, there<br />

were few forest trees alive, little eucalypt regrowth, a dense understorey <strong>of</strong> lantana, and<br />

abundant colonies <strong>of</strong> bell miners (Manorina melanophrys, Meliphagidae). Yellow<br />

bellied gliders and koalas were the only arboreal mammals found at the site. Two years<br />

after treatment the forest contained healthy regrowth, much reduced lantana infestation,<br />

reduced bell miner populations and no measurable change to yellow bellied glider and<br />

koala populations. Tree crown health improved significantly and bell miner populations<br />

decreased.<br />

INTRODUCTION<br />

Eucalypt decline and shrub encroachment in the absence <strong>of</strong> fire is a widespread and increasing<br />

problem in <strong>Australia</strong> (Anon. 2007) and it usually involves a range <strong>of</strong> secondary pests, pathogens and<br />

parasites <strong>of</strong> eucalypts (Jurskis 2005). Near Urbenville in north-eastern NSW, the area <strong>of</strong> declining<br />

forest increased from less than 1,000 ha in 1992 to at least 20,000 ha in 2004 (Jurskis 2005).<br />

Declining forests in this region typically have a dense shrubby understorey <strong>of</strong>ten dominated by<br />

lantana (Lantana camara), and dying overmature, mature and pole stage eucalypts infested with sapsucking<br />

insects called psyllids and high numbers <strong>of</strong> bell miners. This form <strong>of</strong> decline has been<br />

termed Bell Miner Associated Dieback (BMAD). There are basically two divergent models <strong>of</strong> the<br />

decline process (e.g. Stone 1996, 1999, 2005, Jurskis and Turner 2002, Jurskis 2005, Turner and<br />

Lambert 2005, Stone et al. 2008, Turner et al. 2008). Jurskis and others propose that exclusion <strong>of</strong><br />

fire from open forests causes changes in soils which adversely affect the roots <strong>of</strong> established<br />

eucalypts and promote plants and animals that attack or compete with the trees. Stone and others<br />

alternatively consider that disturbance <strong>of</strong> forest structure allows shrub encroachment which provides<br />

nesting habitat for bell miners. Bell miners exclude other psyllid-eating birds resulting in elevated<br />

psyllid populations which eventually kill the trees.<br />

The BMAD specific model proposed by Stone and others cannot explain why decline involving<br />

psyllids in some forest types can occur in the presence or absence <strong>of</strong> disturbance and/or bell miners,<br />

and it cannot explain why a similar process affects a wide range <strong>of</strong> undisturbed forests in most <strong>of</strong><br />

<strong>Australia</strong> where there are no bell miners. Research at Richmond Range and at Jilliby by Stone and<br />

others confirmed many elements <strong>of</strong> the model proposed by Jurskis and others. Canopy decline was<br />

associated with infrequent fire, dense shrubbery, low and wet positions, elevated soil nitrogen, and<br />

areas relatively undisturbed by recent logging, whereas young regrowth stands were healthy. This<br />

and other research also showed that psyllid plagues and decline <strong>of</strong> susceptible eucalypts are not<br />

constantly associated with the presence <strong>of</strong> bell miners (e.g. Moore 1961, Stone et al. 1995, Stone<br />

and Simpson 2006, Dare et al. 2007, Jurskis 2008). It is clear from long term observations <strong>of</strong><br />

eucalypt decline involving bell miners and from the high stockings <strong>of</strong> dead stags in many <strong>of</strong> the<br />

declining stands, that complete canopies in undisturbed, fully stocked stands are gradually reduced<br />

as crowns become thin and die, and under-storey density increases due to increased sunlight and<br />

water availability. Elevated soil N was also correlated with decline in eucalypt plantations near<br />

Kyogle (Mews 2008). To improve tree health, the challenge is to control the understorey density and<br />

soil N content.<br />

1 Forests NSW, NSW Department <strong>of</strong> Primary Industries, PO Box 688, Casino NSW 2470. Email: peters@sf.nsw.gov.au


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 245<br />

Culling <strong>of</strong> bell miners and removal <strong>of</strong> their watering points have been trialled to ameliorate forest<br />

decline. On the South Coast <strong>of</strong> NSW 1500 birds were removed resulting in rapid return <strong>of</strong> other bird<br />

species, and some improvement in tree health, but subsequent re-colonisation by bell miners and<br />

deteriorating tree health. Removal <strong>of</strong> a dam at Toonumbar had no measureable effect on bell miner<br />

numbers (Sommerville pers. com.). Bell miners can accelerate the decline process by excluding<br />

other birds that would have a damping effect on psyllid populations, but they are now considered by<br />

the BMAD Working Group to be a contributor rather than a cause <strong>of</strong> decline. However bell miner<br />

populations are easy to monitor, and may be used as a surrogate for psyllid numbers and an indicator<br />

<strong>of</strong> the effects <strong>of</strong> rehabilitation treatments on the forest decline process.<br />

Combinations <strong>of</strong> harvesting, cultural treatments and fire regimes were trialed to improve and<br />

maintain the health <strong>of</strong> deteriorating stands and to rehabilitate severely declining stands. Objectives<br />

<strong>of</strong> the project were:<br />

1. Lantana cover reduced to less than 15%<br />

2. Increased health <strong>of</strong> retained trees<br />

3. Decrease in abundance <strong>of</strong> bell miners (An indication <strong>of</strong> reduced habitat or food)<br />

4. Maintenance <strong>of</strong> grassy understoreys<br />

5. Restoration <strong>of</strong> severely degraded stands with natural regeneration, supplementary seeding<br />

and enrichment planting <strong>of</strong> native over-storey species<br />

6. Integration <strong>of</strong> harvesting and rehabilitation<br />

METHODS<br />

Site description<br />

The trial was conducted in Mt. Lindesay SF (Cpts 276, 279, 280) in the head <strong>of</strong> the Richmond River<br />

catchment. The trial area <strong>of</strong> 536 ha was considered sufficiently large to minimise edge effects,<br />

accessible, and substantially affected by decline. The Chocolate soils contain about 45% clay and are<br />

derived from Tertiary Lamington Volcanics weathered over the Walloon Coal Measures (sandstone,<br />

siltstone and mudstone) with a small area <strong>of</strong> Marburg sandstone in the north <strong>of</strong> Compartment 279.<br />

Forest decline is prevalent on the Walloon Coal Measures (Jurskis 2004). Annual rainfall is about<br />

1400mm with most falling January to April. Mean maximum temperatures range from 17 0 C in July<br />

to 29 0 C in January and mean monthly minima range from 3 0 C to 16 0 C. Prior to 1950’s the forests<br />

were extensively grazed and burnt to maintain grass with only a small number <strong>of</strong> high quality logs<br />

removed. During 1974-77 large high quality logs (>40cm dub) were selectively removed. The area<br />

was logged again during 1983-84 for low quality logs and durable poles and girders, yielding<br />

1100m 3 (3 m 3 ha -1 ). Parts <strong>of</strong> Cpt 280 were harvested during 1996. The roadside <strong>of</strong> Cpt 276 was<br />

harvested in the mid 80’s for the Mt. Lindesay Highway realignment.<br />

The stands ranged from low to high quality, with variable structure including patches <strong>of</strong> regrowth <strong>of</strong><br />

various age and mature stands. All forest types except Brush Box (Lophostemon confertus) were<br />

declining, though the severity varied widely, and regeneration in most recently disturbed areas was<br />

stifled by lantana. The drier forest types originally had a grassy understorey but had over recent<br />

decades developed an impenetrable lantana understorey (Weaver 2006 pers. com.). There was only<br />

one record <strong>of</strong> fire (in 1994) during the preceding 20 years, burning 40 ha on Hildebrands Rd. Nine<br />

forest types are mapped within the compartments. Areas <strong>of</strong> Narrowleaved White Mahogany-Red<br />

Mahogany-Grey Ironbark occur along ridge tops and upper slopes along with Grey Box, Grey Gum-<br />

Grey Ironbark-White Mahogany, Sydney Blue Gum, Tallowood-Sydney Blue Gum and Rock. Mid<br />

to lower slopes are dominated by Brush Box or Flooded Gum or rainforest, and Forest Red Gum<br />

dominates low flats.<br />

Treatments<br />

Variable forest health and structure as well as a diversity <strong>of</strong> forest types necessitated a variety <strong>of</strong><br />

silvicultural treatments. A Silvicultural Decision Tree (SDT) was developed to allocate treatments<br />

across the trial area (Figure 1). Treatments included a combination <strong>of</strong> harvesting, selective shrub<br />

removal (mainly Lantana and Cissus), ground preparation, burning, planting, weed control, and<br />

reintroduction <strong>of</strong> low intensity fire in appropriate areas. Due to dense lantana understorey it was not


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 246<br />

physically possible to determine in advance the treatment to be applied at a fine scale. The location<br />

<strong>of</strong> all treatments was mapped using GPS and detailed records were maintained for each discrete unit.<br />

Harvestable<br />

trees<br />

YES<br />

No<br />

Harvestable<br />

trees<br />

IS THERE ESTABLISHED REGENERATION?<br />

(>150spha>4m high)<br />

Harvest as<br />

Forest Types Forest Types<br />

per IFOA 46,47,48 60,62,65<br />

Top disposal<br />

LIB = Low Intensity burning<br />

Push, rip,<br />

spray, burn,<br />

plant 750-<br />

1500spha.<br />

LIB<br />

Follow up<br />

Lantana<br />

control<br />

Machine<br />

disturbance<br />

to assist LIB<br />

Enrichment<br />

plant<br />

Figure 1. Silvicultural decision tree used for rehabilitation area at Mt. Lindesay SF in 2007<br />

Monitoring and Methodology<br />

Sixty plots were established in treated and control areas with stratification based on broad forest<br />

types (Dry and Wet as defined Forest Commission <strong>of</strong> NSW Research Note 17 and classified as per<br />

Table 1).<br />

Table 1. Allocation <strong>of</strong> monitoring plots in Mt. Lindesay SF Rehabilitation Trial in 2007<br />

Location Treatment Area Control Area<br />

Dry Forest Types(FT 60, 62 and 65) 20 Plots 10 Plots<br />

Moist Forest Types(FT 46, 47 and 48) 20 Plots 10 Plots<br />

Plots were located on a grid randomly oriented across the area. Plots were not established within<br />

harvest exclusion areas or equivalent areas in controls. Control plots were established in<br />

Compartment 280 immediately east <strong>of</strong> the treatment area to minimise edge effects, facilitate fire<br />

exclusion from the control area, maximise efficiencies for planning, harvesting and rehabilitation<br />

treatments and to allow for the trial to continue for 15 years.<br />

No fuel management or prescribed burning will be undertaken in the control area for the 15 years <strong>of</strong><br />

the trial. Monitoring methodology was according to agreed minimum standards (BMAD WG 2007).<br />

Lantana cover was estimated for the whole plot, individual tree crown score by Stone et al. (2008), a<br />

3 minute bird score on a 30m radius (Craig and Roberts 2002). Fire and harvesting intensities were<br />

estimated from the amount <strong>of</strong> disturbance or residual ground cover vegetation left in the plot.<br />

Mammal surveys were conducted as per Threatened Species Licence protocol.<br />

NO<br />

Seed Trees<br />

8pha<br />

Eucalypts<br />

top disposal<br />

Harvest<br />

additional as<br />

per IFOA<br />

Brushbox<br />

top disposal<br />

as required


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 247<br />

RESULTS<br />

Application <strong>of</strong> Treatments<br />

Harvesting was conducted over the period May to September 2007 and 9260 m 3 were removed from<br />

the net harvest area <strong>of</strong> 425 ha at an average <strong>of</strong> 22 m 3 ha -1 . This yield included 2650 m 3 <strong>of</strong> salvage<br />

logs and 900m 3 <strong>of</strong> logs destined for hardboard manufacture. Gross royalty was $392 000 at an<br />

average <strong>of</strong> $42 m -3 or $922 ha -1 with product royalties ranging from $218 m 3 to $2.40 m 3 .<br />

Herbicide was sprayed along roadsides, snig tracks and drainage depressions to generate drier fuel<br />

for ignition. Herbicide was also used after harvesting and burning to control accessible outbreaks or<br />

residual clumps <strong>of</strong> lantana. Burning was conducted between 16/10/2007 and 25/10/2007. Weather<br />

indices are shown in Table 2.<br />

Table 2. Fire weather conditions at Mt. Lindesay SF during burning operations in 10/ 2007<br />

Date<br />

Oct<br />

2007<br />

Max<br />

Temp<br />

( o C)<br />

Min<br />

RH<br />

(%)<br />

Wind<br />

Direct n<br />

Wind<br />

(km/hr)<br />

16 30 24 W 5 80 9 14<br />

17 24 38 SE 10 82 9 9<br />

22 28 35 NE 5 85 9 10<br />

25 27 40 NE 5 86 9 8<br />

BKDI DF FDI Comment<br />

Western Area<br />

115 ha, 15 plots<br />

Southern area<br />

30 ha, 8 plots<br />

Eastern area<br />

85 ha, 5 plots<br />

Nth Hwy<br />

25 ha, 1 plot<br />

Rain commenced on the evening <strong>of</strong> 25 th October 2007 and continued incessantly for 6 months<br />

eliminating further burning opportunities. Of the 40 plots planned for burning, 11 were not burnt for<br />

various reasons, 21 experienced intense fire and the balance lower fire intensities. Of the 40 plots 8<br />

were not harvested, and 6 had minor harvesting disturbance. These were mostly not burnt. The most<br />

intensively harvested plots experienced higher fire intensities. Rainfall was favourable for the<br />

establishment <strong>of</strong> planted seedlings and natural regeneration <strong>of</strong> vegetation on disturbed sites. In the<br />

period Oct 2007-Feb 2008, 75000 eucalypt or brush box seedlings were established in about 34 ha <strong>of</strong><br />

clearings and about 62 ha <strong>of</strong> sparse stands requiring enrichment planting.<br />

Integration <strong>of</strong> harvesting and rehabilitation<br />

Harvesting disturbance facilitated rehabilitation <strong>of</strong> lantana infested forest by opening up tracks and<br />

creating dry fuel for burning. After harvesting according to the SDT it was possible to identify sites<br />

for selective clearing prior to burning. These areas were burnt at the same time as general logging<br />

slash. Attempts to burn green lantana on lower slopes near drainage lines where harvesting was not<br />

conducted were unsuccessful because the fuels were green and access was difficult.<br />

Lantana cover<br />

Lantana cover was lower in 2008 than in 2007 in all treatments and recovered slightly in 2009 in all<br />

treatments except the controls. The recovery was significant in the moderate and intense fire<br />

treatments and in all treatments combined (Figure 2). There was no significant difference in the<br />

lantana cover in unburnt and minor fire treatments before and after treatment. The more intense fire<br />

treatment had the greatest pre-treatment lantana cover (80%) and following fire had the lowest (5%).<br />

The moderate and intense fire treatments, and all fire treatments combined showed some recovery <strong>of</strong><br />

lantana cover in 2009 compared with 2008 (Figure 2). In 2009 the most intense harvesting treatment<br />

had significantly lower lantana cover than less intense harvesting or treatments that were not<br />

harvested (data not presented). The weighted average lantana cover for all plots was 14.7% in 2009<br />

compared with 66% in 2007 before treatments were applied (Table 3).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 248<br />

Within three months after more intense fire and harvesting treatments there was a proliferation <strong>of</strong><br />

‘fire weeds’ such as inkweed (Phytolacca octandra), yellow weed (Sigesbeckia orientalis),<br />

nightshade (Solanum nigrum), fire fern (Doodia aspera), native grasses, poison (aka native) peach<br />

(Trema aspera), native hibiscus (Hibiscus heterophyllus), bleeding heart (Homalanthus<br />

populifolius), tobacco bush (Solanum mauritianum), wattles, and some eucalypt regeneration. The<br />

exotic weeds scotch thistle (Circium vulgare), ragweed (Ambrosia artemisifoila), noogoora burr<br />

(Xanthium occidentale) and purple-top verbena (Verbena bonariensis) also established. The rapid<br />

growth and density <strong>of</strong> these fire succession species was almost sufficient to eliminate lantana and<br />

cissus from intensely burnt plots.<br />

Average % lantana cover<br />

(incl std error)<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Average % lantana cover according to treatment,<br />

Mt Lindesay State Forest Project Area 2007-2009<br />

Control<br />

Harvested/unburnt<br />

Minor fire<br />

Moderate fire<br />

Treatment type<br />

Intense fire<br />

All fire treatments<br />

2007<br />

Census<br />

2008<br />

Census<br />

2009<br />

Census<br />

Figure 2. The effect <strong>of</strong> fire<br />

treatments on the average lantana<br />

cover following harvesting and fire<br />

treatments imposed after the 2007<br />

census. (n=60)<br />

Table 3. Weighted average <strong>of</strong> lantana cover <strong>of</strong> plots in the Mt. Lindesay project area in March 2007<br />

and 2009 following fire treatments in Oct 2007.<br />

Fire Treatment Nil 1- Minor 2-Moderate 3-Intense Total<br />

#Plots 11 3 4 22 40<br />

Plot % 27.5 7.5 10 55 100<br />

2007 Lantana % 52 46 43 80<br />

2007 weighted % 14.3 3.4 4.3 44 66<br />

2009 Lantana % 28 21 21.5 6<br />

2009 weighted % 7.7 1.6 2.1 3.3 14.7<br />

Health <strong>of</strong> Retained Trees<br />

Tree canopy health improved in both controls and treated plots from 2007 through to 2009. The<br />

health <strong>of</strong> trees in treated plots was lower than trees in control plots at each measure (Figure 3).<br />

Average canopy condition score (incl std<br />

error)<br />

25<br />

20<br />

15<br />

10<br />

(un)Treated<br />

2007<br />

Average canopy condition score for all plots,<br />

Mt Lindesay State Forest Project Area 2007-2009<br />

Treated 2008 Treated 2009 Control 2007 Control 2008 Control 2009<br />

Census date and treatment type<br />

Figure 3. Crown canopy condition score <strong>of</strong><br />

treated and untreated plots as measured in<br />

March 2007 prior to harvesting (April-<br />

September 2007) and burning treatment<br />

(Oct 2007) and measured again in March<br />

2008 and March 2009. Treatments n=40,<br />

Control n=20. Crown score system is that<br />

published by Stone et al. (2008) and was<br />

calculated for all trees retained as at<br />

March 2009 compared with their<br />

corresponding crown score condition in<br />

March 2007 prior to treatments being<br />

applied.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 249<br />

Bell miner numbers<br />

Bell miner numbers were lower than in controls immediately after treatment but had recovered to<br />

similar levels to controls in 2009 (Figure 4). Bell miner populations in the controls were much lower<br />

in 2008 and 2009 compared with 2007. The bell miner score was strongly correlated with the lantana<br />

cover (Figure 5).<br />

Average bell miner bird count<br />

class (incl std error)<br />

Average bell miner bird count class for control and treated<br />

plots, Mt Lindesay State Forest Project Area 2007-2009<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

2007 census 2008 census 2009 census<br />

Census year<br />

control<br />

treated<br />

Figure 4. Average score <strong>of</strong> bell miners in<br />

control and treated plots treated with<br />

harvesting and burning after the 2007<br />

census. Bell miner score 0= nil birds. 1<br />

=1-5 birds, 2 =5-20 birds.<br />

Figure 5. Relationship between<br />

lantana cover and bell miner numbers.<br />

Data averaged for each <strong>of</strong> 6 censuses<br />

where n=60.<br />

Use <strong>of</strong> low intensity burning<br />

There was only about 2 ha <strong>of</strong> grassy forest remaining in the study area. Trees in this area were<br />

healthy, part <strong>of</strong> a Highway visual protection zone and not <strong>of</strong> a size and quality for harvesting. This<br />

area was burnt and maintains its grassy appearance. Some other nearby forests that still exhibit<br />

extensive grassy areas were burnt during 2007 and 2008. A single low intensity fire did not kill<br />

lantana.<br />

Eucalypt and Brush Box Regeneration<br />

Regeneration was stimulated by fire and averaged 800-1000 stems ha -1 and was marginally increased<br />

at the higher level <strong>of</strong> harvesting disturbance (Figure 6).<br />

Number seedlings per hectare (incl<br />

std error)<br />

Average regeneration (stems per hectare) by fire<br />

treatm ent, Mt Lindesay State Forest Project Area 2009<br />

2000<br />

1800<br />

1600<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

0 1 2 3<br />

Fire Intensity<br />

Number seedlings per hectare<br />

(incl std error)<br />

Average regeneration (stems per hectare) by harvest<br />

treatment, Mt Lindesay State Forest Project Area 2009<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

0 1 2 3<br />

Harvest intensity<br />

Figure 6. Effect <strong>of</strong> fire and harvesting treatments in 2007 on counted regeneration <strong>of</strong> eucalypt<br />

seedlings in plots in March 2009. 1=Minor, 2=Moderate, 3=Intense<br />

Regeneration was however variable as indicated by the standard error bars. Highest retained<br />

stocking coincided with lowest regeneration (Table 5). Brush Box regeneration was two orders <strong>of</strong><br />

magnitude greater than the eucalypts (Figure 7). Planted seedlings were found in 6 plots. All<br />

showed good survival and early growth. Planted seedlings in the largely cleared areas ranged in


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 250<br />

stocking from 150 to 1400 stems ha -1 whilst in enrichment planting areas they were difficult to<br />

locate and distinguish from natural regeneration as they were not planted in a grid pattern.<br />

Table 5. New regeneration in plots and retained stocking and basal area <strong>of</strong> trees with<br />

diameter at breast height over bark greater than 10cm for corresponding plots at<br />

Mt. Lindesay SF<br />

Stocking stems ha -1 No<br />

regen<br />

0-100 101-250 251-500 501-1000 >1000 Total<br />

#Plots 13 8 5 2 3 9 40<br />

% plots 32.5 20 12.5 5 7.5 22.5 100<br />

Retained stocking<br />

stems ha -1<br />

202 194 175 88 108 139<br />

Retained BA m 2 ha -1 20.1 12.7 9.7 5.1 22.7 11.2<br />

Number seedlings (log scale)<br />

Total natural regeneration by species following fire, Mt Lindesay State Forest Project Area<br />

2007-2009<br />

1000<br />

100<br />

10<br />

1<br />

Brushbox<br />

Sydney blue gum<br />

White mahogany<br />

Grey gum<br />

Forest redgum<br />

Tallowood<br />

Species (common names)<br />

Red mahogany<br />

Bloodwood<br />

Grey ironbark<br />

Figure 7. Total natural<br />

regeneration by species<br />

in 40 plots <strong>of</strong> aggregate<br />

area <strong>of</strong> 1.6 ha following<br />

treatments in 2007 as<br />

measured in plots in<br />

March 2009. Note log<br />

scale.<br />

Rehabilitation costs<br />

Costs include tractor time, herbicide, plants and planting, fertiliser, fire trail construction, burning,<br />

planning and supervision. In moderate and severely declining stands site preparation is the main<br />

fixed cost and plants/planting and fertiliser cost were a variable cost <strong>of</strong> $1 per seedling. Estimated<br />

total cost for the treatments in this project were $189000, most expense going into the moderately<br />

and severely impacted areas (Table 6).<br />

Table 6. Estimated total rehabilitation cost (AUD) at Mt. Lindesay SF, 2007<br />

Decline Rating Area<br />

(ha)<br />

0 Healthy<br />

Full regeneration<br />

1/2 Slight. Minor Lantana<br />

EnrichmentPlanting<br />

3 Moderate. Lantana.<br />

Enrichment planting<br />

4 Severe. Heavy lantana.<br />

No regeneration<br />

Planted<br />

seedlings/ha<br />

Total<br />

seedlings<br />

Rehab n cost<br />

$/ha<br />

Total<br />

$ cost<br />

229 0 0 100 22900<br />

100 100 10000 200 20000<br />

62 500 31000 1000 62000<br />

34 1000 34000 2500 85000<br />

TOTAL 425 75000 189000<br />

Impacts on mammals<br />

Following harvesting and burning in April to October 2007 numbers <strong>of</strong> arboreal mammals were<br />

similar to numbers recorded in 1999, 2000 and 2007 using the same methods in the same sites<br />

(Table 7).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 251<br />

Table 7. Occurrence <strong>of</strong> yellow bellied gliders and koalas as recorded over a 3 night period in<br />

Mt. Lindesay SF Project Area harvest area. *Koalas sighted whilst measuring plots.<br />

Date recorded July November March March-<br />

1999 2000 2007 April 2008<br />

Yellow bellied glider 14 9 23 16<br />

Koala 0 0 3 2*<br />

DISCUSSION<br />

Lantana cover<br />

Harvesting and herbicide spraying created dry fuels relative to the surrounding untreated areas. This<br />

made lantana combustible at relatively low Fire Danger Indices. In the absence <strong>of</strong> harvesting<br />

disturbance fire intensity was insufficient to consistently kill lantana. After intense fire, dense<br />

lantana cover was replaced with dense ‘fireweeds’, grasses and eucalypt regeneration, however there<br />

was significant recovery <strong>of</strong> lantana in the moderate and intense fire treatments where cover at least<br />

doubled in the year following treatment. Nevertheless, the objective <strong>of</strong> lantana cover being reduced<br />

to less than 15% was met. Growth <strong>of</strong> shrubs including lantana is likely to cause problems with<br />

reintroduction <strong>of</strong> low intensity fire regimes in areas without a complete and even regenerating<br />

eucalypt canopy. The need for costly rehabilitation operations can be avoided by reintroducing low<br />

intensity fire regimes into open forests before understorey development makes this too difficult.<br />

Although a single low intensity fire may not kill lantana, repeated fires will gradually reduce or<br />

eliminate it (e.g. Kington et al 2009) When shrub encroachment is well advanced, integrated<br />

harvesting and rehabilitation operations are likely to be the only feasible solution to forest decline on<br />

a broad scale.<br />

Overstorey recovery<br />

Canopy health in forests in north eastern New South Wales generally improved after 2007 with<br />

ongoing rainfall events. In this project, tree crown health improved in both control and treated areas<br />

but remained lower in treated areas than in controls. There is no evidence that the treatments<br />

improved the health <strong>of</strong> the retained trees at this stage. High intensity fire scorched the crowns and<br />

promoted epicormic foliage which is palatable and nutritious to psyllids. In stands with sparse<br />

regeneration, immediate reintroduction <strong>of</strong> low intensity fire will be necessary to improve the health<br />

<strong>of</strong> retained eucalypts.<br />

Bell Miner abundance<br />

Bell miners can contribute to forest decline and accelerate the process by excluding other birds<br />

which could more effectively damp irruptions <strong>of</strong> psyllids. In a previous project at Donaldson SF it<br />

was found that the most severe treatment (mechanical disturbance and hot fire) produced a greater<br />

reduction in bell miner numbers and increase in other bird species than did either burning alone or<br />

herbicide spraying alone (Mews 2006). In the current study there was a proportionately greater<br />

reduction <strong>of</strong> bell miners in treated areas than controls in 2008, but bellbird numbers had recovered<br />

by 2009 and were not significantly different to controls. It appears that harvesting and burning<br />

resulted in loss <strong>of</strong> nests and fledgings, and loss <strong>of</strong> psyllids (i.e. food resources) in scorched canopies.<br />

Stone et al. (2008) found a strong correlation between bell miner abundance, understorey density<br />

and tree crown health. The current study indicated that reduced bell miner populations in controls<br />

were a result <strong>of</strong> better crown health following ongoing rainfall and consequently reduced psyllid<br />

populations. Reduced lantana cover in the controls was apparently due to increased shading provided<br />

by healthier crowns. Although there was a much greater reduction <strong>of</strong> lantana in treated areas than in<br />

controls (Figure 2), canopy health and bell miner numbers in 2009 were not affected by the<br />

rehabilitation treatments (Figures 3, 4) indicating that canopy health controls psyllid and bell miner<br />

populations, and understorey density. If this is the case, then removal <strong>of</strong> bell miners and poisoning<br />

or burning <strong>of</strong> lantana per se will not improve tree health. The phenomenon <strong>of</strong> linked lantana, psyllid<br />

and bell miner invasions is a consequence <strong>of</strong> poor tree health caused by deteriorating root function<br />

under changing soil conditions in the absence <strong>of</strong> fire as proposed by Jurskis (2005). Frequent low<br />

intensity fire regimes maintain stable soil conditions and forest health in naturally grassy forests


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 252<br />

whilst exclusion <strong>of</strong> fire causes tree decline, shrub invasion and pest outbreaks (Jurskis 2005, Turner<br />

et al. 2008).<br />

Forest rehabilitation and regeneration<br />

A previous fire and mechanical clearing trial in a declining stand <strong>of</strong> Forest Red Gum in Donaldson<br />

SF virtually eliminated lantana and produced a proliferation <strong>of</strong> fire succession species (mainly native<br />

peach and wattle) similar to the more intense fire treatment in this study. There was insufficient<br />

eucalypt regeneration to restore a forest canopy (Shipman 2007). Shipman (2007) reported that 32 %<br />

<strong>of</strong> 60 plots had no eucalypt regeneration, 36 % had a stocking <strong>of</strong> 250 stems ha -1 and only 32 % had<br />

a stocking <strong>of</strong> 500 stems ha -1 or more. In this project about half the plots had regeneration stocking<br />

<strong>of</strong> less than 100 stems ha -1 and about a quarter <strong>of</strong> the plots had more than 1000 stems ha -1 . The low<br />

levels <strong>of</strong> regeneration can be partially attributed to the high stockings and retained basal areas (Table<br />

5) and the absence <strong>of</strong> regeneration where fire was not used (Figure 6).<br />

High intensity fire such as used in parts <strong>of</strong> Donaldson and Mt.Lindesay State Forests resulted in<br />

dense understorey development but achieved large reductions in lantana understoreys, compared<br />

with less intense or no fire. There is need to be cautious with fire treatments as high intensity fire can<br />

exacerbate forest decline by promoting N cycling in the dense nitrophilous understoreys and shading<br />

the ground (e.g. Jurskis et al. 2003). This is true <strong>of</strong> native and other exotic understoreys as well as<br />

lantana. Furthermore, lantana showed significant recovery in moderate and high intensity fire<br />

treatments within a year after treatment. Equally, high intensity fire may scorch crowns and promote<br />

epicormic foliage which is palatable and nutritious to psyllids.<br />

The variability <strong>of</strong> regeneration success in plots with lower residual stockings and basal areas can<br />

primarily be explained by whether or not there was brush box regeneration (Figure 7). The poor<br />

regenerative capacity <strong>of</strong> the eucalypts can be explained by the paucity <strong>of</strong> seed in their debilitated<br />

crowns. Several years <strong>of</strong> normal shoot development are necessary for eucalypts to initiate buds,<br />

flower, and produce mature seed. This cannot occur in trees suffering chronic decline. Where brush<br />

box seed trees were present there was prolific seed fall and establishment, particularly after fire<br />

treatments, irrespective <strong>of</strong> intensity, in harvested areas. From a forest regeneration viewpoint,<br />

retention <strong>of</strong> well spaced brush box seed trees is desirable, however planting <strong>of</strong> eucalypt seedlings is<br />

vital to maintain a natural species composition in mixed stands.<br />

Cost <strong>of</strong> project<br />

Whilst the cost <strong>of</strong> the project was significant, the opportunity cost <strong>of</strong> doing nothing is greater. The<br />

cost <strong>of</strong> rehabilitation was less than the likely loss <strong>of</strong> production if the forest continued to decline and<br />

die (Table8).<br />

Table 8. Estimate <strong>of</strong> the economics <strong>of</strong> BMAD rehabilitation assuming growth rates impacted<br />

by BMAD and actual average royalty rates and rehabilitation costs at Mt. Lindesay.<br />

Decline Net growth Value<br />

Rating rate $/ha @<br />

(m3/ha/yr) $40/m 3<br />

Opportun y<br />

Rehab cost Rehab Net gain<br />

cost (OC) Over 40 yrs cost (RC) production<br />

($/ha/yr) ($) ($/ha/yr) (OC-RC)<br />

($/ha/yr)<br />

0 None 1 40 100 4 0<br />

1/2 Slight 0 0 40 200 25 15<br />

3 Moderate -1 -40 80 1000 50 30<br />

4 Severe -2 -80 120 2500 75 45<br />

IMPLICATIONS OF THE PROJECT<br />

Priority 1. Use <strong>of</strong> practices to limit understorey invasion <strong>of</strong> grassy forests so that costly<br />

rehabilitation is not required. Reintroduction <strong>of</strong> low intensity fire and/or return <strong>of</strong> grazing to forest<br />

areas. Whilst it maybe argued that cattle are not a natural component <strong>of</strong> forest ecosystems, neither is<br />

the exclusion <strong>of</strong> fire or the presence <strong>of</strong> lantana. Cattle grazing provides the opportunity for closer<br />

understorey management through occupier weed control, more fire management capacity and<br />

browsing <strong>of</strong> regenerating shrubs and lantana by cattle.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 253<br />

Priority 2. Use <strong>of</strong> harvesting, burning and planting treatments to restart degraded stands to ensure<br />

rapid growth and dominance <strong>of</strong> the overstorey, thus controlling the understorey. The regenerating<br />

stands <strong>of</strong> drier forest types may begin to decline as they approach pole stage and sapwood basal area<br />

declines (e.g. Jurskis 2005) unless low intensity fire regimes are introduced when forest structure<br />

and fuel permits.<br />

Priority 3. Whilst harvesting is currently not an option for assisting rehabilitation <strong>of</strong> lantana<br />

infestations in National Parks or in State Forests areas reserved from harvesting, wildfires may<br />

present an opportunity to restart patches <strong>of</strong> forest provided seeding or planting <strong>of</strong> overstorey species<br />

can be done before the explosion <strong>of</strong> fire succession species and/or subsequent re-invasion <strong>of</strong> lantana.<br />

ACKNOWLEDGEMENTS<br />

The author acknowledges the Bell Miner Associated Dieback Working Group and members <strong>of</strong> the<br />

Scientific Reference Group (Harry Recher, Ross Florence, Ross Goldingay) for assistance with<br />

design and support <strong>of</strong> this project, partial funding by the National Heritage Trust, Forests NSW staff<br />

who planned, supervised operations and measured plots, especially Casino staff Flavio Bugno, Jayne<br />

Fitzpatrick and Craig Pavy, Lynden Pavy, Jamie Churchill, Ken Wheatley, Dan Hutton, Kevin<br />

Harvey, Paul Meek and Steve Rayson. Vic Jurskis and Doland Nichols for their reviews and<br />

comments, Sarah Nadler and Craig Wall from Department <strong>of</strong> Environment and Climate Change for<br />

database creation and data management, and Adam Kirby <strong>of</strong> Southern Cross University for data<br />

analysis and presentation and Border Ranges Contracting and the Githabul people for planting<br />

operations. The plots will be remeasured in March 2011. Visit www.bmad.com.au for more info.<br />

REFERENCES<br />

Anonymous (2007). Tree decline in the absence <strong>of</strong> fire. Fire Note 13. Bushfire Cooperative Research Centre,<br />

East Melbourne.<br />

BMAD Working Group (2007). Minimum standards for monitoring. DECC C<strong>of</strong>fs Harbour.<br />

Craig, M.D. and Roberts, J.D. (2001). Evaluation <strong>of</strong> the impact <strong>of</strong> time <strong>of</strong> day, weather, vegetation density and<br />

bird movements on outcomes <strong>of</strong> area searches for birds in eucalypt forests <strong>of</strong> south-western <strong>Australia</strong>.<br />

Wildlife Research 28, 33-39.<br />

Dare, A.J., McDonald. P.G. and Clarke. M.F. (2007) The ecological context and consequences <strong>of</strong> colonization<br />

<strong>of</strong> a site by bell miners (Manorina melanophrys). Wildlife Research 34, 616-623.<br />

Forestry Commission <strong>of</strong> NSW (1989) Research Note 17 Forest Types <strong>of</strong> NSW. ISBN: 0 73055649 2<br />

Integrated Forestry Operations Approval<br />

Jurskis, V., (2004). Overview <strong>of</strong> forest decline in coastal New South Wales. In: White, T.C.R., Jurskis, V.<br />

(Eds.), Fundamental Causes <strong>of</strong> Eucalypt Forest Decline and Possible Management Solutions.<br />

Proceedings <strong>of</strong> a Colloquium at Batemans Bay, 18 and 19 November 2003, State Forests <strong>of</strong> NSW,<br />

Sydney, pp. 4–7.<br />

Jurskis, V. (2005) Eucalypt decline in <strong>Australia</strong>, and a general concept <strong>of</strong> tree decline and dieback. Forest<br />

Ecology and Management, 215, 1–20<br />

Jurskis, V., 2008. Drought as a factor in tree declines and diebacks. In: Sanchez, J.M. (Ed.) Droughts: Causes,<br />

effects and predictions. Nova Science Publishers Inc. New York. pp 331-341.<br />

Jurskis, V. and Turner, J. (2002) Eucalypt dieback in eastern <strong>Australia</strong>: a simple model. <strong>Australia</strong>n Forestry<br />

65 (2): 87-98<br />

Kington, W., Kington, D. and Burnham, M. (2009) Guide for planned burning. Open forests and Woodlands<br />

V2. South-east Queensland Parks and Wildlife. Qld Government CD 2009<br />

Mews, J. (2006) Effects <strong>of</strong> understorey modifications trial on bird populations in Bell Miner Associated<br />

Dieback Affected Forest. Undergraduate Project. Southern Cross University, Lismore.<br />

Mews, J. (2008) The role <strong>of</strong> interactions between understorey, soil properties and foliar nutrient status in the<br />

development <strong>of</strong> Bell Miner Associated Dieback (BMAD). Honours Thesis. Southern Cross University.<br />

Moore, K.M., (196)1. Observations on some <strong>Australia</strong>n forest insects. 8. The biology and occurrence <strong>of</strong><br />

Glycaspis baileyi Moore in New South Wales. Proc. Linn. Soc. N.S.W. 86 (2), 185–200.<br />

Shipman, R (2007) Regeneration/treatment <strong>of</strong> native forest bell miner affected areas. Undergraduate Project.<br />

Southern Cross University, Lismore.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 254<br />

Stone, C., (1996), The role <strong>of</strong> psyllids (Hemiptera: Psyllidae) and bell miners (Manorina melanophrys) in<br />

canopy dieback in Sydney Blue Gum (Eucalyptus saligna Sm). <strong>Australia</strong>n Journal <strong>of</strong> Ecology 21.<br />

Stone, C., (1999), Assessment and monitoring <strong>of</strong> decline and dieback <strong>of</strong> forest eucalypts in relation to<br />

ecological sustainable forest management: a review with case study. <strong>Australia</strong>n Forestry 62 (1)<br />

Stone, C., (2005), BMAD at the tree crown scale: A multitrophic process. <strong>Australia</strong>n Forestry 68 (4)<br />

Stone, C., Kathuria, A., Carney, C. and Hunter, J. (2008), Forest canopy health and stand structure associated<br />

with bell miners (Manorina melanophrys) on the central coast <strong>of</strong> New South Wales. <strong>Australia</strong>n<br />

Forestry 71 (4)<br />

Stone, C. and Simpson, J.A. (2006) Leaf, tree and soil properties in a Eucalyptus saligna forest exhibiting<br />

canopy decline. Cunninghamia 9, 507-520.<br />

Stone, C., Spolc, D.and Urquhart, C. A., (1995), Survey <strong>of</strong> Crown Dieback in Moist Hardwood Forests in the<br />

Central and Northern Regions <strong>of</strong> NSW State Forests (Psyllids/Bell miner Research Programe).<br />

Research paper No. 28. Research Division, State Forests <strong>of</strong> NSW. Sydney.<br />

Threatened Species Licence, under the Threatened Species Conservation Act (1995) Upper North East Region<br />

Turner, J. and Lambert, M. (2005), Soil and nutrient processes related to eucalypt forest dieback. <strong>Australia</strong>n<br />

Journal 68 (4)<br />

Turner, J., Lambert, M., Jurskis, V., Bi, H., (2008). Long term accumulation <strong>of</strong> nitrogen in soils <strong>of</strong> dry mixed<br />

eucalypt forest in the absence <strong>of</strong> fire. For. Ecol. Manage. 256, 1133-1142.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 255<br />

USING A PROCESS-BASED MODEL, PAROPSYS,<br />

TO PREDICT THE IMPACT OF CLIMATE CHANGE ON THE<br />

EUCALYPT PLANTATION PEST Paropsis atomaria.<br />

Simon Lawson 1<br />

ABSTRACT<br />

The eucalypt leaf beetle, Paropsis atomaria Olivier, is an increasingly important pest <strong>of</strong><br />

eucalypt plantations in subtropical eastern <strong>Australia</strong>. A process-based model, ParopSys,<br />

was developed using DYMEX TM and was found to accurately predict the beetle<br />

populations. Climate change scenarios within the latest <strong>Australia</strong>n climate model forecast<br />

range were run in ParopSys at three locations to predict changes in beetle performance.<br />

Relative population peaks <strong>of</strong> early generations did not change but shifted to earlier in the<br />

season. Temperature increases <strong>of</strong> 1.0 to 1.5 ºC or greater predicted an extra generation <strong>of</strong><br />

adults at Gympie and Canberra, but not for Lowmead, where increased populations <strong>of</strong><br />

late season adults were observed under all scenarios. Furthermore, an additional<br />

generation <strong>of</strong> late-larval stages was predicted at temperature increases <strong>of</strong> greater than 1.0<br />

ºC at Lowmead. Management strategies to address these changes are discussed, as are<br />

requirements to improve the predictive capacity <strong>of</strong> the model.<br />

INTRODUCTION<br />

Chrysomelid leaf beetles are important insect defoliators <strong>of</strong> young eucalypt plantations in <strong>Australia</strong> in<br />

most growing regions. Damage by these beetles can lower growth rates significantly, with losses <strong>of</strong><br />

up to 30 percent in productivity observed over a ten year period in unprotected trees in Tasmania (Elek,<br />

1997). There, Paropsisterna bimaculata and Pa. agricola are key pests <strong>of</strong> E. nitens and E. globulus<br />

plantations. Pa. bimaculata populations are managed using a sophisticated integrated pest<br />

management system that is based on more than 20 years <strong>of</strong> research data on the biology, ecology and<br />

impact <strong>of</strong> this beetle on tree growth rates (Elliott et al., 2002, Elek, 1997, Elliott et al., 1992).<br />

In the subtropics <strong>of</strong> northern New South Wales and southeast and central Queensland, significant<br />

expansion <strong>of</strong> eucalypt plantations has only occurred since the late 1990’s, with about 60,000 ha now<br />

established in Queensland and 80,000 ha in New South Wales (Gavran and Parsons, 2009). The<br />

chrysomelid beetle, Paropsis atomaria Olivier, has emerged as a key pest <strong>of</strong> several plantation taxa in<br />

the region during this time, including Eucalyptus cloeziana, E. dunnii, E. grandis and its hybrids and<br />

Corymbia citriodora ssp. variegata (Nahrung et al., 2008b, Nahrung et al., 2008a, Nahrung, 2006).<br />

Severe outbreaks have been commonly reported by plantation growers and applications <strong>of</strong> insecticides<br />

have been used to manage beetle populations on an ad hoc basis, not as part <strong>of</strong> an intensive IPM<br />

system (Lawson et al., 2008).<br />

While the biology <strong>of</strong> P. atomaria had been well-studied in south-eastern <strong>Australia</strong> (Carne, 1966a,<br />

Carne, 1966b, Ohmart, 1991, Ohmart et al., 1987), intensive field censuses <strong>of</strong> the beetle and its natural<br />

enemies, and laboratory studies undertaken since 2004 in Queensland have begun to address<br />

deficiencies in knowledge <strong>of</strong> the biology and ecology <strong>of</strong> this species in subtropical plantations. Low<br />

temperature thresholds, temperature-induced mortality rates, and day-degree requirements for<br />

immature stages were obtained in the laboratory using beetles sourced from temperate and subtropical<br />

zones and combined with field mortality estimates <strong>of</strong> these stages to produce a process-based<br />

population model, ParopSys, using DYMEX TM (Maywald et al., 2007). This model has been<br />

successfully validated for sites in central and southeastern Queensland, and for Canberra in temperate<br />

southeast <strong>Australia</strong> (Nahrung et al., 2008b). The model has since been used to predict population<br />

trends in plantations and to optimize timing <strong>of</strong> insecticide applications.<br />

1 Queensland Primary Industries and Fisheries, Department <strong>of</strong> Employment, Economic Development and Innovation, 80<br />

Meiers Rd, Indooroopilly, Qld 4068. Ph: +61 7 38969613 Email: Simon.Lawson@deedi.qld.gov.au


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 256<br />

A feature <strong>of</strong> the DYMEX TM modeling system is the ability to run climate change scenarios and test the<br />

impact <strong>of</strong> these on the population dynamics <strong>of</strong> the organism or system being modeled. Since insects<br />

are poikilothermic, changes in temperature directly affect developmental rates and hence population<br />

dynamics. This module was used to test the effects <strong>of</strong> climate change on the population dynamics <strong>of</strong> P.<br />

atomaria using a series <strong>of</strong> scenarios derived from the most recent projections for <strong>Australia</strong> (Climate<br />

Change in <strong>Australia</strong> Technical Report 2007) for three locations in Queensland and the <strong>Australia</strong>n<br />

Capital Territory that were used in the initial model validation.<br />

MATERIALS AND METHODS<br />

ParopSys<br />

Details <strong>of</strong> the structure and values <strong>of</strong> variables used in ParopSys are detailed by (Nahrung et al.,<br />

2008b). In brief, the model was run using the same climatic data used to validate the model and the<br />

simulations run over the period 28 September 2005 to 10 May 2006, the period when adult beetles are<br />

known to be active in the field. P. atomaria predominantly overwinters in the adult stage and so the<br />

model was initialised with 0.5 pre-reproductive adults on day one, with the same number added on<br />

each <strong>of</strong> the following nine days, giving a total <strong>of</strong> five adults. Pre-reproductive adults are used to<br />

initialise the model because, as they emerge from overwintering, adults must feed for a period <strong>of</strong> time<br />

before they are able to begin reproducing.<br />

Temperature variation in the climate change module can use a number <strong>of</strong> different input factors,<br />

depending on the location <strong>of</strong> the climate data being used. For this study, inputs which raised the<br />

maximum and minimum temperatures for ‘winter’ (defined in the model as March 2 to September 30)<br />

and ‘summer’ (defined as October 1 to March 1) by a defined number <strong>of</strong> degrees were used (see<br />

below).<br />

Regions<br />

The three locations originally used to validate the model and which represent a large part <strong>of</strong> the<br />

distributional range <strong>of</strong> this beetle were chosen to test the climate change scenarios. These were: (1) a<br />

plantation near Gympie, southeast Queensland (26º 06'S; 152º 45'E); (2) Lowmead, central<br />

Queensland (24º 30'S; 151º 42'E); and (3) Canberra, A.C.T (35º 27’S; 149º 12’E).<br />

Meteorological data used in ParopSys were obtained from the Silo Data drill website<br />

(http://www.nrw.qld.gov.au/silo/datadrill/).<br />

Climate Change Scenarios<br />

Variation in temperature was used as the only predictor in these climate change scenarios, since<br />

rainfall, evapotranspiration and other potential climate change variables are not currently used as<br />

inputs to drive ParopSys. The model therefore assumes that beetle access to suitable food resources<br />

(young, expanding foliage for larvae and young or adult foliage for adults) is not limiting. It is<br />

anticipated that future versions <strong>of</strong> ParopSys will incorporate explicit modelling <strong>of</strong> resources such as<br />

presence <strong>of</strong> flush foliage and immigration and emigration processes.<br />

Climate change projections for the regions being modelled were obtained from the Climate Change in<br />

<strong>Australia</strong> Technical Report 2007 website (http://www.climatechangeinaustralia.gov.au). A variety <strong>of</strong><br />

projections are available from this modelling, using high, medium and low carbon/ CO2 emission<br />

levels and 10, 50 and 90 percentile estimates <strong>of</strong> likelihood, with projections to 2030, 2050 and 2070.<br />

For this modelling exercise, we chose what would appear to be the most likely scenarios, i.e., medium<br />

level emissions, 50 percentile likelihood (considered to be the best estimate) for projection to 2030,<br />

2050 and 2070. Across all locations, the 2030 scenario gives summer and winter temperature<br />

increases <strong>of</strong> 0.6 to 1.0 ºC (Fig. 1), for 2050 1.0 to 1.5 ºC for southeast and central Queensland and 1.5<br />

to 2.0 ºC for Canberra, and for 2070 1.5 to 2.0 ºC for southeast and central Queensland and 2.0 to 2.5<br />

ºC for Canberra.<br />

We ran seven climate change scenarios in ParopSys to encompass this forecast range in temperature<br />

variation. Data for the 2005-2006 season for each site were used as the control, with increases <strong>of</strong> 0.25,


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 257<br />

0.5, 1.0, 1.5, 2.0 and 2.5 ºC in mean daily temperature. From these model runs we obtained outputs <strong>of</strong><br />

numbers <strong>of</strong> all Paropsis lifestages over the length <strong>of</strong> the model run (28 September to 10 May). Given<br />

that most damage by this beetle is caused by late stage larvae (third and fourth instars) and adults we<br />

focussed on the effect <strong>of</strong> the climate change scenarios on these stages.<br />

RESULTS<br />

Outputs <strong>of</strong> ParopSys for the seven climate change scenarios simulations are shown in Figure 1 (adult<br />

beetles) and Figure 2 (third and fourth instar larvae).<br />

Numbers <strong>of</strong> Adults Numbers <strong>of</strong> Adults<br />

Numbers <strong>of</strong> Adults<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

28/09/2005<br />

18/10/2005<br />

No Change<br />

0.25<br />

0.5<br />

1.0<br />

1.5<br />

2.0<br />

2.5<br />

7/11/2005<br />

27/11/2005<br />

(a) Gympie<br />

(b) Lowmead<br />

(c) Canberra<br />

17/12/2005<br />

6/01/2006<br />

26/01/2006<br />

15/02/2006<br />

Simulation Date<br />

Figure 1. Comparison <strong>of</strong> adult P. atomaria population responses to seven climate change<br />

scenarios at three sites using the simulation model ParopSys: (a) near Gympie,<br />

southeast Queensland, (b) Lowmead, central Queensland and (c) Canberra, temperate<br />

south-eastern <strong>Australia</strong>.<br />

7/03/2006<br />

27/03/2006<br />

16/04/2006<br />

6/05/2006


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 258<br />

Numbers <strong>of</strong> L3 + L4 Larvae<br />

Numbers <strong>of</strong> L3 + L4 Larvae<br />

Numbers <strong>of</strong> L3 + L4 Larvae<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

3500<br />

3000<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

28/09/2005<br />

18/10/2005<br />

No change<br />

0.25<br />

0.5<br />

1<br />

1.5<br />

2<br />

2.5<br />

7/11/2005<br />

27/11/2005<br />

(a) Gympie<br />

(b) Lowmead<br />

(c) Canberra<br />

17/12/2005<br />

6/01/2006<br />

26/01/2006<br />

15/02/2006<br />

Simulation Date<br />

Figure 2. Comparison <strong>of</strong> late stage (3 rd and 4 th instar) larval P. atomaria population responses to<br />

seven climate change scenarios at three sites using the simulation model ParopSys: (a)<br />

near Gympie, southeast Queensland, (b) Lowmead, central Queensland and (c)<br />

Canberra, temperate south-eastern <strong>Australia</strong>.<br />

Without climate change, ParopSys predicts two beetle generations per year for Canberra, three for<br />

Gympie, and four for Lowmead. (Nahrung et al., 2008b) validated this prediction for Gympie and the<br />

outputs were also shown to fit field observations for Lowmead (Lawson pers. obs.) and for Canberra<br />

(Carne, 1966a).<br />

When temperature was varied according to the six climate change scenarios, adult beetle populations<br />

at each site differed in their response to increases in temperature. The Gympie and Canberra<br />

populations were similar in that an additional generation began to appear in the upper range <strong>of</strong> climate<br />

change scenarios. For the Gympie population, a fourth generation <strong>of</strong> beetles began to emerge with<br />

temperature increases <strong>of</strong> 1.5 ºC and above (Fig. 1a), while the Canberra population added a third<br />

generation at the same elevations in temperature (Fig. 1c). For temperature increases below 1.5 ºC,<br />

7/03/2006<br />

27/03/2006<br />

16/04/2006<br />

6/05/2006


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 259<br />

population peaks were predicted to occur earlier in the season but peak heights did not change. All<br />

increases in temperature resulted in an increased fourth generation peak population for the Lowmead<br />

site, with a more than two-fold increase in peak height being observed at temperatures from 1.5 to 2.5<br />

ºC (Fig. 1b).<br />

Changes in peaks and population levels <strong>of</strong> late stage larvae with increased temperature showed similar<br />

patterns for the Gympie and Canberra sites, where additional late stage larval generations were only<br />

predicted for the top two temperature increases <strong>of</strong> 2.0 and 2.5 ºC (Fig. 2a,c). For the Gympie site,<br />

neither <strong>of</strong> these additional larval population peaks exceeded the previous third generation peak. At the<br />

Canberra site, the peak <strong>of</strong> the third generation late stage larvae was only predicted to exceed that <strong>of</strong> the<br />

second generation for a 2.5 ºC increase. For Lowmead, additional larval population peaks that almost<br />

equalled, or exceeded, the third generation peak occurred at all temperature increases <strong>of</strong> 1.0 ºC and<br />

above.<br />

From Figures 1 and 2 it appears that there is a threshold <strong>of</strong> temperature change below which<br />

population dynamics appear relatively stable and above which there are significant changes in number<br />

<strong>of</strong> generations and size <strong>of</strong> population peaks. This mainly seems to be the case for the southernmost<br />

sites <strong>of</strong> Gympie and Canberra. To test this observation, temperature change was plotted against<br />

population numbers <strong>of</strong> the last peak in the season for both adults and larvae (Fig. 3).<br />

Numbers <strong>of</strong> late season peak adults<br />

Numbers <strong>of</strong> late season peak L3+L4<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

3500<br />

3000<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

Gympie<br />

Low mead<br />

Canberra<br />

Gympie<br />

Low mead<br />

Canberra<br />

0<br />

0.0 0.5 1.0 1.5 2.0 2.5<br />

Temperature change ºC<br />

Figure 3. Relationship between increase in average temperature and predicted numbers <strong>of</strong><br />

adults (top) and late stage larvae (bottom) in late-season population peaks for three<br />

sites: near Gympie, southeast Queensland, Lowmead, central Queensland and<br />

Canberra, temperate south-eastern <strong>Australia</strong>.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 260<br />

For both the Gympie and Canberra sites, a threshold temperature <strong>of</strong> between 1.0 and 1.5 ºC is<br />

predicted to produce an extra generation <strong>of</strong> adults. A similar trend was observed for these two sites<br />

for larvae, but the threshold at which an extra generation <strong>of</strong> these stages occurred was at a higher<br />

temperature <strong>of</strong> between 2.0 and 2.5 ºC. By comparison, no such distinct thresholds were observed for<br />

the Lowmead site, where there the change is restricted to an increase in numbers <strong>of</strong> the already<br />

existing fourth generation.<br />

DISCUSSION<br />

The effects <strong>of</strong> climate change on herbivorous insect population dynamics are complex and involve<br />

interactions between the direct effects <strong>of</strong> temperature on the developmental and mortality rates <strong>of</strong> the<br />

insects themselves, the effects <strong>of</strong> environmental variables on their host plants, and the interactions<br />

between these factors. Effects on host plants are largely mediated through the effects <strong>of</strong> elevated<br />

temperature and enhanced atmospheric levels <strong>of</strong> CO2 on plant growth rates and changes in plant water<br />

relations as a function <strong>of</strong> differing rainfall and evapotranspiration patterns.<br />

Temperature was used as the only climate change variable in this study since the ParopSys model does<br />

not currently use other climate inputs to drive beetle populations. This is an obvious limitation to the<br />

applicability <strong>of</strong> these results given that populations <strong>of</strong> these beetles are also driven by a range <strong>of</strong> other<br />

variables, in particular the availability <strong>of</strong> abundant, expanding juvenile foliage, which with eucalypt<br />

species is closely associated with soil moisture reserves and rainfall during the active growing period.<br />

However, even with this limitation the model has been shown to accurately predict the number, timing<br />

and relative population peak heights for P. atomaria in southern Queensland, and, at a minimum, the<br />

number <strong>of</strong> generations at sites in central Queensland and temperate south-eastern <strong>Australia</strong> (Nahrung<br />

et al., 2008b). Therefore, the trends that emerged from this modelling are potentially useful to forest<br />

managers in predicting potential changes in the population dynamics <strong>of</strong> this beetle over its<br />

distributional range as climate warms.<br />

General outcomes <strong>of</strong> the modelling can be summarised as follows. No changes in relative population<br />

peaks <strong>of</strong> the early generations <strong>of</strong> either beetle stage were observed at any site or with any <strong>of</strong> the<br />

temperature change scenarios, but these generations became time-shifted to earlier in the season (Figs<br />

1&2). These time shifts become more marked as the season progresses, with shifts <strong>of</strong> up to three<br />

weeks earlier for the third generation at Gympie and Lowmead and the second generation at Canberra<br />

being observed. Beetle damage can thus be expected to occur earlier in the season at all sites as<br />

temperature increases.<br />

For the two more southern locations <strong>of</strong> Gympie and Canberra, temperature elevations <strong>of</strong> 1.0 to 1.5 ºC<br />

or greater (equivalent to the temperature changes predicted post 2030 for these sites) predict the<br />

development <strong>of</strong> an extra generation <strong>of</strong> adults but not additional large numbers <strong>of</strong> late stage larvae.<br />

High populations <strong>of</strong> late season adults can cause severe defoliation <strong>of</strong> mature foliage which impacts on<br />

the ability <strong>of</strong> trees to recover in the following spring through reduced photosynthetic capacity. For<br />

these two sites, the model therefore predicts that, up to 2030, damage is largely restricted to that<br />

caused by larval stages (i.e. upper canopy defoliation) and that this damage is progressively shifted<br />

earlier in the season with increased temperature, but post-2030 greater late stage defoliation <strong>of</strong> mature<br />

foliage by adult beetles can be expected.<br />

For the northernmost site <strong>of</strong> Lowmead, the model predicts no extra adult generation even at the<br />

highest temperature increase, but does predict progressive increases in population levels <strong>of</strong> this stage<br />

at all scenarios modelled (a greater than tw<strong>of</strong>old increase in population at the 2.5 ºC elevation). With<br />

late-larval stages an additional generation begins appearing at temperature increases <strong>of</strong> 1.0 ºC and<br />

greater, equivalent to post 2050 scenarios for this site. At Lowmead, the model therefore predicts<br />

increased levels <strong>of</strong> late season adult damage to mature foliage for all temperature increases through to<br />

2070, with an extra larval generation causing additional late season upper canopy defoliation from<br />

2050. From 2050, this effectively means that greater levels <strong>of</strong> whole-tree defoliation may be expected<br />

at this site.<br />

In terms <strong>of</strong> practical management outcomes, the modelling presented here suggests that monitoring <strong>of</strong><br />

beetle populations should commence earlier in the season than currently is practised as temperature


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 261<br />

increases. The development <strong>of</strong> additional generations <strong>of</strong> adults (Gympie, Canberra), and higher<br />

numbers <strong>of</strong> late season adults and an additional larval generation (Lowmead) would indicate that<br />

strategies to reduce beetle populations earlier in the season are going to be increasingly important to<br />

lower the risk <strong>of</strong> highly damaging late-season populations developing. At some sites, multiple control<br />

applications may be required.<br />

As stated above, there are a number <strong>of</strong> limitations to the ParopSys model, including unknown effects<br />

<strong>of</strong> temperature on the survival <strong>of</strong> overwintering stages and high temperature mortality responses.<br />

Elevated winter temperatures could potentially increase survival <strong>of</strong> overwintering adult beetles,<br />

resulting in higher beetle populations earlier in the season. Experimental data is required so that this<br />

potential response to climate change can be incorporated into the model. Whilst P. atomaria is<br />

distributed as far north as tropical Queensland, damaging populations have not been recorded in<br />

plantations there and the beetle appears to be quite rare. High temperatures may therefore be a<br />

limiting factor in development <strong>of</strong> populations <strong>of</strong> this beetle. We have temperature associated mortality<br />

data for the immature stages <strong>of</strong> P. atomaria and this will be included in a future version <strong>of</strong> ParopSys to<br />

improve its predictive power.<br />

Furthermore, the model assumes constant favourability <strong>of</strong> its tree host for survival and reproduction.<br />

Addition <strong>of</strong> a ‘flushing’ module, including tree response to defoliation, would therefore be an<br />

important refinement since the presence <strong>of</strong> flush foliage is required for egg-laying and subsequent<br />

development <strong>of</strong> beetle immature stages.<br />

Other predicted effects <strong>of</strong> climate change on plants are not currently included in the model. Elevated<br />

CO2 levels are predicted to have various effects on plant chemistry and physiology including: (a)<br />

higher plant growth rates, <strong>of</strong>ten leading to higher rates <strong>of</strong> feeding by insects due to higher biomass but<br />

lower foliar nitrogen levels and (b) increased carbon to nitrogen ratio, enabling plants to potentially<br />

produce higher levels <strong>of</strong> carbon-based defensive chemicals, such as phenolics and tannins (important<br />

components in the leaves <strong>of</strong> eucalypts), and lowering their ability to produce nitrogen based defensive<br />

compounds (such as alkaloids and cyanogenic glycosides; (Trumble and Butler, 2009). More complex<br />

responses arise when temperature and CO2 are elevated simultaneously (Zvereva and Kozlov, 2006).<br />

Few specific data exist on the effects <strong>of</strong> climate change on eucalypt leaf chemistry, with the exception<br />

<strong>of</strong> (Gleadow et al., 1998) who examined the effects <strong>of</strong> elevated CO2 levels on nitrogen chemistry<br />

(particulary cyanogenic compounds) in the leaves <strong>of</strong> Eucalyptus cladocalyx. Under normal conditions<br />

leaf nitrogen has not been shown to be a significant limiting factor for development <strong>of</strong> P. atomaria<br />

(Ohmart et al., 1987, Ohmart, 1991) but further specific studies on the performance <strong>of</strong> P. atomaria on<br />

hosts grown under elevated temperature and CO2 are essential to further develop the climate change<br />

predictive ability <strong>of</strong> the ParopSys model.<br />

ACKNOWLEDGEMENTS<br />

Sincere thanks to Dr Valerie Debuse and Dr Manon Griffiths <strong>of</strong> the Department <strong>of</strong> Employment,<br />

Economic Development and Innovation for their help in improving earlier versions <strong>of</strong> this manuscript.<br />

Development <strong>of</strong> the ParopSys model was partly carried out under <strong>Australia</strong>n Research Council<br />

Linkage Projects Program (LP0454856) and in conjunction with Forestry Plantations Queensland<br />

(formerly DPI-Forestry). I gratefully acknowledge all organisations for their support, and especially<br />

Dr Helen Nahrung, Dr Mark Schutze, Dr Anthony Clarke, Michael Duffy, and Elizabeth Dunlop who<br />

were instrumental in the development <strong>of</strong> the ParopSys model.<br />

REFERENCES<br />

Carne, P. B. (1966a) Ecological Characteristics Of The Eucalypt-Defoliating Chrysomelid Paropsis Atomaria.<br />

Aust. J. Zool., 14, 647-72.<br />

Carne, P. B. (1966b) Growth And Food Consumption During The Larval Stages Of Paropsis Atomaria<br />

(Coleoptera : Chrysomelidae). Entomologia Exp. Appl., 9, 105-112 Pp.<br />

Elek, J. A. (1997) Assessing The Impact Of Leaf Beetles In Eucalypt Plantations And Exploring Options For<br />

Their Management. Tasforests, 9, 139-154.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 262<br />

Elliott, H. J., Bashford, R., Greener, A. & Candy, S. G. (1992) Integrated Pest Management Of The Tasmanian<br />

Eucalyptus Leaf Beetle, Chrysophtharta Bimaculata (Olivier) (Coleoptera: Chrysomelidae). Forest<br />

Ecology And Management, 53, 29-35.<br />

Elliott, H. J., Elek, J. A. & Bashford, R. (2002) Developing Pest Management Systems For Eucalypt Plantations.<br />

Forspa Publication, 139-146.<br />

Gavran, M. & Parsons, M. (2009) <strong>Australia</strong>’s Plantations 2009 Inventory Update, National Forest Inventory,<br />

Bureau Of Rural Sciences, Canberra.<br />

Gleadow, R. M., Foley, W. J. & Woodrow, I. E. (1998) Enhanced Co2 Alters The Relationship Between<br />

Photosynthesis And Defence In Cyanogenic Eucalyptus Cladocalyx F. Muell. Plant, Cell &<br />

Environment, 21, 12-22.<br />

Lawson, S. A., Mcdonald, J. M. & Pegg, G. S. (2008) Forest Health Surveillance Methodology In Hardwood<br />

Plantations In Queensland, <strong>Australia</strong>. <strong>Australia</strong>n Forestry, 71, 177-181.<br />

Maywald, G.F., Kriticos, D.J., Sutherst, R.W. & Bottomley, W. (2007) Dymex, Version 3.0. Hearne Scientific<br />

Publishing, Melbourne<br />

Nahrung, H. F. (2006) Paropsine Beetles (Coleoptera: Chrysomelidae) In South-Eastern Queensland Hardwood<br />

Plantations: Identifying Potential Pest Species. <strong>Australia</strong>n Forestry, 69, 270-274.<br />

Nahrung, H. F., Duffy, M. P., Lawson, S. A. & Clarke, A. R. (2008a) Natural Enemies Of Paropsis Atomaria<br />

Olivier (Coleoptera: Chrysomelidae) In South-Eastern Queensland Eucalypt Plantations. <strong>Australia</strong>n<br />

Journal Of Entomology, 47, 188-194.<br />

Nahrung, H. F., Schutze, M. K., Clarke, A. R., Duffy, M. P., Dunlop, E. A. & Lawson, S. A. (2008b) Thermal<br />

Requirements, Field Mortality And Population Phenology Modelling Of Paropsis Atomaria Olivier, An<br />

Emergent Pest In Subtropical Hardwood Plantations. Forest Ecology And Management, 255, 3515-3523.<br />

Ohmart, C. P. (1991) Role Of Food Quality In The Population Dynamics Of Chrysomelid Beetles Feeding On<br />

Eucalyptus. Forest Ecology And Management, 39, 35-46.<br />

Ohmart, C. P., Thomas, J. R. & Stewart, L. G. (1987) Nitrogen, Leaf Toughness And The Population Dynamics<br />

Of Paropsis Atomaria Olivier (Coleoptera: Chrysomelidae) - A Hypothesis. Journal Of The <strong>Australia</strong>n<br />

Entomological Society, 26, 203-207.<br />

Trumble, J. T. & Butler, C. D. (2009) Climate Change Will Exacerbate California's Insect Pest Problems.<br />

California Agriculture, 63, 73-78.<br />

Zvereva, E. L. & Kozlov, M. V. (2006) Consequences Of Simultaneous Elevation Of Carbon Dioxide And<br />

Temperature For Plant-Herbivore Interactions: A Metaanalysis. Global Change Biology, 12, 27-41.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 263<br />

IMPACT OF TREE BORON STATUS ON TREE GROWTH<br />

AND SUSCEPTIBILITY TO QUAMBALARIA SHOOT BLIGHT<br />

(QUAMBALARIA PITEREKA) IN CORYMBIA SPECIES<br />

Timothy Smith 1 and Ge<strong>of</strong>f Pegg 1<br />

ABSTRACT<br />

An experiment was conducted to assess the impact <strong>of</strong> tree boron (B) status on tree growth<br />

and susceptibility to Quambalaria shoot blight caused by the fungal pathogen<br />

Quambalaria pitereka. Increasing tree B status from deficient to sufficient was associated<br />

with increased tree growth and reduced susceptibility to Quambalaria shoot blight.<br />

Corymbia citriodora sb.sp. variegata Woondum provenance seedlings, Corymbia<br />

torelliana seedlings and C. torelliana x C. citriodora sb.sp. variegata hybrid (CTVA 1)<br />

were grown in an aeroponics system under glasshouse conditions and subjected to 6 B<br />

treatments administered as regular root sprays over a period <strong>of</strong> 3.5 months. Increasing B<br />

to roots significantly increased the youngest fully expanded leaf (YFEL) B and provided a<br />

range in tree B status from deficient to sufficient. A foliar inoculation <strong>of</strong> Quambalaria<br />

spores was applied to trees after 3 months and disease infection incidence and severity<br />

assessments were conducted after 14 days. As tree B status increased from deficient (9-12<br />

µg g -1 YFEL B) to sufficient B status (22-24 µg g -1 YFEL B) there were increases in tree<br />

height (107%, P


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 264<br />

Lesion development on immature leaves and stems starts on average 5 days after infection by Q.<br />

pitereka and is visible as discrete chlorotic spots with necrotic centres (Self et al., 2002; Pegg et al.,<br />

2005). These may develop into large, sporulating lesions 10 to 14 days after infection under<br />

favourable conditions. Conidial germination is triggered when Relative Humidity levels exceeded 90<br />

% and commencing within two hours in the presence <strong>of</strong> free water (Pegg et al., 2009). Conidial<br />

germination and hyphal growth occurs on the upper and lower leaf surfaces with penetration occurring<br />

via the stomata or wounds on the leaf surface or juvenile stems. Following penetration through the<br />

stomata, Q. pitereka hyphae grow only intercellularly without the formation <strong>of</strong> haustoria or interaction<br />

apparatus. Instead an interaction zone is evident between host and pathogen cells (Pegg et al., 2009).<br />

Several soils used to support Eucalypt plantations in Eastern <strong>Australia</strong> are limited in their supply <strong>of</strong><br />

boron (B). In Ferrosol soils (as classified using Isbell, 1986), B was found to be the third most<br />

limiting element for growth <strong>of</strong> E. nitens after nitrogen and phosphorus (Smith, 2007). Ferrosols are<br />

characterised by >5% free iron content (Isbell, 1986). The iron and aluminium oxides in these soils<br />

have a high affinity for adsorption <strong>of</strong> B. On this soil type, B deficiency was found to cause poor form<br />

<strong>of</strong> E. nitens (Smith, 2007) and Ccv. (Smith, unpublished data). Similarly, Lambert et al. (1997)<br />

reported that B was the most common micronutrient deficiency limiting growth <strong>of</strong> Radiata pine (Pinus<br />

radiata) in extensive areas <strong>of</strong> <strong>Australia</strong>, New Zealand and Chile. In China, the Philippines and<br />

<strong>Australia</strong>, B deficiency in eucalypts was induced through applications <strong>of</strong> macronutrients, which<br />

accelerated tree growth, but B uptake was inadequate to meet demand (Dell et al., 2008).<br />

The reported roles <strong>of</strong> B in plants include: bonding <strong>of</strong> the primary cell wall, specifically the apiose side<br />

chains <strong>of</strong> rhamnogalacturonan II fractions <strong>of</strong> the pectin network (Matoh, 1997; O’Neill et al., 2004);<br />

stability <strong>of</strong> membranes (Goldbach et al., 2001), possibly through the formation <strong>of</strong> complexes with<br />

membrane glycoproteins (Goldbach and Wimmer, 2007); and anti-oxidative mechanisms associated<br />

with defence (Cakmak and Romheld, 1997). Interestingly B was found to play a role in quorum<br />

sensing <strong>of</strong> the bacterium V. harveyi whereby bonding with B was required to form a biologically<br />

active signal molecule (Chen et al., 2002). The role <strong>of</strong> B in plant defence is rapidly being uncovered<br />

and B has been shown to reduce the severity <strong>of</strong> a variety <strong>of</strong> diseases (Dordas, 2008), but the influence<br />

<strong>of</strong> plant B concentrations on the development and virulence <strong>of</strong> fungi is largely unknown.<br />

This experiment was specifically designed to test the influence <strong>of</strong> root applied B, at several<br />

concentrations, on tree growth and susceptibility <strong>of</strong> treated trees to attack by Q. pitereka.<br />

MATERIALS AND METHODS<br />

Mist culture system<br />

A mist culture B trial was conducted under glasshouse conditions using a randomised complete block<br />

design with 6 B treatments, 3 species, 2 solution tank blocks and 3 replicates (108 trees in total). The<br />

mist culture system used in the experiment is described in detail in Smith (2004). The system was a<br />

modified version <strong>of</strong> earlier mist culture designs used by Carter (1942), Martin and Hendrix (1966) and<br />

Hendrix and Lloyd (1968). The basics <strong>of</strong> the system consists <strong>of</strong> 12 nutrient solution tanks (200 l<br />

double lined food grade polypropylene with screw on lids) which each supply 9 atomiser chambers<br />

(white 25 l high density polyethylene mist chamber buckets) each. The atomiser chambers were<br />

randomly allocated to 3 benches in a temperature controlled glasshouse, with 3 chambers per bench.<br />

The water used in the experiment was purified by passing through a carbon filter, a water s<strong>of</strong>tener, a<br />

reverse osmosis unit (IBC 4000R model), then through an IBC® twin column unit with anion and<br />

cation resin beds to supply bulk purified water with an electrical conductivity <strong>of</strong> less than 0.1 µS cm -1 .<br />

The water was stored in a 3000 l polypropylene tank. It was then passed through an inline UV filter<br />

and a borate specific resin bed (Amberlite 743 resin), before being used in the experiment.<br />

The nutrient solutions were made up as concentrates (x 2000) in 20 l containers. Two concentrate<br />

solutions were used for the basal nutrient solutions, and were termed the ‘red solution’ (which<br />

contained iron and calcium) and the ‘blue solution’. Basal concentrate solutions were made using


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 265<br />

analytical reagent grade salts and quantities are detailed in Table 1. The desired concentrations in<br />

solution were attained by adding 100 ml <strong>of</strong> red solution concentrate, 100 ml <strong>of</strong> blue solution<br />

concentrate and 100 ml <strong>of</strong> the relevant boron treatment solution concentrate per 200 l <strong>of</strong> purified water<br />

(Table 2). Boron treatments were randomly allocated to tanks and associated mist chambers were<br />

randomised in the glasshouse.<br />

Table 1. Essential element concentrations (excluding B) in nutrient solutions atomised on to<br />

tree roots during the 4-month trial period.<br />

Element Elemental<br />

concentration in<br />

nutrient<br />

solutions<br />

Salt used Red solution<br />

concentrate*<br />

g salt l -1<br />

Blue solution<br />

concentrate*<br />

g salt l -1<br />

N<br />

μM<br />

357.5<br />

1414.9<br />

1400<br />

NH4NO3,<br />

KNO3<br />

Ca(NO3)2.4H2O<br />

28.6<br />

2890.8 a<br />

13224 b<br />

2830.8 c<br />

P 129.4 KH2PO4 704.2 d<br />

K 1544.2 KNO3 & KH2PO4 Inc. above a Ca 1400 Ca(NO3)2.4H2O Inc. above<br />

Inc. above b<br />

Mg 411.9 MgSO4.7H2O 4059.2 e<br />

S 411.9 MgSO4.7H2O Inc. above e<br />

Fe 14.3 C10H12N2O8FeNa 210.6<br />

Mn<br />

Zn<br />

1.46<br />

1.23<br />

MnSO4.H2O<br />

C10H12N2O8Zn.4H2O<br />

9.86<br />

23.1<br />

Cu 0.315 C10H12CuN2O8.2Na 5.0<br />

Mo 0.011 Na2MoO4.2H20 0.1<br />

*Concentrate solutions are 2000x the concentration <strong>of</strong> that in final nutrient solutions<br />

Table 2. Boron treatments applied in nutrient solutions atomised on to tree roots during the 4month<br />

trial period.<br />

Boron<br />

treatment<br />

Elemental concentration in<br />

nutrient solutions<br />

μM<br />

c & d<br />

Salt used Boron concentrate*<br />

g salt l -1<br />

1 0 H3BO3 0<br />

2 0.078 H3BO3 0.00966<br />

3 0.3125 H3BO3 0.03864<br />

4 1.25 H3BO3 0.1546<br />

5 5.0 H3BO3 0.6183<br />

6 20.0 H3BO3 2.473<br />

*Concentrate solutions are 2000x the concentration <strong>of</strong> that in final nutrient solutions<br />

The nutrient solutions were pumped using Onga® 442 pressure pumps (240V, 300W, water pressure<br />

100-200kPa) to the atomiser chambers via 25 mm polypropylene pipe (main lines) with 4 mm<br />

polypropylene risers connecting to plastic Netfin® mist nozzles fitted through the side <strong>of</strong> the buckets.<br />

Control boxes fitted with Hager EG400 programmable timers were used to turn on and <strong>of</strong>f pumps (240<br />

V) and in-line solenoid valves (12 V) at programmed intervals.<br />

The main lines circulated surplus solution back into the top <strong>of</strong> tanks, agitating the nutrient solution in<br />

the process. Polypropylene bungs with 2 mm holes drilled in the centre were placed on the end <strong>of</strong><br />

each return line into the nutrient solution tanks. This restricted the flow rate to 5.7 ml s -1 back into the<br />

tank and increased pressure in the line to ensure adequate atomisation <strong>of</strong> nutrient solutions inside the<br />

mist chamber buckets.<br />

Excess nutrient solutions from inside the mist chambers were run to waste to maintain a set nutrient<br />

concentration contacting the roots. A 13 mm black polypropylene drainpipe was inserted at a height <strong>of</strong>


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 266<br />

20mm from the base <strong>of</strong> the buckets for draining excess solution to waste. The nutrient solution in the<br />

base <strong>of</strong> each bucket maintained humidity in the atomiser chamber between sprays.<br />

Tree growth<br />

Two Corymbia species Corymbia torelliana, (open pollinated) (Ct.), C. citriodora subsp. variegata<br />

(Woondum Bulk) (Ccv.) and one hybrid <strong>of</strong> these species Ct. x Ccv. (CTVA 1 = CT2-002 x CV2-018)<br />

were used in this experiment. The trees were propagated as cuttings on the 7 th December 2006 and<br />

grown under glasshouse conditions at an ambient temperature <strong>of</strong> 30 o C for 1 month, prior to being<br />

moved into shade bays to grow to approximately 25cm in height.<br />

On the 1 st February 2007, a selection <strong>of</strong> uniform cuttings was removed from 90 mm plastic tray inserts<br />

(hyco) inserts and potting mix was washed from the plant’s roots using tap water. Plants were placed<br />

in a container <strong>of</strong> deionised water once all <strong>of</strong> the potting mix had been removed. When all 36 plants for<br />

that species had been prepared they were taken to the glasshouse. Individual plants were placed in<br />

foam squares with a slit cut halfway into one side. Plants were then placed in a polypropylene concrete<br />

chair and suspended through holes cut into the lids <strong>of</strong> the atomiser chamber buckets. Where required,<br />

plants were pruned back to a single dominant stem. Medium size white garbage bags were placed over<br />

each plant to produce a humid microclimate for 1 week whilst the trees became established in the<br />

system.<br />

The nutrient solutions were atomised onto the roots for a period <strong>of</strong> 10 seconds at intervals <strong>of</strong> every 30<br />

min between 4-6 am, every 20 min between 6–12 am, every 10 min between 12 am-3 pm, every 20<br />

min between 3–6 pm, every 30 min between 6-8 pm and then every 60 min between 8 pm-4 am<br />

(totalling 55 times a day). Each atomisation event sprayed approximately 95 ml <strong>of</strong> nutrient solution<br />

onto roots in each mist chamber. Therefore, approximately 47 l <strong>of</strong> nutrient solution per day was used<br />

to supply 9 trees. Glasshouse temperatures were set at 25 o C days and 20 o C nights.<br />

The trees were grown in their respective treatments in mist culture and inoculated in the glasshouse<br />

with a single isolate <strong>of</strong> Q. pitereka on the 23 rd April 2007.<br />

Inoculation<br />

Quambalaria pitereka isolate (BRIP 48385) was obtained from single lesions and grown on potato<br />

dextrose agar (PDA) for 2 to 3 weeks in the dark at 25 o C. A spore suspension (1x10 6 spores ml -1 ) was<br />

obtained by washing plates with sterile distilled water (SDW) to which two drops <strong>of</strong> Tween 20 had<br />

been added prior to inoculation.<br />

Foliage <strong>of</strong> trees was inoculated using a fine mist spray (2.9 kPa pressure) generated by a compressor<br />

driven spray gun (Iwata Studio series 1/6 hp; Gravity spray gun RG3, Portland, USA), to the upper<br />

and lower leaf surfaces until run<strong>of</strong>f was achieved. All trees were covered with plastic bags<br />

immediately after inoculation to maintain high humidity levels and to increase the period <strong>of</strong> leaf<br />

wetness. Bags were removed after 48 hours. Sub-samples <strong>of</strong> the spore suspension applied to the trees<br />

were placed onto PDA and incubated at 25 o C for 48 hours to ensure that the spores were viable.<br />

Disease incidence (I) and severity (S) was assessed for all trees 14 days after inoculation, from which<br />

a Quambalaria shoot blight (QSB) score was calculated (I xS/100).<br />

Statistical analysis<br />

An analysis <strong>of</strong> variance (ANOVA) was performed on all data using GenStat® version 9 (Lawes<br />

Agricultural Trust, Rothamsted Experimental Station). For each ANOVA, the mean separation was<br />

determined by calculating the least significant difference (Fisher’s LSD) at P


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 267<br />

RESULTS<br />

As tree B status increased from deficient (9-12 µg g -1 YFEL B) to sufficient B status (22-24 µg g -1<br />

YFEL B) there were increases in tree height (107%), stem diameter (62%), total tree weight (194%)<br />

and root weight (144%)(Tables 3), whilst Quambalaria shoot blight incidence, severity, stem incidence<br />

and disease score, and resultant shoot death decreased (Table 4).<br />

Table 3. Effect <strong>of</strong> root applied B treatment sprays to Corymbia species on foliar B and tree<br />

growth (averaged across Corymbia species)<br />

Treatment YFEL B Tree height Stem diameter Total tree Root weight<br />

(at base)<br />

weight<br />

(μM) (mg kg -1 ) (cm) (mm) (g) (g)<br />

0 11.8 94 1.02 285 95<br />

0.078 8.9 124 1.21 439 126<br />

0.3125 9.6 162 1.47 670 194<br />

1.25 17.4 165 1.21 504 122<br />

5.0 22.2 191 1.53 805 201<br />

20.0 23.5 195 1.65 839 233<br />

P


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 268<br />

Table 5. Effect root applied B treatments sprays to Corymbia species on tree growth.<br />

YFEL B Tree height Total tree weight<br />

(mg kg -1 Treatment<br />

(μM)<br />

) (cm) (g)<br />

Ct. Ct. x Ccv. Ct. Ct. x Ccv. Ct. Ct. x Ccv.<br />

Ccv.<br />

Ccv.<br />

Ccv.<br />

0 16.5 8.5 10.3 93 118 70 361 398 96<br />

0.078 10.5 7.5 8.9 136 143 94 546 560 210<br />

0.3125 11.7 8.8 8.2 186 173 128 949 764 297<br />

1.25 19.9 17.2 15.1 159 176 161 670 459 382<br />

5.0 24.0 18.8 23.8 212 194 167 1182 719 514<br />

20.0 24.7 20.4 25.5 201 200 184 1226 718 572<br />

P 0.012


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 269<br />

loss <strong>of</strong> apical dominance leading to bushy crowns and shrub like growth characteristics, and in severe<br />

cases, tree death (Pegg et al., 2009). Symptom development is rapid and occurs after 7 to 10 days<br />

under conditions <strong>of</strong> high humidity. While the identification <strong>of</strong> host resistance is a key to the further<br />

development <strong>of</strong> spotted gum as a plantation species, risk minimisation strategies are also required to<br />

reduce potential losses. The finding that a balanced tree nutrition approach and specifically B can<br />

reduce the susceptibility to Q. pitereka infections is encouraging and emphasis needs to be placed on<br />

the role <strong>of</strong> nutrition in plant defence as well as tree growth in deciding appropriate nutrient<br />

management systems in plantation forestry.<br />

The Ct. x Ccv. hybrid had reductions in Q. pitereka infection rates due to increasing tree B status from<br />

deficient to sufficient and appreciably lower susceptibility to Q. pitereka infections than Ccv.,<br />

demonstrating a successful outcome <strong>of</strong> the breeding program (Lee, 2007). A breeding strategy that<br />

encompasses selection for disease tolerance and nutrient efficiency as well as tree growth and wood<br />

quality factors will be more suitable for widespread deployment <strong>of</strong> clones across a range <strong>of</strong><br />

environments. However, even with improved progeny, underlying nutrient deficiencies or imbalances<br />

need to be addressed to maintain health and vigour in new and existing plantations.<br />

The function that B plays in plant defence against fungal attack is unclear and further research is<br />

required to determine the mechanism for reduced susceptibility to Q. pitereka in trees supplied<br />

adequate B verses B deficient trees. The treatment levels were insufficient to have a direct fungicidal<br />

effect on the Q. pitereka, however we are uncertain about whether B is playing a direct or indirect role<br />

in plant defence. Pegg et al. (2009) demonstrated that infection and colonisation <strong>of</strong> Corymbia spp. by<br />

Q. pitereka occurs through intercellular growth <strong>of</strong> hyphae following penetration and remains<br />

intercellular until host cell death. A localised interaction zone is apparent at points <strong>of</strong> interaction<br />

between hyphae and host cell walls. It was assumed that Q. pitereka obtains nutrients from the host<br />

cells through these interaction zones. While the influence <strong>of</strong> B on disease development requires more<br />

in-depth studies, B is known to have a direct function in cell wall structure (Matoh, 1997; O’Neill et<br />

al., 2004; Brown et al., 2002). The reduction in cell wall permeability to fungal products produced by<br />

Q. pitereka during the infection process may influence the development <strong>of</strong> disease. Similarly more<br />

research is required to investigate the plant defence implications <strong>of</strong> other reported roles <strong>of</strong> B in the<br />

stability <strong>of</strong> membranes (Goldbach et al., 2001) and anti-oxidative mechanisms (Cakmak and Romheld,<br />

1997).<br />

These studies have shown that B is required for early tree growth <strong>of</strong> Corymbia species and hybrids,<br />

and that increasing tree B to sufficient reduced the susceptibility to Q. pitereka infections in the<br />

establishment phase. Critical foliar B concentrations appear to be in the order <strong>of</strong> 17.2-20.4 mg B kg -1<br />

for the Ct. x Ccv. hybrid and >25 mg B kg -1 for Ccv.. However, more research is required to develop a<br />

better understanding <strong>of</strong> pathogen biology, such as variability in isolate aggressiveness and the effect <strong>of</strong><br />

repeated infection on overcoming host resistance to help formulate a disease management strategy and<br />

to determine the effectiveness <strong>of</strong> increasing tree B status to sufficient on disease rates <strong>of</strong> trees in<br />

plantations that are already infected or have older foliage that may have different host disease<br />

interactions. It must also be stressed that due care needs to be taken with B application rates which<br />

vary depending on soil type and soil B adsorption characteristics (Smith, 2004), to avoid B toxicity<br />

scenarios, particularly on course textured soils. Aiming towards a balanced nutrient management<br />

system in plantation forestry is desirable for plant health and applications <strong>of</strong> B alone are unlikely to<br />

achieve that goal if other essential elements are in limited supply.<br />

ACKNOWLEDGEMENTS<br />

The authors would like to acknowledge the valuable contributions <strong>of</strong> Donna Richardson for<br />

maintaining the trial and assistance with measures and other operations.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 270<br />

REFERENCES<br />

Atkinson D and McKinlay RG 1997 Crop protection and its integration within sustainable farming systems.<br />

Agriculture, Ecosystems and Environment 64, 87-93<br />

Brown PH, Bellaloui N, Wimmer MA, Bassil ES, Ruiz J, Hu H, Pfeffer H, Dannel F, Romheld V 2002, Boron in<br />

plant biology. Plant Biology 4,205-223<br />

Burgess TI and Wingfield MJ 2002, Impact <strong>of</strong> fungi in natural forest ecosystems; A focus on Eucalyptus. In<br />

`Microorganisms in Plant Conservation and Biodiversity’ (Eds. K Sivasithamparam, KW Dixon, RL<br />

Barrett). Kluwer Academic Publishers, Dordrecht, pp. 285-306<br />

Cakmak I and Romheld V 1997, Boron deficiency-induced impairments <strong>of</strong> cellular functions in plants. Plant and<br />

Soil 193, 71-83<br />

Carnegie AJ 2007, Forest health condition in New South Wales, <strong>Australia</strong>, 1996-2005. II. Fungal damage<br />

recorded from eucalypt plantations during forest health surveys and their management. Australasian<br />

Plant Pathology 36, 225-39<br />

Carter W 1942, A method <strong>of</strong> growing plants in water vapour to facilitate examination <strong>of</strong> roots. Phytopathology.<br />

19, 407-412<br />

Chen X, Schauder S, Potier N, Van Dorsselaer A, Pelczer I, Bassler BL, Hughson FM 2002, Structural<br />

identification <strong>of</strong> a bacterial quorum-sensing signal containing boron. Nature 415 (6871), 545-549<br />

Dell B, Hardy G, and Burgess T 2008, Health and nutrition <strong>of</strong> plantation eucalypts in Asia. Southern Forests<br />

70(2), 131-138<br />

Dordas C 2008, Role <strong>of</strong> nutrients in controlling plant diseases in sustainable agriculture. A review. Agronomy<br />

and Sustainable Development 28, 33-46<br />

Drenth A 2004, Fungal epidemics – does spatial structure matter? New Phytologist 163, 4-7.<br />

Goldbach HE, Yu Q, Wingender R, Schulz M, Wimmer M, Findeklee P, and Baluska F 2001, Rapid response<br />

reactions <strong>of</strong> roots to boron deprivation. Journal <strong>of</strong> Plant Nutrition and Soil Science 164, 173-181<br />

Goldbach HE and Wimmer M 2007, Boron in plants and animals: Is there a role beyond cell-wall structure?<br />

Journal <strong>of</strong> Plant Nutrition and Soil Science 170, 39-48<br />

Hendrix JW and Lloyd EH 1968, An improved Washington State University mist chamber for root disease and<br />

root research. Washington Agricultural Experimental Station Bulletin 700, 5<br />

Isbell RF 1986, `The <strong>Australia</strong>n Soil Classification’. CSIRO Publishing, Collingwood, Victoria. pp. 41-44<br />

Lambert MJ, Turner J, and Knott J 1997, Boron nutrition <strong>of</strong> radiate pine plantations in <strong>Australia</strong>. In `Boron in<br />

Soils and Plants’ (Eds. R.W Bell and B. Rerkasem). Kluwer Academic Publishers, Netherlands. pp. 83-88<br />

Lee D 2007, Development <strong>of</strong> Corymbia species and hybrids for plantations in eastern <strong>Australia</strong>. <strong>Australia</strong>n<br />

Forestry 70, 11-16<br />

Martin NE and Hendrix JW 1966, An apparatus for the mist culture <strong>of</strong> wheat. Plant Disease Reporter. 50(6),<br />

369-371<br />

Matoh T 1997, Boron in plant cell walls. Plant and Soil 193, 59-70<br />

Old KM 1990, Diseases caused by fungi. In `Trees for Rural <strong>Australia</strong>’ (Ed. KW Cremer). Inkata Press,<br />

Melbourne, <strong>Australia</strong>.<br />

O’Neill MA, Ishii T, Albersheim P, and Darvill AG 2004, Rhamnogalacturonan II: Structure and function <strong>of</strong> a<br />

borate cross-linked cell wall pectic polysaccharide. Annual Review <strong>of</strong> Plant Biology 55, 109-139<br />

Pegg GS, Drenth A, Wingfield MJ 2005, Quambalaria pitereka on spotted gum plantations in Queensland and<br />

northern New South Wales, <strong>Australia</strong>. International Forestry Review 7, 337.<br />

Pegg GS, O’Dwyer C, Carnegie AJ, Wingfield MJ, Drenth A 2008, Quambalaria species associated with<br />

plantation and native eucalypts in <strong>Australia</strong>. Plant Pathology 57, 702-14<br />

Pegg GS, Webb RI, Carnegie AJ, Wingfield MJ, Drenth A 2009, Infection and disease development <strong>of</strong><br />

Quambalaria spp. on Corymbia and Eucalyptus species. Plant Pathology Doi:10.1111/j.1365-<br />

3059.2009.02087.x<br />

Smith TE 2004, Boron nutrition <strong>of</strong> Hass avocado (Persia americana Mill.). PhD Thesis, the University <strong>of</strong><br />

Queensland, awarded 19 th July 2004. 196 p.<br />

Smith TE 2007, Poor tree form in Eucalyptus nitens linked to boron deficiency. In `Advances in Plant and<br />

Animal Boron Nutrition’ (Eds. F Xu, H.E. Goldbach, PH Brown, RW Bell, T Fujiwara, CD Hunt, S<br />

Goldberg, and L Shei). Springer. pp 163-169<br />

Self NM, Aitken, EAB, Dale MD 2002, Susceptibility <strong>of</strong> provenances <strong>of</strong> spotted gums to Ramularia shoot blight.<br />

New Zealand Plant Protection 55, 68-72.<br />

Walker J, Bertus AL 1971, Shoot blight <strong>of</strong> Eucalyptus spp. caused by an undescribed species <strong>of</strong> Ramularia.<br />

Proceedings <strong>of</strong> the Linnean Society <strong>of</strong> New South Wales. 96, 108-15<br />

Wingfield MJ, Slippers B, Roux J, Wingfield BD 2001, Worldwide movement <strong>of</strong> exotic forest fungi especially<br />

in the tropics and Southern Hemisphere. Bioscience 51, 134-40


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 271<br />

ABSTRACT<br />

THE INFLUENCE OF CLIMATE CHANGE ON<br />

SOIL RESPIRATION WITH EUCALYPTUS SALIGNA<br />

Vivien de Rémy de Courcelles 1 , Bhupinderpal-Singh, Mark Adams<br />

Belowground respiration takes a major part in the cycle <strong>of</strong> carbon between the<br />

atmosphere and the soil; the soil possesses the biggest pool <strong>of</strong> the terrestrial carbon. It<br />

might be influenced by the current rapid increase in atmospheric [CO2]. Within the<br />

Hawkesbury Forest Experiment (HFE) at the University <strong>of</strong> Western Sydney (Richmond),<br />

twelve Whole Tree Chambers (WTC) were installed. Each WTC enclosed a Eucalyptus<br />

saligna tree that was grown from a seedling stage to a maximum <strong>of</strong> 10 m height for up to<br />

2 years. These trees in WTC were subjected to a factorial combination <strong>of</strong> ambient or<br />

elevated atmospheric [CO2] and irrigation or drought treatments. We measured the soil<br />

CO2 efflux from March 2008 to February 2009. First results showed strong inter-seasonal<br />

variations in soil respiration with more respiration occurring during the warmer months.<br />

There was no significant difference between the two CO2 treatments. On the other hand<br />

the irrigation treatment resulted in a significant higher rate <strong>of</strong> soil respiration than the<br />

drought treatment.<br />

INTRODUCTION<br />

The atmospheric concentration <strong>of</strong> carbon dioxide has increased by 36% since the mid 18 th century,<br />

from a pre-industrial value <strong>of</strong> about 280 ppm to 379 ppm in 2005, and continues to increase by an<br />

average <strong>of</strong> 19 ppm/year (Forster et al., 2007). Soil CO2 efflux (SCE), commonly referred to as<br />

belowground or soil respiration, is a major flux <strong>of</strong> carbon dioxide toward the atmosphere. Averaged<br />

over 18 European forests, SCE equaled 69% <strong>of</strong> the total terrestrial ecosystem respiration (Janssens et<br />

al., 2001). In other words, an estimated 10% <strong>of</strong> the carbon dioxide <strong>of</strong> the atmosphere cycles through<br />

soils each year (Raich and Potter, 1995), which represents about 10 times the amount <strong>of</strong> CO2 produced<br />

by burning fossil fuels.<br />

Despite its importance in the carbon cycle, belowground respiration is not as well known as its aboveground<br />

counterpart. In part this is due to the difficulty <strong>of</strong> isolating and quantifying its components.<br />

Belowground soil respiration must be divided between autotrophic and heterotrophic sources.<br />

Autotrophic respiration can be clearly identified as root respiration (Kuzyakov, 2006). On the other<br />

hand, heterotrophic respiration includes respiration from free-living microorganisms in soil (Cisneros-<br />

Dozal et al., 2006). Because <strong>of</strong> the methodological constraints, respiration from rhizosphere<br />

microorganisms, including mycorrhizal fungi, is usually coupled with the autotrophic respiratory<br />

component and the argument for this is that these microorganisms are highly dependent on plants for<br />

their supply <strong>of</strong> recently-fixed carbohydrates via photosynthesis. In the strict sense, soil respiration<br />

from these sources should however be classified as heterotrophic respiration.<br />

The increase in atmospheric [CO2] and the resulting changes in climate can impact on part or all<br />

components <strong>of</strong> soil respiration. Equally, the response in soil CO2 efflux will depend on the land use,<br />

species composition, age class <strong>of</strong> the vegetation as well as disturbances other than climatic (Zak et al.,<br />

2000).<br />

Because <strong>of</strong> the size <strong>of</strong> trees and that <strong>of</strong> their systems, field based experimental designs for studying the<br />

effect <strong>of</strong> climate change on trees and forest require heavy and costly equipment. Some <strong>of</strong> those<br />

devices include open top chambers (OTC) (Leadley and Drake, 1993, Mandl et al., 1973) and closed<br />

top or whole tree chambers (WTC) (Kellomaki et al., 2000, Medhurst et al., 2006) at the tree level,<br />

whereas Free Air CO2 Enrichment (FACE) (Hendrey et al., 1999, Lewin et al., 1994) was used for<br />

studies at the plot level.<br />

1 University <strong>of</strong> Sydney. Email: v.decourcelles@usyd.edu.au.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 272<br />

Here we present results gathered over a year <strong>of</strong> an experiment where Eucalyptus saligna are growing<br />

in WTC under the climate change conditions that are predicted to happen within the next 100 years in<br />

eastern <strong>Australia</strong>.<br />

MATERIALS AND METHOD<br />

Site description<br />

The Hawkesbury Forest Experiment (HFE) is located at the University <strong>of</strong> Western Sydney campus in<br />

Richmond, North-West <strong>of</strong> Sydney, <strong>Australia</strong> (Lat. 33°36’40” S, Lon. 150°44’26.5” E). It sits on the<br />

alluvial floodplain <strong>of</strong> the Hawkesbury River at elevation 25 m a.s.l. The soil is a chromosol, sandy in<br />

texture, with a pH around 5.5 in the first 40 cm, and low in C (


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 273<br />

Two environmental variables were employed inside the chambers: moisture and CO2 concentration.<br />

At the start <strong>of</strong> the experiment all WTC and outside control plots were irrigated with 10 mm <strong>of</strong> water<br />

every third day to ensure a good establishment <strong>of</strong> the trees. A first drought cycle started on 16<br />

February 2008 when all irrigation ceased for half <strong>of</strong> the experimental plots. The drought was broken<br />

after more than 200 days on 18 September 2008 in order to rewet the upper soil level. A second<br />

drought cycle started on 26 October 2008 and lasted until the end <strong>of</strong> the experiment.<br />

The [CO2] was either ambient or elevated and the treatment started at the day <strong>of</strong> planting. The ambient<br />

CO2 varied between 380 ppm and 500 ppm depending on day or night-time, and elevated CO2 was<br />

maintained 240 ppm above the ambient level.<br />

There were 3 replicates <strong>of</strong> each combination <strong>of</strong> water and CO2 treatment, i.e. ambient [CO2]-dry,<br />

ambient [CO2]-irrigated, elevated [CO2]-dry, elevated [CO2]-irrigated.<br />

Six controls including one tree each were also set up to evaluate the WTC effect on the measured<br />

parameters: a frame bearing a plastic sheet was installed 300 to 200 mm over the soil and around the<br />

stem preventing rainfall from reaching the soil. Sprinklers controlled the moisture delivered to the soil<br />

around the trees in accord with the same irrigated and drought treatments applied inside the WTC.<br />

Belowground respiration measurements<br />

Soil respiration measurements were first done using static chambers 204 mm tall and 250 mm outside<br />

diameter for 236 mm internal diameter. Chambers were inserted in the soil in February 2008 within<br />

each WTC as well as in the control plots to a depth <strong>of</strong> about 50mm: Two chambers were placed at<br />

about 500 mm from the stem <strong>of</strong> the trees and 2 more at about 1 metre distance. Measurements were<br />

taken on a monthly basis from March to May, and again in September and October 2008. A 25 ml gas<br />

sample was taken when the lid was closed and immediately transferred to an evacuated 12ml<br />

Exetainer tube. A second sample was taken 30 minutes later. Samples were analysed on a CO2 gas<br />

analyser at the University <strong>of</strong> Western Sydney (UWS). The rate <strong>of</strong> soil respiration was extrapolated<br />

from the slope <strong>of</strong> the concentration line joining the two measurement points. In a separate campaign,<br />

we collected the CO2 gas samples at three intervals (0, 30, 60 min) over a 60 min period in order to<br />

check for linearity in the gas accumulation rate inside the chamber. We always found a strong linearity<br />

(R2 > 0.95) in the gas accumulation rate.<br />

In September 2008, we installed 2 sets <strong>of</strong> cores into the soil <strong>of</strong> each WTC and control plots based on<br />

the design by Heinemeyer et al (Heinemeyer et al., 2007). Each set consists <strong>of</strong> one 50 mm tall x 200<br />

mm diameter ring pressed against the soil. Two 250 mm tall x 200 mm diameter cores were inserted to<br />

a depth <strong>of</strong> 200 mm into the soil. They were equipped with four 50 x 50 mm windows on their side<br />

fitted with a 38-μm nylon mesh or a 1-μm nylon mesh that allows or prevents the in-growth <strong>of</strong><br />

mycorrhizal hyphae inside the cores, respectively (see Heinemeyer et al., 2007 for details). A deeper<br />

700 mm long and 100 mm diameter core without windows was also inserted in each WTC or control<br />

plot. It excludes any roots, including the deeper growing ones. The timeframe for this study did not<br />

allow enough time for roots to decompose inside the mesh-collars, and only the results from the<br />

shallow collars are considered here.<br />

We measured soil respiration out <strong>of</strong> these mesh-collars with a Vaisala carbocap GM 343 mounted on a<br />

5.25 litter lid. The lid-Vaisala-collar association works as a static chamber equipped with its own CO2<br />

analyser. The lid was left on the collars for 10 minutes out <strong>of</strong> which only the last eight were<br />

considered in order to calculate the rate <strong>of</strong> evolution <strong>of</strong> CO2. The recording during the first 2 minutes<br />

was discarded in order to allow the [CO2] to stabilise.<br />

Respiration measurements were carried out in the first week <strong>of</strong> November and December and then<br />

every 3 weeks until 9 February.<br />

In order to compare the respiration results obtained using static chambers-gas sampling and Vaisala<br />

direct measurements, both techniques were used in October. Out <strong>of</strong> ten plots, seven had a soil<br />

respiration measured with the Vaisala within ±20% <strong>of</strong> the respiration measured using the static<br />

chambers; one was about 21% higher and the remaining two were more than 50% lower than those<br />

measured using the static chambers.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 274<br />

Statistical analysis<br />

Belowground respiration between [CO2] treatments, between ambient WTC and outside controls, and<br />

between irrigated and drought treatments were analysed using two-way ANOVA.<br />

RESULTS<br />

Respiration rate over the year<br />

Respiration rate was very variable throughout the year for WTC and control plots. It varied between<br />

158−181 mg CO2/m 2 /h in September or May 2008 to between 416−502 mg CO2/m 2 /h in March 2008<br />

or February 2009, indicating that soil respiration was higher in the warmer than colder months.<br />

Response to the [CO2] treatment<br />

Soil respiration followed a similar annual pattern in the elevated and ambient [CO2] WTC (Figure 1).<br />

Soil CO2 efflux for the ambient plots ranged from a low 181 mg CO2/m 2 /h in September 2008 to a<br />

high 502 mg CO2/m 2 /h in December 2008. The elevated [CO2] plots released a minimum <strong>of</strong> 158 mg<br />

CO2/m 2 /h in September 2008 and a maximum <strong>of</strong> 416 mg CO2/m 2 /h in December 2008 (Figure 2).<br />

The belowground efflux from soil <strong>of</strong> the ambient WTC was greater than the elevated ones except in<br />

March and April 2008. However, there was no significant difference (P>0.05) in belowground<br />

respiration between ambient or elevated plots for any <strong>of</strong> the measurement days, apart from January<br />

2009.<br />

SCE (mgCO2/m 2 /h)<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

1/3/08 1/5/08 1/7/08 1/9/08 1/11/08 1/1/09 1/3/09<br />

Figure 2. Time series <strong>of</strong> Soil CO2 efflux (SCE) at HFE averaged over the 6 control plots (…○…),<br />

the 6 elevated [CO2] WTC (–■–) and the 6 ambient [CO2] (–□–) displayed with<br />

negative standard error bars<br />

Response to irrigation treatment<br />

Across all treatments, the soil <strong>of</strong> the irrigated plots produced significantly more carbon dioxide than<br />

the soil <strong>of</strong> the dry plots, except on three occasions in March, April and October 2008 (Figure 3). In<br />

April, the soil respiration values were not significantly different across the WTC regardless <strong>of</strong> the<br />

treatment, but there was a significant difference between wet and dry treatment in the control plots.<br />

The irrigated control plots showed a significantly (P< 0.05) higher rate <strong>of</strong> soil respiration compared to<br />

the droughted ones in April and from December to January inclusive, with a less marked difference in<br />

October (P=0.1) and November (P=0.057). Inside the ambient [CO2] WTC, during the summer<br />

months, the rate <strong>of</strong> soil respiration was significantly (P


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 275<br />

SCE (mgCO2/m 2 /h)<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

1/3/08 1/5/08 1/7/08 1/9/08 1/11/08 1/1/09 1/3/09<br />

Figure 3. Time series <strong>of</strong> Soil CO2 efflux (SCE) measured at HFE between March 2008 and<br />

February 2008 plotted against the watering treatment. Mean <strong>of</strong> 3 control plots, 3<br />

ambient WTCs and 3 elevated WTCs subjected to drought treatment (receiving no<br />

water) (–■–) or water treatment (irrigated) ( … ○ … )<br />

Interaction between treatments<br />

No interaction between [CO2] and watering treatment was recorded for any measurement day or across<br />

all measurements.<br />

Chamber effect<br />

There was a significant difference in belowground soil respiration rates between the open air control<br />

plots and the ambient WTC plots in March, September and from November 2008 to February 2009.<br />

However in March and September 2008, the soil in the control plots released more carbon dioxide<br />

than the soil inside the ambient WTC whereas the trend was reversed from November 2008 (see<br />

Figure 1).<br />

DISCUSSION<br />

Elevated [CO2] effect<br />

For most <strong>of</strong> the measurements, more carbon dioxide was emitted from soils in the ambient than the<br />

elevated [CO2] WTC, but there was no significant difference except for January 2009. Most studies<br />

noted an increase in carbon dioxide efflux from soils under vegetation exposed to an atmosphere with<br />

elevated [CO2] (Bernhardt et al., 2006, Hamilton et al., 2002, Lin et al., 2001, Pajari, 1995, Wan et al.,<br />

2007, Zak et al., 2000). Others have not recorded any effect <strong>of</strong> elevated atmospheric CO2 enrichment<br />

(Oberbauer et al., 1986, Tingey et al., 2006). At time <strong>of</strong> writing, we are yet to link specific soil<br />

respiration with the growth <strong>of</strong> the tree for each plot. The results will be presented during the<br />

conference.<br />

While root respiration has been evaluated as representing 10 to 90 % <strong>of</strong> total soil respiration in<br />

different studies (Hanson et al., 2000), the development <strong>of</strong> new techniques has allowed to refine these<br />

values. Root and mycorrhizal respiration accounted for up to 56% <strong>of</strong> total soil respiration for the year<br />

following a tree girdling experiment (Hogberg et al., 2001); this difference was increased up to 65% in<br />

the second year i.e. when contributions from stored carbohydrates in roots and from decaying<br />

mycorhizal roots were reduced to a very low value two years after the girdling treatment<br />

(Bhupinderpal et al., 2003). Heinemeyer et al. (2007) found that heterotrophic respiration accounted<br />

for 60% <strong>of</strong> total soil CO2 efflux, whereas ectomycorrhizas and roots contributed about 25% and 15%,<br />

respectively. In our experiment, we didn’t achieve that same level <strong>of</strong> separation <strong>of</strong> the components <strong>of</strong><br />

soil respiration using similar mesh-collar chambers as used by Heinemeyer et al. (2007) in a boreal<br />

forest soil system. However, our 700 mm long collar must have excluded most roots and their


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 276<br />

associated mycorrhiza and therefore the respiration recorded on these collars must be <strong>of</strong> ‘true<br />

heterotrophic’ origin only. Therefore, that portion <strong>of</strong> soil respiration due only to contribution from<br />

roots was, on average, 63% in November and 66% for the two measurements made in December 2008.<br />

There was no difference between CO2 treatments. It should be noted that the floor <strong>of</strong> the WTCs<br />

prevented litter from falling on the ground. We have therefore missed the input <strong>of</strong> microbial<br />

decomposers feeding on the surface litter that would be considered in a natural environment, and thus<br />

these numbers are probably an overestimation <strong>of</strong> the contribution <strong>of</strong> root respiration to the total<br />

belowground CO2 efflux.<br />

Comparison across studies can be deceptive since the intensity <strong>of</strong> stimulation <strong>of</strong> soil respiration varies<br />

with the age or the composition <strong>of</strong> the stands (King et al., 2004, Pregitzer et al., 2006) and the length<br />

<strong>of</strong> exposition to elevated [CO2] (King et al., 2004). Kasurinen et al. (2004) observed a positive effect<br />

<strong>of</strong> elevated [CO2] on soil carbon dioxide efflux under one clone <strong>of</strong> silver birch (Betula pendula)<br />

whereas the effect was negative on the soil respiration under another clone with the difference<br />

increasing after three years. Therefore conclusions should not be drawn before considering the<br />

particular conditions <strong>of</strong> any experimental design.<br />

Irrigation treatment<br />

Our results <strong>of</strong> greater soil CO2 efflux from the watered WTCs than from the droughted WTCs are<br />

consistent with the results <strong>of</strong> many other studies. In Brazil, irrigated clones <strong>of</strong> Eucalyptus grandis x<br />

urophylla produced a higher rate <strong>of</strong> soil respiration than the rainfed counterparts over 2 years <strong>of</strong><br />

experiment (Stape et al., 2008). The drought and rainy periods <strong>of</strong> the year were matched by<br />

respectively low and high rates <strong>of</strong> SCE in eucalypts forest (Stape et al., 2008) as well as in conifers,<br />

broadleaf mixed and evergreen broadleaf forests (Tang et al., 2006).<br />

Jassal et al. (2008) have found that water stress in a Douglas fir forest significantly reduced carbon<br />

dioxide emissions from soils. Furthermore, under a threshold <strong>of</strong> 11 m 3 .m -3 <strong>of</strong> soil water at 4 cm depth,<br />

belowground respiration was not affected by soil temperature, but was linearly dependent on soil<br />

moisture content. The drought treatment in our experiment achieved a similar reduction in soil CO2<br />

efflux most <strong>of</strong> the time. The only exceptions occurred right at the start <strong>of</strong> both drought cycles. The<br />

most notable one was the October measurement following irrigation <strong>of</strong> the dry plots. They showed a<br />

greater, though not significant, belowground respiration rate than did their irrigated counterparts.<br />

The increase in soil respiration rate occurring after the rewatering <strong>of</strong> soils previously submitted to a<br />

lengthy dry episode, is called “Birch effect” named after H.F. Birch, one <strong>of</strong> the first to describe this<br />

phenomenon (Jarvis et al., 2007). Four reasons are commonly listed to explain the Birch effect:<br />

1. Successive drying and wetting <strong>of</strong> soils make organic matter available for<br />

decomposition by breaking soil aggregates.<br />

2. The dry spells kill soil micro-organisms which are subsequently decomposed when<br />

wetting occurs, hence releasing nutrients.<br />

3. Wetting triggers a sudden growth <strong>of</strong> microbial biomass and fungal hyphae.<br />

4. Microbes respond to the osmotic shock created by the rewetting by releasing carbon<br />

rich solutes that accumulated in their cytoplasm during the dry periods (Jarvis et al.,<br />

2007).<br />

Recent studies have shown both an increase in yearly forest soil CO2 emissions (Jarvis et al., 2007)<br />

and no extra release <strong>of</strong> CO2 (Muhr et al., 2008) following the wetting <strong>of</strong> previously dry soils. In our<br />

case, the increase <strong>of</strong> SCE in the dry plots following watering, led to a higher rate <strong>of</strong> respiration than<br />

the irrigated plots but not significantly so. Moreover, the effect <strong>of</strong> the rewatering was cancelled by the<br />

following month measurement. These results would tend to indicate that over a year no significant<br />

amount <strong>of</strong> extra CO2 would be emitted from our soils because <strong>of</strong> the Birch effect.<br />

The <strong>Australia</strong>n climate <strong>of</strong> the 21 st century is predicted to bring longer dry spells broken by heavier<br />

rainfall episodes (CSIRO and BOM, 2007). These irregular wetting-drying events make prediction <strong>of</strong><br />

the pattern <strong>of</strong> future belowground respiration a very hard task.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 277<br />

Chamber effect<br />

Similar chamber effects (i.e. greater soil respiration in the control plots than in the WTC) as the one<br />

we recorded in March and September 2008 have been observed by Niinisto et al (2004) in their WTC<br />

experiment. However, over the 4 years <strong>of</strong> their experiment, there was no reversal <strong>of</strong> the trend <strong>of</strong><br />

greater soil respiration in the control plots than in the ambient WTC to lower respiration in the control<br />

plots than in the WTCs.<br />

Belowground respiration is linked positively to the temperature measured at the soil surface (Niinisto<br />

et al., 2004, Rustad et al., 2001, Bronson et al., 2008, Pajari, 1995) as well as the soil moisture (Keith<br />

et al., 1997, Tang et al., 2006, Orchard and Cook, 1983). Our measurements <strong>of</strong> soil temperatures<br />

(result not presented) show that for both the September and October measurements, soils in the outside<br />

control plots were on average 1.4ºC warmer than in the ambient WTC, a significant difference,<br />

whereas the difference ranged from a 0.1 ºC cooler to a 0.9 ºC warmer for the other months. Except in<br />

May, soil moisture was always lower in the control than the WTC plots. Both these factors play a role<br />

in the soil CO2 efflux and their respective importance might change according to the seasons: soil<br />

moisture can become a limiting factor in summer whereas temperature would be a limiting factor in<br />

winter. Their interaction is a determinant for soil respiration and would explain most variations. The<br />

WTC experimental design at HFE has an effect on soil conditions that varies between seasons; this<br />

could explain the observed switch from greater to lower soil respiration in the control plots compared<br />

to ambient WTC.<br />

CONCLUSIONS<br />

Increase in the atmospheric [CO2] from an ambient level to an elevated level did not induce significant<br />

changes in the soil CO2 efflux from a Eucalyptus saligna plantation. However, compared to the<br />

irrigated treatment, the drought treatment led to a significant decrease in soil CO2 efflux. .<br />

Considering the predictions <strong>of</strong> increasing dry spells in future climate change scenarios in <strong>Australia</strong>,<br />

soil respiration could be predicted to decrease in the next few years; however, irregular rainfall events<br />

may lead to greater soil respiration than in areas subjected to long dry spells. Because <strong>of</strong> the chamber<br />

effect that we observed and the relatively short duration <strong>of</strong> the present study, extrapolation <strong>of</strong> these<br />

results to field and ecosystem scales should be undertaken with caution.<br />

Long-term, studies are needed to better understand the future effect <strong>of</strong> climate change on soil<br />

respiration.<br />

REFERENCES<br />

BERNHARDT, E. S., BARBER, J. J., PIPPEN, J. S., TANEVA, L., ANDREWS, J. A. & SCHLESINGER, W.<br />

H. (2006) Long-term effects <strong>of</strong> free air CO2 enrichment (FACE) on soil respiration. Biogeochemistry, 77,<br />

91-116.<br />

BHUPINDERPAL, S., NORDGREN, A., LOFVENIUS, M. O., HOGBERG, M. N., MELLANDER, P. E. &<br />

HOGBERG, P. (2003) Tree root and soil heterotrophic respiration as revealed by girdling <strong>of</strong> boreal Scots<br />

pine forest: extending observations beyond the first year. Plant Cell and Environment, 26, 1287-1296.<br />

BRONSON, D. R., GOWER, S. T., TANNER, M., LINDER, S. & VAN HERK, I. (2008) Response <strong>of</strong> soil<br />

surface CO2 flux in a boreal forest to ecosystem warming. Global Change Biology, 14, 856-867.<br />

CISNEROS-DOZAL, L. M., TRUMBORE, S. & HANSON, P. J. (2006) Partitioning sources <strong>of</strong> soil-respired<br />

CO2 and their seasonal variation using a unique radiocarbon tracer. Global Change Biology, 12, 194-204.<br />

CSIRO & BOM (2007) Climate change in <strong>Australia</strong>-Technical report 2007. <strong>Australia</strong>n Bureau <strong>of</strong> Meteorology<br />

(BOM)<br />

FORSTER, P., RAMASWAMY, V., ARTAXO, P., BERNTSEN, T., BETTS, R., FAHEY, D. W., HAYWOOD,<br />

J., LEAN, J., LOWE, D. C., MYHRE, G., NGANGA, J., PRINN, R., RAGA, G., SCHULZ, M. &<br />

DORLAND, A. R. V. (2007) Changes in Atmospheric Constituents and in Radiative Forcing. IN<br />

SOLOMON, S., QIN, D., MANNING, M., CHEN, Z., MARQUIS, M., AVERYT, K. B., TIGNOR, M. &<br />

MILLER, H. L. (Eds.) Climate Change 2007: The Physical Science Basis. Contribution <strong>of</strong> Working<br />

Group I to the Fourth Assessment Report <strong>of</strong> the Intergovernmental Panel on Climate Change. Cambridge,<br />

United Kingdom and New York, NY, USA, Cambridge University Press.<br />

HAMILTON, J. G., DELUCIA, E. H., GEORGE, K., NAIDU, S. L., FINZI, A. C. & SCHLESINGER, W. H.<br />

(2002) Forest carbon balance under elevated CO2. Oecologia, 131, 250-260.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 278<br />

HANSON, P. J., EDWARDS, N. T., GARTEN, C. T. & ANDREWS, J. A. (2000) Separating root and soil<br />

microbial contributions to soil respiration: A review <strong>of</strong> methods and observations. Biogeochemistry, 48,<br />

115-146.<br />

HEINEMEYER, A., HARTLEY, I. P., EVANS, S. P., CARREIRA DE LA FUENTE, J. A. & INESON, P.<br />

(2007) Forest soil CO2 flux: uncovering the contribution and environmental responses <strong>of</strong><br />

ectomycorrhizas. Global Change Biology, 13, 1786-1797.<br />

HENDREY, G. R., ELLSWORTH, D. S., LEWIN, K. F. & NAGY, J. (1999) A free-air enrichment system for<br />

exposing tall forest vegetation to elevated atmospheric CO2. Global Change Biology, 5, 293-309.<br />

HOGBERG, P., NORDGREN, A., BUCHMANN, N., TAYLOR, A. F. S., EKBLAD, A., HOGBERG, M. N.,<br />

NYBERG, G., OTTOSSON-LOFVENIUS, M. & READ, D. J. (2001) Large-scale forest girdling shows<br />

that current photosynthesis drives soil respiration. Nature, 411, 789-792.<br />

JANSSENS, I. A., LANKREIJER, H., MATTEUCCI, G., KOWALSKI, A. S., BUCHMANN, N., EPRON, D.,<br />

PILEGAARD, K., KUTSCH, W., LONGDOZ, B., GRUNWALD, T., MONTAGNANI, L., DORE, S.,<br />

REBMANN, C., MOORS, E. J., GRELLE, A., RANNIK, U., MORGENSTERN, K., OLTCHEV, S.,<br />

CLEMENT, R., GUDMUNDSSON, J., MINERBI, S., BERBIGIER, P., IBROM, A., MONCRIEFF, J.,<br />

AUBINET, M., BERNHOFER, C., JENSEN, N. O., VESALA, T., GRANIER, A., SCHULZE, E. D.,<br />

LINDROTH, A., DOLMAN, A. J., JARVIS, P. G., CEULEMANS, R. & VALENTINI, R. (2001)<br />

Productivity overshadows temperature in determining soil and ecosystem respiration across European<br />

forests. Global Change Biology, 7, 269-278.<br />

JARVIS, P., REY, A., PETSIKOS, C., WINGATE, L., RAYMENT, M., PEREIRA, J., BANZA, J., DAVID, J.,<br />

MIGLIETTA, F., BORGHETTI, M., MANCA, G. & VALENTINI, R. (2007) Drying and wetting <strong>of</strong><br />

Mediterranean soils stimulates decomposition and carbon dioxide emission: the "Birch effect". Heron<br />

Publishing.<br />

JASSAL, R. S., BLACK, T. A., NOVAK, M. D., GAUMONT-GUAY, D. & NESIC, Z. (2008) Effect <strong>of</strong> soil<br />

water stress on soil respiration and its temperature sensitivity in an 18-year-old temperate Douglas-fir<br />

stand. Global Change Biology, 14, 1305-1318.<br />

KASURINEN, A., KOKKO-GONZALES, P., RIIKONEN, J., VAPAAVUORI, E. & HOLOPAINEN, T. (2004)<br />

Soil CO2 efflux <strong>of</strong> two silver birch clones exposed to elevated CO2 and O-3 levels during three growing<br />

seasons. Global Change Biology, 10, 1654-1665.<br />

KEITH, H., JACOBSEN, K. L. & RAISON, R. J. (1997) Effects <strong>of</strong> soil phosphorus availability, temperature and<br />

moisture on soil respiration in Eucalyptus pauciflora forest. Plant and Soil, 190, 127-141.<br />

KELLOMAKI, S., WANG, K. Y. & LEMETTINEN, M. (2000) Controlled environment chambers for<br />

investigating tree response to elevated CO2 and temperature under boreal conditions. Photosynthetica, 38,<br />

69-81.<br />

KING, J. S., HANSON, P. J., BERNHARDT, E., DEANGELIS, P., NORBY, R. J. & PREGITZER, K. S.<br />

(2004) A multiyear synthesis <strong>of</strong> soil respiration responses to elevated atmospheric CO2 from four forest<br />

FACE experiments. Global Change Biology, 10, 1027-1042.<br />

KUZYAKOV, Y. (2006) Sources <strong>of</strong> CO2 efflux from soil and review <strong>of</strong> partitioning methods. Soil Biology &<br />

Biochemistry, 38, 425-448.<br />

LEADLEY, P. W. & DRAKE, B. G. (1993) Open top chamber for exposing plant canopy to elevated CO2<br />

concentration and for measuring net gas-exchange Vegetatio, 104, 3-15.<br />

LEWIN, K. F., HENDREY, G. R., NAGY, J. & LAMORTE, R. L. (1994) Design and application <strong>of</strong> a free air<br />

carbon-dioxide enrichment facility. Agricultural and Forest Meteorology, 70, 15-29.<br />

LIN, G., RYGIEWICZ, P. T., EHLERINGER, J. R., JOHNSON, M. G. & TINGEY, D. T. (2001) Timedependent<br />

responses <strong>of</strong> soil CO2 efflux components to elevated atmospheric [CO2] and temperature in<br />

experimental forest mesocosms. Plant and Soil, 229, 259-270.<br />

MANDL, R. H., WEINSTEIN, L. H., MCCUNE, D. C. & KEVENY, M. (1973) A Cylindrical, Open-Top<br />

Chamber for the Exposure <strong>of</strong> Plants to Air Pollutants in the Field. J Environ Qual, 2, 371-376.<br />

MEDHURST, J., PARSBY, J., LINDER, S., WALLIN, G., CESCHIA, E. & SLANEY, M. (2006) A whole-tree<br />

chamber system for examining tree-level physiological responses <strong>of</strong> field-grown trees to environmental<br />

variation and climate change. Plant Cell and Environment, 29, 1853-1869.<br />

MUHR, J., GOLDBERG, S. D., BORKEN, W. & GEBAUER, G. (2008) Repeated drying-rewetting cycles and<br />

their effects on the emission <strong>of</strong> CO2, N2O, NO, and CH4 in a forest soil. Wiley-V C H Verlag Gmbh.<br />

NIINISTO, S. M., SILVOLA, J. & KELLOMAKI, S. (2004) Soil CO2 efflux in a boreal pine forest under<br />

atmospheric CO2 enrichment and air warming. Global Change Biology, 10, 1363-1376.<br />

OBERBAUER, S. F., OECHEL, W. C. & RIECHERS, G. H. (1986) Soil respiration <strong>of</strong> Alaskan toundra at<br />

elevated atmospheric concentration. Plant and Soil, 96, 145-148.<br />

ORCHARD, V. A. & COOK, F. J. (1983) Relationship between soil respiration and soil moisture. Soil Biology<br />

& Biochemistry, 15, 447-453.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 279<br />

PAJARI, B. (1995) Soil respiration in a poor upland site <strong>of</strong> Scots pine stand subjected to elevated temperatures<br />

and atmospheric carbon concentration. Plant and Soil, 168-169, 563-570.<br />

PREGITZER, K., LOYA, W., KUBISKE, M. & ZAK, D. (2006) Soil respiration in northern forests exposed to<br />

elevated atmospheric carbon dioxide and ozone. Oecologia, 148, 503-516.<br />

RAICH, J. W. & POTTER, C. S. (1995) Global patterns <strong>of</strong> carbon-dioxide emissions from soils. Global<br />

Biogeochemical Cycles, 9, 23-36.<br />

RUSTAD, RUSTAD, L., CAMPBELL, CAMPBELL, J., MARION, MARION, G., NORBY, NORBY, R.,<br />

MITCHELL, MITCHELL, M., HARTLEY, HARTLEY, A., CORNELISSEN, CORNELISSEN, J.,<br />

GUREVITCH, GUREVITCH, J., GCTE, N. & GCTE, N. (2001) A meta-analysis <strong>of</strong> the response <strong>of</strong> soil<br />

respiration, net nitrogen mineralization, and aboveground plant growth to experimental ecosystem<br />

warming. Oecologia, 126, 543-562.<br />

STAPE, J. L., BINKLEY, D. & RYAN, M. G. (2008) Production and carbon allocation in a clonal Eucalyptus<br />

plantation with water and nutrient manipulations. Forest Ecology and Management, 255, 920-930.<br />

TANG, X. L., ZHOU, G. Y., LIU, S. G., ZHANG, D. Q., LIU, S. Z., LI, J. & ZHOU, C. Y. (2006) Dependence<br />

<strong>of</strong> soil respiration on soil temperature and soil moisture in successional forests in southern China. Journal<br />

<strong>of</strong> Integrative Plant Biology, 48, 654-663.<br />

TINGEY, D. T., LEE, E. H., WASCHMANN, R., JOHNSON, M. G. & RYGIEWICZ, P. T. (2006) Does soil<br />

CO2 efflux acclimatize to elevated temperature and CO2 during long-term treatment <strong>of</strong> Douglas-fir<br />

seedlings? New Phytologist, 170, 107-118.<br />

WAN, S., NORBY, R. J., LEDFORD, J. & WELTZIN, J. F. (2007) Responses <strong>of</strong> soil respiration to elevated<br />

CO2, air warming, and changing soil water availability in a model old-field grassland. Global Change<br />

Biology, 13, 2411-2424.<br />

ZAK, D. R., PREGITZER, K. S., KING, J. S. & HOLMES, W. E. (2000) Elevated atmospheric CO2, fine roots<br />

and the response <strong>of</strong> soil microorganisms: a review and hypothesis. New Phytologist, 147, 201-222.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 280<br />

BUSHFIRE GOVERNANCE IN AUSTRALIA IN 2009:<br />

A NOTE TO A FUTURE HISTORIAN<br />

Roger Underwood 1<br />

ABSTRACT<br />

Historians at the end <strong>of</strong> the 21st century will look back with bewilderment to the<br />

repeated bushfire disasters in <strong>Australia</strong> during the period 1983-2009. They will be<br />

perplexed as to why an educated, civilised, forest-loving, prosperous and science-based<br />

community adopted a forest management system that made effective bushfire<br />

management impossible. This paper will provide one perspective. I focus on the<br />

deficiencies in public policy, the lack <strong>of</strong> leadership and accountability, the capacity for<br />

non-accountable pressure groups to influence political decision-making, and the<br />

transfer <strong>of</strong> bushfire responsibilities from land management to emergency services<br />

agencies; in short the failure <strong>of</strong> governance at all levels.<br />

INTRODUCTION<br />

This is a note to a future bushfire historian. I am writing it because I feel sure he or she will look back<br />

on the late 20 th and early 21st century with bewilderment. How did it happen, the question will be<br />

asked, that the <strong>Australia</strong>ns <strong>of</strong> those times developed an excellent bushfire management system..... and<br />

then discarded it, thus needlessly exposing themselves to death and destruction by fire.<br />

The answer? Read on.<br />

A POTTED HISTORY<br />

Serious bushfires in <strong>Australia</strong>n forests are nothing new. Their occurrence goes back 200 years or so to<br />

the early 19 th century, when European settlers rejected the bushfire management approach <strong>of</strong><br />

Aboriginal <strong>Australia</strong>ns and replaced it with an imported European approach.<br />

There were a few decades <strong>of</strong> transition, and there were even a few European <strong>Australia</strong>ns who could<br />

see that the Aboriginal approach made sense. But this didn’t last. The insertion <strong>of</strong> fire vulnerable<br />

townships, homesteads, crops, fencing and stock into bushland designed by nature to burn regularly,<br />

and then the failure to burn this bush regularly, did not seem at the time to be a particularly silly idea.<br />

Even when it produced the inevitable result: bushfire disasters. They became milestones in our rural<br />

history, with successive disasters named after days <strong>of</strong> the week. Black Friday, Ash Wednesday,<br />

Incinerated Tuesday and so on.<br />

The Empire fought back. Harking back to their British heritage, the new <strong>Australia</strong>ns remembered the<br />

concept <strong>of</strong> the fire brigade, and introduced it to their new land. Bushfire brigades, progressively better<br />

organised and resourced over the years, at first voluntary but increasingly paid, came into being. Vast<br />

paramilitary organisations were developed. Their efforts were underpinned by European ecophilosophy:<br />

bushfires were not inevitable, but were an anomaly. The default position for the<br />

<strong>Australia</strong>n bush was one in which no fires occurred. It was the bushfire equivalent <strong>of</strong> Terra Nullius.<br />

Those espousing this philosophy believed in an ideal world, forests without fire. They convinced<br />

themselves that if a sufficiently well-trained and well-equipped force <strong>of</strong> fire-fighters could be set up,<br />

1 Roger Underwood is Chairman <strong>of</strong> the Bushfire Front Inc. He is a former district and regional forester and senior land<br />

management administrator in Western <strong>Australia</strong> and a Fellow <strong>of</strong> the <strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong>. He was the recipient <strong>of</strong> the<br />

<strong>Institute</strong>'s NW Jolly Medal in 2008. Email: yorkgum@westnet.com.au


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 281<br />

the ideal could be achieved. All fires would be pounced upon the moment they started, and<br />

extinguished. Serious fires would become a thing <strong>of</strong> the past. The war on fire was declared.<br />

Early <strong>Australia</strong>n foresters were firm devotees <strong>of</strong> this philosophy, indeed they provided the scientific<br />

underpinning, based on their knowledge <strong>of</strong> northern hemisphere temperate hardwood forests. They<br />

saw fire as an ecological disaster. In this view they were encouraged to observe that the US Forest<br />

Service had adopted the very same philosophy. Wildfire was also to be expunged from North<br />

American forests, providing an example to the rest <strong>of</strong> the world.<br />

There was only one problem with this philosophy: it didn’t work. It certainly didn’t work in the USA.<br />

It especially didn’t work in <strong>Australia</strong> where fire fighting resources were so limited and our eucalypt<br />

forests were so fire-loving. Over time, keen observers started to spot the trend: the longer fires were<br />

excluded from the bush, the more fire fuels accumulated. The heavier the fuels became, the more<br />

fierce the inevitable fire. And the fiercer the fire, the less able were fire-fighters to suppress it and the<br />

more damage was done.<br />

No-one much remembers the 1950s these days, but they were a decade <strong>of</strong> appalling forest fires all<br />

over the country, especially in my home state <strong>of</strong> Western <strong>Australia</strong>, where they brought to an end<br />

three decades <strong>of</strong> attempted fire exclusion in the jarrah forest. And just as happened in Victoria a halfcentury<br />

later, it all culminated in disaster. Towns were burnt, hundreds <strong>of</strong> thousands <strong>of</strong> hectares <strong>of</strong><br />

forest and farmland were severely burned, and there was a Royal Commission.<br />

THE PENNY DROPS<br />

The decades <strong>of</strong> attempted fire exclusion and the subsequent fire disasters in the 1950s and 1960s<br />

signified a new era in <strong>Australia</strong>n fire management. The penny finally dropped. <strong>Australia</strong> is not<br />

Europe. It is not even the USA. A new generation <strong>of</strong> forest ecologists who studied the bush rather<br />

than English and French text books realised that eucalypt forests actually thrive under a regime <strong>of</strong><br />

frequent mild fire. They were pleasantly surprised when anthropologists pointed out that the<br />

Aborigines had understood this for thousands <strong>of</strong> years. For the first time, foresters in the districts were<br />

encouraged to start putting into practice the practical wisdom they had been accumulating during the<br />

bad years. They well understood that in forests which are periodically burned under mild conditions,<br />

fuel levels remain low, and fire suppression is easier and safer. With a prescribed burning program,<br />

fires could not be eliminated, but the number, and the damage caused by large high intensity forest<br />

fires declined significantly. What was better, the frequently burned forest was healthier and more<br />

beautiful and better able to sustain the whole range <strong>of</strong> forest values.<br />

Later, practical bush wisdom was supported by practical fire research. These put numbers to fire<br />

behaviour, and enabled rates <strong>of</strong> spread and intensity to be predicted, which in turn allowed the<br />

development <strong>of</strong> prescribed burning guides. The innovation <strong>of</strong> aerial burning revolutionised prescribed<br />

burning operations, as maximum use could be made <strong>of</strong> optimum conditions.<br />

For the first time since the Aborigines lost control <strong>of</strong> the bush, <strong>Australia</strong>ns had a home-grown fire<br />

management system that was ecologically sound and highly effective in minimising the occurrence <strong>of</strong><br />

landscape-level high-intensity fires. The new philosophy drew on medical rather than military<br />

science. “Prevention is better than cure” became the catchcry; fuel reduction by prescribed burning<br />

was regarded as immunising the forest against a wildfire disease epidemic. People came from all over<br />

the world to see what we were doing, and were filled with admiration....even the Americans.<br />

THE PENNY IS PICKED UP AGAIN<br />

The trouble with successful bushfire management is that it is its own worst enemy. Success breeds<br />

apathy and apathy breeds mismanagement. Mismanagement leads to system degrade. And in the case<br />

<strong>of</strong> bushfire management, system degrade leads to serious bushfires. This is exactly the situation<br />

witnessed right across forested <strong>Australia</strong> over the last 15 years.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 282<br />

There is a theory to explain all this, called Chaos Theory. Systems always come apart once they lose<br />

energy. The admirable fire management systems developed for <strong>Australia</strong>n forests in the 1960s and<br />

1970s, and evolving successfully through the 1970s and 1980s, suddenly had the energy sucked from<br />

them, and in the 1990s, they came to pieces.<br />

There were several reasons for this, but the fundamental problem was incompetent government. The<br />

1990s was the time in <strong>Australia</strong>n history when urban pressure groups took over policy-making in land<br />

management. It was the time when academics who were still wedded to the idea that <strong>Australia</strong> was<br />

Europe, began to wield serious influence over political processes and bureaucrats. It was the time<br />

when people who had not started at the bottom took over at the top, when for the first time our land<br />

management agencies were being led by people with nil bushfire experience. In the blink <strong>of</strong> an eye<br />

the system flip-flopped; the <strong>Australia</strong>n philosophy was thrown out the window and the discredited<br />

European philosophy and its American counterpart were reinstated.<br />

The result was that all over forested <strong>Australia</strong>, land managers began to turn away from prescribed<br />

burning and to look to the fire brigades as their bushfire panacea. The tired old idea that “all fires are<br />

bad” and must be immediately extinguished again took hold. The intelligentsia emerged from their<br />

leafy campuses to preach a message about the irreparable ecological damage caused by periodic mild<br />

fire. Prescribed burning, it was said, was leading to fauna extinctions, to dieback, to soil infertility, to<br />

weed infestations, to killer smoke hazes, even (horror <strong>of</strong> horrors) to tainted wine. It was the late 19 th<br />

and early 20 th century all over again, the only difference being that now we had jazzier fire fighting<br />

equipment and our fire chiefs had sexier uniforms.<br />

But, lo and behold! The fire brigade approach still did not work! Not only did it fail to eliminate large<br />

intense bushfires, their number increased. Luckily there was an excuse: unstoppable bushfires were<br />

the result <strong>of</strong> global warming, <strong>of</strong> Mother Nature counter-attacking!<br />

I think not. The decline in bushfire management in <strong>Australia</strong> in the last 15 years has a much less<br />

palatable explanation: the incompetence <strong>of</strong> our governance, and the failure <strong>of</strong> our leadership. This is<br />

readily demonstrated.<br />

ANALYSING GOVERNANCE – A CASE STUDY<br />

During 2003, my colleagues and I in The Bushfire Front developed a model for Best Practice in<br />

Bushfire Governance for a jurisdiction like WA. The model set out ten basic requirements which a<br />

government could adopt either as goals during a period <strong>of</strong> system development, or as performance<br />

measures against which progress and achievements could later be gauged.<br />

The ten Best Practice requirements are:<br />

1. Western <strong>Australia</strong> has a State Bushfire Policy providing leadership, guidance and overarching<br />

direction to all government agencies, land managers and local government.<br />

2. There is clear accountability for bushfire management and for fire outcomes at Ministerial<br />

level, and within the land management and emergency services bureaucracies.<br />

3. There is a unified and consistent approach to prevention, preparedness, damage mitigation,<br />

suppression, fire recovery and community education across government and local<br />

government. This approach is spelled out in writing and signed <strong>of</strong>f by the government, and<br />

assigned to a single Minister for implementation and review.<br />

4. Land management, bushfire and emergency services legislation is up-to-date and assigns<br />

accountability, responsibility, priorities, standards and an efficient hierarchy <strong>of</strong><br />

administration, leadership and control to Ministers and to agencies.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 283<br />

5. Bushfire management on crown lands is the responsibility <strong>of</strong> a competent land management<br />

agency, not the emergency service.<br />

6. The land management agency has a published Bushfire Strategy that focuses on bushfire<br />

prevention and damage mitigation as well as suppression, incorporates an effective program<br />

<strong>of</strong> fuel reduction burning, and is supported by ongoing appropriate research.<br />

7. Performance outcomes, standards and targets for bushfire management on crown lands<br />

(directed at minimising the number and size <strong>of</strong> high intensity wildfires) have been developed<br />

and published.<br />

8. There is a public report at the end <strong>of</strong> every fire season based on independent monitoring <strong>of</strong><br />

bushfire outcomes compared with performance standards and targets, and listing towns and<br />

settlements at the urban/bushland interface which are most at risk from bushfires in the<br />

coming summer due to proximity <strong>of</strong> heavy long-unburnt fuels.<br />

9. The government has developed and is implementing a pr<strong>of</strong>essional communications strategy<br />

to educate the community (including schoolchildren) about fire science, fire preparedness,<br />

damage mitigation and action in the event <strong>of</strong> a fire.<br />

10. The government has empowered rural and semi-rural communities to practice effective<br />

bushfire management, and has funded the placement <strong>of</strong> pr<strong>of</strong>essional fire management<br />

specialists within local government authorities.<br />

We found that the Western <strong>Australia</strong>n government failed utterly (in fact scored zero) on points 1, 2, 3,<br />

4, 7, 8, 9 and 10. The most critical failures were:<br />

• Policy. WA has no overarching Bushfire Policy. Instead individual agencies and local<br />

governments each have their own policies, developed independently and usually in conflict<br />

with each other;<br />

• Accountability. There is no single Minister in the government with responsibility for fire and<br />

who can be held accountable for bushfire outcomes. Instead we found six Ministers with a<br />

finger in the bushfire pie, and accountability was so diffuse that no one Minister could ever<br />

be held accountable for any outcome.<br />

• Agency mismanagement. We found that within the Department <strong>of</strong> Environment and<br />

Conservation (DEC), the agency responsible for management <strong>of</strong> national parks, regional<br />

parks, State forests, nature reserves and Unallocated Crown Land in WA, fire management<br />

did not have the status <strong>of</strong> “core function”.<br />

Our protocol was rejected out <strong>of</strong> hand by the government. Apparently, bushfires were not a problem.<br />

We were told to mind our own business, that bushfire management was in good hands, and that they<br />

were very satisfied. Every subsequent attempt we made to get the government to focus on system<br />

deficiencies was thwarted by grossly over-confident and poorly informed Ministers, no doubt guided<br />

by their political advisers and public servants.<br />

The change in this attitude after the Victorian disaster in February this year was dramatic. All <strong>of</strong> a<br />

sudden we detected an awful realisation in the corridors <strong>of</strong> power that if a similar disaster occurred in<br />

WA they could not say they had not been warned.<br />

WHAT ARE THE PROSPECTS FOR CHANGE?<br />

The ghastly Victorian disaster has had one positive outcome. Almost everyone now acknowledges<br />

that the current approach to bushfire management in <strong>Australia</strong> is a failure. Putting all the eggs in the<br />

suppression basket only works for small fires under mild conditions, i.e., the sort <strong>of</strong> fires that do little<br />

damage anyway.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 284<br />

Realising that there is a problem is the first step towards solving it. I am encouraged to think that most<br />

sensible people have now moved past the point <strong>of</strong> denial. I detect that many politicians now accept<br />

that they have been seriously misled over the years – misled by their political advisers, misled by the<br />

environmentalists and the whining academics, even misled by their fire chiefs. If this process <strong>of</strong><br />

facing up to reality survives subsequent elections, it will mean that the lives and assets lost in the<br />

Victorian fires will not have been completely in vain.<br />

However, I can see no prospect for change unless <strong>Australia</strong>n governments start to govern. This will<br />

require four things:<br />

1. Governments must reject and reverse the concept <strong>of</strong> placing responsibility for bushfire<br />

management with the emergency services rather than the land managers. Fire management<br />

must become holistic and focus on prevention and damage mitigation, not just suppression.<br />

2. Each jurisdiction must design and face up to an appropriate Best Practice protocol, and get on<br />

with the job <strong>of</strong> building an effective fire management system, starting with policy and ending<br />

with monitoring and reporting <strong>of</strong> outcomes.<br />

3. There can be no more gutless hiding behind the pathetic excuse <strong>of</strong> “global warming”. If the<br />

temperature does go up a degree or two, then let’s be ready for it, not use it to excuse<br />

incompetent leadership and outdated philosophies.<br />

4. Significant cultural change must be induced within the senior ranks <strong>of</strong> our land management<br />

agencies. The people in charge <strong>of</strong> our national parks and State forests must be made to realise<br />

that unless they manage fire, they cannot manage for anything else, and that attempts to<br />

manage fire by the fire brigade approach will always fail. They must summon up the courage<br />

to reject the false prophets from academia.<br />

None <strong>of</strong> this will happen under the current system <strong>of</strong> diffuse Ministerial responsibility and<br />

accountability. Each state needs a Bushfire Supremo, one person clearly running the show, whose job<br />

depends on performance — the degree to which effective fire management systems are designed,<br />

implemented and energised. This person must sit outside and above both the land management<br />

agencies and the emergency services, and report directly to Parliament.<br />

So, dear historian <strong>of</strong> the future, this is what happened back in the years 1983-2009. Basically we<br />

stuffed things up. But not so badly that some <strong>of</strong> us did survive, and we still have an opportunity to put<br />

things right. We do not see ourselves as “powerless in the face <strong>of</strong> unstoppable fires caused by global<br />

warming” but as intelligent people who love our forests, value our community assets and recognise<br />

our duty <strong>of</strong> care to our fellow-humans. What’s more, we actually know already what has to be done!<br />

If the technology <strong>of</strong> 2109 allows it, get back to me with your review, will you? I’m keen to know how<br />

we go.<br />

September 2009


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 285<br />

ABSTRACT<br />

NATIONAL FIRE BEHAVIOUR PREDICTION SYSTEM<br />

Miguel Cruz 1,2 , Jim Gould 1,2<br />

The estimation <strong>of</strong> fire behaviour is an important component <strong>of</strong> any fire management<br />

approach, allowing the determination <strong>of</strong> the impacts <strong>of</strong> fire on ecosystem components and<br />

supporting forest fire management decision-making. Fire behaviour prediction combines<br />

quantitative and qualitative information based on experience and scientific principles <strong>of</strong><br />

describing the combustion and behaviour <strong>of</strong> fire influence by topography, weather and<br />

fuel. Predictions are based on mathematical models that integrate important factors in a<br />

consistent way. The National Fire Behaviour Prediction (NFBP) system will consist <strong>of</strong><br />

four primary components (fuel models, fuel moisture models, wind models, and fire<br />

behaviour models) to predict fire characteristics (e.g., rate <strong>of</strong> spread, flame height, fireline<br />

intensity, onset <strong>of</strong> crowning spotting potential, etc). This paper will focus on the fire<br />

behaviour component <strong>of</strong> the NFBP system. This component integrates a suite <strong>of</strong> models<br />

covering the main fuel types <strong>of</strong> <strong>Australia</strong>, eucalyptus forests, exotic pine plantations,<br />

grasslands, shrublands and Mallee-heath.<br />

The desired accomplishments <strong>of</strong> the proposed National Fire Behaviour Prediction<br />

Systems is to provide fire managers with better operating models to implement prescribed<br />

burning programs, suppression resources, risk and biodiversity management programs.<br />

The fuel type specificity <strong>of</strong> the fire models, its greater accuracy and updated calculation<br />

methods allow also for more accurate simulations <strong>of</strong> the impact <strong>of</strong> hypothetical climate<br />

change scenarios on fire potential and risk in <strong>Australia</strong>.<br />

INTRODUCTION<br />

Recent extreme bushfire seasons have increased the focus on how rural fire and land management<br />

activities determine landscape level fuel dynamics, fire regimes, bushfire risk and bushland urban<br />

planning. There is need to improve ecosystems health and to reduce the likelihood <strong>of</strong> catastrophic<br />

fires. Recommendations from the Council <strong>of</strong> <strong>Australia</strong>n Governments (COAG) National Inquiry on<br />

Bushfire Mitigation and Management (Ellis et al. 2004) that the provision <strong>of</strong> additional resources<br />

jointly by the <strong>Australia</strong>n Government and the state and territory governments to accelerate the research<br />

necessary for the characterisation <strong>of</strong> fuel loads and dynamics for <strong>Australia</strong>n ecosystems (both natural<br />

and exotic), the characterisation <strong>of</strong> fire behaviour and ecological responses, the development <strong>of</strong><br />

‘burning guides’ from this information, and the compilation <strong>of</strong> this information and knowledge in<br />

nationally accessible databases.<br />

Current weather patterns, society perception <strong>of</strong> fire, and land management agencies policies and<br />

constrains have increased the complexity and magnitude <strong>of</strong> the wildland fire problem. Information is<br />

the key in sound decision making to mitigate detrimental effects <strong>of</strong> wildland fire. The ability to predict<br />

and comprehend fire behaviour in relation to its drivers is fundamental to a safe, effective and<br />

ecosystem enhancing fire management decision-making. Fire behaviour, “the manner in which fuel<br />

ignites, flame develops, and fire spread and exhibits other related phenomena as determined by the<br />

interaction <strong>of</strong> fuels, weather and topography” (Merrill and Alexander 1987) is a vital component on<br />

the decision making process related to prescribed fire use planning and execution, dispatching, firefighting<br />

safety, definition <strong>of</strong> bushfire suppression strategies and tactics, evaluating fuel treatments<br />

effectiveness and recurrence.<br />

1 Bushfire Dynamics and Applications, CSIRO Sustainable Ecosystems and CSIRO Climate Adaptation Flagship,<br />

GPO Box 284, Canberra, ACT 2601, <strong>Australia</strong>.<br />

2 Bushfire Cooperative Research Centre, East Melbourne, VIC, 3002, <strong>Australia</strong>.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 286<br />

A review <strong>of</strong> the pertinent fire behaviour research carried out in <strong>Australia</strong> in the past 30 years (e.g.,<br />

Cheney et al 1992, Burrows 1994, Marsden-Smedley and Catchpole 1995, McCaw 1997, Gould et al.<br />

2007, Cheney and Sullivan 2008, Burrows et al. 2009) and the tools currently used by fire managers<br />

reveals a large disconnection between the knowledge available and the knowledge used.<br />

This is partially due to the lack <strong>of</strong> effective technology-transfer programs and a considerable<br />

technology gap regarding the format decision support tools are available to users. A lack <strong>of</strong><br />

sophistication is manifest in the broad reliance on slide rules introduced by McArthur (1966, 1967)<br />

and tables (Sneeuwjagt and Peet 1985) to support fire management decision-making. It is paradoxical<br />

that although the cost <strong>of</strong> fire fighting, our understanding <strong>of</strong> fire phenomena and the complexity <strong>of</strong> the<br />

fire manager job increased significantly through time, the tools that are used to predict fire behaviour<br />

and support decision making at any temporal and spatial scale are based in technology introduced in<br />

the 1950’s.<br />

The use <strong>of</strong> slide rules, tables and other field guides have a role in producing first approximations <strong>of</strong><br />

fire behaviour characteristics and intended mainly to be used as a field reference guides. Nonetheless,<br />

precise fire behaviour predictions, combining spatially explicit fuel and topography information with<br />

detailed weather forecast, are best made using computing-based representation <strong>of</strong> fire behaviour<br />

models. The integration <strong>of</strong> new technologies, such as, Geographical Information Systems, satellite<br />

information, large computing power, internet and handheld devices, to harness available datasets and<br />

conduct fire behaviour simulations, either at the local or landscape level, is episodic, and unfortunately<br />

not the norm.<br />

Furthermore, the overly reliance on the old methods to predict fire behaviour results in a lack <strong>of</strong><br />

integration <strong>of</strong> recent research on fuel moisture (e.g., Matthews et al. 2007), fuel dynamics (Gould et al<br />

2007) and wind speed (Sullivan and Knight 2001), decreasing the certainty <strong>of</strong> fire predictions.<br />

The situation in <strong>Australia</strong> contrasts with the evolution <strong>of</strong> fire management decision support systems in<br />

the US and Canada. It is well accepted that in the 1960’s these three countries had the most advanced<br />

fire behaviour research programs in the world. At that time the technologies used by fire practitioners<br />

were comparable, although the development <strong>of</strong> the grassland and forest fire danger meters (McArthur<br />

1966, 1967) gave <strong>Australia</strong>n fire managers a tool still inexistent to others in North America, the<br />

capacity to predict rate <strong>of</strong> spread and intensity from information on weather and fuel characteristics.<br />

The advent <strong>of</strong> reliable fire behaviour models in both the USA and Canada, in the mid-seventies and<br />

late eighties respectively, was accompanied by the development <strong>of</strong> computer s<strong>of</strong>tware <strong>of</strong> various<br />

levels <strong>of</strong> complexity that allow fire managers to simulate fire propagation at a range <strong>of</strong> spatial and<br />

temporal scales, e.g., BEHAVE, FARSITE, PROMETHEUS, FSPro (Andrews and Finney 2007,<br />

Tymstra et al. 2006). These systems not only form the backbone <strong>of</strong> effective fire management<br />

programs in these countries but are also extensively used in research applications.<br />

In <strong>Australia</strong> there have been attempts to provide users with computer-based simulation systems<br />

(Colman and Sullivan 1996) although they seemed to not have been adopted by the fire management<br />

community.<br />

CONCEPT<br />

This paper presents a proposal for a National Fire Behaviour Prediction (NFBP) system aimed at<br />

addressing the needs <strong>of</strong> fire managers in regards to fire behaviour information, namely quantifying<br />

fuel hazard, assessing fire danger and predicting site specific fire behaviour. The purpose <strong>of</strong> the<br />

system is to integrate all available and peer-reviewed fire behaviour knowledge into a set <strong>of</strong> userfriendly<br />

tools that can be used by fire managers to deal with fire management issues over a range <strong>of</strong><br />

spatial and temporal scales, and complexity levels. The availability <strong>of</strong> such a system will improve fuel<br />

management programs, lead to more effective and safe firefighting, enhance protection <strong>of</strong> rural<br />

communities and reduce detrimental effects to natural resources.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 287<br />

Figure 1. Flow diagram <strong>of</strong> the National Fire Behaviour Prediction System (NFBPS) illustrating<br />

the links between fire drivers, the knowledge base (models and data) and output<br />

systems.<br />

The heart <strong>of</strong> the model system is composed by two distinct components (Figure 1), the core fire<br />

models module and a fire behaviour knowledge base module. The core fire models module is the<br />

engine <strong>of</strong> the system, providing simulations <strong>of</strong> fire behaviour aimed at respond to a large array <strong>of</strong> fire<br />

management questions. Main physical fire behaviour quantities incorporated in the model are:<br />

• Sustainability <strong>of</strong> fire spread<br />

• Rate <strong>of</strong> surface fire spread and intensity;<br />

• Flame dimensions (height, depth and angle) and residence time<br />

• Spotting potential<br />

• Onset <strong>of</strong> crowning and crown fire behaviour<br />

• Coarse woody fuel consumption<br />

• Initial fire development potential (area and perimeter assuming point source ignition)


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 288<br />

Along with the above described fire behaviour models, an assortment <strong>of</strong> sub-models will provide<br />

accessory information related to first order fire effects:<br />

• Burn patchiness;<br />

• Scorch height;<br />

• Emissions;<br />

Figure 2. Prediction <strong>of</strong> Cheney et al (1998) natural pasture fire-spread model as a function <strong>of</strong><br />

wind speed against experimental data. Data has been normalized for the effect <strong>of</strong> fuel<br />

moisture. The 68 and 95% prediction intervals are shown. Figure from Cheney et al.<br />

(1998).<br />

The fire behaviour knowledge base module provides a setting where fire behaviour simulations can be<br />

compared with real world data. This module will incorporate all available fire behaviour data in<br />

<strong>Australia</strong>n ecosystems and will allow users to visualize their simulations within the context <strong>of</strong> data,<br />

illustrating the uncertainty in the predictions and the limitations <strong>of</strong> the datasets. A grassfire rate <strong>of</strong><br />

spread model (Cheney et al 1998) output as a function <strong>of</strong> wind speed (data corrected for fuel moisture<br />

effect) provides an example <strong>of</strong> such concept (Figure 2). The figure not only provides information on<br />

how rate <strong>of</strong> spread increases approximately linearly with wind speed, but also indicates the variability<br />

in observed fire spread.<br />

For the area where the bulk <strong>of</strong> the data exists (wind speeds between 15 and 25 km/h) the rates <strong>of</strong><br />

spread seem to spread +/-33% <strong>of</strong> the predicted value, with a few cases showing extreme variability.<br />

The graph also provides the bounds <strong>of</strong> the dataset used to build models, leading to caution by users<br />

when their simulations are required for conditions not covered by the existent datasets. The system<br />

considers that users can input their own datasets (collected in prescribed burns or wildfires) to verify<br />

simulations and strength the predictive capacity <strong>of</strong> the system. This analysis will provide users with<br />

an understanding <strong>of</strong> the predictive capacity <strong>of</strong> models, something that is only subconsciously acquired<br />

after a long career at predicting fire behaviour.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 289<br />

MODELLING UNCERTAINTY<br />

The operational prediction <strong>of</strong> fire spread to support fire management operations relies on a<br />

deterministic approach where a single “best-guess” forecast is produced from the best estimate <strong>of</strong> the<br />

environmental conditions driving the fire. Although fire can be considered a phenomenon <strong>of</strong> low<br />

predictability and the estimation <strong>of</strong> input conditions for fire behaviour models are fraught with<br />

uncertainty, no error component is associated with these predictions. The NFBPS will incorporate<br />

modelling techniques that allow describing the inherent uncertainty in fire behaviour into the<br />

simulation process. Ensemble methods that consider the variability <strong>of</strong> model inputs and Monte Carlo<br />

sampling (Cruz, in press) will be integrated in the core fire modelling component. This will allow to<br />

describe prediction statistics and estimate the likelihood that extreme phenomena, e.g., onset <strong>of</strong><br />

crowning, will occur. These outputs allow users to ascertain with some confidence what will and what<br />

will not happen. The probabilistic outputs from the ensemble method extend the range <strong>of</strong> questions<br />

that can be answered by fire behaviour models, enabling linkages between fire behaviour models and<br />

quantitative risk analysis.<br />

In situations where users provide observed fire data to the system, data assimilation methods will be<br />

used to improve the prediction potential <strong>of</strong> the model system.<br />

FUELS<br />

The NFBPS will cater for a large variety <strong>of</strong> fuel types existent in <strong>Australia</strong>, incorporating research<br />

carried out since the 1980s.<br />

• Grasslands<br />

o Natural grasslands (Cheney et al 1998)<br />

o Grazed grasslands (Cheney et al 1998)<br />

o Eaten out grasslands (Cheney et al 1998)<br />

o Grassland/woodland (Cheney and Andrews 2008)<br />

o Spinifex grasslands (Burrows et al. 2009)<br />

• Shrubland<br />

o Coastal heaths (Catchpole et al, in preparation)<br />

o Buttongrass Moorlands (Marsden-Smedley and Catchpole 1995)<br />

o Mallee-heaths (McCaw 1997, Cruz et al, in preparation)<br />

o Mallee-spinifex<br />

• Forest<br />

o Exotic pine plantations (Sneeuwjagt and Peet 1985, Cruz et al 2008a);<br />

o Dry sclerophyll eucalypt forest (Sneeuwjagt and Peet 1985, Burrows 1994, Cheney et<br />

al. 1992, Ellis 2000, Gould et al 2007);<br />

• Logging slash<br />

OUTPUT COMPONENTS<br />

The NFBPS is comprised <strong>of</strong> several output modules that aim respond to the requirements <strong>of</strong> distinct<br />

applications <strong>of</strong> fire behaviour information (Figure 1). The four module systems are:<br />

• Fuel Hazard Assessment System<br />

• Fire Danger Rating System<br />

• Fire Behaviour Prediction System<br />

• Bushfire Forecasting System<br />

Fuel Hazard Assessment System (FHA)<br />

Fuel hazard assessment aims to describe qualitatively the contribution <strong>of</strong> fuel complex structure to<br />

potential fire behaviour (McCarthy et al 1998, Gould et al. 2007). It allows integrating information<br />

from the various fuel strata into an overall hazard. Within the NFBPS fuel hazard assessment is linked


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 290<br />

to the quantitative component <strong>of</strong> fire behaviour modelling and is used to support landscape level fuel<br />

management, namely evaluating the effectiveness <strong>of</strong> fuel modification treatments, prescribed burn<br />

planning and evaluate WUI (Wildland Urban Interface) specific fuel hazards and propose mitigation<br />

actions.<br />

Fire Danger Rating System (FDR)<br />

Fire Danger is a process systematically evaluating and integrating the individual and combined factors<br />

influencing fire potential (e.g., ignition, rate <strong>of</strong> spread, difficulty <strong>of</strong> control, fire impact) at a large<br />

spatial scale (Merrill and Alexander 1987). Fire danger is calculated to optimize pre-suppression<br />

planning, resource allocation, and initial attack fire fighting tactics. It is also a vital component <strong>of</strong><br />

public awareness to potential fire hazard drives public policy related to certain fire issues. In NFBPS<br />

fire danger is intimately linked to fire behaviour. By integrating spatial information this adjusts fire<br />

potential to the specificity <strong>of</strong> a region vegetation cover and topography. This approach makes use <strong>of</strong><br />

available datasets to provide detailed information on the local conditions, necessary for some<br />

applications, while still allowing the scale up <strong>of</strong> the results to regional <strong>of</strong> state wide fire danger<br />

analysis.<br />

Fire Behaviour Prediction System (FBP)<br />

The fire behaviour prediction system aims to provide the site specific fire behaviour information that<br />

is necessary to plan and conduct prescribed burns, support wildfire suppression strategies and tactics,<br />

advise firefighters <strong>of</strong> safety concerns (e.g., red flag warnings) and gauging fuel management<br />

effectiveness. The temporal and spatial scope <strong>of</strong> the simulations depend on the information available<br />

(point data vs. GIS database) and the intended use <strong>of</strong> the simulation output.<br />

Bushfire Forecasting System (BF)<br />

The Bushfire Forecasting System is a module that aims to extend the temporal and spatial scale <strong>of</strong> the<br />

FBP system to look at long range fire potential (5 to 20 days). This system will combine spatial<br />

information on vegetation and topography with long-term weather forecasts and climate data and<br />

produce probabilistic outputs <strong>of</strong> fire impact zones and impact severity.<br />

FORMATS<br />

The system will be available in different formats. The core system is s<strong>of</strong>tware that has the capacity to<br />

be used at multiple spatial and temporal scales. The scale used for a particular simulation will depend<br />

solely upon the availability <strong>of</strong> the necessary inputs and the scope <strong>of</strong> the simulation. A user can carry<br />

out simple simulations with only a few basic input parameters (weather, fuel type, slope), extend such<br />

predictions for a 12- or 24-h period if forecasted weather is available, or simulate fire propagation<br />

across the landscape over complex topography under variable fuels and evolving weather conditions.<br />

This later simulation will require spatially explicit information <strong>of</strong> fuels, topography and weather.<br />

Aside from stand alone computers running the s<strong>of</strong>tware, it is envisioned that the system should also be<br />

able to run in a centralized server (localized in a Fire Behaviour Service Center) with the results being<br />

able to be visualised in mobile devises (laptops, smart phones, etc) through the internet. This will<br />

allow processor intensive simulations to be carried out quickly and the results could be available to a<br />

suit <strong>of</strong> users that would not have the time or training to run the simulations.<br />

The system will be also available in different formats suitable for quick verification in field conditions<br />

namely slide rules, tables and nomograms. In these formats simplified version <strong>of</strong> the core models will<br />

be used to produce first approximations <strong>of</strong> fire behaviour characteristics and intended mainly to be<br />

used as field reference guides.<br />

CONCLUDING REMARKS<br />

Fire is a multifaceted phenomena, and knowledge <strong>of</strong> its behaviour is necessary to comprehend its<br />

likely impact on human live, property and natural resources, develop proactive fire management plans<br />

and establish effective fire suppression strategies and tactics. The increase in complexity <strong>of</strong> the land<br />

and fire manager job in face <strong>of</strong> new climate scenarios, weather extremes and conflicting land<br />

management objectives has not been accompanied by the development <strong>of</strong> decision support tools that<br />

allow fire prediction using the most up to date fire behaviour science and technologies.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 291<br />

We propose a decision support system to predict fire behaviour and impacts based on the integration<br />

<strong>of</strong> the latest fire behaviour and peer-reviewed science within state <strong>of</strong> the art IT tools. The success <strong>of</strong><br />

fire behaviour prediction tools in supporting fire management decision-making in North America<br />

resulted in part from the extensive training received by users to understand how to use the model<br />

systems, and most importantly, the limitations associated with then and how they should be taken into<br />

consideration. There are still large areas <strong>of</strong> uncertainty in fire behaviour prediction and it is <strong>of</strong> utmost<br />

importance that users understand those. Failure to do so can lead to erroneous interpretations <strong>of</strong> output<br />

results, which at best can lead to a decrease in the trust users put on the models, and at worst can put<br />

lives and human property at risk. The development <strong>of</strong> the NFBPS needs to be accompanied by a well<br />

supported technology transfer program that ensures users are aware <strong>of</strong> system limitations and<br />

application bounds.<br />

REFERENCES<br />

Andrews, P. and Finney, F. (2007). Predicting wildfires. Scientific American, August issue. 47-55.<br />

Burrows, N.D. (1994). Experimental development <strong>of</strong> a fire management model for jarrah (Eucalyptus marginata<br />

Donn ex Sm.) forest. Aust. Natl. Univ., Canberra, <strong>Australia</strong>n Capital Territory. Ph.D. thesis. 293 p.<br />

Burrows, N.D., Ward, B., Robinson, A. (2009). Fuel dynamics and fire spread in spinifex grasslands <strong>of</strong> the<br />

western desert. Pages 69-76 in (Tran C. Ed.) The Proceedings <strong>of</strong> the Royal Society <strong>of</strong> Queensland –<br />

Bushfire 2006 conference. Brisbane 6-9 June 2006.<br />

Cheney, N.P., Gould, J.S., Catchpole, W.R. (1998). Prediction <strong>of</strong> fire spread in grasslands. International Journal<br />

<strong>of</strong> Wildland Fire 8(1) 1-13.<br />

Cheney, P., Sullivan, A. L. (2008). Grassfires: fuel, weather and fire behaviour. 2 nd Ed, CSIRO Publishing,<br />

Melbourne, <strong>Australia</strong>. 150p.<br />

Cheney, N.P., Gould, J.S., Knight, I. (1992). A prescribed burning guide for young regrowth forests in silvertop<br />

ash. Technical report. Forestry Commission <strong>of</strong> New South Wales., Sydney.<br />

Coleman, J., Sullivan, A. L. (1996). A real-time computer application for the prediction <strong>of</strong> fire spread across<br />

<strong>Australia</strong>n landscape. Simulation 67,230-240.<br />

Cruz, M.G., Alexander, M.E., Fernandes, P.A.M. (2008a). Development <strong>of</strong> a model system to predict wildfire<br />

behaviour in pine plantations. <strong>Australia</strong>n Forestry 71, 113-121.<br />

Cruz, M.G. xxxx. Application <strong>of</strong> a Monte Carlo based ensemble method to predict the rate <strong>of</strong> spread <strong>of</strong> grassland<br />

fires. Int. J. Wildland Fire, in press.<br />

Ellis, S, Kanowski, P, Whelan, R. (2004). National Inquiry on Bushfire Mitigation and Management.<br />

Commonwealth <strong>of</strong> <strong>Australia</strong>, Canberra.<br />

Ellis, P.F. (2000). The aerodynamic and combustion characteristics <strong>of</strong> eucalyptus bark – a firebrand study. Ph.D.<br />

Thesis, <strong>Australia</strong>n National University, Canberra, <strong>Australia</strong>. 187 p.<br />

Gould, J.S., McCaw, W.L., Cheney, N.P., Ellis, P.F., Knight, I.K., Sullivan, A.L. (2007). Project Vesta : fire in<br />

dry eucalypt forest : fuel structure, fuel dynamics and fire behaviour. (Ensis-CSIRO, Canberra ACT, and<br />

Department <strong>of</strong> Environment and Conservation, Perth WA)<br />

Matthews S., McCaw, W.L., Neal J.E., Smith R.H. (2007) Testing a process-based fine fuel moisture models in<br />

tow forest types. Can J. For. Res. 37:23-35.<br />

Marsden-Smedley, J.B., Catchpole, W.R. (1995). Fire behaviour modelling in Tasmanian buttongrass moorlands<br />

II. Fire behaviour. Int. J. Wildland Fire 5:215-228.<br />

McArthur, A.G. (1966). Weather and grassland fire behaviour. Forestry and Timber Bureau <strong>of</strong> <strong>Australia</strong> Leaflet<br />

1000.<br />

McArthur, A.G. (1967). Fire behaviour in dry eucalypt forest. Forestry and Timber Bureau <strong>of</strong> <strong>Australia</strong> Leaflet<br />

107.<br />

McCaw, W.L. (1997). Predicting fire spread in Western <strong>Australia</strong>n mallee heath shrubland. PhD thesis.<br />

Canberra, <strong>Australia</strong>: University <strong>of</strong> New South Wales, University College, School <strong>of</strong> Mathematics and<br />

Statistics. 235 p.<br />

Merrill, D.F., Alexander, M.E. (1987). Glossary <strong>of</strong> fire management terms. Canadian Committee on Forest Fire<br />

Management, NRCC Nº26516, Ottawa, Canada. 91 p.<br />

Sneeuwjagt, R.J., Peet, G.B. (1985). Forest fire behaviour tables for Western <strong>Australia</strong>. Department <strong>of</strong><br />

Conservation and Land Management, Perth, Western <strong>Australia</strong> 59p.<br />

Sullivan, A.L., Knight, I.K. (2001). Estimating error in wind speed measurements for experimental fires. Can. J.<br />

For. Res. 31:401-409.<br />

Tymstra, C,, Bryce, R.W., Wotton, B.M., Armitage, B. (2006). Prometheus – the Canadian Wildland Fire<br />

Growth Model: Model Development and Validation. Inf. Rep. In preparation. Alberta Sustainable<br />

Resource Development/Nat. Resour. Can., Can. For. Serv., North. For. Cent.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 292<br />

WOODY FUEL CONSUMPTION AND CARBON<br />

IN THE CHANGING CLIMATE OF AUSTRALIA<br />

Jennifer Hollis 1,3 and Lachlan McCaw 2,3<br />

ABSTRACT<br />

Woody fuel consumption was assessed at prescribed burns in four different forest types<br />

across southern <strong>Australia</strong>n eucalypt forests using variations <strong>of</strong> the line intersect<br />

technique. Research into the effect <strong>of</strong> climate change in fire potential in <strong>Australia</strong>n<br />

ecosystems indicated that future climate scenarios will result in an increase in the severity<br />

<strong>of</strong> fire seasons and fire intensity. The study aimed to quantify the link between fireline<br />

intensity and the fraction <strong>of</strong> coarse woody fuel consumed and carbon released in<br />

bushfires. Results from 39 prescribed burns ranging in intensity between 53 and 5000<br />

kW/m were analysed for the relationship between fireline intensity and proportion <strong>of</strong><br />

woody fuel consumed. Without the ability to isolate the effect <strong>of</strong> fireline intensity from<br />

other variables such as fuel moisture, it was difficult to assess its direct relationship with<br />

woody fuel consumption. While fireline intensity appears to effect the consumption <strong>of</strong><br />

fine fuels and to a lesser extent the small woody fuels (


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 293<br />

New South Wales. This included detailed assessment <strong>of</strong> woody fuel moisture, density and wood<br />

decay. The authors found a strong relationship between woody fuel consumption and fire intensity and<br />

noted that the greater the degree <strong>of</strong> decay, the greater the proportion <strong>of</strong> consumption. This suggests<br />

that CWD age may also be an important variable affecting fuel consumption.<br />

From 184 studies available in literature (Mackensen et al. 2003) found that in 57% <strong>of</strong> all cases, the<br />

calculated lifetime <strong>of</strong> CWD (time to when 95% <strong>of</strong> mass is lost) is longer than 40 years (the median <strong>of</strong><br />

the distribution was 49 years and the mean 92 years). In fact, coarse woody debris in some species<br />

such as jarrah (Eucalyptus marginata) were found to have a lifetimes <strong>of</strong> up to 120 years.<br />

CWD is significantly affected by fire disturbances making accurate accounting for its’ contribution to<br />

the carbon stock <strong>of</strong> a forest ecosystem particularly complex. In <strong>Australia</strong>n forests where CWD<br />

contributes approximately 18% <strong>of</strong> the total forest above-ground biomass and carbon stock in<br />

<strong>Australia</strong>n dry sclerophyll forests and 16% in wet sclerophyll forests (Woldendorp et al. 2002), fire<br />

can significantly modify CWD volume resulting in changes to greenhouse gas emissions and carbon<br />

stocks. This will vary from forest to forest and due to differences in the conditions under which they<br />

are burnt. One <strong>of</strong> the studies by (Hingston et al. 1980) in the jarrah forest (Eucalyptus marginata) <strong>of</strong><br />

southwest Western <strong>Australia</strong>, found coarse woody debris proportion <strong>of</strong> above ground biomass was<br />

32%, significantly higher than the average found in the culmination <strong>of</strong> studies by (Woldendorp et al.<br />

2002).<br />

In the event <strong>of</strong> a fire, prescribed or otherwise, consumed CWD will directly add to carbon emissions<br />

(Apps et al. 2006), however, the event may not necessarily result in the equivalent reduction in the<br />

coarse woody debris fuel load. As a result <strong>of</strong> the fire, newly fallen trees and branches will be<br />

transferred from overstorey and understory trees to a new coarse woody debris load (Waterworth and<br />

Richards 2008) which could even be as high as the pre-fire load depending on variables such as<br />

species, stand structure, fire behaviour (e.g., fire intensity, residence time) and termite damage.<br />

As the mass <strong>of</strong> carbon in coarse woody debris is a function <strong>of</strong> carbon concentration (%) and density,<br />

uncertainties exist in determining the contribution <strong>of</strong> coarse woody debris to carbon stocks. This can<br />

be attributed to the scarcity <strong>of</strong> information on coarse woody debris volume and decay rates (Brown et<br />

al 1996), and the impact <strong>of</strong> decay on wood density in <strong>Australia</strong>n forests (Grierson et al. 1992;<br />

Mackensen et al. 2003). As a result, reliable calculations <strong>of</strong> biomass, carbon stock values and CO2<br />

emissions are currently fraught with uncertainty until further data is collected through long-term<br />

studies <strong>of</strong> coarse woody debris in different environments (Mackensen et al. 2003). Current estimates<br />

<strong>of</strong> carbon in coarse woody debris mass range from 45 to 50 percent (Woodwell et al. 1978); Tilman et<br />

al. 2000; Mackensen and Bauhus 1999). For the purposes <strong>of</strong> this report, the conversion factor<br />

currently used by the <strong>Australia</strong>n Greenhouse Office (Mackensen and Bauhus 1999) will be used where<br />

the average carbon content is 50% <strong>of</strong> the coarse woody debris biomass.<br />

Woody fuels and climate change<br />

Fires have long been part <strong>of</strong> the <strong>Australia</strong>n bush and have played a major role in shaping vegetation<br />

structure and species composition (Bradstock et al. 2002; Burrows 2008; Gill et al. 1981). <strong>Australia</strong>n<br />

fire behaviour and regimes are closely related to meteorological conditions including temperature,<br />

humidity, temporal and seasonal rainfall patterns (and their affect on fuel moisture content) and wind<br />

speed (Catchpole, 2002; Cheney 1981; Luke and McArthur 1977; McAurthur 1973; McCaw et al.<br />

2003) as well as opportunities for ignition from lightning or man-made sources. Climate change<br />

projections for <strong>Australia</strong> indicate that each <strong>of</strong> these variables are likely to be impacted by climate<br />

change including a rise in annual mean temperatures as well as local variations in rainfall patterns that<br />

include precipitation regimes that have longer dry spells broken by heavier rainfall events (Lucas et al.<br />

2007).<br />

Climate change has the potential to alter a number <strong>of</strong> components <strong>of</strong> the fire regimes including fire<br />

frequency, size, intensity and length <strong>of</strong> the fire season including the period suitable for prescribed<br />

burning (Lucas et al. 2007). These changes have the potential to result in an increase in the area burnt<br />

by wildfires (Amiro et al. 2001; Gould and Cheney 2007) and as a result, increase instantaneous<br />

releases <strong>of</strong> greenhouse gases linked to biomass (including coarse woody debris). Understanding the<br />

effects that these changes could have on the amount and quality <strong>of</strong> coarse woody debris and improving


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 294<br />

our ability to predict fuel consumption and the resulting implications for carbon stocks and cycles<br />

within a forest ecosystem is becoming increasingly important.<br />

In this paper we assess the particular impact <strong>of</strong> fire intensity on woody fuel consumption and carbon<br />

release at prescribed burns in 4 different forest types across <strong>Australia</strong>. This includes:<br />

1. jarrah (Eucalyptus marginata) forest in the south-west <strong>of</strong> Western <strong>Australia</strong><br />

2. blue gum (Eucalyptus globulus)/ manna gum (Eucalyptus viminalis) forest in northern-central<br />

Victoria<br />

3. mountain gum (Eucalyptus dalrympleana) / narrow-leaf peppermint (Eucalyptis radiata)<br />

forest in south-eastern New South Wales<br />

4. stringybark (Eucalyptus obliqua) forest Tasmania<br />

We test the hypothesis that fireline intensity is a significant variable determining woody fuel<br />

comsumption. This hypothesis will be tested over the spectrum <strong>of</strong> woody fuel size classes.<br />

METHODOLOGY<br />

Woody Fuel Consumption Project (WFCP): Behind the Flaming Zone<br />

Woody fuel consumption was assessed during three prescribed burns in southwest Western <strong>Australia</strong><br />

including Wilga, Quilben and Hester blocks (Figure 1) in 2007 and 2008. Woody fuel consumption<br />

was also assessed in the Tallarook State Forest located in the Victorian northern central district (Figure<br />

1). Multiple plots were burnt under the same weather, season and fuel conditions at the Hester site in<br />

order to determine the specific effect <strong>of</strong> varying fireline intensity. This was achieved by varying fire<br />

direction in each plot, i.e. head-fire, backing fire and flank fires.<br />

Woody fuel loads (>0.6cm) were determined pre and post fire using Van Wagner’s line intersect<br />

method (Van Wagner 1968) with four fixed 100m transects (400m total transect length) placed within<br />

the burning plots. Five size classes were adopted including; Size 1: 0.6-2.5cm, Size 2: 2.5-7.5cm, Size<br />

3: 7.5-22.5cm, Size 4: 22.5-50.0cm, Size 5: >50.0cm. Fuels crossing each transect were counted and<br />

measured for diameter prior to the fire and wires were tied around the circumference <strong>of</strong> the fuels<br />

greater than 2.5cm to determine change in volume during post-fire assessment. Each fuel item was<br />

scored on a scale <strong>of</strong> 1-5 for decay, suspension, arrangement, charing and bark hazard. When possible,<br />

the species was also recorded. Pre and post-fire woody fuel loads were calculated using Brown’s<br />

Woody Material formula (Brown 1974).<br />

Rates <strong>of</strong> spread were calculated using thermologgers and the time intervals at which the fire passed<br />

known grid points at 320°C within the plots. Pr<strong>of</strong>ile, near surface and surface fine fuel moisture<br />

(


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 295<br />

Figure 1. Location <strong>of</strong> woody fuel consumption research sites across <strong>Australia</strong> including Wilga,<br />

Quilben, Hester and McCorkhill (Project Aquarius) in Western <strong>Australia</strong>, Tallarook<br />

in Victoria, Tumbarumba in New South Wales and the Warra LTER site in Tasmania<br />

Project Aquarius<br />

Woody fuel consumption was assessed at 32 experimental fires ignited under dry summer conditions<br />

at McCorkhill block in the southwest <strong>of</strong> Western <strong>Australia</strong> in 1983 (Figure 1).<br />

Woody fuel loads (>1.0cm) were determined pre and post fire using Van Wagner’s line intersect<br />

method (Van Wagner 1968) with 20m transects placed at regular intervals along established grid lines<br />

placed 100m apart within each fire plot. The total transect length for each plot varied according to plot<br />

size. Fuels in eight size classes were counted including; 10-25mm, 25-50mm, 50-75mm, 75-100mm,<br />

100-150mm, 150-200mm, 200-300mm and >300mm where the fuels horizontal diameter was also<br />

measured. These were subsequently converted to those used in the Woody Fuel Consumption Project<br />

to enable comparison across projects. Pre and post-fire woody fuels loads were calculated using<br />

Brown’s Downed Woody Material formula (see above Equation 1 and 2 where the fuel diameter is<br />

known).<br />

Patterns <strong>of</strong> fire development within the plots was measured from periodic mapping with an infra red<br />

line (IR) scanner. The digital data from the scanner was corrected for scale, scene compression, jitter<br />

by aircraft movements and geometrically rectified onto a known planar map to enable fire spread rates<br />

to be calculated.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 296<br />

Pr<strong>of</strong>ile and surface fine fuel moisture (7.0cm. Brown’s equation (1974) was adopted to calculate fuel load which incorporates an angle<br />

correction factor as well as a slope correction factor from Brown and Roussopoulus (1974).<br />

The quadratic mean diameter (QMD) for size classes 2.5-5.0cm and 5.0-7.0cm was determined during<br />

field sampling by recording diameters within each size class and using Van Wagner’s equation to<br />

calculated the QMD (Van Wagner 1982).<br />

Species densities were calculated using the water displacement method (TAPPI, 1994) and were<br />

means <strong>of</strong> at least 20 branches in each diameter class for each species.<br />

The ignition method used, centre fire ignition (convection), did not allow detailed measurements <strong>of</strong><br />

fire behaviour properties. Flame heights were estimated to be 4 m. In the absence <strong>of</strong> fire behaviour<br />

information fireline intensity was estimated based on the Gould et al (2007) relationship between<br />

mean flame height and mean fire intensity.<br />

Tumbarumba<br />

Woody fuel consumption research was undertaken in February 2004 within the Maragle State Forest,<br />

Tumbarumba, located in south-eastern New South Wales.<br />

Detailed background, methodology and fire behaviour from this research was given by Tollhurst et al.<br />

(2006). Pre and post fire woody load (>2.5cm diameter) was determined using the Van Wagner (1968)<br />

line intersect method with 3 x 30m transects in each <strong>of</strong> the plots (i.e. 90m transect lengths). The<br />

diameter <strong>of</strong> fuels intersecting the transect was recorded along with a rating for decay and suspension<br />

class. Each fuel circumference was tied with 2mm wire to determine volume consumed. Four size<br />

classes were adopted including 2.6-7.5cm, 7.6-22.5cm, 22.6-50cm and >50cm diameter.<br />

Woody fuel moisture was assessed across experimental blocks. Three measures <strong>of</strong> fuel moisture were<br />

obtained for each wood sample including two by oven determination; from ‘inner’ and ‘outer’<br />

locations <strong>of</strong> the sample, and one by electronic moisture meter (T-H Fine Fuel Moisture Meter (Chatto<br />

and Tolhurst 1997)) from the saw dust generated during the cutting <strong>of</strong> the sample .<br />

Woody density was measured in a range <strong>of</strong> diameter and decay classes by determining the volume <strong>of</strong><br />

wood samples from the weight increase after submersion in water.<br />

Plot ignition was from a 100m continuous line and on the up-wind, down-slope position <strong>of</strong> the block<br />

with a final burn area approximately 4ha. Fire intensity was calculated using Byram’s (1959) fireline<br />

intensity equation as described previously.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 297<br />

Table 1. Site and Burn Characteristics<br />

Burn ID Av. Annual<br />

Rainfall<br />

(mm)<br />

Dominant Species Burn Type Ignition Technique<br />

WCFP - Wilga 830 Eucalyptus marginata Silvicultural Long Line<br />

WCFP - Quilben 1012 Eucalyptus marginata Ecological / Fuel Reduction Long Line<br />

WCFP - Hester 830 Eucalyptus marginata Ecological / Fuel Reduction Long Line<br />

WCFP -<br />

595 Eucalyptus globulus Ecological / Fuel Reduction Long Line<br />

Tallarook<br />

Eucalyptus viminalis<br />

Project Aquarius 1140 Eucalyptus marginata Ecological / Fuel Reduction Long Line/ Multiple<br />

- McCorkhill<br />

Ignition Point<br />

Warra LTER 883 Eucalyptus obliqua Silvicultural Central Ignition<br />

Tumbarumba 975 Eucalyptus dalrympleana<br />

Eucalyptus radiate<br />

Ecological / Fuel Reduction Long Line<br />

RESULTS AND DISCUSSION<br />

Pre-fire fuel load distribution<br />

The pre-fire fuel load distribution by size class is similar across sites with the exception <strong>of</strong> the Warra<br />

LTER sites which appear to have significantly higher fuel loads in each <strong>of</strong> the size classes (Figure 2).<br />

The Quilben site in Western <strong>Australia</strong> also has a higher than average fuel load for fuels greater than<br />

50cm (size class 5). At each site, the larger size classes, particularly sizes 4 and 5, form much <strong>of</strong> the<br />

overall woody fuel load (on average 35 and 30% respectively).<br />

This highlights the importance <strong>of</strong> the larger fuels to overall woody fuel consumption. This is an<br />

important characteristic <strong>of</strong> the problem under analysis, namely when we attempt to investigate and<br />

understand the effect <strong>of</strong> different environmental variables and fire behaviour characteristics, such as<br />

fireline intensity, on fuel consumption.<br />

Fuel Load (t/ha)<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

Wilga<br />

Quilben<br />

Hester<br />

Tallarook<br />

Aquarius<br />

Warra LTER<br />

Tumbarumba<br />

Pre-fire Fuel Load Distrubution by Size Class<br />

Fine Fuel (0-0.6cm) Size 1 (0.6-2.5cm) Size 2 (2.5-7.5cm) Size 3 (7.5-22.5cm) Size 4 (22.5-50cm) Size 5 (>50cm)<br />

Size Class<br />

Figure 2. Pre-fire fuel load distribution by size class.<br />

Fire behaviour and fuel condition<br />

A wide range <strong>of</strong> weather and seasonal influences are represented in the dataset including burns<br />

conducted under typical spring and autumn prescribed burning conditions as well as those<br />

characteristic <strong>of</strong> dry summer wildfires (see Table 2 where burning conditions have been averaged for<br />

each site). The broad range <strong>of</strong> burning conditions is also reflected by fireline intensity which ranged<br />

from 53 to 5000 kW/m and in the fine (pr<strong>of</strong>ile) and woody fuel (log average) moisture contents,<br />

ranging from 8-72% and 33-56% respectively (Table 3).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 298<br />

Table 2. Summary <strong>of</strong> average fire behaviour characteristics across burn sites. Range across<br />

multiple burns (minimum to maximum) in italics.<br />

Site/Mean<br />

Characteristics<br />

No.<br />

<strong>of</strong><br />

fires<br />

RH %<br />

Temp<br />

( o C)<br />

10m Open<br />

Wind<br />

Speed<br />

(kph)<br />

KBDI SDI<br />

ROS<br />

(m/hr)<br />

Residence<br />

Time<br />

(s)<br />

Fireline<br />

Intensity<br />

(kW/m)<br />

WFCP -<br />

Wilga<br />

1 27.5 24.9 7.8 42.9 97.7 93.7 299.0<br />

WFCP -<br />

Quilben<br />

1 68.5 21.1 5.5 84.9 52.2 26.9 209.9<br />

WFCP -<br />

Hester<br />

4 56.4<br />

(48-63)<br />

24.8<br />

(23-27)<br />

13.7<br />

(11-17.5)<br />

147.5<br />

(148-<br />

148)<br />

105.0<br />

(15-217)<br />

21.6<br />

(10-40)<br />

348.7<br />

(53-678)<br />

WFCP -<br />

Tallarook<br />

2 50.1<br />

(34-66)<br />

16.5<br />

(13-20)<br />

8.6<br />

(8.2-8.9)<br />

36.8<br />

(14-60)<br />

139.5<br />

(136-<br />

143)<br />

52.3<br />

(20-85)<br />

77.5<br />

(28-127)<br />

234.0<br />

(76-393)<br />

Project<br />

18 45.7 24.5 5.2 139.1 372.8<br />

not<br />

measured<br />

1680.5<br />

Aquarius<br />

(20-61) (18-33) (2.5-24)<br />

(129-<br />

163)<br />

(153-<br />

774)<br />

(585-3304)<br />

Warra<br />

LTER<br />

11 67.4<br />

(52-90)<br />

18.1<br />

(17-19)<br />

not<br />

measured<br />

51.0<br />

(51-51)<br />

not<br />

measure<br />

d<br />

not<br />

measured<br />

5000.0<br />

(5000-<br />

5000)<br />

Tumba-<br />

2 32.5 27.0 7.3 122.0 369.9<br />

not<br />

measured<br />

2430.6<br />

rumba<br />

(20-45) (26-28) (6.5-8)<br />

(122-<br />

122)<br />

(122-<br />

618)<br />

(955-3906)<br />

Table 3. Summary <strong>of</strong> fuel moisture conditions and fuel consumption outcomes across burn<br />

sites. Range across multiple burns (minimum to maximum) in italics.<br />

Site/<br />

Mean<br />

Characteristics<br />

Pr<strong>of</strong>ile<br />

Fuel<br />

Moisture<br />

Content<br />

(%)<br />

Log<br />

Moisture<br />

Content<br />

>0.6cm<br />

(%)<br />

Total Prefire<br />

Fine<br />

Fuel Load<br />

0.6cm<br />

(t/ha)<br />

Woody<br />

Fuel<br />

Consum<br />

ption<br />

(%)<br />

Carbon<br />

Release<br />

(t/ha)<br />

WFCP -<br />

Wilga<br />

11.6 38.7 5.9 0.2 42.3 22.1 47.6 10.1<br />

WFCP -<br />

Quilben<br />

24.6 37.3 7.8 3.0 175.0 121.7 30.5 26.7<br />

WFCP - 16.6 32.9 6.6 0.6 93.8 47.3 49.4 23.3<br />

Hester (16.6-16.6) (33-33) (6.0-7.0) (0.1-1.3) (76-106) (40-62) (42-57) (17-30)<br />

WFCP - 51.9 45.0 8.1 1.4 52.2 31.6 39.7 10.3<br />

Tallarook (32.3-71.5) (35-56) (7.3-8.9) (0.3-2.6) (49-55) (28-36) (36-43) (10-11)<br />

Project<br />

10.4 not measured 8.8 0.0 60.5 27.1 54.9 16.7<br />

Aquarius (8.3-13.2) (6.3-12.0) (0-0) (33-107) (5-54) (33-90) (8-36)<br />

Warra not measured not measured 44.1 9.0 599.5 336.8 46.4 131.4<br />

LTER (30-53) (6.7-10.6) (226-1322) (71-795) (9-69) (31-263)<br />

Tumba-<br />

11.0 47.2 12.8 0.0 86.3 52.3 44.2 17.0<br />

rumba (11.0-11.0) (47-47) (12-13) (0-0) (49-123) (22-83) (33-56) (14-20)<br />

Woody fuel consumption and fireline intensity<br />

At a plot/fire specific level woody fuel (>0.6cm) consumption ranged from 9 to 90%. However the<br />

average for each site ranged from 31 – 55% (Table 3). Using the amount <strong>of</strong> fuel consumed at each plot<br />

this equates to a range <strong>of</strong> carbon release between 8 and 263t/ha, for an average <strong>of</strong> 50t/ha. At a plot/fire<br />

level woody fuel consumption appears to increase with fireline intensity up to approximately 700<br />

kW/m (Figure 3) after which no definable relationship appears to exist. This is illustrated in the weak


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 299<br />

regression relationship for site fuel consumption with fireline intensity (sizes 1-5 combined) (R 2 =<br />

0.01) in Table 4 below.<br />

By following the diameter reduction relationships <strong>of</strong> individual fuel items at each <strong>of</strong> the Woody Fuel<br />

Consumption Project (WFCP) burns, it has been possible to assess woody fuel consumption by size<br />

class and intensity. Figure 4 illustrates the scatter <strong>of</strong> the consumption (by intensity) across the WFCP<br />

sites. There appears to be a weak relationship between fireline intensity and wood fuel consumption <strong>of</strong><br />

the fine (R 2 = 0.48) and size class 5 (R 2 = 0.41). The regression explanatory power for the other classes<br />

was weaker, with an R 2 <strong>of</strong> 0.23 obtained for size class 1, and R 2 below 0.1 for sizes 2, 3 and 4.<br />

This supports the hypothesis that fireline intensity is an influencing variable in the consumption <strong>of</strong> fine<br />

fuels and possibly the small woody fuels (i.e. size class 1 (0.6-2.5cm). It also appears that fireline<br />

intensity may influence the consumption <strong>of</strong> fuels greater than 50cm (size 5). This may suggest that the<br />

consumption <strong>of</strong> larger proportions <strong>of</strong> the fine and small woody associated with higher fireline<br />

intensities, is required to ignite and consume the fuels greater than 50cm.<br />

Woody Fuel (>0.6cm) Consumption (%)<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

Woody Fuel Consumption & Fireline Intensity<br />

0<br />

0 1000 2000 3000 4000 5000<br />

Fireline Intensity (kW/m)<br />

Figure 3. Scatterplot <strong>of</strong> consumption and fireline intensity across all prescribed burns.<br />

Woody Fuel (>0.6cm) Consumption (%)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Fine Fuel Consumption & Fireline Intensity<br />

0<br />

0 100 200 300 400 500 600 700<br />

Fireline Intensity (kW/m)<br />

Fine Fuel (50cm)<br />

Fine Fuel<br />

Size 1<br />

Size 2<br />

Size 3<br />

Size 4<br />

Size 5<br />

Project Aquarius<br />

T umbarumba<br />

WFCP WA<br />

WFCP Tallarook<br />

Warra LTER<br />

Figure 4. Scatterplot and regression (y=a ln(x)) analysis <strong>of</strong> consumption and fireline intensity by<br />

size class at the Woody Fuel Consumption Project (WFCP) sites.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 300<br />

Table 4. Prediction equations for woody fuel consumption by size class using fireline intensity<br />

at the Woody Fuel Consumption Project (WFCP) sites.<br />

Fuel size class Regression Equation (linear) R²<br />

Fine fuel (50cm) 13.6 + 0.0731 Fireline Intensity 0.41<br />

Site (class 1-5 combined) *<br />

* Includes data across entire dataset<br />

51.7 - 0.00081 Fireline Intensity 0.01<br />

It is noted that the variety <strong>of</strong> conditions under which each plot and site have been burnt will also<br />

contribute significantly to the varied consumption outcomes. Some <strong>of</strong> the variables are also likely to<br />

be highly correlated, for example woody fuel moisture – log decay, and fireline intensity - Soil<br />

Dryness Index. Given this, the best example <strong>of</strong> the effect <strong>of</strong> fireline intensity on woody fuel<br />

consumption was found at the Hester site burns which were burnt concurrently (same day, same time).<br />

Through this method we were able to burn the various plots under the same fuel moisture conditions,<br />

and variable fireline intensity. This was achieved by burning distinct areas <strong>of</strong> the plot by head, back<br />

and flank fires. For this prescribed burn woody fuel consumption ranged from 42.2 – 56.9% with the<br />

highest consumption occurring within the plot with the highest fireline intensity (678 kW/m)(Table 5).<br />

Table 5. Woody fuel consumption by size class at the Hester site burns.<br />

Fireline<br />

Intensity<br />

(kW/m)<br />

Fine<br />

Fuels<br />

Fuel Consumption by Size Class (%)<br />

1 2 3 4 5<br />

Woody Fuel<br />

>0.6cm<br />

(sizes 1-5<br />

combined)<br />

Hester 1 678 99.0 76.7 42.6 52.4 62.5 56.5 56.9<br />

Hester 2 380 94.5 57.8 48.3 33.5 46.4 67.3 55.2<br />

Hester 3 284 92.4 55.8 48.0 50.8 35.0 47.4 42.2<br />

Hester 4 53 81.3 68.2 45.0 46.2 56.5 16.7 43.5<br />

Within each <strong>of</strong> the size classes, fine fuel consumption increased with increasing intensity however the<br />

same relationship was not clear in any <strong>of</strong> the woody fuel size classes greater than 0.6cm. It appears<br />

that the low intensity (53 kW/m) <strong>of</strong> the Hester 4 burn may have had an effect on the consumption <strong>of</strong><br />

the woody fuels greater than 50cm (size 5) which was minimal (16.7%). This could be the result <strong>of</strong> the<br />

small energy quantity being released by the surface fire not meeting the energy requirements to ignite<br />

the large fuels.<br />

Climate change implications associated with fireline intensity<br />

Climate change has the potential to affect fire regimes by modifying fire intensity. While the<br />

relationship between fine fuel consumption and fireline intensity appears to support theories that fire<br />

intensity influences the burn patchiness and proportion <strong>of</strong> fine fuels consumed (e.g., Moreno &<br />

Oechel, 1989), the effect on woody fuel consumption and carbon release is unclear. This could be<br />

because there are so other variables, some <strong>of</strong> them correlated, that affect woody fuel consumption,<br />

making it difficult to isolate the relationship between fireline intensity and woody fuel consumption.<br />

It may also be that the relationship between woody fuel consumption and fireline intensity is very<br />

weak, possibly less important than other variables such seasonal dryness and associated large fuel<br />

moisture content. It is reasonable to expect that under the future altered fire climate, with longer and<br />

more severe (drier) fire seasons, the consumption <strong>of</strong> woody fuels will increase as a greater proportion<br />

<strong>of</strong> areas are burned under drier, more intense (high fire danger indices) wildfire conditions. This is an<br />

important aspect to take into consideration when planning and conducting prescribed burns.<br />

It is possible that fireline intensity plays a role in influencing the consumption <strong>of</strong> large size 5 fuels<br />

(>50cm), and further research is required to look at this relationship as they form a large proportion <strong>of</strong>


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 301<br />

the CWD fuel load (on average 30%). Their ignition and consumption are important processes<br />

responsible for the release <strong>of</strong> large quantities <strong>of</strong> stored carbon.<br />

CONCLUSIONS<br />

The relationship between fireline intensity and the consumption <strong>of</strong> woody fuels (>0.6cm) and<br />

associated carbon release in southern <strong>Australia</strong>n eucalypt forests is unclear making it difficult to assess<br />

the affect <strong>of</strong> potential climate change scenarios if only fireline intensity is considered.. While fireline<br />

intensity appears to effect the consumption <strong>of</strong> fine fuels and to a lesser extent the small woody fuels<br />

(


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 302<br />

Mackensen J, Bauhus J, Webber E (2003) Decomposition rates <strong>of</strong> coarse woody debris-A review with particular<br />

emphasis on <strong>Australia</strong>n tree species. <strong>Australia</strong>n Journal <strong>of</strong> Botany 51, 27-37.<br />

Marsden-Smedley JB, Slijepcevic A (2001) Fuel characteristics and low intensity buring in Eucalyptus obliqua<br />

wet forest at the Warra LTER site. Tasforests 13, 261-280.<br />

Moreno, J.M, Oechel, W.E. 1989. A simple method for estimating fire intensity after a burn in California<br />

chaparral. Acta OEcologica 10(1):57-68.<br />

Nelson RM 2003 Reaction times and burning rates for wind tunnel headfires*. International Journal <strong>of</strong> Wildland<br />

Fire 12, 195-211<br />

Ottmar RD (1987) Prescribed Fire and Fuel Consumption in Uncured Slash--Preliminary Results. In<br />

'Proceedings <strong>of</strong> the Ninth Conference on Fire and Forest Meteorology'. San Diego, CA<br />

Ottmar RD, Burns MF, Hall JN, Hanson AD (1993) 'Consume Users Guide.' USDA Forest Service Pacific<br />

Northwest Research Station.<br />

Ottmar RD, Prichard SJ, Nihanek RE, Sandberg DV, Bluhn A (2006) 'Modification and Validation <strong>of</strong> Fuel<br />

Consumption Models for Shrub and Forested lands in Southwest, Pacific Northwest, Rockies, Midwest,<br />

Southeast and Alaska.' Fire and Environmental Applications Team, Pacific Northwest Research Station,<br />

Pacific Wildland Fire Sciences Laboratory, Seattle, Washington.<br />

Sandberg DV, Ottmar RD (1983) Slash burning and fuel consumption in the Douglas-fire subregion. In 'Seventh<br />

Conference on Fire and Forest Meteorology Symposium Proceedings'. Fort Collins, CO pp. 90-93.<br />

(American Meteorological Society)<br />

Slijepcevic A (2001) Loss <strong>of</strong> carbon during controlled regeneration burns in Eucalyptus obliqua forest.<br />

Tasforests 13.<br />

Sullivan AL, Knight IK, Cheney NP (2002) Predicting the radiant heat flux from burning logs in a forest<br />

following a fire. <strong>Australia</strong>n. Forestry 65, 59–67.<br />

TAPPI 1994. Basic density and moisture content <strong>of</strong> pulpwood. Test Method T268 om-94. TAPPI, Atalanta,<br />

Georgia, USA. 5p.<br />

Tilman D, Reich P, Phillips H, Menton M, Patel A, Vos E, Peterson D, Knops J (2000) Fire Suppression and<br />

Ecosystem Carbon Storage. Ecology 81.<br />

Tollhurst KG, Anderson WR, Gould J (2006) Woody Fuel Consumption experiments in an undisturbed forest. In<br />

'Conference Proceedings <strong>of</strong> the V International Conference on Forest Fire Research'. (Ed. DX Viegas)<br />

Van Wagner CE (1968) The line intersect method in forest fuel sampling. Forest Science 14, 20-26.<br />

Van Wagner CE (1982) Notes: Graphical Estimation <strong>of</strong> Quadratic Mean Diameters in the Line Intersect Method.<br />

Forest Science 28, 852-855.<br />

Woodwell GM, Whittaker RH, Reiners WA, Likens GE, Delwiche CC, Botkin DB (1978) The Biota and the<br />

World Carbon Budget. Science 199, 141-146.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 303<br />

ABSTRACT<br />

VICTORIAN BUSHFIRES 2009 -<br />

A PLANTATION COMPANY’S EXPERIENCE<br />

Malcolm Tonkin 1<br />

February 7 th 2009 was a day <strong>of</strong> tragedy in Victoria on which many lives were lost, along<br />

with houses and other community assets. The fire weather conditions reached and<br />

exceeded predicted abnormal extremes and HVP Plantations lost around 16,500 ha <strong>of</strong> its<br />

plantation estate, while 7,800 ha <strong>of</strong> its native forest was also burnt.<br />

As the drought continues beyond 12 years, the level <strong>of</strong> Company fire preparedness<br />

increases each year, the fire seasons get longer and their intensity increases; betweenseason<br />

relief for staff gets shorter and the potential for asset loss is more sustained. This<br />

is a reflection <strong>of</strong> the broader Victorian community experience. Although the drought<br />

conditions and fuels were similar to the much larger 2006 Alpine fires, the severity <strong>of</strong> the<br />

weather and the geographic context <strong>of</strong> the fires on February 7 th provided very different<br />

outcomes for the community and the company.<br />

INTRODUCTION<br />

HVP Plantations (HVP) is <strong>Australia</strong>’s largest, private, timber plantation company, with assets <strong>of</strong> over<br />

$800 million. The company is owned jointly by <strong>Australia</strong>n and US superannuation and investment<br />

funds and manages 245,000 hectares <strong>of</strong> land in Victoria including 170,000 hectares <strong>of</strong> plantations<br />

These plantations supply over 3 million tonnes <strong>of</strong> wood to rural and regionally based processing<br />

industries, including sawmills, paper mills, newsprint mills and panel plants, and generating<br />

employment for over 3,000 people.<br />

In the current continuing drought, fire is a significant risk to the plantations and to the industries and<br />

communities dependent on the timber resources they provide. To manage this risk, HVP is a<br />

substantial contributor to the prevention and suppression <strong>of</strong> fire in rural Victoria. HVP lost over 2000<br />

ha <strong>of</strong> plantation in each <strong>of</strong> the large 2003 and 2006 Alpine and associated fires and 16,500 ha in the<br />

recent January /February 2009 fires. These events have significant Company and community impacts.<br />

The most recent year in which there was widespread, above-average rain in Victoria was 1996. At<br />

some locations, such as Melbourne, there have been 12 consecutive years <strong>of</strong> below-average rainfall.<br />

In this area, twelve-year rainfall totals have been around 20% below the 1961-90 average, and 10-13%<br />

below the lowest on record for any twelve-year period prior to 1996.<br />

Overlayed on the seasonal preconditions for fire generated by the drought, there was a specific build<br />

up <strong>of</strong> warm to hot weather conditions in the 11 days leading to February 7 th 2009.<br />

The weather conditions for this day were accurately forecast well in advance <strong>of</strong> the day itself, and<br />

clearly indicated a day <strong>of</strong> abnormally high risk with Forest Fire Danger Indices over most <strong>of</strong> the state<br />

exceeding 100.<br />

HVP ROLE IN FIRE PREVENTION AND SUPPRESSION<br />

Wildfire in the rural landscape is a community issue and as a rural landowner HVP plays a role both<br />

within and around its estate to protect life and assets; in particular its plantation assets.<br />

HVP maintains seven Forest Industry Brigades (FIBs) under the legislative framework <strong>of</strong> the Country<br />

Fire Authority Act (1958). These HVP FIBs are located, managed, equipped, maintained, structured<br />

1 rd<br />

General Manager, Stewardship and Risk, HVP Plantations,3 Floor, 517 Flinders Lane, Melbourne, Victoria 3000.<br />

Ph: 03 9289 1400 Email: MTonkin@hvp.com.au


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 304<br />

and financed by the company but, like other Country Fire Authority (CFA) brigades, they operate<br />

legally in accordance with the Standing Orders <strong>of</strong> the Chief Officer <strong>of</strong> the CFA.<br />

HVP exceeds its statutory obligations in relation the provision <strong>of</strong> trained brigade members and<br />

equipment, providing crews to man 20 large (3,000-4,000 litre) forest fire tankers and 38 light (400<br />

litre) first-attack slip-on units. The company also places two first-attack helicopters on standby during<br />

key periods in the summer.<br />

In response to the fire emergency <strong>of</strong> early 2009, HVP provided more than 200 trained and experienced<br />

fire-fighters, including more than 50 drawn from other operations interstate and overseas.<br />

The company enjoys strong relationships with the Department <strong>of</strong> Sustainability and Environment<br />

(DSE) and the CFA at both the local and corporate level, as a brigade within the CFA and as a major<br />

neighbour <strong>of</strong> DSE, with whom it shares about 70% <strong>of</strong> its property boundaries in Victoria.<br />

HVP radio communications are compatible with DSE and CFA and all <strong>of</strong> the company’s fire vehicles<br />

have real time GPS radio tracking for additional safety during fires.<br />

HVP crews attend fires within what it considers to be its area <strong>of</strong> interest as plantation losses<br />

commonly occur from fire entering a plantation from outside, rather than from fire igniting within the<br />

company’s plantations. As FIBs within the CFA, HVP operates at fires within the control plan for that<br />

incident as determined by the responsible fire control agencies (DSE and CFA). HVP chooses not to<br />

provide staff for roles in the Incident Control Teams managing fires, but prefers to provide a liaison<br />

person who is able to emphasise the areas <strong>of</strong> strategic importance to HVP assets and ensure that our<br />

crews and adequate equipment are deployed largely under our control to those areas. The role <strong>of</strong><br />

liaison <strong>of</strong>ficers has proven to be very effective in sharing information and co-ordinating operations.<br />

In large fires the company benefits from managing its own logistics. Even though HVP works within<br />

the Incident Action Plan, Company fire-fighters in debriefing sessions reinforce the need to maintain<br />

our own logistic functions where possible. This is to ensure efficient service rather than be subject to<br />

logistics provided by the large emergency organisational arrangements which can become less<br />

efficient due to their scale. HVP has at all major fires successfully managed its own logistics,<br />

including equipment, meals, accommodation, fuel distribution mapping services, together some<br />

additional planning functions.<br />

HVP Regions each have a fire plan and an annual works program which schedules fire prevention<br />

maintenance works and additional summer fire crews and equipment availability. Contractors are<br />

required to have specified fire equipment and HVP has a graduated forest closure schedule to<br />

minimise internal fire ignition risk which ultimately shuts down all operations in the plantation on bad<br />

days when the FFDI is predicted to be 45 or above.<br />

As part <strong>of</strong> the wider forest industry, HVP has worked co-operatively with other plantation companies<br />

in fire prevention and suppression for many years. The aggregated resources and skill levels <strong>of</strong> the<br />

plantation industry in Victoria are a significant private contribution to the state’s fire suppression<br />

resources and are the prime example <strong>of</strong> a public-private partnership in the provision <strong>of</strong> emergency<br />

services.<br />

FEBRUARY 7 TH 2009<br />

On Saturday February 7 th , after several weeks <strong>of</strong> temperatures in the mid to high 30’s, the forecast for<br />

Ballarat (as an example) was for 41 o C, NNW winds <strong>of</strong> 65 km/hour gusting to 85 km/hour, with a dry<br />

wind change to the SW in the afternoon at 60 km/hr. The Forest Fire Danger Index (FFDI) was<br />

forecast to be 185. While it is not clear what FFDI’s <strong>of</strong> this magnitude really mean, the day was<br />

clearly forecast in advance to be abnormally severe.<br />

As a precursor to February 7 th , the first Gippsland fires occurred on 29 th January. The Delburn fire<br />

(6,400ha) was caused by multiple deliberate lights (a suspect has been charged by police) on the<br />

afternoon <strong>of</strong> 29 th January 2009. It burnt around 2,700 ha <strong>of</strong> HVP plantation (eucalypt and pine) and<br />

other assets including 29 houses in the Boolara area. The fact that this 6,000 ha fire did not run again<br />

a week later, under the most extreme conditions parallel to the Churchill fire into the Strzelecki


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 305<br />

Ranges, is a great credit to those CFA, DSE and HVP crews who worked on it both before and on that<br />

day.<br />

Fires which were controlled but not yet safe at Delburn in Gippsland, and several lightning strikes in<br />

the North East, were priority tasks identified by HVP in the lead up to February 7 th , as potential fires<br />

to threaten plantation assets.<br />

On Saturday February 7 th four major fires burnt into HVP plantations. All four fires were managed by<br />

joint incident control between the DSE and the CFA. As a Forest Industry Brigade within the CFA,<br />

HVP resources were tasked by the Incident Control Team for each fire and HVP liaison <strong>of</strong>ficers<br />

placed in key Incident Control Centres.<br />

The Churchill fire (24,500 ha) in Gippsland started around 1.30 pm. It was allegedly deliberately lit<br />

and a suspect has been charged by police. It reportedly travelled 6km in 10 minutes in the HVP pine,<br />

eucalypt plantations and native forest <strong>of</strong> the Strzelecki Ranges. Substantial loss <strong>of</strong> life, community<br />

property and 8,000ha <strong>of</strong> HVP plantations occurred under the NW wind and then after the SW wind<br />

change.<br />

The Murrindindi fire <strong>of</strong> unknown cause (suspected to be deliberate) started at around 3pm and<br />

travelled quickly through 15km <strong>of</strong> native forest to spot into HVP plantations near Buxton within an<br />

hour or so. Substantial loss <strong>of</strong> life, community property and 3000 ha <strong>of</strong> HVP plantations occurred<br />

under the NW wind and also after the SW wind change. This included the devastation <strong>of</strong> the town <strong>of</strong><br />

Marysville. The mill <strong>of</strong> a long-time HVP customer at Narbethong, was also lost, as was a harvesting<br />

contractor’s equipment.<br />

The Kilmore East fire is alleged to have been started by fallen power lines around 11.30pm and<br />

travelled quickly through 25km <strong>of</strong> the Mt Disappointment forest and burnt HVP’s 980 ha Kinglake<br />

West plantation. Substantial loss <strong>of</strong> life and community property occurred both under the NW wind<br />

and also after the SW wind change including the town <strong>of</strong> Kinglake, some 35km from the origin. A<br />

cable harvesting tower and other harvesting equipment was lost in the fire. The Kilmore East and<br />

Murrindindi fires eventually merged with a total area <strong>of</strong> over 250,000 ha.<br />

The Beechworth fire (31,000ha) is alleged to have been started by fallen power lines at 6.20pm,<br />

burning a number <strong>of</strong> HVP plantations totalling 1,900 ha under the NW wind and subsequently on the<br />

SW wind change the next day.<br />

There was little effective work which could be done to save plantations on the day <strong>of</strong> February 7 th<br />

Many HVP staff were involved in saving houses and other assets. Fortunately, all except one <strong>of</strong> our<br />

staff could return to a home, although many staff homes were threatened. There were no serious<br />

injuries and in the subsequent weeks no mills ran out <strong>of</strong> logs to compound the community distress.<br />

Sadly there are many in our communities who were not so fortunate.<br />

Support from the timber industry was critical in enabling HVP staff to get some rest, while still<br />

ensuring protection <strong>of</strong> plantation assets. A crew <strong>of</strong> 9 from Hancock Forest Management in NZ<br />

assisted on the Delburn fire and were held over in anticipation <strong>of</strong> the bad weather forecast for Saturday<br />

7 th February and thereafter a second crew <strong>of</strong> 8 also came to assist.<br />

On subsequent days there were many other <strong>of</strong>fers <strong>of</strong> assistance from the timber industry and crews<br />

came from plantation companies from both within and outside the state.<br />

COMPARISON BETWEEN THE 2006 ALPINE FIRES AND FEBRUARY 7 TH 2009<br />

The 2006 Alpine fires and the 2009 fires both became major fire campaigns within the state. The<br />

comparison is <strong>of</strong> interest because while they were both destructive and disruptive events for the state<br />

and for HVP business, the circumstances were quite different and both scenarios need to be<br />

incorporated in future plantation management, risk and response strategies.<br />

Following serious winter/spring rainfall deficits, the fires in 2006 commenced with multiple (80+)<br />

lightning strikes in remote alpine locations in north-eastern Victoria. These fires had great potential to<br />

run and coalesce, but it was not until 5 days later that the fire weather caused significant expansion.<br />

These fires then continued for about 2 months, burning over 1 million hectares (5% <strong>of</strong> Victoria).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 306<br />

These enlarged fires exceeded the capacity <strong>of</strong> fire-fighting resources to contain them. No lives and<br />

few houses were lost. In 2003, 1.2 million ha was burnt similarly in the Alpine regions.<br />

While again following serious winter/spring rainfall deficits, the fire scenario in 2009 was quite<br />

different. The fires were limited in number but attributed to arson with their ignition occurring under<br />

extreme fire weather conditions. They were large fires and again exceeded the capacity <strong>of</strong> the firefighting<br />

resources to contain them; however, almost all the destruction happened in the first 12 hours,<br />

Some 173 persons died and about 2,000 houses destroyed.<br />

HVP Plantations lost around 2000ha <strong>of</strong> plantations in 2006. However, due to the nature <strong>of</strong> the fires,<br />

the company could use its fire-fighting resources strategically and was thus able to avoid greater loss.<br />

On the day <strong>of</strong> February 7 th 2009, when HVP lost about 14,000 ha <strong>of</strong> plantations, the company could do<br />

little to avoid loss under horrendous conditions. On subsequent days, only relatively minor losses were<br />

incurred due to the success <strong>of</strong> suppression action. In addition to the plantation area, some 7,800 ha <strong>of</strong><br />

managed native forest was also burnt.<br />

Many HVP plantations are widely dispersed around the northern and southern foothills <strong>of</strong> the Alpine<br />

Region. The sheer size <strong>of</strong> the 2006 fires resulted in different Company plantations coming under<br />

threat over an extended period and this placed great pressure on staff and resources. With what seems<br />

to be an increasing regime <strong>of</strong> large fires in Victoria, the risks to Company resources and the strain on<br />

available staff are a matter <strong>of</strong> great Company concern.<br />

COMPANY RECOVERY<br />

These fire events have created massive trauma to affected Victorian communities. They also caused a<br />

massive disruption for HVP. For a three week period almost all staff were diverted for some or all <strong>of</strong><br />

their time on fire related tasks. With tremendous support from contractors and customers, no<br />

processing plants ran out <strong>of</strong> logs and this continuity was critical factor in ensuring rural communities<br />

were not subject to further trauma through loss <strong>of</strong> income.<br />

The four main components <strong>of</strong> the company’s recovery phase are considered to be: salvage, reestablishment,<br />

staff recovery, and legal processes.<br />

Salvage<br />

Approximately 13,000ha <strong>of</strong> the burnt plantation was too young, or unviable, for salvage operations.<br />

This land will be cleared for replanting in subsequent years. The 3,500ha remainder were stands older<br />

than 20 years <strong>of</strong> age (2,500 ha s<strong>of</strong>twood and 1000 ha hardwood) which might, depending on the fire<br />

damage, produce logs <strong>of</strong> merchantable quality.<br />

These older stands were classified as either:<br />

• Conventional salvage i.e. logs with burnt bark but no burnt fibre and thus could be supplied to<br />

existing customers; burnt butt logs go to export);<br />

• Black salvage (logs containing burnt fibre, which could not be delivered to existing customers<br />

but are deemed economic to harvest for sale “out <strong>of</strong> specification” or to export markets);<br />

• Uneconomic salvage (not scheduled for harvest due to terrain, stand quality or market<br />

constraints). Where applicable some will be harvested to reduce silviculture costs for reestablishment.<br />

S<strong>of</strong>twood (Pinus radiata) salvage was planned to be mostly completed by June 2009 due to the<br />

anticipated progressive reduction in log quality caused by blue stain, however some salvage operations<br />

will continue until there are no viable markets remaining. The ability to salvage many stands is<br />

reduced due to depressed market conditions and a current market surplus <strong>of</strong> chips and roundwood.<br />

Opportunities for bio-fuels remain open and are being investigated.<br />

Hardwood salvage is likely to continue for at least 2 years. Priorities to harvest Eucalyptus globulus<br />

and E.regnans stands will be primarily set by seeking to return the most pr<strong>of</strong>itable sites to production<br />

quickly.<br />

Indications are that around 600,000 tonnes <strong>of</strong> s<strong>of</strong>twood and 400,000 tones <strong>of</strong> hardwood may be<br />

salvaged.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 307<br />

Re-establishment<br />

The HVP Board stated after the fires that “HVP is committed to the Victorian timber industry and its<br />

people” and “We will … be replanting as quickly as possible.” A strategy has been developed to<br />

replant the fire affected areas over 5 years. This represents a 40% increase in the establishment<br />

program and requires a build-up <strong>of</strong> production in Company nurseries to around 10 million plants per<br />

year. The strategy for replanting is primarily based on such considerations as spreading the workload<br />

across the regions, weed management, pr<strong>of</strong>itability <strong>of</strong> sites, and capacity to access appropriate planting<br />

stock.<br />

The Company’s research program will need to be reviewed as 330 ha <strong>of</strong> research plots were lost.<br />

Staff<br />

Many HVP employees had their personal homes at risk at some stage during the fires and left the fire<br />

line in order to protect their own assets and families.<br />

A significant number <strong>of</strong> employees have been personally touched by the fires, through the loss <strong>of</strong> life<br />

or loss <strong>of</strong> house <strong>of</strong> a friend or family member, or through the tragic scenes they witnessed. There is<br />

likely to be an ongoing impact on the mental and emotional welfare <strong>of</strong> staff.<br />

HVP employed a counsellor to tour the regions and provide group and one-on-one counselling where<br />

appropriate. Reflecting the broader affected community, there were a number <strong>of</strong> issues <strong>of</strong> concern<br />

such as insomnia, exhaustion and fatigue; survivor guilt; anxiety about future fires; and anxiety about<br />

the future <strong>of</strong> their employment given the significant loss. It is expected that some <strong>of</strong> these issues will<br />

linger and ongoing counselling service is <strong>of</strong>fered for employees and their families.<br />

Legal<br />

A number <strong>of</strong> legal processes arise from such major events. Because <strong>of</strong> the extensive loss <strong>of</strong> resource,<br />

force majeure clauses in a number <strong>of</strong> contracts had to be reviewed, due to a possible inability to meet<br />

contract commitments. The operation <strong>of</strong> these clauses can be quite complex in long term contracts<br />

where the impacts may not be felt for a number <strong>of</strong> years. This is new ground for HVP.<br />

An arrest has been made and a prosecution is pending for the deliberate lighting <strong>of</strong> the Churchill Fire<br />

which started on HVP land, and the Victorian Police have sought company information to assist their<br />

investigative process. Evidence tendered to coronial enquiries will inevitably involve HVP.<br />

Documentation, maps, ground and aerial photos have been collated in anticipation <strong>of</strong> enquiries which<br />

may continue for a number <strong>of</strong> years.<br />

The Victorian State Government established a Royal Commission into the Bushfires on 16 February<br />

with very broad terms <strong>of</strong> reference, and instructions to deliver an interim report by 17 August 2009,<br />

and a final report by 31 July 2010. HVP lodged a submission and will follow the deliberations closely<br />

in anticipation <strong>of</strong> changes to fire prevention and suppression measures in Victoria which may have<br />

implications for the company.<br />

Two fires which caused plantation loss to HVP, the Kilmore East and the Beechworth are alleged to<br />

have been ignited by powerlines in the electricity distribution network. There is likely to be a class<br />

action and other damages claims in response to these fires.<br />

HVP must manage these legal processes while ensuring they do not become a distraction from the<br />

company business.<br />

LESSONS<br />

The extended drought in southern <strong>Australia</strong> continues to require review and re-assessment <strong>of</strong> the risk<br />

<strong>of</strong> fire to the plantation industry and the management <strong>of</strong> that risk. There have been a number <strong>of</strong><br />

lessons:<br />

1. The longer and more intense fire seasons provide greater stress and less time for fire<br />

suppression personnel to recover. This season has shown that fire fighting forces can still<br />

perform at a high level despite signs <strong>of</strong> fire season fatigue, and they continue to provide a<br />

strong commitment and focus on fire suppression. However given the increasing burden both


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 308<br />

physically and emotionally, innovative schemes must be created to reduce the pressure and<br />

intrusion on personal lives over such a long summer in order to retain skills and experience.<br />

2. Fire-fighting has inherent dangers, but with few fire-related lost time injuries this past<br />

summer, HVP is confident that well trained and experienced staff and crews can minimise the<br />

safety risks while maintaining an aggressive and effective approach to fire control.<br />

3. The extended drought has raised the cost <strong>of</strong> fire risk management. Suppression costs to save<br />

forest plantations have increased due to larger fires burning for longer periods, and smaller<br />

fires being more threatening. In the drier conditions, the focus for fire suppression extends<br />

further afield from the Company’s forest boundary, with more fires being attended as a result.<br />

Fire prevention costs increase as additional on-ground maintenance works are programmed<br />

and additional measures are put in place, such as the two first-attack helicopters on standby for<br />

key summer periods.<br />

4. Fuel reduction burning has been shown again to have great value for managing wildfire.<br />

There are good examples <strong>of</strong> fuel reduction burning carried out in recent years by the<br />

Department <strong>of</strong> Sustainability and Environment which were instrumental in containing key<br />

sectors <strong>of</strong> the Beechworth fire and in turn saving plantation assets.<br />

5. Good strategy development very early in the fire event is a key factor in containing a fire and<br />

thereby protecting all assets. In recognising this, HVP places a liaison person in the Incident<br />

Control Centre to ensure strategy relating to HVP assets and the deployment <strong>of</strong> HVP firefighting<br />

resources are aimed at containment <strong>of</strong> the sections <strong>of</strong> fire relevant to the minimisation<br />

<strong>of</strong> plantation loss. Fire suppression resources at these early stages <strong>of</strong> large fires can <strong>of</strong>ten be<br />

diverted to the immediate protection <strong>of</strong> houses and buildings while insufficient focus is placed<br />

on medium term strategy development for fire containment.<br />

6. Working within the fire control structures and the Incident Action Plan <strong>of</strong> the responsible fire<br />

authorities is critical for ensuring Company fire operations are both legally protected, safe and<br />

effectively co-ordinated. To work independently in a potentially dangerous environment<br />

would be an unacceptable risk, and would not attract the mutual support required to achieve<br />

the plantation protection objective.<br />

CONCLUSION<br />

Extended drought creates a higher, more sustained, level <strong>of</strong> fire risk which must be managed as it<br />

places a strain on fire-fighting personnel beyond previous norms. A high level <strong>of</strong> co-operation is<br />

required with the responsible fire authorities together with an aggressive strategic approach to fire<br />

prevention and control by the forest owner to ensure that timber resources remain available to<br />

processors. Growers and processors are significant employers in the rural and regional communities<br />

in which they are based and days such as February 7 th 2009 are destructive and distracting for both the<br />

industry and their communities. Effective recovery plans are crucial to return community and industry<br />

confidence.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 309<br />

SILVICULTURE FOR A CLIMATE OF CHANGE<br />

David Doley 1<br />

ABSTRACT<br />

Climatic changes within the life span <strong>of</strong> a tree may lead to species extinctions from their<br />

present environments. Withdrawal <strong>of</strong> silviculture from native forests follows from the<br />

assumption that natural processes will achieve the desired conditions. Benign neglect<br />

will not ensure success in a changed environment. The nature and extent <strong>of</strong> physical<br />

changes and the biological attributes <strong>of</strong> species must be understood in order to apply<br />

adaptive measures. Most plant species will grow satisfactorily beyond their natural<br />

distribution ranges provided various growth stages are facilitated and physical factors<br />

and competition are managed appropriately. Silviculture can and should be used to<br />

maintain species in desired locations and to introduce them to suitable new locations.<br />

Native forest silviculture is complex and our understanding <strong>of</strong> most species is limited,<br />

but we must attempt to overcome the effects <strong>of</strong> past disturbances and realise the many<br />

benefits that forests can provide.<br />

INTRODUCTION<br />

Climate may be interpreted broadly as the total environment or domain within which something exists.<br />

Changes in the forest domain include the meteorological climate, the physical situation <strong>of</strong> forests in<br />

the landscape, economics and societal expectations. All these factors impact on the ways in which<br />

forests function, how they could be managed and how they will actually be managed. The problem is<br />

complex but it can be addressed by examining the threats and identifying the conditions that are<br />

needed to secure the stability <strong>of</strong> the forest. These are the objectives <strong>of</strong> silviculture in a climate <strong>of</strong><br />

change.<br />

This review will show that the maintenance <strong>of</strong> healthy and productive forests in <strong>Australia</strong> is a complex<br />

task, <strong>of</strong>ten undertaken in the absence <strong>of</strong> detailed information about the species involved, and<br />

sometimes without a clear resolution <strong>of</strong> the objectives that may be held for a particular area <strong>of</strong> forest.<br />

If forests have experienced no human influence and they exist in a stable physical environment, then<br />

over the lifespan <strong>of</strong> a tree (say 100 to 1,000 years), it can be expected that the required combination <strong>of</strong><br />

conditions will result in sufficient regeneration events to replace the species and that growing season<br />

conditions will enable the trees to develop to their genetic potential.<br />

Most <strong>Australia</strong>n forests have been impacted to some degree by humans and are not necessarily in a<br />

state <strong>of</strong> ecological equilibrium. One view <strong>of</strong> forest land management is that human intervention should<br />

be minimised, whereupon natural processes will restore the forest structures and functions that existed<br />

before the arrival <strong>of</strong> European settlers, the ‘benign neglect’ described by Brown (1996).<br />

Another view is that disturbances to the forest have caused changes that may not be rectified by<br />

natural processes. These conditions may apply where humans have altered the forest or the<br />

surrounding environments, where exotic or pest species have been introduced, or when rapid and<br />

systematic climate change is occurring. Sadly, almost all <strong>of</strong> the native forests <strong>of</strong> <strong>Australia</strong> have<br />

suffered these disturbances to some extent, as well as the changes that have followed commercial<br />

timber removals. If natural processes cannot ensure the recovery <strong>of</strong> forests to the pre-European<br />

settlement condition, continued intervention may be essential to ensure that future changes in the<br />

forest are in the directions that will result in the pre-European or some other desired forest condition<br />

(Doley 1991).<br />

It does not take long to realise that both views <strong>of</strong> the responses <strong>of</strong> forests to disturbance may be<br />

correct. Small human-induced changes within a large area <strong>of</strong> forest may have less impact than natural<br />

disturbance events such as fire and cyclones. Extensive or radical human-induced changes are likely to<br />

1 The University <strong>of</strong> Queensland, Centre for Mined Land Rehabilitation, St Lucia Qld 4072, <strong>Australia</strong>


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 310<br />

have greater effects than natural disturbance, especially if the characteristics <strong>of</strong> the surrounding lands<br />

have been altered.<br />

Here is the dilemma. How can forests be maintained in a constant condition while the surrounding<br />

environment is changing? If constant forest conditions are desired while the forests and their<br />

environments are not at equilibrium, then we can not turn our backs on continuing management inputs.<br />

Silvicultural techniques must be accepted by the community if they are to be applied for the long term.<br />

In the past, forest authorities were left to their own devices in state forests, but increasing demands for<br />

public participation in land use determinations resulted in the questioning <strong>of</strong> forest practices and the<br />

transfer <strong>of</strong> substantial areas <strong>of</strong> former commercial forest to reserves (Dovers 2003). At about the same<br />

time, administrators adopted a market approach to resource management, so that the new objectives no<br />

longer accommodated some <strong>of</strong> the silvicultural practices <strong>of</strong> the earlier times (Dovers 2003). Despite<br />

the commercial pressures to simplify procedures and disperse skills, there has been progress towards<br />

the development <strong>of</strong> silvicultural systems, ranging in Tasmania from ‘benign neglect’ <strong>of</strong> large areas<br />

reserved for conservation to intensive management <strong>of</strong> forests in locations more suited to production<br />

forestry (Hickey and Brown 2003).<br />

It is important to focus on the aspects <strong>of</strong> forest functioning that may require or may benefit from<br />

intervention, especially in a changing environment. To do that, we must appreciate what is changing,<br />

how the changes might affect forests, whether the changes can be mitigated and what might be done to<br />

enhance adaptation to changes that cannot be avoided.<br />

THE NATURE OF ENVIRONMENTAL CHANGE<br />

The present climate and physical situation <strong>of</strong> the forest can be described with some certainty and these<br />

parameters are widely used to predict the growth rates <strong>of</strong> growth <strong>of</strong> trees and the rates <strong>of</strong> production <strong>of</strong><br />

dry matter or wood in forests. Because <strong>of</strong> the recorded and projected changes in climate in recent<br />

times (IPCC 2007), it is appropriate to consider whether the changes in the production functions for<br />

forest goods and services call for changes in the silvicultural practices in <strong>Australia</strong>n forests.<br />

Projected climatic changes include increasing atmospheric concentrations <strong>of</strong> carbon dioxide and other<br />

greenhouse gases, increasing mean temperature and, depending on the locality, decreasing or<br />

increasing mean precipitation, increasing seasonality <strong>of</strong> precipitation, more extreme weather events<br />

(IPCC 2007). Throughout the world, the concentration <strong>of</strong> carbon dioxide in the air is increasing by<br />

about 1.9 ppm yr -1 and the recent increases have been greater than those <strong>of</strong> earlier decades (IPCC<br />

2007). For eastern <strong>Australia</strong>, the rate <strong>of</strong> increase in mean surface temperature is predicted to be about<br />

0.25 o C per decade, resulting in mean summer temperatures rising by between 0.6 and 1.5 o C across<br />

<strong>Australia</strong> by 2030 (CSIRO 2009) (Figure 1a). At the same time there has been a general poleward shift<br />

in the high pressure systems, resulting in longer periods for which dry air flows over the <strong>Australia</strong>n<br />

continent, leading to reduced winter rainfall and hotter summer weather with higher extreme<br />

temperatures. Slight decreases in annual precipitation occurred over eastern <strong>Australia</strong> between 1979<br />

and 2005, but there were slight increases over much <strong>of</strong> western <strong>Australia</strong> (IPCC 2007). In contrast,<br />

CSIRO (2009) predictions to 2030 indicate less rainfall in the west <strong>of</strong> the continent, with little change<br />

on parts <strong>of</strong> the east coast and in the north (Figure 1b).Mean annual relative humidity is predicted to<br />

decrease in the west and south <strong>of</strong> <strong>Australia</strong> (Figure 1c), with an associated increase <strong>of</strong> up to 4% in<br />

predicted potential evapotranspiration rates by 2030 (CSIRO 2009) (Figure 1d).<br />

These and other physical changes impact on biological processes that affect forest growth. For<br />

example, a change in mean annual temperature <strong>of</strong> 1 o C is equivalent to a latitudinal shift towards the<br />

equator <strong>of</strong> approximately 2 o (Charles-Edwards et al. 1986). This rate <strong>of</strong> change means that, for the<br />

condition shown in Figure 1a, a given temperature regime will be found up to 3 o latitude<br />

(approximately 330 km) farther from the equator after 40 years. At the same time, an increase in<br />

temperature <strong>of</strong> 1 o C is equivalent to an increase in altitude associated with a given environment <strong>of</strong><br />

approximately 100 m. This would suggest that altitudinal temperature boundaries in <strong>Australia</strong> could<br />

rise by up to 150 m between 1990 and 2030. What might this mean for the functioning <strong>of</strong> trees or<br />

other types <strong>of</strong> plants in the forest?


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 311<br />

(a) (b)<br />

(c) (d)<br />

Figure 1. Predictions <strong>of</strong> mean changes in (a) summer temperature ( o C), (b) annual rainfall (%),<br />

(c) annual relative humidity (%) and (d) potential evapotranspiration (%) for<br />

<strong>Australia</strong> between 1990 and 2030, assuming a high greenhouse gas emission scenario.<br />

From CSIRO (2009).<br />

RESPONSES TO ENVIRONMENTAL CHANGE<br />

Biological responses<br />

For plant species, the responses <strong>of</strong> many functions to changing temperature can be described by a<br />

curve with a broad optimum and limiting upper and lower temperatures at which the function ceases or<br />

the plant dies. For example, Battaglia et al. (1996) showed that in summer, Eucalyptus nitens leaves<br />

function almost uniformly between about 10 and 30 o C, but in winter the optimum temperature is<br />

between 10 and 15 o C (Figure 2). In contrast, E. globulus showed an optimum temperature close to 20<br />

o o<br />

C in spring and summer, and between 10 and 15 C during winter. It would appear that Eucalyptus<br />

nitens might be more suited to higher environmental temperatures than E. globulus. This characteristic<br />

has not prevented selected genotypes <strong>of</strong> E. globulus from being planted far beyond its original range.<br />

Slatyer (1977) showed that when snow gum (Eucalyptus pauciflora) plants were grown at different<br />

temperatures, their optimum temperatures for photosynthesis varied in a predictable manner. Over<br />

temperature regimes from about 6 to 30 o C there was substantial but not complete adjustment <strong>of</strong> the<br />

optimum temperature (Figure 3). It is interesting that the summer sunny day maximum temperature for<br />

each site studied by Slatyer and Morrow (1977) was lower than the optimum temperature for the<br />

provenance. Does this mean that snow gum would grow faster at lower altitude than in its present


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 312<br />

range? The prediction from these photosynthesis measurements alone would be ‘yes’, but there is<br />

more to growth than photosynthesis.<br />

Net photosynthesis (μmol/m 2 /s)<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 10 20 30 40<br />

Temeprature ( o C)<br />

E. globulus spring<br />

E. globulus summer<br />

E. globulus winter<br />

E. nitens summer<br />

E. nitens winter<br />

Figure 2. Variation in rates <strong>of</strong> photosynthesis with temperature <strong>of</strong> plantation-grown Eucalyptus<br />

globulus and E. nitens leaves during different seasons in Tasmania. Redrawn from<br />

Battaglia et al. (1996).<br />

Optimum temperature ( o C)<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Waste Point 915 m<br />

Daners Gap 1645 m<br />

T opt = 0.27T gr + 20.64<br />

T opt = 0.37T gr + 15.02<br />

0 10 20 30 40<br />

Mean growth temperature ( o C)<br />

Figure 3. Relationship between the optimum temperature for photosynthesis in two provenances<br />

<strong>of</strong> snow gum (Eucalyptus pauciflora) from different altitudes and mean growth regime<br />

temperature in a phytotron. The intersection <strong>of</strong> the regression line for a provenance<br />

and the 1:1 line for optimum temperature vs. growth temperature is defined as the<br />

preferred temperature. The arrows above and below the regression lines indicate the<br />

sunny summer day maximum air temperatures for the two locations. Redrawn from<br />

Slatyer (1977).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 313<br />

Temperature response functions <strong>of</strong> tree growth are difficult to organise and measure at the field scale.<br />

However, growth can be predicted with reasonable precision by models based on the responses <strong>of</strong><br />

physiological processes such as photosynthesis and transpiration. Several <strong>of</strong> these models have been<br />

developed in <strong>Australia</strong>, by Landsberg and Waring (1987), McMurtrie et al. (1992), Kirschbaum (1999)<br />

and Battaglia et al. (2004).<br />

Considering all the factors associated with likely changes in temperature and CO2 concentrations,<br />

Kirschbaum (2000) concluded that an increase in temperature <strong>of</strong> 2 o C would cause small changes in<br />

wood production, but larger and more variable changes would accompany an increase in CO2<br />

concentration from 350 to 700 ppm. By comparison with seasonal variations, the temperature<br />

differences <strong>of</strong> 1 or 2 o C being considered in connection with climate change are small and there may<br />

be very effective adaptation to the new regimes (Kirschbaum 2000). However, increasing mean<br />

temperature is associated with an increased tendency for water loss (Figure 1), leading to stomatal<br />

closure and a reduction in photosynthesis.<br />

At the local scale, comprehensive climate modelling has predicted that the areas available for many<br />

tropical rainforest plant communities will contract because they are associated with certain altitudinal<br />

bands within the wet tropics <strong>of</strong> Queensland (Hilbert et al. 2001). The changes in forest environment<br />

are predicted to impact on the habitats for threatened fauna such as the golden bower bird (Prionodura<br />

newtonia) to an extent that may cause the extinction <strong>of</strong> the species (Hilbert et al. 2004).<br />

These habitat contractions are associated chiefly with changes in precipitation, including the frequency<br />

<strong>of</strong> cloud or mist at higher altitudes, which has been identified as one factor influencing the distribution<br />

<strong>of</strong> mammal species (Kanowski et al. 2001). An increase in air temperature <strong>of</strong> 1 o C will raise the<br />

altitude <strong>of</strong> the cloud base by close to 100 m. This lifting <strong>of</strong> the cloud base could have much more<br />

dramatic consequences for species survival than the associated change in temperature. On the Main<br />

Range <strong>of</strong> south-east Queensland, throughfall resulting from the interception <strong>of</strong> cloud by a large tree<br />

crown increased the annual precipitation beneath that tree by more than one-quarter over the rainfall in<br />

the open (Hutley et al. 1997).<br />

The distribution <strong>of</strong> this cloud interception through the dry season could be especially critical to the<br />

survival <strong>of</strong> drought-sensitive species in marginal rainforest environments. Climate-change induced<br />

movement <strong>of</strong> distribution limits <strong>of</strong> European beech (Fagus sylvatica) has been described for the upper<br />

temperature and dry limits (Jump et al. 2006). Both small saplings and large trees from Central Europe<br />

(where droughts are uncommon) have been shown to be sensitive to water availability and relative<br />

humidity (Jump et al. 2006, Lendzion and Euschner 2008) while in Greece populations exposed to<br />

more frequent droughts were not so affected (Fotelli et al. 2009). This finding emphasises the<br />

conclusions from <strong>Australia</strong>n studies that ecotypic variation is important in determining the adaptability<br />

<strong>of</strong> eucalypt species to environmental change (Ladiges 1974, Li et al. 2000).<br />

Defining the management objectives<br />

This catalogue <strong>of</strong> threats might indicate that catastrophic changes to forest communities are imminent.<br />

Much <strong>of</strong> the concern relating to environmental change, whether natural or anthropogenic, is associated<br />

with its likely impact on biodiversity. While predictable changes could result in many adverse effects,<br />

including species extinctions, it does not follow that the threats cannot be mitigated by carefully<br />

planned human intervention (Jones et al. 2007). For forests, the intervention could be in the form <strong>of</strong><br />

silvicultural practices that assist in the maintenance <strong>of</strong> species in desired areas, the introduction <strong>of</strong><br />

species to areas where they may be expected to survive in future and in the maintenance <strong>of</strong> forest<br />

structures that may be crucial to the survival <strong>of</strong> dependent species.<br />

The maintenance <strong>of</strong> biodiversity is commonly the principal function <strong>of</strong> native forest reserves, and it is<br />

an important function <strong>of</strong> native forests managed for commercial timber production. Even if it were<br />

possible to resolve the constantly varying objectives for native forest management, simple and cheap<br />

procedures for securing and maintaining regeneration <strong>of</strong> particular species may no longer be adequate.<br />

Ferguson (1996) observed that most forest management activities are based on perceptions <strong>of</strong> market<br />

value. The monetary valuation <strong>of</strong> conservation is extremely difficult and a satisfactory valuation<br />

mechanism has eluded managers up to this time. Despite these difficulties, we are confronted with<br />

changing conditions that call for planning and action at the physical level so that management options


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 314<br />

are still available in 50 or 100 years’ time. One <strong>of</strong> the critical decisions in relation to forest<br />

management for both production and conservation is a determination <strong>of</strong> where species may be able to<br />

exist and under what conditions, that is, their ecological niche.<br />

The ecological niche is the environmental space occupied by a species (Austin and Smith 1989) and it<br />

has two forms. The fundamental niche is described as the range <strong>of</strong> physical conditions within which<br />

the species can survive. The realised niche is the actual space occupied by the species. These two<br />

niches may be identical for species in extreme environments containing very few individuals, but for<br />

most species they are not congruent and the optimum environment for the realised niche may not even<br />

be close to the optimum for the fundamental niche (Austin and Smith 1989). This difference can be<br />

seen in a species such as Pinus radiata, which has a restricted natural range (realised niche) but has<br />

been planted successfully across a much wider environmental range (fundamental niche).<br />

Figure 4 presents a hypothetical example <strong>of</strong> how two species with the same optimum environmental<br />

condition but different ranges and maximum values for some attribute might be distributed if that<br />

attribute was critical for their survival and performance. It shows that natural distributions do not<br />

necessarily indicate whether a species could survive in an environment that differs slightly from its<br />

existing environment or if management makes the new environment available to that species.<br />

Relative performance<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0 5 10 15 20<br />

Environmental value<br />

Figure 4. A hypothetical example <strong>of</strong> differences between fundamental ecological niches (1F and<br />

2F) and realised ecological niches (1R and 2R) for two species with identical optimum<br />

conditions for an environmental value but different ranges and maximum values for<br />

some attribute (e.g. photosynthesis or growth). Relative performance is the amount <strong>of</strong><br />

an attribute or the rate <strong>of</strong> a process for a species.<br />

Realising the biological potential for change<br />

From the point <strong>of</strong> view <strong>of</strong> determining whether a species will adapt to a new environment, growth<br />

models can predict the rates <strong>of</strong> dry matter accumulation under conditions that are not represented by<br />

the present geographical range <strong>of</strong> a species. This capacity is most important for indicating the likely<br />

fate <strong>of</strong> species with very restricted distributions. For example, Hughes et al. (1996) found that 210<br />

Eucalyptus species had narrow ecological niches as indicated by a range <strong>of</strong> mean annual temperature<br />

<strong>of</strong> less than 1 o C. Although a narrow range and a small population might indicate limited capacity for<br />

evolutionary change (Skelly et al. 2007), this does not necessarily mean that the physiological<br />

tolerance <strong>of</strong> a species is limited to exactly the location <strong>of</strong> their present occurrence because the<br />

distribution may be limited by edaphic requirements or by competition from other species.<br />

Kirschbaum (2000) showed that conservative assumptions about the extent <strong>of</strong> adaptation <strong>of</strong><br />

photosynthesis to changing temperature and the response <strong>of</strong> a species to water deficits can be applied<br />

to growth models in order to predict the potential range <strong>of</strong> a species.<br />

Nevertheless, restricted species distributions may mean that the natural rates <strong>of</strong> dispersion are slower<br />

than the rates <strong>of</strong> environmental translation through climate change (Lindenmayer and Franklin 2002;<br />

1F<br />

2F<br />

1R<br />

2R


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 315<br />

IPCC 2007). As a result, the preferred thermal environment for the species may not occur at a<br />

contiguous location with suitable nutrient or water availabilities and the future existence <strong>of</strong> the species<br />

may be threatened if it is not able to adapt rapidly to environmental change or cannot be relocated to a<br />

suitable new environment.<br />

One aspect <strong>of</strong> the ecological niche may be illustrated by the requirements for seed germination. Weed<br />

species can <strong>of</strong>ten germinate speedily and completely over a wide temperature range, whereas species<br />

from a forest understorey may have slower germination over a much more limited temperature range.<br />

However, once germinated and established, the plant may be able to grow successfully and reproduce<br />

under a different or a wider range <strong>of</strong> conditions. Therefore, the realised niche for a species may be<br />

manipulated to ensure the survival <strong>of</strong> the species or to enhance the yield <strong>of</strong> a desired product. This is<br />

the basis <strong>of</strong> artificial regeneration practices and they should be considered for conservation as well as<br />

production objectives.<br />

The physical requirements for growth, based on the environmental constraints on photosynthesis,<br />

reflect the fundamental niche <strong>of</strong> the species (Booth et al. 1988). Typically, plants have a broad<br />

tolerance <strong>of</strong> temperature and light regimes although some tropical species are sensitive to temperatures<br />

below about 10 o C and species from humid and shaded environments are subjected to photoinhibition<br />

and heat stress when exposed to high light and dry atmospheres. Compared with these extreme<br />

changes, climate change is predicted to cause small changes in average temperature. Therefore, it is<br />

the likely occurrence <strong>of</strong> extreme conditions <strong>of</strong> high temperature and low humidity that are most likely<br />

to result in damage to plants.<br />

One aspect <strong>of</strong> plant growth that may be sensitive to relatively small environmental changes is<br />

flowering and seed production. In several important forest species, seed set is not uniform from year to<br />

year and the triggers for mast seeding are not fully understood. One factor that is important is the<br />

accumulation <strong>of</strong> sufficient reserves, but there is good evidence in hoop pine (Araucaria cunninghamii)<br />

that rainfall and seasonal variation in temperature are also important in cone initiation. Seed<br />

production <strong>of</strong> hoop pine at coastal or humid locations in Queensland (e.g. Imbil) is sporadic and<br />

generally limited, but in the more continental environment near the inland limit <strong>of</strong> the species<br />

distribution (Yarraman) seed set is more regular and prolific. As a result, the more productive seed<br />

orchards are located at Yarraman rather than at Imbil (Nikles 1996). Despite these limitations in seed<br />

production, hoop pine grows naturally over a wide latitudinal and climatic range. Therefore, it is<br />

possible to establish the species successfully despite some barriers to seed production. The critical<br />

issue is that natural regeneration is not relied upon in plantations and the same approach could be<br />

applied to the regeneration <strong>of</strong> hoop pine and other desired species in natural forests.<br />

Abundant natural regeneration <strong>of</strong> a species may occur rarely. If the life span <strong>of</strong> a tree is 1000 years,<br />

then theoretically an individual needs to be replaced only once in 1000 years. However, the probability<br />

<strong>of</strong> any tree surviving for 1000 years is very small and it might be necessary for millions <strong>of</strong> seeds to<br />

produce hundreds <strong>of</strong> thousands <strong>of</strong> seedlings and hundreds <strong>of</strong> young trees in order to overcome the<br />

risks <strong>of</strong> individual deaths.<br />

If seedlings are placed deliberately and then managed to minimise the risk <strong>of</strong> loss, the number <strong>of</strong><br />

individuals that need to be introduced to a site can be reduced dramatically, possibly to ten instead <strong>of</strong><br />

thousands. The management <strong>of</strong> the seedlings – their placement and tending – then becomes the key to<br />

successful regeneration.<br />

Enrichment planting in native forest is not undertaken very <strong>of</strong>ten, principally because <strong>of</strong> the cost and<br />

because the established vegetation may prevent the newly planted seedling from obtaining the required<br />

resources for growth. Shading and water shortages are the principal recognised causes <strong>of</strong> death in<br />

newly planted seedlings but in rainforests, burial under litter or total removal by foraging birds and<br />

animals (Song 2007) may be the most frequent causes <strong>of</strong> death. For high-value plants, the use <strong>of</strong> tree<br />

guards may be necessary to secure seedling establishment.<br />

SILVICULTURE FOR CHANGE MANAGEMENT<br />

With some notable exceptions, the recovery plans for many endangered species do not consider<br />

extending their geographic ranges and could be described more accurately as triage documents. If<br />

small populations are considered to be expanding from a few individuals and are therefore new, then


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 316<br />

the environmental range <strong>of</strong> sites at which they might be expected to survive may be greater than that<br />

<strong>of</strong> the sites <strong>of</strong> their present occurrence. If dispersal <strong>of</strong> a species was removed as a restriction to its<br />

spread, introductions <strong>of</strong> species to potentially suitable habitats would be possible. The resulting<br />

increase in the number <strong>of</strong> habitats should reduce the risk <strong>of</strong> catastrophic loss in one habitat.<br />

Introductions <strong>of</strong> species to national parks may not be desired, but there should be less concern over the<br />

introduction <strong>of</strong> species into managed forests. In this way, production forests could greatly enhance the<br />

opportunities for biological conservation. If the small populations <strong>of</strong> plants are regarded as old and<br />

contracting (e.g. Wollemia nobilis and Idiospermum australiense), then it is even more appropriate to<br />

consider ex situ conservation.<br />

The effort to understand the biology <strong>of</strong> Wollemi pine (Wollemia nobilis) (NSW Department <strong>of</strong><br />

Environment and Conservation 2006) and to capitalise on its rarity is an example <strong>of</strong> commercial<br />

conservation that follows in the footsteps <strong>of</strong> the plant hunters <strong>of</strong> the eighteenth century who carried<br />

unusual specimens back to Europe for cultivation in garden environments far different from the<br />

original habitats <strong>of</strong> the species. Should Wollemia be re-introduced into the wild? Why not? It has<br />

already been introduced deliberately to areas far beyond its present natural range.<br />

Idiospermum australiense (ribbonwood), one <strong>of</strong> the more primitive angiosperms, is a rainforest<br />

canopy tree with a large and poisonous seed that is dispersed only by gravity. It occurs naturally along<br />

a few coastal creeks in the wet tropics <strong>of</strong> Queensland. If left to its own devices, this species is bound<br />

for extinction as dispersal occurs only seawards and sea levels are predicted to rise. Should this<br />

species be allowed to become extinct, or should it be relocated to higher positions in the landscape?<br />

Humankind has a poor record <strong>of</strong> just leaving things alone and Idiospermum australiense has already<br />

been introduced to a number <strong>of</strong> locations outside its natural range. One such planting, by then-<br />

Governor-General Bill Hayden, commemorated the initiation <strong>of</strong> the Cooperative Research Centre for<br />

Tropical Rainforest Ecology and Management. This action may be taken as giving vice-regal<br />

endorsement <strong>of</strong> species translocation in the interests <strong>of</strong> conservation. It is a small step to consider<br />

management <strong>of</strong> established plants for the same end. Despite this, there is no recovery plan in place or<br />

in preparation for Idiospermum australiense.<br />

Clearly, for each species there needs to be a comprehensive understanding <strong>of</strong> all <strong>of</strong> the conditions<br />

associated with its success in different environments. If a new environment conforms to the central<br />

portion <strong>of</strong> the fundamental niche or even if it resembles the former realised niche, then manipulation<br />

<strong>of</strong> other stresses (competition) in the new environment could enable the species to survive and<br />

possibly flourish. Silviculture is the manipulation <strong>of</strong> a forest in order to secure the optimal<br />

development <strong>of</strong> desired individuals <strong>of</strong> desired species. The application <strong>of</strong> silviculture to a long list <strong>of</strong><br />

species is much more complex than plantation silviculture <strong>of</strong> a single species, but it is not impossible.<br />

Rainforest silviculture must consider individual trees and their immediate environment and decisions<br />

must be made about favouring one individual over another. While this is requires skill and is timeconsuming<br />

and therefore expensive, it can be done. If endangered species are accorded a high value,<br />

then it follows that the resources necessary for their survival in environments appropriate to their<br />

fundamental niches must be available.<br />

Most plant species will grow satisfactorily beyond their natural distribution ranges, which may be<br />

constrained more by the complex interactions <strong>of</strong> climate and soil on the processes <strong>of</strong> reproduction and<br />

natural establishment, rather than on vegetative growth. The ability to cultivate species is already<br />

highly developed and this knowledge can be applied to the maintenance <strong>of</strong> species in desired<br />

locations, or the introduction <strong>of</strong> species to new locations in order to ensure their continuation on earth.<br />

We are <strong>of</strong>ten told that the speed <strong>of</strong> environmental change in the last 50 years has been equal to that<br />

normally experienced over thousands <strong>of</strong> years (Alley et al. 2003). If that is so, then the processes <strong>of</strong><br />

translocation and establishment <strong>of</strong> species and their cultivation, even in natural areas, need to be<br />

accelerated accordingly. This is the time for further and resolute intervention, to restore stability to our<br />

forest environments.<br />

Hickey and Brown (2003) described a range <strong>of</strong> silvicultural techniques that had been tested in the<br />

mountain ash forests <strong>of</strong> Tasmania. Identification <strong>of</strong> the appropriate mixture <strong>of</strong> techniques and scapes<br />

<strong>of</strong> application is critical for the optimisation <strong>of</strong> the different and <strong>of</strong>ten conflicting outputs that are<br />

expected from these forests. There has been a move from large-scale radical and uniform treatments <strong>of</strong>


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 317<br />

felling and regeneration areas to a variety <strong>of</strong> opening size, proportion and characteristics <strong>of</strong> retained<br />

trees and other vegetation types. One important factor in these treatments is the need for skilled design<br />

and execution <strong>of</strong> the working plans, which represents a need for pr<strong>of</strong>essionals who understand the<br />

needs <strong>of</strong> the tree species as well as those <strong>of</strong> forest-dwelling fauna.<br />

The need for intervention in Tasmanian national parks was indicated by Marsden-Smedley and<br />

Kirkpatrick (2000) so that fire regimes more closely resembled those <strong>of</strong> pre-European times. Hickey<br />

and Brown (2003) considered it essential to attack bushfires threatening the very fire sensitive King<br />

Billy pine (Athrotaxis selaginoides) in Tasmanian reserves. Following such fires, further intervention<br />

in the form <strong>of</strong> artificial planting may also be warranted in order to secure regeneration <strong>of</strong> the species<br />

when natural processes have failed.<br />

In Victoria, the maintenance <strong>of</strong> habitats for Leadbeater’s possum (Gymnobelideus leaadbeateri) has<br />

been a critical issue in the montane ash forest type. Squire (1993), Attiwill (1994a, b, 1995) and<br />

Lindenmayer and Franklin (2002) have outlined in some detail practices that might contribute to the<br />

maintenance <strong>of</strong> commercially productive forests and robust populations <strong>of</strong> Leadbeater’s possum.<br />

Suggested measures include the retention <strong>of</strong> certain densities <strong>of</strong> trees that are suitable as nesting sites<br />

now or some time in the future, the provision <strong>of</strong> varied age structures in the forest and the avoidance<br />

<strong>of</strong> salvage operations following disasters. Unanimity regarding the desired or required measures is yet<br />

to be reached but the existence <strong>of</strong> varied silvicultural approaches in the montane ash forests indicates<br />

that the essential experimentation is occurring.<br />

As the size <strong>of</strong> a forest management unit decreases, the intensity <strong>of</strong> management is likely to increase<br />

and the appropriateness <strong>of</strong> the use <strong>of</strong> heavy machinery diminishes. For many <strong>of</strong> silvicultural<br />

operations, hand working may be the most effective method. Unfortunately, cost structures in<br />

<strong>Australia</strong> almost prohibit manual tending <strong>of</strong> the forest, so there may be nothing between the extremes<br />

<strong>of</strong> extensive disruption by machinery and neglect.<br />

FORESTS FOR CARBON STORAGE<br />

Climate change effects and responses in forests cannot be divorced from the issue <strong>of</strong> carbon storage.<br />

Forests are viewed as providing the simplest land-based mechanism for ameliorating the effects <strong>of</strong><br />

carbon dioxide released from fossil fuel burning (Garnaut 2008). Increased carbon accumulation in<br />

forests is expressed mainly in the form <strong>of</strong> an increase in the quantity <strong>of</strong> wood that is retained in the<br />

forest. An important assumption in the modelling <strong>of</strong> the contribution <strong>of</strong> forests is the price <strong>of</strong> carbon,<br />

which is predicted to increase from about $20 per tonne CO2 equivalent (CO2-e) (in 2005 <strong>Australia</strong>n<br />

dollars) at present to nearly $600 per tonne by 2100.<br />

For the first half <strong>of</strong> the century, however, the price <strong>of</strong> carbon is not expected to exceed $100 per tonne<br />

so we should concentrate on the near rather than the very distant view. Polglaise et al. (2008) estimate<br />

that the wooded lands <strong>of</strong> <strong>Australia</strong> could remove up to 143 Mt CO2-e per year and Mackey et al.<br />

(2008) estimated that the south-eastern eucalypt forests alone could remove 136 Mt CO2-e per year at<br />

an average rate <strong>of</strong> approximately 9 t CO2-e ha -1 yr -1 . At a price <strong>of</strong> $20 per tonne or $180 ha -1 yr -1 , it<br />

could be more pr<strong>of</strong>itable than some present activities. In lower rainfall environments, the potential<br />

returns from carbon accumulation will be more modest and the risk <strong>of</strong> loss may be much greater.<br />

Mackey et al. (2008) argued that if the eucalypt forests <strong>of</strong> south-eastern <strong>Australia</strong> were left<br />

undisturbed they would accumulate much more carbon in ‘green’ or non-tree biomass than has been<br />

considered in the National Carbon Accounting System (Kesteven et al. 2004). For the wettest eucalypt<br />

forests that are likely to burn only rarely, it may be appropriate to use these much greater estimates <strong>of</strong><br />

above-ground carbon storage capacity (up to 640 tonne ha -1 ).<br />

However, in the aftermath <strong>of</strong> the Victorian fires <strong>of</strong> 2009, the durability <strong>of</strong> non-tree biomass in much <strong>of</strong><br />

the south-eastern <strong>Australia</strong>n eucalypt forests must be questioned, especially as these forests are<br />

expected to become drier in future (IPCC2007, CSIRO 2009). It may be more prudent to aspire to<br />

lesser goals and to accept that forest management for the accumulation <strong>of</strong> wood is a lower risk<br />

alternative to the maximisation <strong>of</strong> non-wood carbon stocks in forests.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 318<br />

CONCLUSIONS<br />

Society accepts readily the need to intervene in the management <strong>of</strong> species that are on the brink <strong>of</strong><br />

extinction. This is demonstrated by the number <strong>of</strong> recovery plans that have been approved throughout<br />

the country (examples). For species that occur in forested landscapes, it would seem to be more<br />

effective to develop comprehensive landscape (forest) management plans that take into account the<br />

needs <strong>of</strong> all the species. <strong>Foresters</strong> have been pilloried for concentrating in the past only on the<br />

commercial timber species in their forests, but modern management plans are already far more<br />

cognisant <strong>of</strong> the needs for conservation <strong>of</strong> other species, even though almost nothing may be known<br />

about the requirements for their regeneration and growth. These considerations persuade me that<br />

continuing, but appropriate, management is required for the whole forest estate. The specific<br />

objectives <strong>of</strong> management may vary from place to place and from time to time but the overall<br />

objective must be to maintain the landscape in the most stable and healthy condition possible, subject<br />

to the ability to pay for the work.<br />

National parks may be left alone so that the effects <strong>of</strong> climate change can be observed with minimal<br />

interference from humans. However, there is an important role for forests outside these reserves to act<br />

as repositories for threatened species as well as sites where resources with commercial value can be<br />

produced sustainably. Hickey and Brown (2003) indicated several directions in which more complex<br />

management systems were being applied to the forests <strong>of</strong> Tasmania. This complexity increases the<br />

price <strong>of</strong> any product, but if forest management is to be tested against market based criteria, then all<br />

products and services should be accorded appropriate values. Unfortunately, the timber market is<br />

global, and <strong>Australia</strong>n products compete against those from countries that do not have the same<br />

environmental goals. If an open market for wood products is to be retained, then it may not be possible<br />

for the costs <strong>of</strong> matrix silviculture to be met from the sale <strong>of</strong> produce and they must be made up from<br />

other sources such as general taxation.<br />

Despite the extent <strong>of</strong> disturbance caused by timber harvesting, commercial forestry does not appear to<br />

have been associated with the loss <strong>of</strong> plant species. Therefore, it can be argued that past silvicultural<br />

practices have not been totally detrimental to biodiversity in forest areas and the considered<br />

application <strong>of</strong> silviculture should provide many opportunities for the cultivation <strong>of</strong> species that may be<br />

threatened with extinction due to climate change.<br />

Silviculture has a crucial role to play in the maintenance <strong>of</strong> forested landscapes in a changing<br />

environment. Far from withdrawing from the forest, there should now be a redoubling <strong>of</strong> effort to<br />

overcome the effects <strong>of</strong> past disturbances and to optimise the many benefits that forests can provide.<br />

REFERENCES<br />

Alley, R.B., Marotzke, J., Nordhaus, W.D., Overpeck, J.T., Peteet, D.M., Pielke, R.A., Pierrehumbert, R.T.,<br />

Rhines, P.B., Stocker, T.F., Talley, L.D. and Wallace, J.M.(2003) Abrupt climate change. Science 299,<br />

2005-2010.<br />

Attiwill, P.W. (1994a) The disturbance <strong>of</strong> forest ecosystems: the ecological basis for conservative management.<br />

Forest Ecology and Management 63, 247-300.<br />

Attiwill, P.W. (1994b) Ecological disturbance and the conservative management <strong>of</strong> eucalypt forests in <strong>Australia</strong>.<br />

Forest Ecology and Management 63, 301-346.<br />

Attiwill, P.W. (1995) Managing Leadbeater’s possum in the mountain ash forests <strong>of</strong> Victoria, <strong>Australia</strong> - Reply.<br />

Forest Ecology and Management 74, 233-237.<br />

Austin,M.P. and Smith, T.M. (1989) A new model for the continuum concept. Vegetatio 83, 35-47.<br />

Battaglia, M.C., Beadle, C. and Loughhead, S. (1996) Photosynthetic temperature response <strong>of</strong> Eucalyptus<br />

globulus and Eucalyptus nitens. Tree Physiology 16, 81-89.<br />

Battaglia, M., Sands, P., White, D. and Mummery, D. (2004) CABALA: a linked carbon, water and nitrogen<br />

model <strong>of</strong> forest growth for silvicultural decision support. Forest Ecology and Management 193, 251-282.<br />

Booth, T.H., Nix, H.A., Hutchinson, M.F. and Jovanovic, T. (1988) Niche analysis and tree species introduction<br />

Forest Ecology and Management 23, 47-59.<br />

Brown, M.J. (1996) Benign neglect and active management in Tasmania’s forests: a dynamic balance <strong>of</strong><br />

ecological collapse? Forest Ecology and Management 85, 279-289.<br />

Charles-Edwards, C.A., Doley, D. and Rimmington, G.M. (1986) Modelling Plant Growth and Development.<br />

Academic Press, Sydney.<br />

CSIRO (2009) Climate Change <strong>Australia</strong> at http://www.climatechangeinaustralia.gov.au/nattemp1.php.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 319<br />

Doley, D. (1991) The C.T. White Lecture for 1990: The future forest: management or curation. The Queensland<br />

Naturalist 30, 100-109.<br />

Dovers, S. (2003) Are forests different as a policy challenge? In D.B. Lindenmayer and J.F. Franklin (eds.)<br />

Towards Forest Sustainability, pp 15-29. CSIRO Publishing, Melbourne.<br />

Ferguson, I.S. (1996) Sustainable Forest Management. Oxford University Press, Melbourne.<br />

Fotelli, M.N., Nahm, M., Radoglou, K.m Rennenberg, H., Halvyopoulos, G. and Matzarakis, A., (2009).<br />

Seasonal and interannual ecophysiological responses <strong>of</strong> beech (Fagus sylvatica) at its south-eastern<br />

distribution limit in Europe. Forest Ecology and Management 257, 1157–1164.<br />

Garnaut, R. (2008) The Garnaut Climate Change Review: Final Report. Cambridge University Press, Port<br />

Melbourne.<br />

Hickey, J.E. and Brown, M.J. (2003) Towards ecological forestry in Tasmania. In D.B. Lindenmayer and J.F.<br />

Franklin (eds.) Towards Forest Sustainability, pp 31-46. CSIRO Publishing, Melbourne.<br />

Hilbert, D.W., Bradford, M., Parker, T. and Westcott, D.A. (2004) Golden bowerbird (Prionodura newtonia)<br />

habitat in past, present and future climates: predicted extinction <strong>of</strong> a vertebrate in tropical highlands due<br />

to global warming. Biological Conservation 116, 367-377.<br />

Hilbert, D.W., Osterndorf, B. and Hopkins, M.S. (2001) Sensitivity <strong>of</strong> tropical forests to climate change in the<br />

humid tropics <strong>of</strong> north Queensland. Austral Ecology 26, 590-603.<br />

Hughes, L., Cawsey, E.M. and Westoby, M. (1996) Climatic range sizes <strong>of</strong> Eucalyptus species in relation to<br />

future climate change. Global Ecology and Biogeography Letters 5, 23-29.<br />

Hutley, L.B., Doley, D., Yates D. J. and Boonsaner A.. 1997. Water balance <strong>of</strong> an <strong>Australia</strong>n subtropical<br />

rainforest at altitude: the ecological and physiological significance <strong>of</strong> intercepted cloud and fog.<br />

<strong>Australia</strong>n Journal <strong>of</strong> Botany 45, 311-329.<br />

IPCC 2007 Climate Change 2007: The Physical Science Basis. Contribution <strong>of</strong> Working Group I to the Fourth<br />

Assessment Report <strong>of</strong> the Intergovernmental Panel on Climate Change [Solomon, S., D. Qin, M.<br />

Manning, Z. Chen, M. Marquis, K.B. Averyt, M. Tignor and H.L. Miller (eds.)] Cambridge University<br />

Press, Cambridge, UK, 996 pp.<br />

Jones, R.N., Dettmann, P., Park, G., Rogers, M. and White, T. (2007) The relationship between adaptation and<br />

mitigation in managing climate change risks: a regional response from North Central Victoria, <strong>Australia</strong>.<br />

Mitigation and Adaptation Strategies for Global Change 12, 685-712.<br />

Jump, A., Hunt, J.M. and Penuelas, J. (2006) Rapid climate change-related growth decline at the southern range<br />

edge <strong>of</strong> Fagus sylvatica. Climate Change Biology 12, 2163–2174.<br />

Kanowski, J., Hopkins, M.S., Marsh, H. and Winter, J. (2001) Ecological correlates <strong>of</strong> folivore abundance in<br />

north Queensland rainforests. Wildlife Research 28, 1-8.<br />

Kesteven, J., Landsberg, J. and URS <strong>Australia</strong> (2004) Developing a National Forest Productivity Model.<br />

National Carbon Accounting System - Technical Report No. 23. <strong>Australia</strong>n Greenhouse Office, Canberra.<br />

Kirschbaum, M.U.F. (1999) CenW, a forest growth model with linked carbon, energy, nutrient and water cycles.<br />

Ecological Modelling 181, 17-59.<br />

Kirschbaum, M.U.F. (2000) Forest growth and species distribution in a changing climate. Tree Physiology 20,<br />

309-322.<br />

Ladiges, P.Y. (1974) Variation in drought tolerance in Eucalyptus viminalis Labill. <strong>Australia</strong>n Journal <strong>of</strong> Botany<br />

22, 489-500.<br />

Landsberg, J.J. and Waring, R.H. (1997) A generalised model <strong>of</strong> forest productivity using simplified concepts <strong>of</strong><br />

radiation-use efficiency, carbon balance and partitioning. Forest Ecology and Management 95, 209-228.<br />

Lendzion, J. and Leuschner, C. (2008). Growth <strong>of</strong> European beech (Fagus sylvatica L.) saplings is limited by<br />

elevated atmospheric vapour pressure deficits. Forest Ecology and Management 256, 648–655.<br />

Li, C., Berninger, F., Koskela, J. and Sonninen, E. (2000) Drought responses <strong>of</strong> Eucalyptus microtheca<br />

provenances depend on seasonality <strong>of</strong> rainfall in their place <strong>of</strong> origin. <strong>Australia</strong>n Journal <strong>of</strong> Plant<br />

Physiology 27, 231–238.<br />

Lindenmayer, D.B. and Franklin, J.F. (2002) Conserving Forest Biodiversity: A Comprehensive Multiscale<br />

Approach. Island Press, Washington.<br />

Mackey, B.G., Keith, H., Berry, S.L. and Lindenmayer, D.B. (2008) Green Carbon: The role <strong>of</strong> natural forests in<br />

carbon storage. Part 1. A green carbon account <strong>of</strong> <strong>Australia</strong>’s south-eastern eucalypt forests, and policy<br />

implications. <strong>Australia</strong>n National University, Canberra.<br />

Marsden-Smedley, J.B. and Kirkpatrick, J.B. (2000) Fire management in Tasmania’s Wilderness World Heritage<br />

Area: ecosystem restoration using indigenous-style fire regimes? Ecological Management and Restoration<br />

1, 195-203.<br />

McMutrtrie, R.E., Comins, H.N., Kirschbaum, M.U.F.and Wang, Y.-P. (1992) Modifying existing forest growth<br />

models to take account <strong>of</strong> effects <strong>of</strong> elevated CO2. <strong>Australia</strong>n Journal <strong>of</strong> Botany 40, 657-677.<br />

Nikles, D.G. (1996) The first 50 years <strong>of</strong> the evolution <strong>of</strong> forest tree improvement in Queensland. In Dieters,<br />

M.J., Matheson, A.C., Nikles, D.G., Harwood, C.E. and Walker, S.M. (eds.) Tree Improvement for


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 320<br />

Sustainable Tropical Forestry. Proceedings <strong>of</strong> QFRI-IUFRO Conference, Caloundra, Queensland,<br />

<strong>Australia</strong>. 27 October-1 November 1996.<br />

NSW Department <strong>of</strong> Environment and Conservation (2007) Wollemi Pine (Wollemia nobilis) Recovery Plan.<br />

Department <strong>of</strong> Environment and Conservation, Hurstville, New South Wales.<br />

Polglaise, P., Paul, K., Hawkins, C., Siggins, A., Turner, J., Booth, T., Crawford, D., Jovanovic, T., Hobbs, T.,<br />

Opie, K., Almeida, A. and Carter, J. (2008) Regional Opportunities for Agr<strong>of</strong>orestry Systems in<br />

<strong>Australia</strong>: A Report for the RIRDC/L&WA/FWPA/MDBC Joint Venture Agr<strong>of</strong>orestry Program, RIRDC,<br />

Barton, ACT.<br />

Skelly, D.K., Joseph, L.J., Possingham, H.P., Freidenburg, L.K., Farrugia, T.J., Kinnison, M.T. and Hendry, A.P.<br />

(2007) Evolutionary responses to climate change. Conservation Biology 21, 1353-1355.<br />

Slatyer, R.O. (1977) Altitudinal variation in the photosynthetic characteristics <strong>of</strong> snow gum, Eucalyptus<br />

pauciflora Sieb. ex Spreng. VI. Comparison <strong>of</strong> field and phytotron responses to growth temperature.<br />

<strong>Australia</strong>n Journal <strong>of</strong> Plant Physiology 4, 901-916.<br />

Slatyer, R.O. and Morrow, P.A. (1977) Altitudinal variation in the photosynthetic characteristics <strong>of</strong> snow gum,<br />

Eucalyptus pauciflora Sieb. ex Spreng. I. Seasonal changes under field conditions in the Snowy<br />

Mountains area <strong>of</strong> south-eastern <strong>Australia</strong>. <strong>Australia</strong>n Journal <strong>of</strong> Botany 25, 1-20.<br />

Song, G-Z.M. (2007) Patterns <strong>of</strong> Regeneration <strong>of</strong> Tree Species in a Subtropical Rainforest. PhD Thesis, The<br />

University <strong>of</strong> Queensland.<br />

Squire, R.O. (1993) The pr<strong>of</strong>essional challenge <strong>of</strong> balancing sustained wood production and ecosystem<br />

conservation in the native forests <strong>of</strong> south-eastern <strong>Australia</strong>. <strong>Australia</strong>n Forestry 56, 2237-248.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 321<br />

IMPROVING GREY GUMS TO SEQUESTER CARBON ON<br />

MARGINAL SITES IN SUBTROPICAL AUSTRALIA<br />

Paul Warburton 12 , Paul Macdonell 2 ,<br />

Jeremy Brawner 2 , John Huth 3 , David Lee 3,4<br />

ABSTRACT<br />

Expected decreasing rainfall in subtropical <strong>Australia</strong>, driven by elevated atmospheric<br />

carbon dioxide levels, will present a range <strong>of</strong> challenges for the forest plantation<br />

industry. One challenge for forestry research providers is identifying species with<br />

adaptive plasticity for use over a wide range <strong>of</strong> sites. The grey gums—Eucalyptus<br />

longirostrata, E.biturbinata (syn. punctata), E.propinqua, E.major, E.caniculata and<br />

E.grisea—have high density wood suitable for poles, sleepers, flooring and other uses<br />

that require strong, durable timber (Boland et al 2006). The performance <strong>of</strong> some <strong>of</strong><br />

these species in taxa and provenance trials indicates that they have potential as<br />

hardwood plantation species. As these are dense-wooded species (up to 1070 kg·m –3 )<br />

they will exhibit higher levels <strong>of</strong> carbon sequestration for the same piece size as lower<br />

density trees and may also have an important role as carbon sinks to mitigate the<br />

effects <strong>of</strong> elevated CO2 levels in a changing environment. Results from three year old<br />

trials, established across south-east and central Queensland and northern New South<br />

Wales by CSIRO and Queensland Primary Industries and Fisheries (QPIF), indicate<br />

there are significant differences between species and provide early indications <strong>of</strong> their<br />

potential in marginal environments (below 800 mm mean annual rainfall). Results<br />

indicate that E. longirostrata (14 provenances, 165 families) and E. biturbinata (five<br />

provenances, 67 families) are the most productiive <strong>of</strong> the grey gum species and that<br />

E. longirostrata has the greatest potential across most sites. We also report on the<br />

establishment <strong>of</strong> a new E. longirostrata breeding population as part <strong>of</strong> a collaborative<br />

improvement program to develop improved germplasm for timber production and for<br />

<strong>of</strong>fsetting carbon.<br />

INTRODUCTION<br />

Identifying species with adaptive plasticity for use over a wide range <strong>of</strong> sites, particularly those <strong>of</strong><br />

lower rainfall and poorer soils, is an immediate challenge for forestry research providers. It is<br />

predicted that elevated atmospheric carbon dioxide levels will result in increases in temperature<br />

between 1.0°C and 6.0°C and changes in annual rainfall from –35% to +10% in most regions <strong>of</strong><br />

<strong>Australia</strong> by 2070, which will result in marked changes in evapo-transpiration rates and soil moisture<br />

balances (CSIRO 2001, Henry et al. 2002).<br />

The grey gums—E.longirostrata, E.biturbinata (syn. punctata), E.propinqua, E.major, E.caniculata<br />

and E.grisea— are members <strong>of</strong> the sub-genus Symphyomyrtus. They have a natural distribution that<br />

primarily covers coastal regions from central New South Wales to south-east Queensland although<br />

populations <strong>of</strong> E.longirostrata and E.major can be found further west in Queensland in the Carnarvon<br />

Ranges and Blackdown Tablelands. These species occur on a variety <strong>of</strong> sites including hills, ridges<br />

and slopes and on moderately fertile to poor soils (Boland et al. 2006).<br />

It was reported that in 2004, commercial forestry in <strong>Australia</strong> accounted for 43.7 million tonnes <strong>of</strong><br />

carbon dioxide removed from the atmosphere (FWPRDC, date unknown). There are therefore<br />

opportunities for hardwood plantations to play a significant role in greenhouse abatement. Growth in<br />

trials established in <strong>Australia</strong> and South Africa indicates that some grey gum have excellent potential<br />

1<br />

Corresponding Author<br />

2<br />

CSIRO Plant Industry, Forest Biosciences, Queensland Biosciences Precinct, Brisbane, <strong>Australia</strong><br />

3<br />

Queensland Primary Industries and Fisheries, Horticulture and Forestry Science, Fraser Road, Gympie<br />

4<br />

University <strong>of</strong> Sunshine Coast, Sippy Downs. Queensland


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 322<br />

as hardwood plantations species. With high density (up to 1070 kg·m -3 in natural stands), hard and<br />

durable wood suitable for poles, sleepers, flooring and other uses requiring heavy grade timber, grey<br />

gums should play an important role as carbon sinks to help mitigate the effects <strong>of</strong> elevated carbon<br />

dioxide levels in a changing environment.<br />

This paper reports the overall performance <strong>of</strong> grey gum species in three year old trial established by<br />

CSIRO and QPI&F with closer investigation <strong>of</strong> growth data <strong>of</strong> E. longirostrata and E. biturbinata.<br />

MATERIALS AND METHODS<br />

Figure 1 (A-F) illustrates the geographical locations <strong>of</strong> the original seed sources established in all three<br />

trials, the trial locations and the natural distribution <strong>of</strong> the species included.<br />

Figure 1. Location <strong>of</strong> CSIRO and QPIF trials, species’ natural distribution and geographical<br />

locations <strong>of</strong> the original seed sources <strong>of</strong> A. E.longirostrata B. E.biturbinata/E.punctata,<br />

C. E.propinqua, D. E.major, E. E.moluccana, F. E.grisea.<br />

CSIRO trials<br />

Two grey gum species – provenance trials were established in May and September 2006 (Table 1).<br />

The first trial planted in May at Mt Alma (29 km west by south <strong>of</strong> Calliope) in central Queensland,<br />

contained 261 individual families from the five grey gum species. Gum-topped box (E.moluccana)<br />

and Dunn’s white gum (E.dunnii) were planted as controls in this trial. The trial was established in a<br />

randomised complete block design with four tree line plots replicated 5 times and a spacing <strong>of</strong> 4 m ×<br />

2.8 m to give a nominal stocking density <strong>of</strong> 990 stems per hectare. Fertiliser was applied to the site at<br />

the following rates; 54.3 kg/ha Amsul, 11.2 kg/ha muriate <strong>of</strong> potash, 34.4 kg/ha MAP, 5.1 kg/ha DAP,<br />

1.77 kg/ha zinc oxysulphate, 6.88 kg/ha boronate, 4.86 kg/ha copper oxysulphate. Trees were<br />

supplemented with 500ml saturated water-retaining gel in the root zone at time <strong>of</strong> planting. The Mt<br />

Alma trial site is a free-draining alluvial flat that had been planted with improved pasture prior to the<br />

trial establishment.<br />

A subset <strong>of</strong> one hundred and seventy-five <strong>of</strong> the seedlots from the Mt Alma trial were established in a<br />

second trial in September 2006 at Pagan’s Flat (47 km west-south-west <strong>of</strong> Casino) in northern New


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 323<br />

South Wales (Figure 1). This trial was established in a similar design as the Mt Alma trial with five<br />

tree line plots replicated 5 times and a spacing <strong>of</strong> 4.5 m × 2 m to give a stocking density <strong>of</strong> 1111 stems<br />

per hectare. Fertiliser was applied in the rip line at a depth <strong>of</strong> 40cm during site preparation at a rate <strong>of</strong><br />

100kg/ha <strong>of</strong> N:P:K:S (13:14:15:1). No additional treatment was applied. Pagan’s Flat is characterised<br />

by rolling low hills with a large alluvial plain associated with the Clarence River. The soil at this trial<br />

site is a very dark, strongly-structured clay loam over medium clay to a depth <strong>of</strong> 150–200 cm.<br />

Table 1. Treatment details <strong>of</strong> CSIRO grey gum trials, established at Mt Alma, Queensland and<br />

Pagan’s Flat, New South Wales, 2006.<br />

No. entries<br />

Species Provenance Latitude Longitude Mt Alma Pagan’s Flat<br />

Ballon (20899) 26° 27´ 150° 49´ 18 12<br />

Barakula (20404, 20898) 26° 19´ 150° 41´ 22 12<br />

Blackdown Tablelands (20008) 23° 46´ 149° 04´ 8 6<br />

Chinchilla (16008) 26° 22´ 150° 27´ 5 0<br />

Coominglah (a) (19312) 24° 48´ 150° 57´ 4 2<br />

Coominglah (b) (20464) 24° 49´ 150° 58´ 7 4<br />

Goodger (20943) 26° 38´ 151° 49´ 6 5<br />

Kroombit Tops (20954) 24° 20´ 150° 56´ 3 2<br />

Nanango (20942) 26° 38´ 152° 00´ 5 4<br />

E. longirostrata Starkvale Creek (20007) 25° 20´ 149° 15´ 8 5<br />

Chaelundi (19812) 29° 57´ 152° 22´ 6 4<br />

Girad (19809) 28° 58´ 152° 15´ 6 4<br />

Jimna (20405) 26° 37´ 152° 24´ 5 3<br />

Maryland (20406) 28° 28´ 152° 11´ 5 3<br />

E. biturbinata Mt Colliery (20407) 28° 17´ 152° 19´ 5 3<br />

Blackdown Tablelands (20009, 20491) 23° 45´ 149° 04´ 12 10<br />

Boxvale (20959) 25° 21´ 148° 38´ 3 2<br />

E. major Gympie (15603) 26° 18´ 152° 49´ 5 4<br />

Brooweena (20967) 33° 49´ 150° 23´ 5 2<br />

Blackdown Tablelands (20492) 23° 47´ 149° 04´ 5 4<br />

C<strong>of</strong>fs Harbour (15145) 30° 26´ 152° 58´ 7 6<br />

Goomeri (20949, 20950) 26° 13´ 152° 00´ 10 6<br />

Kin Kin (20969) 26° 14´ 152° 54´ 3 3<br />

Nanango (20961) 34° 45' 150° 11´ 4 3<br />

Taylors Arm (18674) 30° 47´ 152° 39´ 6 5<br />

E. propinqua Unumgar (20499) 28° 25´ 152° 42´ 6 5<br />

Curryall (20149) 32° 07´ 149° 53´ 6 3<br />

Kedumba Valley (19280) 33° 49´ 150° 23´ 7 4<br />

Nullo S.F. (19797) 32° 45´ 150° 12´ 6 3<br />

E. punctata Wingello (19352) 34° 45´ 150° 11´ 6 4<br />

Ballon (15877; n=20) 26° 27´ 150° 49´ 1 1<br />

Biloela (20951) 24° 15´ 150° 32´ 5 4<br />

Calliope (20484) 23° 57´ 151° 06´ 5 4<br />

Coominglah (20771) 24° 49´ 150° 58´ 6 4<br />

Cooyar (15221; n=10) 26° 52´ 152° 00´ 1 1<br />

Crediton S.F. (20010) 21° 15´ 148° 29´ 6 5<br />

Gunnawarra (20960) 18° 10´ 145° 17´ 3 2<br />

Long Mile Range Creek (20408) 29° 59´ 152° 35´ 5 3<br />

Monto (15225; n=8) 24° 50´ 150° 57´ 1 1<br />

Mt Garnett (20957) 17° 43´ 145° 02´ 5 4<br />

Running Creek (20955) 25° 54´ 152° 19´ 5 3<br />

Tairo (20965) 25° 45´ 152° 35´ 4 3<br />

Tomoulin (20958) 17° 32´ 145° 26´ 5 4<br />

E. moluccana Wondai S.F. (17752; n=3) 26° 19´ 151° 54´ 1 1<br />

South African SSO n.a n.a 1 0<br />

E. dunnii Yabbra S.F. (20850; n=30) 28° 35´ 152° 30´ 1 1<br />

Unknown 2 1


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 324<br />

QPIF trial<br />

Trial 519b grey gum species – provenance trial (Table 2) was established at Pechey (31 km north-east<br />

Toowoomba), Queensland (see Figure 1) by QPIF in February 2002 and contains 13 bulked<br />

provenance seedlots from three species. Five provenances <strong>of</strong> E. longirostrata (Barakula, Coominglah,<br />

Starkvale Creek, Chinchilla and Blackdown Tablelands) and five provenances <strong>of</strong> E.biturbinata<br />

(Chaleundi, Girad, Jimna, Maryland and Mt Colliery) had seedlots in-common with the CSIRO trials.<br />

The QPIF trial was established as an incomplete block design <strong>of</strong> 12 tree line-plots with three<br />

replications at a spacing <strong>of</strong> 4 m × 2 m, giving a planting density <strong>of</strong> 1250 stems per hectare. The trial<br />

was thinned to 300 stems per hectare at age four years. The trial was established in a plantation<br />

previously planted with Pinus patula. The soil is described as a red ferrosol with a pH <strong>of</strong> 5.5 in the A<br />

horizon. The site has an elevation <strong>of</strong> 720 m a.s.l.<br />

Table 2. Treatment details <strong>of</strong> QPIF grey gum trial 519b, established at Pechey, Qld, 2001.<br />

Species Provenance Latitude Longitude No. Parent trees<br />

E. biturbinata Chaelundi (19812) 29° 57´ 152° 22´ 10<br />

Girad (19809) 28° 58´ 152° 15´ 15<br />

Jimna (20405) 26° 37´ 152° 24´ 5<br />

Maryland (20406) 28° 28´ 152° 11´ 5<br />

Mt Colliery (20407) 28° 17´ 152° 19´ 5<br />

E. grisea Mt M<strong>of</strong>fat (19702) 25° 05´ 148° 07´ 10<br />

E .longirostrata Barakula (20404) 26° 19´ 150° 41´ 8<br />

Blackdown Tablelands (20008) 23° 46´ 149° 04´ 8<br />

Chinchilla (16008) 26° 22´ 150° 27´ 5<br />

Coominglah (a) (19312)) 24° 48´ 150° 57´ 5<br />

Coominglah (b) (20464) 24° 49´ 150° 58´ 7<br />

Monto (a) (15227) 24° 52´ 150° 58´ 27<br />

Monto (b) (9662) 24° 52´ 150° 57´ 6<br />

Starkvale Creek (20007) 25° 20´ 149° 15´ 9<br />

Climate<br />

Interpolated climate data for all trial sites, including temperature, rainfall and accumulated monthly<br />

vapour pressure deficit (VPD) are presented in Table 3.<br />

Table 3. Temperature and rainfall data for Alarm Creek, Pagan’s Flat and Pechey trial sites<br />

(source: Datadrill©, Queensland Dept <strong>of</strong> Natural Resources and Management, 2009)<br />

Trial<br />

Location Rainfall Temperature VP deficit<br />

Latitude Longitude<br />

M.A.R<br />

(long<br />

term)<br />

M.A.R<br />

(since<br />

planting)<br />

Mean<br />

maximum<br />

(hottest month)<br />

Average accrued<br />

monthly VPD<br />

(since planting)<br />

Mt Alma 24° 01´ 150° 56´ 865 675 31.7°C (Jan) 14.62 hPa<br />

Pagan’s Flat 28° 56´ 152° 34´ 780 1004 30.6°C (Jan) 11.75 hPa<br />

Pechey 27° 18´ 152° 04´ 897 746 27.7°C (Jan) 10.94 hPa<br />

Assessment<br />

The Mt Alma, Pagan’s Flat and Pechey 519b trials were assessed for survival, height and stem<br />

diameter at breast height over bark (DBHOB) at 32 months, 27 months and 36 months respectively.<br />

Due to different assessment ages across the trials, it was necessary to set an arbitrary age for<br />

comparison purposes. Therefore the assessments <strong>of</strong> Mt Alma, Pagan’s Flat and Pechey will be treated<br />

as one age class (nominally age three years). Trees with dead tops, or tops blown out and trees blown<br />

over were omitted from the growth calculations but were included for survival. Tree height and<br />

DBHOB measurements were used to calculate volume indexes (VI). As no volume equation is<br />

available for these species, the following formula was used:<br />

VI (m 3 ) = 1/3 × Height (m) × Basal Area (m 2 )<br />

The Giant Wood Moth (Endoxyla cinerea) was present at Mt Alma and the incidence <strong>of</strong> attack on<br />

trees in the trial was recorded. Trees were scored for the presence or absence <strong>of</strong> an entry or exit hole.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 325<br />

Analysis<br />

The data from the three grey gum trials were analysed separately to evaluate the performance <strong>of</strong><br />

species and provenances from south-east Queensland and northern New South Wales. Analysis <strong>of</strong><br />

variance <strong>of</strong> the data for each trial was undertaken using Genstat version 11.1. The statistical model<br />

used was;<br />

yijk = + Ri + Sj + Pjk + Eijkl<br />

where, μ= overall mean; Ri is the effect <strong>of</strong> the i th replicate; Sj is the effect <strong>of</strong> the j th species; Pjk is the<br />

effect <strong>of</strong> the k th provenance within the j th species and Eijkl is the random error associated with the ijkl th<br />

plot mean. Transformation (Log, Sqrt and arcsin) used on survival data but did not alter results <strong>of</strong> the<br />

analysis and therefore results <strong>of</strong> the analysis <strong>of</strong> non-transformed data are presented here. Post hoc<br />

Fischer’s Protected Least Significant Difference tests were applied to plot means if significant<br />

differences were found by the ANOVA.<br />

RESULTS<br />

Mean survival was good across all species and sites, ranging from 82.5% to 96% (Table 4). Only Mt<br />

Alma showed a significant difference in survival (P


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 326<br />

Based on their performance at the species level, provenance performance data <strong>of</strong> E.longirostrata and<br />

E.biturbinata was selected for separate analysis (Table 5). There was little difference in survival<br />

between provenances at all three sites and only Pechey exhibited any significant differences. The<br />

Chaelundi provenance <strong>of</strong> E. biturbinata had significantly lower survival (P


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 327<br />

DISCUSSION<br />

In <strong>Australia</strong>, the National Agriculture and Climate Change Action Plan (2006–2009) identifies forests<br />

as carbon sinks (Natural Resource Management Ministerial Council 2006). Schemes such as the<br />

Greenhouse Gas Reduction Scheme (GGAS) allow abatement certificates for sequestering carbon to<br />

be generated by plantation growers. Provision in the GGAS Carbon Sequestration Rule recognises<br />

permanent carbon storage whilst rotating harvests from plantations in the sequestration pool.<br />

Assuming that net carbon dioxide flux is proportional to the biomass stocks in a given area, there are<br />

clear opportunities for greenhouse gas abatement via plantations and examples <strong>of</strong> plantings for carbon<br />

sequestration can be found in many countries (Christie and Scholes 1995, Singh et al. 2000). The<br />

main hardwood species currently dominating solid wood plantations in subtropical <strong>Australia</strong> are<br />

Corymbia citriodora ssp. variegata, E.grandis x E.camaldulensis hybrids, E.cloeziana, E.dunnii and<br />

E.argophloia. Current expansion <strong>of</strong> plantations into new areas, expected changes in our climate and<br />

the lack <strong>of</strong> a species that would be a panacea for <strong>Australia</strong>n wood production drive the need for<br />

genetic diversity within the northern plantation estate. Numerous species trials in South Africa have<br />

shown that E.longirostrata has potential on wet sites in Zululand, ranking alongside E. grandis for<br />

growth and interest is now turning to its adaptability to drier sites (Gardner et al. 2007, Gardner and<br />

Little 2007). Lee et al. (2001; 2009) identified that this species has potential in Queensland and<br />

Henson et al. (2008) suggested E. longirostrata had potential for marginal sites in New South Wales.<br />

In this study, E.longirostrata and E.biturbinata consistently ranked well for growth, quantified as<br />

volume index (VI), against the other species established in these trials. E. biturbinata had a greater VI<br />

than E.longirostrata at Pechey although the difference was not significant. The best VI across all trials<br />

was shown by E.longirostrata at Mt Alma. At age three, the mean annual increment (MAI) for<br />

E.longirostrata was 7.7 m 3 /ha at Mt Alma and 4.3 m 3 /ha at Pagan’s Flat which was slightly lower than<br />

the top performer, E.dunnii (5.1 m 3 /ha). It should be noted however, that while E.dunnii performed<br />

better at Pagan’s Flat than the grey gums as a species, some E. longirostrata provenances (e.g. Ballon<br />

and Coominglah) had a similar VI at the provenance level. E.biturbinata achieved a MAI <strong>of</strong> 6.0 m 3 /ha<br />

at Pechey.<br />

At the time <strong>of</strong> planting, E.biturbinata and E.punctata were distinct species. Recent taxonomic<br />

changes have placed E. biturbinata in synonomy with E. punctata (CPBR 2006). We have retained<br />

the distinction for the purposes <strong>of</strong> this study but the affiliation between E.biturbinata and E.punctata is<br />

reflected in the fact that there was no significant difference in VI between the species at either Mt<br />

Alma or Pagan’s Flat. The slightly lower mean values for E.punctata may be due to the more southern<br />

origins <strong>of</strong> the seedlots with respect to the climatic differences with the trial sites. With the lowest<br />

VPD and mean hottest temperature and highest long-term MAR, the Pechey trial site appears to<br />

provide the most conducive conditions to good tree growth however, three-year growth was generally<br />

greatest at Mt Alma, followed by Pechey and Pagan’s Flat. This is despite Mt Alma receiving the<br />

lowest rainfall and reaching the highest mean maximum temperatures during the trial period. Gardner<br />

et al. (2007) reported the potential <strong>of</strong> E.longirostrata on productive wet sites and these results<br />

demonstrate that E.longirostrata also has potential on sites receiving


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 328<br />

for volume along with E. longirostrata Monto (a) and Coominglah (P


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 329<br />

are susceptible to damage by Giant Wood Moth. Early indications <strong>of</strong> differences in susceptibility<br />

between species and provenances within species require thorough investigation at more advanced<br />

ages. A large progeny trial has been established by CSIRO and QPIF to further improve the growth<br />

characteristics and wood properties <strong>of</strong> E. longirostrata. The objectives for this trial is to make<br />

selections <strong>of</strong> best individuals for breeding, the production <strong>of</strong> improved seed and selection <strong>of</strong> clonal<br />

germplasm for distribution to the commercial hardwood plantation industry.<br />

ACKNOWLEDGEMENTS<br />

We thank many staff from CSIRO (David Boden, Tim Vercoe and Darren Morrow) and at the<br />

Department <strong>of</strong> Employment, Economic Development and Innovation and its predecessor departments<br />

(Murray Johnson, Alan Ward, Tony Burridge and John Oostenbrink) for their assistance in<br />

establishing and measuring the trials and collating data. Thanks are also due to Integrated Tree<br />

Cropping and Great Southern Limited for their support and commitment in hosting the CSIRO trials.<br />

REFERENCES<br />

Boland, D.J., Brooker, M.I.H., Chippendale, G.M., Hall, N., Hyland, B.P.M., Johnston, R.D., Kleinig, D.A.,<br />

McDonald, M.W and Turner, J.D. (2006) Forest Trees <strong>of</strong> <strong>Australia</strong>. 5 th ed. CSIRO Publishing,<br />

Collingwood, Victoria, <strong>Australia</strong>. 736 p.<br />

Christie, S.I. and Scholes, R.J. (1995) Carbon storage in Eucalyptus and pine plantations in South Africa.<br />

Environmental Monitoring and Assessment 38:231–241<br />

CPBR 2006. Euclid – eucalypts <strong>of</strong> <strong>Australia</strong>, 3 rd edition. Centre for Plant Biodiversity and Research, Canberra.<br />

DVD.<br />

CSIRO (2001) www.dar.csiro.au/publications/projections2001.pdf<br />

Darrow, W.M. 1997 Eucalypt site-species trials in Zululand. ICFR Bulletin Series 3 – 97. i–26<br />

FWPRDC (Forest and Wood Products Research and Development Corporation) pub date unknown. Forests,<br />

Wood and <strong>Australia</strong>’s Carbon Balance.<br />

Gardner, R.A.W. (2001) Alternative Eucalypt species for Zululand: Seven year results <strong>of</strong> site : species<br />

interaction trials in the region. Southern African Forestry Journal. 190: 79-88<br />

Gardner, R.A.W. (2006) Early performance <strong>of</strong> promising cold-tolerant and sub-tropical eucalypt species in the<br />

warm temperate climate zone <strong>of</strong> KwaZulu-Natal. ICFR Bulletin Series 13/2006 i+ 21pp.<br />

Gardner, R.A.W., and Little, K.M. (2007). Investigating the commercial potential <strong>of</strong> alternative Eucalyptus and<br />

Corymbia species for northern, coastal Zululand. In IUFRO WG 2.08.03 Improvment and culture <strong>of</strong><br />

Eucalypts (Durban, Republic <strong>of</strong> South Africa: <strong>Institute</strong> for Commercial Forestry Research), pp. 1-14.<br />

Gardner, R.A.W., Little, K.M., Arbuthnot, A. (2007) Wood and fibre productivity potential <strong>of</strong> promising new<br />

eucalypt species for coastal Zululand, South Africa. <strong>Australia</strong>n Forestry. 70:37–47.<br />

Henry, B.K., Danaher, T. McKeon G.M. and Burrows, W.H. (2002) A review <strong>of</strong> the potential role <strong>of</strong><br />

greenhouse abatement in native vegetation management in Queensland’s rangelands. Rangelands Journal.<br />

24(1): 112–132.<br />

Henson, M., Smith, H.J. and Boyton, S. (2008) Eucalyptus longirostrata: a potential species for <strong>Australia</strong>’s<br />

tougher sites? New Zealand Journal <strong>of</strong> Forestry Science. 38(1): 227–238.<br />

Lawson, S. (2003) Susceptibility <strong>of</strong> eucalypt species to attack by longicorn beetles (Phoracantha spp.) in<br />

Queensland. Hardwoods Queensland Report No. 10. Queensland Forestry Research <strong>Institute</strong>, Agency for<br />

Food and Fibre Sciences, DPI. 11??pp<br />

Lee, D.J., Nikles, D.G., and Dickinson, G.R. (2001). Prospects <strong>of</strong> eucalypt species, including interspecific<br />

hybrids from South Africa, for hardwood plantations in marginal subtropical environments in Queensland,<br />

<strong>Australia</strong>. Southern African Forestry Journal 190, 89-94.<br />

Lee, D.J., Huth, J.R., Osborne, D.O., and Hogg, B.W. (2009). Selecting Hardwood Varieties for Fibre<br />

Production in Queensland's Subtropics. In Australasian Forest Genetics Conference (Fremantle, West<br />

<strong>Australia</strong>: Forest Products Commission, Western <strong>Australia</strong>).<br />

Natural Resource Management Ministerial Council (2006) National Agriculture and Climate Change Action<br />

Plan, 2006–2009. Department <strong>of</strong> Agriculture, Fisheries and Forestry. 14pp<br />

Singh, T.P., Varalakshmi, V. and Ahluwalia, S.K. (2000) Carbon sequestration through farm forestry: case<br />

from India. Indian Forester. 126(12): 1257–1264.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 330<br />

THE UTILISATION OF RED MAHOGANY FOR HIGH VALUE<br />

PLANTATION <strong>FORESTRY</strong> IN THE TROPICS<br />

Jeremy Brawner 1 2 , David Bush 3 , Paul Macdonell 2 ,<br />

David Boden 2 , Simon Potter 4 Paul Warburton 2 , and Paul Clegg 5<br />

ABSTRACT<br />

The expansion <strong>of</strong> red mahogany (Eucalyptus pellita) plantations in the north <strong>of</strong><br />

<strong>Australia</strong> has increased interest in its domestication. A review <strong>of</strong> an overseas breeding<br />

population was undertaken to enhance our current understanding <strong>of</strong> genetic parameters,<br />

which are essential for the development <strong>of</strong> advanced generation breeding programs.<br />

The large genetic differences between and within provenances, as well as the moderate<br />

heritability estimates for growth and form traits, imply that extensive tree breeding can<br />

create improved breeds and reduce variability in plantation forests. Our analysis<br />

suggests that breeding value predictions <strong>of</strong> parents tested in first generation<br />

provenance/progeny trials are less indicative <strong>of</strong> genetic merit than those from second<br />

generation trials. The low heritability estimate in the first generation relative to the<br />

second generation, and the reduced inter-generational correlations will influence the<br />

level <strong>of</strong> genetic gain that can be realised. The implications <strong>of</strong> these findings for the<br />

management <strong>of</strong> E. pellita breeding populations and the production <strong>of</strong> improved seed for<br />

plantation forestry in the tropics are discussed.<br />

INTRODUCTION<br />

Eucalyptus pellita is a forest tree <strong>of</strong> the humid and subhumid tropics, which occurs naturally in<br />

Southern New Guinea and Northern Queensland (Harwood 1998). Red mahogany is sympatric with<br />

Acacia mangium, A. crassicarpa, A. aulacocarpa, Xanthostemon sp., Stenocarpus sp., Syzygium sp.<br />

and Eucalyptus brassiana between rainforest and savannah woodland (Vercoe and McDonald 1991).<br />

The species has been grown in plantations to a limited extent in <strong>Australia</strong> with a recent expansion <strong>of</strong><br />

the plantation estate as Managed Investment Schemes move into the tropical north.<br />

A tree improvement program using a wide range <strong>of</strong> New Guinea and Queensland provenances<br />

established in replicated progeny trials was initiated by the Commonwealth Scientific and Industrial<br />

Research Organisation and the Queensland Department <strong>of</strong> Primary Industries and Fisheries in 1991<br />

and 1992. This program has led to the establishment <strong>of</strong> many second generation seedling seed<br />

orchards and the recent establishment <strong>of</strong> a third generation progeny trial in 2007.<br />

In these second generation trials, one year after planting the mean height <strong>of</strong> the second generation<br />

orchard families was 10 percent greater than the mean <strong>of</strong> natural-provenance controls (Harwood et al.<br />

1997). The species has also been established in collaborative trials in the Northern Territory <strong>of</strong><br />

<strong>Australia</strong> as well as across 22 sites in Brazil, in Dongmen, China, South Kalimantan, Indonesia and<br />

Sabah, Malaysia (Harwood 1998).<br />

1 Corresponding Author<br />

2 CSIRO Plant Industry, Forest Biosciences, Queensland Biosciences Precinct, Brisbane, Qld, <strong>Australia</strong><br />

3 CSIRO Plant Industry, Forest Biosciences, Yarralumla, ACT, <strong>Australia</strong><br />

4 CSIRO Plant Industry, Forest Biosciences, Clayton, Vic, <strong>Australia</strong><br />

5 Toba Pulp Lestari, Medan, Sumatra, Indonesia


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 331<br />

Figure 1. – Native range <strong>of</strong> Eucalyptus pellita<br />

Toba Pulp Lestari (TPL) is one <strong>of</strong> the smaller companies that form part <strong>of</strong> the APRIL group, which<br />

has a pulp and paper production focus. The company has several mills and produces pulp and paper<br />

for both the domestic (Indonesian) and export markets. TPL’s plantation forest estate is located<br />

around Lake Toba, Sumatra, Indonesia at elevations from 900 to nearly 2,100 metres above sea level.<br />

The soils are mostly volcanic ash from the Toba eruption which occurred most recently 50,000 years<br />

ago. The soil is classed locally as ‘Toba Tuff’, which is a typic Andisol <strong>of</strong> volcanic origin with<br />

characteristics <strong>of</strong> very high permeability, high Al content with resulting low N and P status. An<br />

annual rainfall <strong>of</strong> about 2,000 to 3,000 mm occurs throughout the year, with lower rainfall from May<br />

to August. Temperatures range from about 17-22 o C. The environment <strong>of</strong> the TPL plantation estate is<br />

dramatically different from that encountered in the lowlands <strong>of</strong> Sumatra, where APRIL has established<br />

extensive Acacia mangium plantations.<br />

The E. pellita breeding population is a major part <strong>of</strong> TPL’s eucalyptus tree improvement program,<br />

which has focused on increasing per hectare pulp yield. The company has invested significant<br />

resources on the genetic improvement <strong>of</strong> eucalypts for use in the high-elevation plantation estate and<br />

has advanced generation breeding populations for other eucalypt species (E. grandis and E. urophylla)<br />

at various stages <strong>of</strong> domestication. Although E. pellita may not be used extensively for reforestation<br />

as a pure species in the highlands <strong>of</strong> Sumatra, it has considerable value as: 1) a hybrid parent, 2) an<br />

alternative for Acacia mangium on the Sumatran non-peat soils <strong>of</strong> the lowlands and 3) a base<br />

population for TPL’s sister companies operating in other regions.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 332<br />

For a variety <strong>of</strong> reasons, interest in alternative species for the lowland Acacia mangium estate has<br />

increased recently. This has prompted the company’s tree improvement program to establish an E.<br />

pellita progeny trial and a range <strong>of</strong> other genetic material in the lowland estate <strong>of</strong> its sister company<br />

Riau Andalan Pulp and Paper (RAPP), which is located around the town <strong>of</strong> Pekanbaru in Sumatra.<br />

These tropical lowland environments are typically not suitable for plantations <strong>of</strong> eucalypt species due<br />

to very high pest and disease pressure (Harwood 1998). If trial and pilot scale plantations can cope<br />

with the disease pressure common to the wet lowlands and yield is satisfactory, the creation <strong>of</strong> an E.<br />

pellita based breed for the lowlands could greatly increase the utility <strong>of</strong> the species for operational<br />

reforestation across the tropics.<br />

MATERIALS AND METHODS<br />

The foundation <strong>of</strong> TPL’s E. pellita breeding program is a base population contained within a<br />

first generation progeny trial established in 1995. This trial was assessed to three years <strong>of</strong> age and<br />

each eight-tree row-plot was phenotypicaly thinned to leave the best tree and create a seedling seed<br />

orchard (Eldridge et al 1993). The resulting open-pollinated seedlots were collected from the best<br />

individuals in each family for second generation trials. One second generation progeny trial was<br />

established close to the first generation trial, and although these trials were planted several years apart,<br />

they have experienced very similar growing environments. Within the upland estate, the rainfall and<br />

temperature patterns are generally constant within and across years. The two upland field trials sites<br />

were selected to be relatively flat with homogenous, very deep and well-drained ‘Toba tuff’ soils. The<br />

first and second generation trials were established at a different spacing to reflect the changing<br />

operational practices <strong>of</strong> TPL with 1333 (2.5 X 3 metres) and 2222 (1.5 X 3 metres) trees per hectare<br />

planted respectively. Supplementary genetic material from the native range (19 families) was infused<br />

into the E. pellita base population at the establishment <strong>of</strong> the second generation upland and lowland<br />

trials.<br />

Additionally, a smaller second generation progeny trial was established in the lowlands <strong>of</strong><br />

Sumatra. This trial was established to evaluate the productivity <strong>of</strong> E. pellita in the lowlands and<br />

provide an indication <strong>of</strong> the stability <strong>of</strong> family performance across the contrasting lowland and upland<br />

environments. Based on previous experience and the findings <strong>of</strong> other research (Harwood et al 1997,<br />

Pegg and Wang 1994, Werren 1991) showing Queensland sources were less productive than New<br />

Guinea sources when established in the wet tropics. It has a very pronounced dry season)., the openpollinated<br />

families established in the lowland trial were exclusively from New Guinea. A seedlot that<br />

was a bulk <strong>of</strong> selected Queensland families identified in the first generation progeny trial was included<br />

to assess the merit <strong>of</strong> expanding the representation <strong>of</strong> Queensland germplasm in future trials.<br />

Table 1. Description <strong>of</strong> Eucalyptus pellita field trials established in Sumatra, Indonesia by<br />

Toba Pulp Lestari<br />

Trial Location<br />

EB1 EB15A EB15B<br />

Near Lake Toba Near Lake Toba Near Pekanbaru<br />

Highland<br />

Highland<br />

Lowland<br />

Generation First Second Second<br />

Date Established July 1995 February 2003 March 2003<br />

Soil Toba Tuff Toba Tuff Lowland clay<br />

Spacing 2.5 X 3 m 1.5 X 3 m 1.5 X 3 m<br />

Replications 8 8 8<br />

Design 8-tree row-plots 10-tree row-plots 10-tree row-plots<br />

Treatments 66 73 29<br />

DBH Assessed last 36 months 30 months 30 months<br />

Genetic material assessed in field trials<br />

The majority <strong>of</strong> the germplasm established in the three progeny trials <strong>of</strong> E. pellita were<br />

established by the TPL breeding program from seeds collected within the species native range for the<br />

first generation and from seeds collected within the thinned first-generation progeny trial for the<br />

second generation. Additionally, controls <strong>of</strong> locally produced clones and bulk seedlots were planted<br />

within these trials. Descriptive statistics were generated for each trial individually, using data sets that


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 333<br />

included all controls and native range E. pellita families. For these descriptive statistics, an additional<br />

country code labeled ‘TPL’ was included to congregate the various controls that were used in these<br />

trials. A description <strong>of</strong> the genetic material established and the data analyzed for each trial is included<br />

in Table 2.<br />

Table 2. Description <strong>of</strong> provenances and number <strong>of</strong> trees within each <strong>of</strong> the Eucalyptus pellita<br />

field trials established by Toba Pulp Lestari<br />

Source EB1 EB15A EB15B<br />

<strong>Australia</strong><br />

Abergowrie<br />

Bloomfield<br />

56<br />

16<br />

80<br />

80<br />

Cardwell 8 80<br />

Daintree 24 80<br />

Helenvale 64 240<br />

Kirrama Range 64 80<br />

Kuranda 1120 3240<br />

Tinaroo Creek 16<br />

Wonga – Daintree 64 80<br />

Indonesia<br />

Papua New Guinea<br />

Toba Pulp Lestari<br />

Bupul Muting 64 80 60<br />

Papua (Province <strong>of</strong> Indonesia) 312 860 670<br />

Keru to Kumbalusi 64 80 60<br />

Keru to Mata 56 60<br />

Serisa 56<br />

South <strong>of</strong> Kiriwo 16 80 60<br />

Tokwa 16 270 290<br />

Control Clone 1 80 80<br />

Control Clone 2 80 80<br />

Control E.grandis 24 60 60<br />

Control E.pellita Lowland Seed 70 180<br />

Control E.pellita QLD selects 80 60<br />

Control E.urophylla 8 60 60<br />

Total trees 2048 5760 1720<br />

Statistical Analyses<br />

Data were analyzed in two stages; descriptive statistics were generated to provide general information<br />

on the performance <strong>of</strong> the different material included in the trials prior to undertaking across site<br />

analyses to produce genetic parameters. Data cleaning, standardisation, pedigree file production and<br />

the generation <strong>of</strong> descriptive statistics within each trial were completed using SAS (SAS 1991). To<br />

generate descriptive statistics for each site, a mixed model including the following factors was used to<br />

produce least square means for normally distributed traits: overall trial mean, replication, plot within<br />

replication, country <strong>of</strong> origin, locality within country, country by replication, locality by replication<br />

interaction and residual error. For these single site analyses, all factors were fixed, with the exception<br />

<strong>of</strong> the plot within replication (family by replication) and the error terms.<br />

Estimates <strong>of</strong> genetic parameters for diameter growth at latest assessment<br />

A single pooled-site analysis was used to generate breeding value predictions and variance component<br />

estimates using the latest-age, pre-thinning diameter at breast height (DBH) data available for each<br />

trial. Data was standardized to have a mean <strong>of</strong> zero and a variance <strong>of</strong> one within each trial to remove<br />

scale effects that contribute to genotype by environment interactions and facilitate convergence <strong>of</strong> the<br />

Restricted Maximum Likelihood process. All controls, treatments that were not open pollinated<br />

seedlots <strong>of</strong> E. pellita, were removed to ease variance component estimation and clarify their genetic<br />

interpretation. An individual-tree model was implemented in ASREML with standard errors for<br />

genetic parameters calculated using the Taylor series expansion method within ASREML (Gilmour et<br />

al. 2005). For the pooled-site analysis, the following tri-variate mixed-effects model was used to


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 334<br />

obtain estimates <strong>of</strong> variance and covariance components for diameter growth at each <strong>of</strong> the three trial<br />

sites:<br />

y = Xb + Z p p + Zii<br />

+ e ,<br />

where y i is the vector <strong>of</strong> observations <strong>of</strong> the latest DBH assessment undertaken by TPL,<br />

b is the vector <strong>of</strong> fixed effects including overall mean, and replicate within trials and X is an<br />

incidence matrix relating the observations in y to the fixed effects in b , p is the vector <strong>of</strong> random<br />

2<br />

plots nested within replicate where a separate plot variance ( P) was estimated for each site and Zp is<br />

a known incidence matrix relating the observations in y to the plot effects, i is the vector <strong>of</strong> random<br />

individual tree effects ~MVN ( 0, G ⊗ A , where,<br />

)<br />

2 ⎡σ<br />

1<br />

⎢<br />

G = ⎢σ<br />

12<br />

⎢<br />

⎣σ<br />

13<br />

σ 12<br />

2<br />

σ 2<br />

σ 23<br />

σ ⎤ 13<br />

⎥<br />

σ 23 ⎥<br />

2<br />

σ ⎥<br />

3 ⎦<br />

A = the numerator relationship matrix and Zi is a known incidence matrix<br />

2 2 2<br />

relating the observations in y to effects in i, 1, 2 and 3 are the genetic<br />

variances for trial 1, 2, and 3 respectively with covariances between the trials<br />

on the <strong>of</strong>f-diagonal, e is random vector <strong>of</strong> residuals where a separate error<br />

term,<br />

2<br />

E, was estimated for each trial to account for heterogeneous residual errors between trials. It<br />

was assumed that there was no correlation <strong>of</strong> residuals or plot effects between trial sites.<br />

The mixed model equations were modified by including genetic groups to implicitly fit fixed effects<br />

for ‘provenances’ or ‘localities’ (Westell et al. 1988, Quass, 1988, Gilmour et al. 2005). The primary<br />

reason for using genetic groups in this analysis was to differentiate between the common environment<br />

(or pollen source) <strong>of</strong> the seed collected from the native range and seed collected from within the first<br />

generation trial. Each the 16 native range provenances included in the progeny trials were assigned to<br />

a different genetic group. Using volatile leaf oil composition, Doran et al (1995) concluded there was<br />

no clear evidence to support the subdivision <strong>of</strong> E. pellita from northern Queensland, Cape York and<br />

southern New Guinea into sub-populations. In a separate study using isozymes House and Bell (1996)<br />

determined there was no support for the separation <strong>of</strong> New Guinea and Cape York populations as a<br />

distinct grouping from the more southern populations <strong>of</strong> E. pellita. It was therefore decided to create<br />

‘finer’ genetic groups based on provenances or localities <strong>of</strong> seed collection as reported by TPL rather<br />

than group the populations arbitrarily by region or country. An additional genetic group was created to<br />

represent the common pollen parent <strong>of</strong> all seedlots collected from within the first generation for<br />

establishment in the second generation trials. For comparative purposes, a pedigree file assigning<br />

groups back to the provenance <strong>of</strong> origin was also used to determine the effect <strong>of</strong> this new genetic<br />

group on genetic parameter estimates.<br />

Individual-tree narrow-sense heritability and inter-site genetic correlation estimates for diameter<br />

growth at the latest age <strong>of</strong> assessment were taken from the single pooled-site analyses following<br />

Falconer and Mackay (1996):<br />

V<br />

2<br />

2 A ˆ<br />

σ G<br />

h = ≈<br />

is the estimated individual-tree narrow-sense heritability from the<br />

2 2 2<br />

VP<br />

σ G + σ P + σ E<br />

pooled-site analysis, where VA is and estimate <strong>of</strong> the additive genetic variance, VP is and estimate <strong>of</strong><br />

2 2 2 2 2<br />

the phenotypic variance, G is the sum <strong>of</strong> the genetic variance estimates ( 1, 2, 3), P is the<br />

sum <strong>of</strong> the within plot variance estimates and<br />

2 E is the sum <strong>of</strong> the estimated error variances. As the<br />

genetic analysis included seedlots from both natural stands and the first generation progeny trial, the<br />

coefficient <strong>of</strong> relationship used to estimate <strong>of</strong> additive genetic variance was not altered from 0.25 (Luo<br />

et al 2006 and House and Bell 1996).<br />

V σ<br />

= is the estimated individual tree narrow-sense heritability from each site<br />

hˆ 2<br />

i<br />

Ai<br />

VPi<br />

≈ 2<br />

σ P i<br />

2<br />

G i<br />

2 2<br />

+ σ P + σ i E i<br />

2<br />

Gi is the genetic variance from site i,<br />

i, where<br />

variance for site i.<br />

2 Pi is the plot error for site i, and<br />

2 Ei is the residual


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 335<br />

The genetic correlation between trial sites for growth was calculated as: rˆ<br />

ADD =<br />

σ g<br />

2 2<br />

σ gσ<br />

g<br />

, with<br />

covariance ( g) between sites (ie. 1, 2 and 3) and genetic variances defined as above (Burdon 1977,<br />

Gilmour et al. 2005).<br />

RESULTS<br />

There were significant differences between the various origins <strong>of</strong> material established in these progeny<br />

trials at all ages for the assessed growth traits. Survival was similar at each site with 71, 88 and 83<br />

percent <strong>of</strong> the trees remaining in EB1, EB15A and EB15B respectively at the final assessment. The<br />

locally produced TPL material was consistently superior across all ages <strong>of</strong> assessment in each trial,<br />

with the exception <strong>of</strong> the 36 month DBH assessment at the first generation trial (EB1). Within this<br />

first generation trial, the material from <strong>Australia</strong> performed particularly well at 36 months <strong>of</strong> age and<br />

was significantly larger than the PNG (p


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 336<br />

Table 4. Descriptive statistics <strong>of</strong> provenances, localities and controls included in three Toba Pulp Lestari E. pellita field trials<br />

EB1 EB15A EB15B<br />

DBH<br />

Origin HT 12 HT 24 DBH 24 DBH 36 HT 18 DBH 18 DBH 30 HT 18 18 DBH 30<br />

AUS Abergowrie 3.1 9.2 8.9 12.6 8.5 7.2 9.6<br />

AUS Bloomfield 2.6 7.9 7.2 9.7 6.7 5.8 7.9<br />

AUS Cardwell 3.0 8.8 8.1 10.7 6.8 5.8 7.8<br />

AUS Daintree 3.1 9.2 8.6 12.9 6.5 5.9 8.1<br />

AUS Helenvale 3.0 8.7 8.5 12.4 7.1 6.1 8.6<br />

AUS Kirrama Range 2.9 9.0 8.6 13.0 9.2 7.8 10.5<br />

AUS Kuranda 3.5 9.7 9.5 13.3 8.1 6.8 9.1<br />

AUS Tinaroo Creek 3.3 10.9 10.1 13.5<br />

AUS Wonga – Daintree 3.3 9.1 8.2 11.6 9.0 6.8 9.5<br />

IND Bupul Muting 3.2 8.4 8.2 11.7 6.4 5.7 7.8 9.3 6.9 8.6<br />

IND Irian Jaya 3.2 8.7 8.1 11.1 8.0 6.8 9.6 10.9 7.6 9.6<br />

PNG Keru to Kumbalusi 2.7 8.0 7.5 10.6 9.9 8.3 11.7 11.9 8.7 11.1<br />

PNG Keru to Mata 2.3 6.7 6.4 8.7 9.0 6.4 7.8<br />

PNG South Kiriwo 3.1 8.6 8.6 11.6 7.1 6.3 8.9 9.9 6.9 8.6<br />

PNG Serisa 2.5 7.7 7.2 10.2<br />

PNG Tokwa 3.0 7.9 8.3 11.0 6.9 5.8 7.5 9.7 6.6 8.1<br />

TPL E.pellita Lowland Seed 8.1 6.8 9.3 11.5 7.9 9.9<br />

TPL E.pellita QLD selects 7.7 6.8 9.1 10.3 7.1 9.0<br />

TPL E.grandis 4.0 11.4 9.8 12.2 9.9 7.8 11.1 9.6 6.2 8.7<br />

TPL E.urophylla 3.5 10.6 8.1 11.2 8.3 7.2 10.2 10.9 7.4 8.8<br />

TPL Clone 1 High 11.0 8.4 11.3<br />

TPL Clone 2 High 10.2 7.9 11.5<br />

TPL Clone 1 Low 11.1 8.1 9.5<br />

TPL Clone 2 Low 13.1 9.3 12.4<br />

SED 0.49 0.80 0.92 1.25 0.50 0.41 0.62 0.53 0.46 0.64<br />

Pr > F 0.0378 0.0012 0.0006 0.0012


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 337<br />

The germplasm originating from the TPL breeding program that was used as controls generally<br />

performed well in these trials with substantial increases in relative performance from the first<br />

generation to the second. The drop in the growth <strong>of</strong> the E. grandis and E. urophylla controls relative<br />

to the other controls when established in the second generation lowland trial is likely due to fungal leaf<br />

pathogens reported to affect these species when established in the wet tropics (Dionese et al. 1984,<br />

Harwood et al. 1997, Werren 1991). While the survival <strong>of</strong> the E. grandis and E. urophylla controls in<br />

the second generation highland trial was 82 and 85 percent respectively, survival was just 13 and 50<br />

percent in the lowland trial. This compares to an overall second generation trial survival in the<br />

highland trial <strong>of</strong> 88 percent and an overall trial survival in the lowland trial <strong>of</strong> 83 percent. The<br />

differential impact <strong>of</strong> disease was less apparent in the E. pellita controls from the lowland and<br />

highland sources; material selected in a lowland orchard was significantly larger than a bulk <strong>of</strong><br />

selected Queensland families (9.9 cm vs 9.1 cm) and survival was 79 and 68 percent respectively<br />

when planted in the lowland trial. Within both second generation trials the clonal controls grew very<br />

well, showing significant superiority to most <strong>of</strong> the families originating from the native range.<br />

When comparing the first generation and second generation trials in the upland region, improvements<br />

in form and stem defects are visually apparent. In the first generation trial top-breakage was a<br />

problem with a trial average <strong>of</strong> 4.9 percent <strong>of</strong> the trees being broken. This varied greatly across<br />

families in the trial, ranging from 0 percent to as high as 44 percent for some families. Although<br />

confounded with differences between the first and second generation trial sites, the effect <strong>of</strong> selecting<br />

against top-breakage by removing individuals with broken tops prior to seed collection for the second<br />

generation appeared to be very effective. The incidence <strong>of</strong> top breakage was nearly halved in the<br />

second generation trial EB15A when compared to EB1 with a reduction from nearly 5 to 2.5 percent.<br />

The <strong>of</strong>fspring <strong>of</strong> the individuals selected from within a highly damaged first generation family, which<br />

had 44 percent <strong>of</strong> the tops broken out, reduced to 13.3 percent top-breakage in the second generation<br />

trial.<br />

Table 5. Genetic parameters and standard errors <strong>of</strong> estimates from the latest diameter<br />

assessment taken in the three trials<br />

EB1 EB15A EB15B<br />

Heritability 0.15 (0.06) 0.33 (0.07) 0.24 (0.09)<br />

Genetic Correlation 0.54 (0.31) 0.49 (0.59)<br />

Genetic Correlation 0.79 (0.16)<br />

The across site heritability estimate from the three trials was 0.24 (standard error <strong>of</strong> 0.05), which was<br />

very similar to the average <strong>of</strong> the individual site heritability estimates detailed in table 4. The<br />

heritability estimates increased markedly between the first and second generation trials, indicating a<br />

substantial rise in the ability <strong>of</strong> differentiate between families in the second generation. The principal<br />

reason the across-site analysis was undertaken was to estimate genetic correlations between the trials,<br />

specifically between the first and second generation trials. Family rankings from the first generation<br />

trial, from which seed was collected for the second generation trials, did not correlate well with<br />

rankings in either second generation trial. The standard errors <strong>of</strong> these correlation estimates were<br />

large and indicated little precision in their estimation. This contrasted with the relatively high<br />

correlation between the two second generation trials, which were established in strikingly different<br />

environments.<br />

The assumption that an additional genetic group was required to account for the common pollen parent<br />

was checked by allocating second generation parents to the original provenance from which the seed<br />

was collected. This alternative approach generated very similar heritability estimates for both across<br />

and individual site analyses. However, correlations between the first and second generation trials were<br />

considerably lower (0.45 (±0.32) for EB1-EB15A, 0.42 (±0.61) for EB1-EB15B) while the correlation<br />

between the two second generation trials was similar (0.79 (±0.16)).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 338<br />

DISCUSSION<br />

Establishing a range <strong>of</strong> Eucalyptus plantations in the highlands <strong>of</strong> Sumatra has been successful for a<br />

range <strong>of</strong> species. Comparisons <strong>of</strong> locally produced seedlots and clones indicate that significant<br />

genetic improvement can be achieved relative to the E. pellita families in these trials. The utility <strong>of</strong><br />

the species has increased recently following the demonstration <strong>of</strong> its resistance to a range <strong>of</strong> leaf fungi,<br />

which typically cause eucalypt plantations in the wet tropics to fail (Werren 1991, Harwood 1998), as<br />

well as the species ability to produce productive hybrids (Luo et al. 2006). While E. pellita seedlings<br />

may not be the primary source <strong>of</strong> material for deployment within the Sumatran highlands estate, the<br />

worldwide E. pellita plantation estate has expanded in recent years and derived hybrids have been<br />

identified by various tree improvement programs in the tropics.<br />

As previously reported, the most productive source in the first generation trial was from Kuranda (Luo<br />

et al. 2006 and Harwood 1998) with nearby Tinaroo Creek also performing well. Kuranda has been<br />

identified as having a particularly high outcrossing rate when compared to other natural populations<br />

from Cape York and Irian Jaya (House and Bell 1996). Although these trials were rather<br />

comprehensive in the number <strong>of</strong> sources included, many <strong>of</strong> the provenances that were compared to<br />

Kuranda were represented with few families and their true genetic potential may not have been<br />

indicated. The relative superiority <strong>of</strong> Kuranda diminished in the second generation trial, possibly<br />

because the growth advantage resulting from naturally high outcrossing rates was not present in the<br />

second generation trial where other families had equal chances <strong>of</strong> outcrossing. Low outcrossing rates<br />

<strong>of</strong> eucalypt species have been shown to increase within-family variation, because many selfed or<br />

inbred individuals perform poorly as a consequence <strong>of</strong> inbreeding depression (Eldridge et al 1993).<br />

Increased within-family variation would tend to bias genetic parameter estimates downwards if<br />

differential selfing rates are prevalent in the families included in these trials. On the other hand, related<br />

matings resulting in a larger number <strong>of</strong> full-sib <strong>of</strong>fspring within families would tend to bias heritability<br />

estimates upwards (Eldridge et al 1993).<br />

The main focus <strong>of</strong> the genetic analysis was on the estimation <strong>of</strong> genetic parameters within the first and<br />

second generation trials for comparative purposes. The heritability estimate <strong>of</strong> the first generation trial<br />

was low and standard error was high when compared to both second generation trials. The impact <strong>of</strong><br />

differential outcrossing rates on heritability estimates has been discussed at length by Eldridge et al<br />

(1993). The distribution <strong>of</strong> E. pellita, as reviewed by Harwood (1998), has been described as ‘heavily<br />

dissected’ in the Fly-Diogel shelf <strong>of</strong> New Guinea by Paijmans (1976), ‘small scattered populations’ in<br />

Southern Papua, ‘extensive tall open forests’ from Bupul to Muting in Irian Jaya, ‘discontinuous and<br />

scattered’ in ‘narrow bands’ in Northern Queensland with a ‘major disjunction’ between populations.<br />

Due to the dispersed occurrence <strong>of</strong> individuals within parts <strong>of</strong> the native range, it is expected that<br />

selfing would be prevalent in some <strong>of</strong> the families included in these trials. While an increase in<br />

related mating would tend to bias heritability upwards due to an increase in the true coefficient <strong>of</strong><br />

relationship, a downward bias in the heritability estimate <strong>of</strong> the first generation progeny trial could be<br />

caused by differential levels <strong>of</strong> selfing inflating the within family variance. In addition to the<br />

differential selfing rates, other reasons the heritability estimate was found to be lower in the first<br />

generation trial than in the second generation trials could be: fewer trees per plot, reduced silvicultural<br />

intensity or any cause <strong>of</strong> increased experimental error. The reduced heritability in the first generation<br />

implies the rankings <strong>of</strong> families are not well predicted and there is therefore an expectation that<br />

rankings would differ in subsequent generations.<br />

Genetic correlations have provided an indication <strong>of</strong> the relative importance <strong>of</strong> between-generation and<br />

between-site differences on the ranking <strong>of</strong> families <strong>of</strong> E. pellita. Low correlation estimates between<br />

the first and second generation trials indicate that substantial changes in ranking occurred following<br />

the thinning that was used to convert the progeny trial into a seedling seed orchard. Although the<br />

genetic correlation between the first and second generation highland trials is effected by both the<br />

climatic and management differences as well as outcrossing, it was expected that the correlation<br />

between the highland and lowland trials would be lower than the correlation between the two highland<br />

trials. Nevertheless, the correlation between the two second generation trials was found to be much<br />

greater than the correspondence <strong>of</strong> family rankings in the first and both <strong>of</strong> the second generation trials.<br />

The selection <strong>of</strong> improved families for tropical highlands and deployment <strong>of</strong> these same families in the


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 339<br />

lowlands appears to be a possible given the results <strong>of</strong> these trials. Although the statistics and genetic<br />

theory imply a single population would generate more genetic gain by focusing resources, the<br />

differences in climate and soils in the lowland and highland estate are so dramatic that intuition does<br />

not lead prudent tree breeders into forming a single population for deployment into these two<br />

environments.<br />

CONCLUSIONS<br />

The results <strong>of</strong> this study will provide some guidance for other organizations initiating tree<br />

improvement programs for Eucalyptus species. First generation trials <strong>of</strong> eucalypts provide a first<br />

indication <strong>of</strong> the relative merit <strong>of</strong> populations and families, but results from these trials should be<br />

treated with caution (Eldridge et al. 2003, White et al. 2007). Culling families at the early stages <strong>of</strong><br />

breeding population development could eliminate individuals that would have been valuable once<br />

given the opportunity to outcross. Relatively simple and extensive breeding strategies that quickly<br />

move through the evaluation <strong>of</strong> native-range populations and do not purge poorly performing families<br />

may be an effective path to creating diverse populations for advanced generation breeding.<br />

REFERENCES<br />

Burdon, R.D. 1977. Genetic correlation as a concept for studying genotype-environment interaction in forest<br />

tree breeding. Silvae Genet. 26: 168-175.<br />

Dionese, J.C., Haridasan, M. and Moraes, T.S. de A. 1984 Tolerance to “mal do Rio Doce”, a major disease <strong>of</strong><br />

Eucalyptus in Brazil, Tropical Pest Management 30(3), 247-252<br />

Doran, J.C. Williams, E.R. and Brophy, J.J. 1995. Patterns <strong>of</strong> variation in the leaf oils <strong>of</strong> Eucalyptus urophylla,<br />

E.pellita and E.scias. <strong>Australia</strong>n Journal <strong>of</strong> Botany 43, 327-336<br />

Eldridge, K., Davidson, J., Harwood, C. and van Wyk, G. (1993): Eucalypt Domestication and Breeding. Oxford<br />

University Press. Oxford, UK.<br />

Falconer, D. S. and Mackay, T. F. C. (1996): Introduction to Quantitative Genetics. 4th Edition. Addison Wesley<br />

Longman Ltd. Essex, UK.<br />

Gilmour, A.R., Gogel, B.J., Cullis, B.R., and Thompson, R. 2005. ASReml User Guide Release 2.0. VSN<br />

International Ltd., Hemel Hempstead, HP1 1ES, UK. 267 pp.<br />

Harwood, C.E., 1998. Eucalyptus pellita – An annotated Bibliography. CSIOR Forestry and Forest Products,<br />

Canberra. 70 pp.<br />

Harwood, C.E., Nikles, D.G., Pomroy, P.C. and Robson, K.J. 1997. Genteic improvement <strong>of</strong> Eucalyptus pellita<br />

in north Queensland, <strong>Australia</strong>. Pp 219-226, Vol1 in IUFRO Converence on silviculture and improvement<br />

<strong>of</strong> Eucalypt, 1997, Slavador. EMBRAPA Centro Nacional de Pesquia de Florestais<br />

House, A.P.N. and Bell, J.C. 1996. Genetic diversity and systematic relationships in two red mahoganies,<br />

Eucalyptus pellita and Eucalyptus scias. <strong>Australia</strong>n Journal <strong>of</strong> Botany 44, 157-174<br />

Luo, J. Arnold, R.J. and Aken, K. 2006. Genetic variation in growth and typhoon resistance in Eucalyptus pellita<br />

in south-western China. <strong>Australia</strong>n Forestry. Vol 69, No 1:38-47<br />

Pegg, R.E., and Wang, G.X. 1994. Results <strong>of</strong> Eucalyptus pellita trials at Dongmen, China. In Brown, A.G. (ed.)<br />

<strong>Australia</strong>n tree species research in China: Proceedings <strong>of</strong> and international workshop held at Zhangzhou,<br />

Fujian Province, PRC, 2_5 November 1992. ACIAR Proceedings no. 48. ACIAR Canberra, 226p<br />

Paijmans, K. 1976) New Guinea Vegetation. <strong>Australia</strong>n National University Press, Canberra, 213 pp.<br />

Quass, R.L. 1988. Additive genetic model with groups and relationships. J. Dairy Sci. 71:1338-1345<br />

SAS <strong>Institute</strong>, Inc. 1991 SAS Language, SAS <strong>Institute</strong> Inc., Cary, NC. USA.<br />

Vercoe, T.K. and McDonald, M.W. 1991. Eucalyptus pellita F.Muell and Acacia seed collections in New<br />

Guinea, September – October 1990. Forest Genetic Resources Information 19, 38-42<br />

Werren, M. 1991 Eucalyptus plantation development in Indonesia. Pp 1160-1166 in AGP Schonau, (ed.)<br />

Symposium on Intensive Forestry: The role <strong>of</strong> Eucalypts. IUFRO and Souther African <strong>Institute</strong> <strong>of</strong><br />

Forestry, Pretoria, South Africa<br />

Westell, R.A. Quass, R.L., and van Vleck, L.D.. 1988. Genetic groups in animal models. J. Dairy Sci, 71:1310-<br />

1318<br />

White TL, Adams WT, Neale DB (2007) Forest genetics. CABI, Wallingford, p 682<br />

Yamada, Y. 1962. Genotype by environment interaction and genetic correlation <strong>of</strong> the same trait under<br />

different environments. Jap. J. Genet. 37: 498-509.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 340<br />

GROWING TEAK FOR THE ROOTS – THE JIFFY SOLUTION<br />

ABSTRACT<br />

Don Willis 1<br />

Have you ever managed your plantation for what you do not see? What if you were<br />

able to produce Teak seedlings in the nursery in a shorter time period and plant out a<br />

younger root system that is ground-ready?<br />

S<strong>of</strong>t-walled pellets mimic nature <strong>of</strong>fering unrestricted root growth symmetry and<br />

superior lateral root growth. Through air pruning and proper water management in the<br />

nursery, Jiffy pellets allow roots to maximize stability in Teak plantations, resulting in<br />

improved wood quality. Maximizing Teak seedling crop management includes: 100%<br />

tray stocking at anytime, easy grading for quality at anytime, shipping at anytime, and<br />

water conservation while simultaneously producing a long-term viable root system for<br />

the life <strong>of</strong> the plantation.<br />

The Jiffy Pellet System is a value added, quality seedling container and media in one<br />

solution that simplifies nursery management. In an era <strong>of</strong> environmental economics,<br />

businesses that lead us into the future choose superior products that lead the way<br />

towards zero waste. The Jiffy Pellet System is the next generation.<br />

1 Don Willis RPF, Global Forestry Product Manager, Jiffy Products, 125 Industrial Park, Shippagan, NB E8S 3H1, Canada,<br />

Ph (+1) 506 – 336 – 2284 Web: www.jiffypot.com.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 341<br />

LESSONS LEARNED FROM IMPLEMENTATION OF REDD –<br />

A COLLECTION OF ANALOGIES, METAPHORS AND CLICHÉS<br />

Zoe Harkin 1<br />

ABSTRACT<br />

Fauna & Flora International, together with the Macquarie Group, have partnered to<br />

form a ‘Carbon Forests Taskforce’ (CFT), with the aim <strong>of</strong> implementing at least six<br />

projects for Reducing Emissions from Deforestation and Degradation (REDD), located<br />

in threatened landscapes spread across south east Asia, South America and Africa. It is<br />

planned for avoided emissions generated from the projects to be verified and sold on<br />

voluntary carbon markets (i.e. markets driven by corporate social responsibility and<br />

speculators). It is also hoped that the resultant REDD credits will be accepted by<br />

international compliance-based (i.e. formal) markets for REDD, if and when they<br />

become operational.<br />

The paper provides a summary <strong>of</strong> observations and lessons learned by the CFT Forest<br />

Carbon Specialist, following one year’s experience in the implementation <strong>of</strong> REDD.<br />

To facilitate a common understanding <strong>of</strong> the sometimes complex concepts associated<br />

with REDD, the paper is presented as a series <strong>of</strong> analogies, metaphors and clichés.<br />

REDD: Calm before the storm... but clearly the next gold rush<br />

In his 2006 ‘Review <strong>of</strong> the Economics on Climate Change’, the head <strong>of</strong> the UK Economic Service,<br />

Lord Nicholas Stern, proclaimed that “curbing deforestation is a highly cost-effective way <strong>of</strong> reducing<br />

greenhouse gas emissions.” Stern’s economic models suggested that significant quantities <strong>of</strong> REDD<br />

credits could be developed for less than USD 5 per tonne <strong>of</strong> CO2 (compared with marginal abatement<br />

costs for Carbon, Capture and Storage (CCS) <strong>of</strong> up to USD 270 per tonne CO2).<br />

Stern’s conclusions grabbed the attention <strong>of</strong> the developed nations <strong>of</strong> the world - hungry for cheap,<br />

proven ‘technologies’ that could be mobilised rapidly to mitigate climate change. REDD had the<br />

added appeal <strong>of</strong> aligning with the objectives <strong>of</strong> other international treaties such as the UN Human<br />

Development goals, Convention to Combat Desertification, and the Convention on Biological<br />

Diversity. Likewise, forested developing nations, long unable to reign in their rampant deforestation<br />

rates, saw REDD as a mechanism to attract compensation for efforts to reduce deforestation, and for<br />

what they perceived to be foregone development opportunities.<br />

In the international policy arena, REDD came into the international spotlight at the 13 th Conference <strong>of</strong><br />

the Parties (COP) to the United Nations Framework Convention on Climate Change (UNFCCC), held<br />

in Bali in December 2007. Here, the ‘Bali Roadmap’ paved the way for REDD to be included within<br />

a post-Kyoto international climate change agreement, to be implemented from 2013 onwards. By all<br />

appearances, REDD had the unanimous support <strong>of</strong> all signatory countries to the UNFCCC, including<br />

particularly strong support from <strong>Australia</strong>.<br />

This international in-principle endorsement <strong>of</strong> REDD spawned a flurry <strong>of</strong> activity in the budding<br />

voluntary carbon market. Environmental groups rushed to stake out their claim on areas <strong>of</strong> forest<br />

considered at risk <strong>of</strong> conversion to other land uses. Investors looked to partner with the<br />

environmental groups, in the hope <strong>of</strong> securing cheap credits ahead <strong>of</strong> the rush, and perhaps hoping to<br />

polish their corporate images at the same time. The Governments <strong>of</strong> forested developing countries<br />

looked on, sensing a significant opportunity while at the same time wary <strong>of</strong> being ripped <strong>of</strong>f, yet<br />

again, in a market they did not fully understand. The World Bank, the United Nations Collaborative<br />

1 Dr Zoe Harkin, Fauna & Flora International. Ph: 03 9416 5220 Email: zoe.harkin@fauna-flora.org.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 342<br />

Programme on Reducing Emissions from Deforestation and Forest Degradation in Developing<br />

Countries (UN-REDD) and other donor countries formed massive funds to help fund REDD<br />

‘readiness’ activities. And Indigenous Peoples human rights activists grew increasingly weary <strong>of</strong><br />

waiting for the aforementioned REDD proponents to consult with them. And so the ‘REDD gold<br />

rush’ was born....<br />

At the time <strong>of</strong> writing, this ‘REDD gold rush’ is in a state <strong>of</strong> calm before the storm, awaiting formal<br />

endorsement <strong>of</strong> REDD at COP 15 in Copenhagen in December this year. If the draft negotiating text<br />

for the meeting is any indication, it appears almost certain that REDD will be included within the<br />

post-2012 UNFCCC agreement. If and when this occurs, carbon market analysts suggest that<br />

demand, and subsequently prices for REDD credits, will increase rapidly.<br />

REDD: the ultimate horse/cart, chicken/egg problem<br />

In our numerous presentations on REDD, we have used linear or circular process diagrams, such as<br />

the one depicted below, in order to explain the ‘REDD implementation process’.<br />

Figure 1 Schematic depicting the REDD implementation process<br />

Unfortunately this schematic is an over-simplification. Implementation <strong>of</strong> REDD is not quite the<br />

linear process we had envisioned. For example: in commencing consultation with local stakeholders<br />

and indigenous groups, one <strong>of</strong> the first things they want to know is an indication <strong>of</strong> the benefits they<br />

might receive from the proposed REDD project. In order to develop a reasonable estimate <strong>of</strong> the<br />

benefits <strong>of</strong> the REDD project, it is necessary to conduct a forest carbon inventory. However investors<br />

may be reluctant to provide a significant investment in a carbon inventory, unless they have the inprinciple<br />

support <strong>of</strong> the local communities and Government, which can only be acquired following<br />

extensive consultation. Furthermore, under existing standards in the voluntary carbon market, the<br />

REDD project cannot be verified unless the project developer has secure tenure over the site.<br />

Securing tenure can be an expensive business, particularly if this involves buying out existing<br />

commercial stakeholders or paying hefty application fees. Therefore an investor might reasonably<br />

want assurance that the project is verifiable before going to the considerable effort and expense <strong>of</strong><br />

securing tenure... and so goes the REDD chicken-egg-horse-cart merry-go-round.<br />

Ultimately, it appears that REDD is an integrated assessment in the truest sense <strong>of</strong> the term. It<br />

requires pursuit <strong>of</strong> all facets <strong>of</strong> the project – carbon inventory, community consultation, government


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 343<br />

endorsement and tenure security – at the same time. The project will fail if just one <strong>of</strong> these aspects is<br />

not achieved, and the inter-relationship between all facets largely precludes a sequential, linear<br />

approach.<br />

Forecasting baselines: smoke and mirrors, or clear as day?<br />

In order to quantify avoided emissions due to implementation <strong>of</strong> REDD, current accounting rules in<br />

the voluntary carbon market typically require the project developer to forecast the forest carbon<br />

pr<strong>of</strong>ile under the project (or REDD intervention) scenario. This is then compared to a forecast <strong>of</strong> the<br />

forest carbon pr<strong>of</strong>ile under the baseline (or ‘business-as-usual’) scenario (Figure 2).<br />

Figure 2 Forecast <strong>of</strong> baseline and project scenarios to estimate avoided emissions<br />

This rule defines eligible avoided emissions are those that occur in excess <strong>of</strong> ‘business as usual’<br />

activity, which are said to be ‘additional’. The intention <strong>of</strong> this accounting rule is to “incentivise”<br />

genuine behavioural changes, rather than provide credits for activities that would have occurred in the<br />

absence <strong>of</strong> carbon markets anyway.<br />

During the very early stages <strong>of</strong> REDD market development, forecasting <strong>of</strong> the baseline carbon pr<strong>of</strong>ile<br />

could be described as little more than crystal ball gazing. Under this ‘smoke and mirrors’ approach, a<br />

few references to the scientific literature, supplemented with some commentary on the deforestation<br />

drivers affecting the site, appeared to suffice. As the voluntary carbon market has matured, however,<br />

the technically prescriptive standards such as the Voluntary Carbon Standard (VCS) are requiring<br />

baseline forecasts to be developed using more advanced methodologies. These methodologies are<br />

tailored to different baseline scenarios, depending on whether the baseline deforestation/degradation<br />

is planned (i.e. sanctioned by the government, such as oil palm conversion licences); or unplanned<br />

(i.e. unsanctioned by the government, such as expansion <strong>of</strong> slash and burn agriculture, clearing for<br />

grazing, or illegal logging).<br />

In cases <strong>of</strong> planned deforestation, the REDD project developer is now required to collate an extensive<br />

audit trail including documentation such as regional development plans, zoning maps, management<br />

plans and licences. This documentation needs to convince the auditors <strong>of</strong> the exact time and location<br />

that the deforestation or degradation event was planned to have occurred. In the case <strong>of</strong> unplanned<br />

deforestation, the methodologies required to forecast the rate and location <strong>of</strong> deforestation are far<br />

more complex. These typically require the use <strong>of</strong> GIS based s<strong>of</strong>tware to calculate the historical rate <strong>of</strong><br />

land use conversion in a suitable reference (or analogous) area, and then correlate this conversion rate<br />

with specific indicator variables such as distance to roads, population centres, sawmills, nearest<br />

distance to clearing, etc.<br />

Regardless <strong>of</strong> which methodology is used to forecast baselines, auditors are demanding more detailed<br />

analysis, documentation and support from the scientific literature in order to approve baseline


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 344<br />

forecasts. It is also important to keep in mind that it may eventuate that baselines are not required in<br />

future, if and when REDD tends towards national level accounting. This is because avoided<br />

emissions would be sold relative to the collective national reference level, regardless <strong>of</strong> additionality<br />

at the project level.<br />

Additionality is a double-edged sword<br />

As described above, additionality refers to the requirement that eligible avoided emissions are those<br />

that would have occurred in excess <strong>of</strong> business-as-usual activity. This additionality requirement is<br />

currently included in almost all forest carbon standards on the voluntary carbon market today 2 .<br />

In order to demonstrate additionality, the project developer is required to show pro<strong>of</strong> that the forest<br />

was under imminent threat <strong>of</strong> deforestation or degradation. Since forward crediting is not allowed<br />

under the VCS, then REDD credits can only be claimed at the time at which the deforestation or<br />

degradation would have occurred. The combined effect <strong>of</strong> these rules is that forests considered most<br />

at risk <strong>of</strong> deforestation or degradation are most attractive as a REDD project.<br />

This approach has obvious advantages. Primarily, the additionality rule attracts REDD investment to<br />

the forests that are most at threat. On the flipside, forests that are most at threat are <strong>of</strong>ten the hardest<br />

(and arguably the most expensive) to save. Intervening just prior to deforestation is likely to require<br />

compensation <strong>of</strong> a whole suite <strong>of</strong> stakeholders, all <strong>of</strong> whom have invested significant time and energy<br />

in pursuit <strong>of</strong> the business-as-usual activity. The more advanced the stage <strong>of</strong> this development, the<br />

more costs that are likely to have been incurred, and the more committed these parties are to the<br />

activity. Averting this course <strong>of</strong> action ‘at the last minute’ can be extremely tricky, and potentially<br />

quite expensive.<br />

REDD project developers clearly need to weigh up the pros and cons <strong>of</strong> this ‘double edged sword’<br />

phenomenon – assessing the risk <strong>of</strong> investing in projects with strong additionality, against the<br />

attractiveness <strong>of</strong> the more immediate revenue stream that can be delivered from such projects.<br />

Unravelling tenure issues is like peeling the layers <strong>of</strong> an onion<br />

Lack <strong>of</strong> clarity on forest tenure is <strong>of</strong>ten cited as one <strong>of</strong> the fundamental drivers <strong>of</strong> deforestation or<br />

degradation. This is because uncertainty around who is eligible to do what in the forest creates a<br />

‘tragedy <strong>of</strong> the commons’ scenario. Each stakeholder ultimately feels that if they don’t utilise the<br />

forest to their maximum utility, then someone else will. The negative impacts <strong>of</strong> the<br />

deforestation/degradation event will be shared to some extent by all, and particularly those living<br />

closest to the forest. Therefore the marginal utility <strong>of</strong> deforestation for an individual is generally<br />

greater than the negative impacts they might experience, even if the overall negative impacts shared<br />

by society exceed this benefit.<br />

Lack <strong>of</strong> clarity around forest tenure in many developing countries can be due to an ‘onion-like’<br />

layering <strong>of</strong> land rights and use claims. The uppermost layer might typically involve some sort <strong>of</strong> legal<br />

permit to clear the land. However multiple levels <strong>of</strong> government, and multiple ministries within the<br />

same government, can lead to a situation where the same piece <strong>of</strong> forest land has been licensed to<br />

multiple parties for overlapping, and <strong>of</strong>ten conflicting, land uses. A common example is approval <strong>of</strong> a<br />

mine on an area also approved for commercial forestry. In part, the ‘onion-like’ land tenure situation<br />

arises due to corruption, whereby government <strong>of</strong>ficials receive financial benefits from each successive<br />

licence approval. The government <strong>of</strong>ficial may have little interest in which party ultimately benefits<br />

from their licence. If the multiple licensees are aware <strong>of</strong> each other’s presence, this can lead to a ‘race<br />

against time’ to undertake their licensed activities before the other licensees.<br />

Beneath these government-endorsed tenure arrangements <strong>of</strong>ten lie customary or historical land use<br />

claims. These may or may not be formally recognised by the government. And the customary land<br />

use claims can also overlap, for example where different indigenous groups or local communities<br />

claim ownership or customary use <strong>of</strong> the same area <strong>of</strong> land, or when the ’bundle’ <strong>of</strong> forest use rights<br />

is divided between or shared amongst multiple forest users. This can be particularly complex where<br />

indigenous groups have been displaced, and migrants may have moved in to the area. Such events<br />

2 With the exception <strong>of</strong> the Chicago Climate Exchange.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 345<br />

may have occurred many years in the past, however the indigenous group may still have historical<br />

links to the area, while the ‘newer’ migrants feel a sense <strong>of</strong> ownership <strong>of</strong> the forest because <strong>of</strong> their<br />

daily interaction with and dependence on the forest.<br />

This ‘onion-like’ layering <strong>of</strong> tenure issues <strong>of</strong>ten leads to a hierarchy <strong>of</strong> drivers <strong>of</strong> deforestation. If the<br />

primary driver <strong>of</strong> deforestation is addressed, then the secondary drivers <strong>of</strong> deforestation may become<br />

dominant. This tenure situation is similar to Liebig’s ‘barrel analogy’ <strong>of</strong> forest nutrition (Justus von<br />

Liebig (1803-1873), whereby forest growth is theoretically constrained by the most limiting factor.<br />

Likewise, deforestation is determined by the most dominant driver.<br />

What this means is that implementation <strong>of</strong> REDD requires a holistic approach, where each successive<br />

driver <strong>of</strong> deforestation is identified, and ultimately addressed.<br />

The ‘baby out with the bathwater’ problem<br />

The commercial viability <strong>of</strong> REDD is highly sensitive to assumptions around the price <strong>of</strong> carbon.<br />

During these early stages <strong>of</strong> REDD market development, the assumed carbon price is quite low. As a<br />

result, marginal increases in project development costs can quickly make a potential REDD project<br />

unviable, or considered too risky. This creates a ‘baby out with bathwater’ problem – if the cost <strong>of</strong><br />

developing a REDD project becomes too high, then many good prospective REDD projects will be<br />

abandoned before they even start.<br />

Alongside this cost sensitivity, the proliferation <strong>of</strong> technologies to measure and monitor forest carbon<br />

are growing in sophistication (and cost). It is currently unclear as to which <strong>of</strong> these methodologies and<br />

technologies will be required for implementation <strong>of</strong> REDD, either in the voluntary or compliance<br />

market. In the face <strong>of</strong> this uncertainty, a prudent REDD project developer might seek to adopt ‘best<br />

practices’ carbon inventory and baseline development practices, in order to reduce the risk that the<br />

project is considered non-compliant when REDD methodologies are clarified. This might involve<br />

utilisation <strong>of</strong> advanced remote sensing technologies, complex GIS-based land use change forecasting<br />

algorithms, combined with a high density <strong>of</strong> field sample plots. There is a danger that the cost <strong>of</strong> such<br />

‘best practices’ methodologies may become prohibitive, if they become a requirement for all REDD<br />

projects. A more sensible approach might be for project developers to commit to ‘continuous<br />

improvement’; developing procedures for retrospective back-calculation <strong>of</strong> avoided emissions, thus<br />

allowing the project developer to apply the best practice technologies consistently across the time<br />

series once they can be afforded.<br />

This ‘baby out with bathwater’ problem has been recognised by a number <strong>of</strong> stakeholders involved in<br />

REDD methodology development. For example, a US-based group, Avoided Deforestation Partners,<br />

adopted a ‘de minimus’ approach in development <strong>of</strong> their methodologies for submission to the VCS.<br />

This is based on the principle that REDD methodologies should be based on the minimum effort<br />

required in order to achieve best practices.<br />

Robbing Peter to pay Paul: the leakage problem<br />

One <strong>of</strong> the major criticisms <strong>of</strong> REDD has been its potential to cause ‘leakage’ issues. This relates to<br />

the concern that attempts to reduce deforestation or degradation in one area, may simply displace<br />

these activities to another area, resulting in little or no net benefit to the atmosphere. In effect, poorly<br />

designed REDD projects could effectively be ‘robbing Peter to pay Paul’.<br />

There are two main solutions to the leakage problem. The first is to conduct carbon accounting on a<br />

national basis, thereby capturing all leakage effects within a single national reporting system. The<br />

second is to address the fundamental drivers <strong>of</strong> deforestation, rather than simply addressing the<br />

symptoms. The example most relevant to foresters is timber supply. Timber is extracted for two<br />

reasons: 1) as a source <strong>of</strong> income generation for those extracting it; and 2) as a supply <strong>of</strong> wood<br />

products for paper, construction, furniture and other uses. If a REDD project cuts <strong>of</strong>f timber supply<br />

without providing alternative income sources for the local timber industry, or without implementing<br />

demand-side abatement measures, then the project will likely result in leakage.<br />

Within reason, there is a danger that demand-side abatement measures could quickly result in perverse<br />

greenhouse outcomes – if low greenhouse intensity wood products were displaced by more


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 346<br />

greenhouse intensive substitutes such as steel, aluminium or concrete, then the REDD project could<br />

ultimately result in more greenhouse gas emissions than it avoided.<br />

Attempts to find alternative income sources for the local timber industry are not always readily<br />

identifiable; do not necessarily guarantee a reduction in pressure on the forest; and local communities<br />

may resent any project that severely restricts their access to the forest. As a result, in many cases it<br />

appears that implementation <strong>of</strong> some form <strong>of</strong> sustainable forestry regime may be one <strong>of</strong> the best ways<br />

to address leakage issues. It allows a continued supply <strong>of</strong> timber from the forest, provides a continued<br />

stream <strong>of</strong> income for local people, and adoption <strong>of</strong> reduced impact logging techniques can avoid<br />

significant greenhouse gas emissions when compared with conventional logging. In this way, REDD<br />

is likely to become an important tool in achieving sustainable forest management objectives,<br />

particularly if coupled with certification and measures to reduce illegal logging.<br />

Light at the end <strong>of</strong> the tunnel...<br />

From the stream <strong>of</strong> issues and problems presented above, one might reasonably surmise that the<br />

obstacles involved in implementation <strong>of</strong> REDD are insurmountable. However, a quick ‘back <strong>of</strong><br />

envelope’ calculation on avoided forest emissions reveals why many investors are interested in<br />

REDD. The numbers on avoided emissions multiply very quickly, creating a ‘light at the end <strong>of</strong> the<br />

tunnel’ that is sure to reward project developers and investors with the persistence and tenacity to<br />

overcome the numerous problems that REDD presents. Plus the level <strong>of</strong> international enthusiasm for<br />

implementation <strong>of</strong> REDD is astounding. Very rarely does a proposal get such unanimous agreement<br />

from all countries in the world.<br />

Finally, the ability for REDD to assist in addressing a suite <strong>of</strong> other global issues that have long<br />

eluded policy-makers, such as global poverty alleviation, biodiversity conservation, sustainable timber<br />

production and agriculture, causes one to resort to a final cliché: REDD is a true a win-win (win, win,<br />

win....) situation.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 347<br />

DEVELOPING A REDD SCHEME FOR POST 2012: THE<br />

KALIMANTAN FORESTS AND CLIMATE PARTNERSHIP<br />

Grahame Applegate 1<br />

ABSTRACT<br />

The Kalimantan Forests and Climate Partnership (KFCP) is a demonstration for reducing<br />

greenhouse gas (GHG) emissions from avoided deforestation and forest degradation<br />

(REDD), with a focus on peat swamp forests in Indonesia. Incentive-based approaches<br />

will be adopted that can help pave the way for large scale financing <strong>of</strong> avoided<br />

deforestation and degradation with methodologies provided for the UNFCCC<br />

negotiations for a post 2012 climate change agenda. KFCP is identifying the enabling<br />

conditions for supporting environmental governance and institutional requirements,<br />

incentives, payment mechanisms and processes which link national, provincial, district<br />

and village level payments to performance based outcomes. KFCP is also developing and<br />

implementing appropriate systems for peat swamp forest and GHG measurements and<br />

accounting at the project and national level, taking account <strong>of</strong> additionality, leakage and<br />

permanence. KFCP is linked to the national carbon accounting system being developed<br />

by the Government <strong>of</strong> Indonesia.<br />

INTRODUCTION<br />

Peatlands cover 3 per cent <strong>of</strong> earth’s land area and store a large fraction <strong>of</strong> the world’s terrestrial<br />

carbon; up to 528 000 million tonnes, equivalent to 70 times the current annual global emissions from<br />

burning <strong>of</strong> fossil fuel (Hooijer et al. 2006). Peat consists <strong>of</strong> plant debris which has accumulated over<br />

thousands <strong>of</strong> years in water-logged, acidic conditions that prevent it from decomposing and releasing<br />

the stored carbon into the atmosphere. Peatlands emit around 800 million t CO2 annually, three<br />

quarters <strong>of</strong> which are from South-East Asia (Parish et al. 2007).<br />

Indonesia has around 22.5 million ha <strong>of</strong> peatlands, (Page and Banks 2007) which store 55± 10 Gt <strong>of</strong><br />

carbon (Jaenicke et al. 2008). The carbon stored in below ground biomass (peat) is 18.6 times higher<br />

than intact Indonesian peat swamp forests which have an above ground carbon content <strong>of</strong> 140.5 t C /ha<br />

(Uryu et al.2008). The peat swamp forests are the habitat <strong>of</strong> a large range <strong>of</strong> plants and animals,<br />

including endangered species such as orangutans. Over 26% <strong>of</strong> peatland in Indonesia is found on the<br />

island <strong>of</strong> Borneo, where 3.1 million ha, much <strong>of</strong> it originally covered in peat swamp forests, is found<br />

in Central Kalimantan (Hooijer et al.2006).<br />

Tropical peatlands are increasingly being cleared and drained for logging, farming and plantations <strong>of</strong><br />

oil palm and fast growing tree plantations, with the result that almost half <strong>of</strong> Indonesia’s peat swamp<br />

forests have been converted for other uses (Hooijer et al. 2006). Clearing and draining peatlands<br />

speeds up the decomposition <strong>of</strong> the peat and exposes it to fire, both <strong>of</strong> which result in large amounts <strong>of</strong><br />

carbon dioxide emissions. Hooijer et al. (2006) estimated that annual emissions from deforestation and<br />

burning <strong>of</strong> tropical peatlands are 2 billion tonnes CO2, with over 90% coming from Indonesia, the<br />

country’s single largest source <strong>of</strong> greenhouse gas emissions. Fires in peatlands were a major cause <strong>of</strong><br />

CO2 emissions resulting from the burning <strong>of</strong> 2.1 million ha <strong>of</strong> peatland during the large fires in 1997-<br />

98 that blanketed large parts <strong>of</strong> South-East Asia in haze (Page et al. 2002).<br />

Peat Swamp Forests and Factors Causing Deforestation and Degradation<br />

As shown in Figure 1, rivers <strong>of</strong>ten demarcate the boundaries <strong>of</strong> large continuous areas <strong>of</strong> peat.<br />

Furthermore, each continuous area <strong>of</strong> peat has its own hydrological unit, and can form a ‘dome’ up to<br />

fifty km in diameter (see Figure 2).<br />

1<br />

Consultant, AusAID, <strong>Australia</strong>n Embassy, Jl. H.R. Rasuna Said, Kav. C15-16, Jakarta, 12940, Indonesia. Email<br />

grahame.applegate@iafcp.or.id


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 348<br />

The Ex Mega Rice Project Master Plan (Euroconsult et al. 2008) provides a strong argument for<br />

managing the peatlands at a landscape level and taking a “whole-<strong>of</strong>-dome” approach based on<br />

hydrological boundaries. This approach and understanding is fundamental to developing solutions for<br />

managing peat to reduce GHG emissions from avoided deforestation and degradation. It is important<br />

to understand that affecting a change in water level in one area <strong>of</strong> the dome will impact over time on<br />

the remainder <strong>of</strong> the dome. Draining areas <strong>of</strong> shallow peat containing forest on the edge <strong>of</strong> the dome,<br />

for example for agricultural purposes, may lower the water table in the central part <strong>of</strong> the dome, which<br />

may result in drying-out areas <strong>of</strong> intact forest (making them susceptible to fire) and possibly dieback<br />

due to a lack <strong>of</strong> water. Although further research is required to quantify these relationships, there is<br />

evidence that the construction <strong>of</strong> canals to drain peat swamp forest areas in Indonesia has adversely<br />

impacted on them and continues to impact a much larger area in which the canals are located<br />

(Euroconsult et al. 2008).<br />

Source: Bappenas- National Strategy and Action Plan for Sustainable Management <strong>of</strong> Peatlands, 2006<br />

Figure 1. Cross section <strong>of</strong> a peat dome showing Peat Swamp Forest<br />

While additional research is required on these phenomena, the risk <strong>of</strong> not taking a whole-<strong>of</strong>-dome<br />

approach to avoiding deforestation or degradation is too high, and could jeopardise the success and<br />

longevity <strong>of</strong> any interventions designed to maintain forest cover.<br />

Restoring and Protecting Peat Swamp Forests<br />

Most <strong>of</strong> the peat swamp forests in Indonesia and elsewhere in S.E Asia were cleared by first<br />

constructing large canals which intersected across the ‘peat dome’ and drained the water to adjacent<br />

streams. This then allowed access to the area for logging and subsequent clearing and burning. Many<br />

<strong>of</strong> the peat swamp forests which were logged but not cleared, also had canals constructed for the<br />

extraction <strong>of</strong> the logs. These canals were usually small and excavated by chain saw and were<br />

particularly prevalent for illegal logging activities in logged-over or expired timber production<br />

concessions. See Figure 2.<br />

One <strong>of</strong> the best known examples <strong>of</strong> peatland degradation was the ‘Mega Rice Project’ in Central<br />

Kalimantan, which cleared 1 million ha <strong>of</strong> peat swamp forest for rice production in the mid 1990s. The<br />

peatlands were found to be unsuitable for growing rice and the project was abandoned resulting in<br />

large areas <strong>of</strong> degraded and exposed peatland which continue to emit large quantities <strong>of</strong> GHGs from<br />

decomposition and annual fires. The Indonesian Government is now committed to restoring these<br />

deforested and degraded peat lands (Presidential Instruction 2007).<br />

In order to maintain existing peat swamp forests requires blocking the canals to further restrict the<br />

drainage and the oxidation and to protect the forests from further fire and exploitation. The degraded<br />

peat swamp forests can be hydrologically restored by raising the water table back to original levels (by<br />

blocking the canals previously used to drain them) and reintroducing vegetation to hasten the<br />

rehabilitation and hydrological restoration process.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 349<br />

CO2<br />

Canals<br />

Source: Delft Hydraulics<br />

Figure 2. Drainage <strong>of</strong> the peat dome<br />

River<br />

In order to maintain existing peat swamp forests requires blocking the canals to further restrict the<br />

drainage and the oxidation and to protect the forests from further fire and exploitation. The degraded<br />

peat swamp forests can be hydrologically restored by raising the water table back to original levels (by<br />

blocking the canals previously used to drain them) and reintroducing vegetation to hasten the<br />

rehabilitation and hydrological restoration process.<br />

INTRODUCTION TO KALIMANTAN FORESTS AND CLIMATE PARTNERSHIP<br />

The International Forest Carbon Initiative (IFCI) is <strong>Australia</strong>’ contribution to the global effort on<br />

reduced emissions from avoided deforestation and degradation (REDD). The IFCI aims to<br />

demonstrate that REDD can be part <strong>of</strong> equitable and effective global climate change activities, with a<br />

central element <strong>of</strong> this work developing REDD demonstration activities with Indonesia. The<br />

Kalimantan Forests and Climate Partnership (KFCP) is the first demonstration activity being<br />

implemented under IFCI through the Indonesia- <strong>Australia</strong> Forest Carbon Partnership (IAFCP), part <strong>of</strong><br />

the bilateral cooperation between <strong>Australia</strong> and Indonesia.<br />

Goal<br />

The goal <strong>of</strong> the KFCP is:<br />

• to demonstrate a credible, equitable, and effective approach to reducing greenhouse gas<br />

emissions from deforestation and forest degradation, including from the degradation <strong>of</strong> peat<br />

swamp forest lands,<br />

• inform a future international climate change framework, and<br />

• enable Indonesia’s meaningful participation in future international carbon markets.<br />

Location and Site Description<br />

The REDD demonstration site represents a complete dome and covers an area <strong>of</strong> approximately<br />

120,000 hectares in the northern part <strong>of</strong> the Ex Mega Rice Project Area (Euroconsult et al. 2008) area<br />

in Central Kalimantan (centre is approximately 2 o South and 115 o East). See Figure 3. The site is<br />

bordered by the Kapuas River to the west and south-west, and the Mantangai River to the east and<br />

south east. The demonstration site lies completely within Kapuas District. Kapuas District is<br />

administered from Kuala Kapuas, some 100 km by river to the south.<br />

Demography<br />

The area has a low population density relative to much <strong>of</strong> Indonesia, with most villages located along<br />

the Kapuas River. Around a dozen villages are likely to have some connection (either currently or in<br />

the past) with the demonstration site, primarily for the use <strong>of</strong> natural resources from the forest or river<br />

systems, harvesting rattan or use <strong>of</strong> agricultural land. The villages primarily contain members <strong>of</strong> the<br />

local populations <strong>of</strong> Dayak. There are no plans for families to be settled in the area under Indonesia’s<br />

transmigration scheme.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 350<br />

Land use<br />

Other than community land, which consists <strong>of</strong> a strip <strong>of</strong> land (3-5 km) along the river adjacent to each<br />

village, the vast majority <strong>of</strong> the site is government owned land, under control <strong>of</strong> the Ministry <strong>of</strong><br />

Forestry, but administered by the Province.<br />

The northern half <strong>of</strong> the demonstration site is heavily forested which was logged using elevated<br />

railways and the ‘kuda kuda’ system 2 . While there are some areas <strong>of</strong> remaining forest in the southern<br />

half <strong>of</strong> the site, the majority <strong>of</strong> this area was drained using an intensive network <strong>of</strong> canals, and has<br />

largely been deforested and burnt.<br />

The land adjacent to the Kapuas River is used by communities for agricultural activities, including<br />

food crops and commercial rubber. Fishing is a major activity for both income and subsistence<br />

purposes, (including a range <strong>of</strong> native fish species and freshwater prawns). Forest areas are similarly<br />

important for a range <strong>of</strong> commercial products, including jelutung and gemor, and a range <strong>of</strong><br />

subsistence products used traditionally for buildings, food (both plants and animals), medicines and<br />

handicrafts.<br />

Illegal logging occurs across the site, although the extent and severity has reduced over recent years.<br />

Illegal logging is likely to be undertaken by people from both the local communities and others from<br />

further a field (most likely further downstream, where there is an abundance <strong>of</strong> small to medium scale<br />

sawmilling operations).<br />

Social characteristics<br />

Although the Kapuas River is used as the main transport route in the area, the impacted villages are<br />

relatively remote and have limited public infrastructure, including power supply. The level <strong>of</strong> access to<br />

and quality <strong>of</strong> both health and education services is relatively poor. The isolation also limits the range<br />

<strong>of</strong> economic opportunities available, and the deforestation <strong>of</strong> large areas <strong>of</strong> land since 1996 has greatly<br />

reduced these opportunities. Although specific information is not available on the incidence <strong>of</strong> poverty<br />

in the impacted villages, there is a high incidence <strong>of</strong> poverty in peat areas in general within Central<br />

Kalimantan.<br />

Ex-Mega Rice Project Central Kalimantan<br />

Figure 3. Location <strong>of</strong> the Kalimantan Forests and Climate Partnership demonstration area<br />

in Central Kalimantan<br />

2 This is a logging system used in peat swamp forests in SE Asia which involves logs being manually hauled in sledges on<br />

wooden skids constructed on the forest floor. The logs are usually manually skidded to small rail heads or river landings<br />

from where they are transported to the processing mills.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 351<br />

KFCP IMPLEMENTATION STRATEGY AND COMPONENTS<br />

KFCP is designed to implement four components to achieve the goal. These are: emissions reduced<br />

from avoiding forest /peatland deforestation and degradation includingcapturing and communicating<br />

knowledge from this REDD demonstration; forest and GHG baseline measurement and monitoring;<br />

developing and implementing effective incentive based payment mechanisms; and REDD<br />

management, including capacity building and readiness at the provincial and district and village levels.<br />

Component 1: Emissions Reduced from Avoided Deforestation and Degradation<br />

This is a key component <strong>of</strong> KFCP activities and provides a framework in which the other components<br />

are integrated. The community based activities implemented at the village level are designed to be<br />

coordinated with the information dissemination <strong>of</strong> REDD at the village level, changes in behaviour,<br />

peatswamp forest restoration and hydrological restoration; GHG monitoring and incentive based<br />

payment mechanisms. As the REDD is a new concept with many new ideas being introduced and<br />

tested, it will bring with it a high level <strong>of</strong> risk from the social and political and technical perspective.<br />

However it does provide an opportunity to significantly improve tropical forest management<br />

(Applegate and Smith 2000).<br />

Community engagement: Essential to the success <strong>of</strong> REDD, is the involvement <strong>of</strong> the local villages,<br />

as gaining support from these communities is a precondition for emissions reduction from the peat<br />

swamp forests. It will be essential that the local people have the ability to replace lost income from the<br />

limitations placed on them on the destructive use <strong>of</strong> forest resources under some form <strong>of</strong> tenure and<br />

land use right on village lands. Hence, the village engagement is designed to be flexible, participatory<br />

and provide for informed consent and adaptive in nature. It will ensure that improved livelihoods are<br />

compatible with REDD and <strong>of</strong>fer real income opportunities which do not exacerbate gender and social<br />

differences. The activities will be planned within the Government <strong>of</strong> Indonesia village level planning<br />

process, which enable these plans to be integrated into the higher levels <strong>of</strong> government spatial<br />

planning.<br />

Peat swamp forest rehabilitation and hydrologic restoration: The hydrologic restoration process is<br />

based on the whole-<strong>of</strong>-dome approach as the protection strategy. An intact patch <strong>of</strong> forest will be<br />

severely impacted in both the short and long term and become degraded and eventually destroyed if<br />

water levels are not maintained by blocking any canals located downstream and rehabilitating the<br />

degraded forest areas. The dams them selves will not raise levels sufficiently in the short term to<br />

inundate the higher parts <strong>of</strong> the peat dome, but this first step is essential to stop further drying <strong>of</strong> peat<br />

in the vicinity <strong>of</strong> the canals, and therefore reducing the rate <strong>of</strong> GHG emissions. Dams along the canals<br />

will assist in reducing access to areas by people which also has the added benefit <strong>of</strong> reducing the risk<br />

<strong>of</strong> fire and by increasing moisture levels <strong>of</strong> the peat close to the canals. Rehabilitation <strong>of</strong> degraded<br />

peat swamp forests down stream from intact forests, either by promoting natural regeneration or<br />

replanting with appropriate species, will ‘kick-start’ the ecological processes essential for keeping the<br />

peat surface moist and reducing water run<strong>of</strong>f from the upper sections <strong>of</strong> the dome. Reforestation<br />

(natural or artificial) in this context is not carried out for the sake <strong>of</strong> establishing more trees in the<br />

traditional sense <strong>of</strong> reforestation, but is essential for controlling water flows and starting the ecological<br />

process essential for protecting the remaining forests upstream. The same rationale exists for<br />

improving livelihoods <strong>of</strong> local communities as part <strong>of</strong> a fire prevention strategy designed to protect the<br />

forests and reduce GHG emissions. Fire is one <strong>of</strong> the main causes <strong>of</strong> deforestation and degradation, so<br />

preventing fires from starting in degraded forests has a major influence on determining the integrity <strong>of</strong><br />

the remaining peat swamp forest.<br />

The rehabilitation <strong>of</strong> the degraded peat swamp forest will involve planting on a demonstration basis to<br />

test a new range <strong>of</strong> suitable tree species . This will be scaled up to a size which allows techniques and<br />

procedures for measuring and monitoring the impacts to be credible. Rehabilitation will demonstrate<br />

various approaches (e.g., full planting, partial planting, natural regeneration) for reforesting near the<br />

central part <strong>of</strong> the dome in areas where land tenure is not in dispute. Orangutan food species and those<br />

appropriate for forest protection and conservation will be planted and promoted near the centre <strong>of</strong> the<br />

dome and species with non-timber values to be utilised by the communities towards the edge closer to<br />

their villages. Community groups or individuals will be encouraged to develop nurseries with<br />

technical assistance, and sell seedlings for the rehabilitation activity. This commerce will provide an


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 352<br />

opportunity for women to benefit financially and enhance skills in the early stages <strong>of</strong> the<br />

demonstration. Local people will be hired (as part <strong>of</strong> the incentive mechanism) to plant trees, either<br />

directly, or by arrangement with implementing partners or private contractors. Trees must be planted<br />

to coincide with dam construction to enable access for planting material and people. A rehabilitation<br />

and strategy will guide the planning process for the rehabilitation which will focus on promoting<br />

natural regeneration as well as replanting where required. The trees may have a direct economics value<br />

as non-timber forest products on designated community land, food source for orang-utans in the<br />

Protection Forests, as well as forming ‘dams’ on the peat to reduce run-<strong>of</strong>f and thus keep the peat from<br />

drying out and oxidising.<br />

Construction <strong>of</strong> appropriate dams on the canals could reduce emissions quickly by blocking access and<br />

reducing the water draining from the site and stabilizing the water table. The basic approach is to start<br />

dam construction at the centre <strong>of</strong> the dome and work outwards, spacing dams relatively closely (20 cm<br />

-50 cm vertical elevation) to avoid putting too much hydraulic pressure on each dam (this has caused<br />

dam failure in the past). Options for developing dams at different vertical intervals and cost<br />

implications will be part <strong>of</strong> the rehabilitation and restoration strategy. Methods for constructing more<br />

cost-efficient dams will be explored including ways <strong>of</strong> achieving economies <strong>of</strong> scale through bulk<br />

purchase and transport <strong>of</strong> materials and use <strong>of</strong> contractors in agreed areas.<br />

Livelihood interventions and incentives: One <strong>of</strong> the key aspects <strong>of</strong> a successful REDD mechanism<br />

will include the development <strong>of</strong> appropriate incentives to adopt sustainable land-use practices. These<br />

incentives and how the mechanisms will function are a key part <strong>of</strong> the overall strategy to mitigate<br />

against the further deforestation and degradation, as farming and land use preparation methods depend<br />

on fire for land clearing and preparation. Incentives to encourage sustainable practices will<br />

include: 1) input-based payments, which comprise payment or other direct benefits for adopting and<br />

undertaking interventions, such as dam construction, tree planting, provision <strong>of</strong> dam-building and treeplanting<br />

materials, or fire suppression; 2) performance-based payments for maintaining the<br />

interventions so as to achieve the desired results, such as maintaining dams in order to keep water<br />

levels high, protecting forest from encroachment, or reducing the incidence and extent <strong>of</strong> fire; 3)<br />

outcome-based payments commensurate with the level <strong>of</strong> GHG emissions reductions. These will<br />

initially be a proxy for a future forest carbon market which will later be based on tradeable credits in a<br />

real market.<br />

Measurement and monitoring: In order to effectively monitor the impact <strong>of</strong> the interventions on the<br />

reducing emissions and reducing the loss <strong>of</strong> peat swamp forests and their degradation, initial<br />

measurements will be undertaken on a number <strong>of</strong> parameters prior to any interventions in order for<br />

monitoring and evaluation to have a sound basis. These measurements will be also used to determine<br />

the various baselines and current emission levels and Reference Emission Level (REL). It is important<br />

to recognise what parameters need to be measured prior to the commencement <strong>of</strong> the interventions.<br />

For example, social and village baselines and GHGs emissions for determining the REL from the<br />

current area <strong>of</strong> the peat swamp forests in the KFCP area is well underway and needs to be carried out<br />

before interventions are implemented. In terms <strong>of</strong> the REL, KFCP will determine the carbon content<br />

<strong>of</strong> the various components <strong>of</strong> above and below ground component <strong>of</strong> the peat swamp forests and the<br />

methodologies by which the changes in avoided emissions and changes in forest quality and area will<br />

be estimated. In terms <strong>of</strong> GHGs, the KFCP will undertake a scanning Light Detection and Ranging<br />

(LIDAR)study prior to the interventions to determine the elevation <strong>of</strong> the peat in the peat swamp<br />

forests, but will then need to undertake research into carbon content <strong>of</strong> the below (peat) and above<br />

ground forest biomass (no data at present), carbon content <strong>of</strong> peat at different depths and position in<br />

the dome (determined by the forest type when peat was developed), bulk density which varies with<br />

depth and position in the dome, size <strong>of</strong> the project area and peat depth, in order to determine estimate<br />

avoided emissions from the interventions and the impact <strong>of</strong> the interventions.<br />

KFCP as a demonstration activity is designed to capture the knowledge and lessons learned and<br />

providing this information to a range <strong>of</strong> stakeholders and policy makers. The audience for this<br />

information are local communities, the Government <strong>of</strong> Indonesia and <strong>Australia</strong> and the international<br />

community such as the UNFCCC. These groups which will be kept informed by implementation <strong>of</strong><br />

both a communications strategy and a knowledge capture strategy. The work will include


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 353<br />

identification <strong>of</strong> the KFCP audiences, their information needs and the best media for communicating<br />

with them. The strategy to capture knowledge will be developed to collect, and store the information<br />

coming from multiple sources within and outside the KFCP and then disseminate the “lessons learned”<br />

and other knowledge-based results to target audiences.<br />

Component 2: Emissions Estimation and Monitoring<br />

The current international agreements concerned with forest carbon emissions do not cover reductions<br />

<strong>of</strong> emissions from deforested and degraded peat swamp forest, so KFCP will endeavour to provide<br />

information on the methodologies to estimate and monitor emissions from peat and peat swamp forests<br />

to enable them to be included in a future climate change agreement, and whether it would be<br />

incorporated into future action on REDD. Hence this work will be focused improving and informing<br />

the international debate and providing lessons learned through contribution to UNFCCC discussions<br />

through 2009, in areas such as: research required to develop the methodologies required for estimating<br />

changes in GHG emissions from land use changes and measurement and monitoring <strong>of</strong> peatland<br />

characteristics, GHG emissions. To facilitate this process, the KFCP has convened a group <strong>of</strong><br />

specialists for establishing a Project (KFCP) specific emissions baseline (REL) and determine<br />

scientific methodologies and implementation guidelines for the measurement and monitoring changes<br />

in peat swamp forests.<br />

To facilitate the implementation <strong>of</strong> this component, the work is divided into two distinct, but<br />

interrelated tasks:<br />

• developing, testing, and validating a GHG emissions baseline and monitoring system that<br />

estimates emissions changes against interventions and will be accepted as scientifically<br />

valid in climate change negotiations (i.e. methods to be between Tier 2 and Tier 3); and<br />

• operationalise GHG estimating and monitoring through remote sensing and direct ground<br />

measurement in ways that will meet the requirements <strong>of</strong> a future REDD carbon market<br />

and can be integrated into the Indonesian National Carbon Accounting System (NCASI).<br />

Component 3: Practical and Effective REDD GHG Payment Mechanisms<br />

One <strong>of</strong> the crucial aspects <strong>of</strong> a successful compliance based REDD scheme is the development <strong>of</strong> an<br />

equitable and effective payment mechanism for the potential REDD credits. The KFCP provides the<br />

opportunity and flexibility to test different approaches, guided by experience with payment<br />

mechanisms for other environmental services globally and other services within Indonesia. Testing<br />

payment systems will therefore benefit from having an overarching mechanism to pay for emission<br />

reduction incentives and have to be closely linked to monitoring <strong>of</strong> GHG emissions and socioeconomic<br />

impact in order to verify reductions, create a system that is credible, and inform the<br />

development <strong>of</strong> NCASI.<br />

The KFCP will demonstrate payment mechanisms, and in doing so, will inform the<br />

development <strong>of</strong> a national REDD system by; testing accountable, transparent and equitable<br />

payment mechanisms that create positive incentives for achieving emissions reductions;<br />

calculates the cost involved in implementation, impact mitigation, monitoring and regulating<br />

(including nominal costs per tonne <strong>of</strong> CO2), and compares whether KFCP implementation costs are<br />

lower than opportunity costs, the potential market value <strong>of</strong> the emission reductions it has realised<br />

(viability) and develops arrangements for revenue allocation between all stakeholders that takes<br />

sufficient account <strong>of</strong> equity considerations as well as capacity to address causes <strong>of</strong> deforestation. The<br />

mechanism will also involve keeping a balance between cost recovery and an attractive investment<br />

climate for REDD activities with effective incentives to ensure that communities are motivated to<br />

implement and a change in behavior that supports REDD.<br />

Component 4: REDD Management/Technical Capacity and Readiness Developed<br />

KFCP will integrate REDD into governance arrangements at the province and district levels by<br />

assisting to develop management institutions, a legal framework, and provide a technical capacity to<br />

support demonstration activities which will allow local integration into a REDD carbon market. The<br />

current levels <strong>of</strong> support at the province level for a REDD Working Group will be further supported at


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 354<br />

the district level through consultation and awareness raising. The Ministry <strong>of</strong> Forestry and the EMRP<br />

Master Plan support the development <strong>of</strong> management units based on hydrological characteristics. The<br />

Forest Management Unit concept which would encompass the KFCP site provides a promising way to<br />

improve the institutional and technical capacity within the district to manage REDD and improve land<br />

management.<br />

CHALLENGES AND NEXT STEPS<br />

The multidimensional partnerships involved in a REDD demonstration activity such as KFCP is a<br />

management challenge. This becomes even more interesting by the fact that KFCP is essentially a<br />

community development project superimposed by a science and learning, socialization, policy<br />

development, and multiple audience communications project. In addition, there are the ever present<br />

surrounding international issues; additionality where REDD will be required to reduce deforestation<br />

below a “business as usual” baseline; leakage which involves avoiding increased deforestation<br />

elsewhere in the region as a result <strong>of</strong> REDD, and permanence, which involves maintaining GHG<br />

reductions for many years.<br />

KFCP is also aware that as REDD is new to many, we are on a steep learning curve, so there is a need<br />

to share with others what is learned. The REDD process does not work in isolation and needs to<br />

consider a holistic approach, with the host (community, national and sub-national governments and<br />

other stakeholders) leading the process. Finally the international community may be able to assist in<br />

the process as the outcome is designed to inform further UNFCCC processes leading to a post 2012<br />

climate change agenda.<br />

ACKNOWLEDGEMENT<br />

This paper is based on the Design for the Kalimantan Forests and Climate Partnership in Indonesia funded<br />

through the International Forest and Carbon Initiative. Contributions to the Design were provided by consultants<br />

to AusAID, and the Government <strong>of</strong> <strong>Australia</strong> and Indonesia. The views expressed in this paper are solely those<br />

<strong>of</strong> the author.<br />

REFERENCES<br />

Applegate, G. and Smith, J. (2000), Could Trade in Forest Carbon Contribute to Improved Tropical Forest<br />

Management. Centre for International Forestry Research (CIFOR).<br />

Bappenas. (2006), National Strategy and Action Plan for Sustainable Management <strong>of</strong> Peatlands,<br />

Euroconsult, Mott MacDonald and Deltares (2008), Master Plan for the Rehabilitation and Revitalisation <strong>of</strong> the<br />

Ex Mega Rice project area in Central Kalimantan, Indonesia, October 2008.<br />

Hooijer, A., Silvius, M., Wösten, H. and Page, S. (2006), Peat CO2: Assessment <strong>of</strong> Co2 emissions from drained<br />

peatlands in SE Asia. Delft Hydraulics and Wetlands International.<br />

Jaenicke, J., Rieley, J.O., Mott, C., Kimman, P. and Siegert, F. (2008), Determination <strong>of</strong> the amount <strong>of</strong> carbon<br />

stored in Indonesia’s peatlands. Geoderma 147, 151-159.<br />

Page, S. E. and Banks, C. (2007), Tropical peatlands: distribution, extent and carbon storage- uncertainties and<br />

knowledge gaps. Peatlands International 2007, (2), 6-27.<br />

Page, S.E., Siegert, F., Rieley, J.O., Boehm, H.-D.V. and Jaya, A. (2002), The amount <strong>of</strong> carbon released from<br />

peat and forest fires in Indonesia in 1997. Nature 420:61-65.<br />

Parish, F., Sinn, A., Charman, D., Joosten, H., Minayeva, T. and Silvius, M. (Eds), (2007), Assessment <strong>of</strong><br />

Peatlands, Biodiversity and climate change: Executive Summary. Global Environment Centre, Kuala<br />

Lumpur and Wetlands International, Wageningen.<br />

Presidential Instruction 2007 (No 2), Acceleration <strong>of</strong> the Rehabilitation and Revitalisation <strong>of</strong> the Ex-Mega Rice<br />

Project Area. Government <strong>of</strong> Indonesia, 2008.<br />

Uryu, Y., Mott, C., Foead, N., Yulianto, K., Budiman, A., Takakai, F., Nursamsu, Sunarto, Purastuti, E.,<br />

Fadhli, N., Hutajulu, C.M.B., Jaenicke, J., Hatono, R., Siegert, F., and Stüwe, M. (2008), Deforestation,<br />

Forest Degradation, Biodiversity Loss and CO2 Emissions in Riau, Sumatra, Indonesia. WWF Indonesia,<br />

Technical Report Jakarta, Indonesia, 74pp.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 355<br />

REDUCING EMISSIONS FROM<br />

DEFORESTATION AND DEGRADATION (REDD)<br />

IN LAO PEOPLE’S DEMOCRATIC REPUBLIC<br />

Majella Clarke 1<br />

ABSTRACT<br />

This paper sets out to synthesize the current activities and strategies for Reducing<br />

Emissions from Deforestation and Degradation (REDD) in Lao PDR, and discusses the<br />

potential implications <strong>of</strong> REDD strategies for rural livelihoods and sustainable<br />

development. The drivers <strong>of</strong> deforestation and degradation within the country are<br />

discussed, and the correlation between canopy density thresholds and deforestation drivers<br />

is reviewed. It is argued that canopy density threshold has an important impact on the<br />

potential allocation <strong>of</strong> REDD emission reduction credits, and plays an important role in the<br />

REDD strategy formulation process. Previous Lao forest strategies for avoiding<br />

deforestation in Lao PDR are discussed, noting the improvements made and challenges<br />

ahead.<br />

INTRODUCTION<br />

The objective <strong>of</strong> this paper is to present a synthesis <strong>of</strong> the challenges and opportunities for capacity<br />

development and strategy formulation for reducing emissions from deforestation and degradation<br />

(REDD) in Lao PDR. The paper will introduce the present forest landscape in Lao PDR and discuss the<br />

potential and recognized drivers <strong>of</strong> deforestation and degradation within the country. The correlation<br />

between canopy density thresholds and drivers <strong>of</strong> deforestation will also be reviewed in this respect. The<br />

paper will argue that the canopy density threshold parameter has an important impact on the potential<br />

REDD emission reduction credits allocated and plays an important role in the REDD strategy formulation<br />

process. Finally, a presentation <strong>of</strong> REDD strategies to overcome the main drivers <strong>of</strong> deforestation and<br />

degradation in Lao PDR will be discussed.<br />

Lao PDR has one <strong>of</strong> the most intact forest areas in South East Asian region with over 40% <strong>of</strong> its land area<br />

classified as forest. Some 80% <strong>of</strong> the population live in rural areas and are dependent on forest functions<br />

and resources for their livelihoods. Non-Timber Forest Products (NTFPs) are crucial for meeting<br />

subsistence needs and perform an important role in providing food security to the majority <strong>of</strong> rural Laos.<br />

While there have been few studies conducted, Foppes and Ketpanh (1997) found that NTFPs provide on<br />

average 55% <strong>of</strong> family cash income compared with 15% <strong>of</strong> farm income and 30% <strong>of</strong> household incomes<br />

are derived from livestock, while income generation from NTFPs has slightly declined (more recent<br />

estimates state 50%, Ingles (2006)), it is still representative <strong>of</strong> their important role in food security and<br />

income generation in rural livelihoods in Laos.<br />

As part <strong>of</strong> efforts to conserve its forest resources and biodiversity, Lao PDR established a National<br />

Protected Area (NPA) system designed to preserve natural resources, and protect nature and preserve the<br />

natural landscape. These NPAs are now National Biodiversity Conservation Areas (NBCAs) and cover<br />

22% <strong>of</strong> the country’s land area (that is, 5.3 million ha) and contains 20 NBCAs in addition to two<br />

corridors.<br />

Production forests play an important role in timber export revenues contributing to some 15% <strong>of</strong> GDP,<br />

and to meet ambitious forest sector targets, the government intended to establish 400,000 ha <strong>of</strong><br />

plantations from 1993. However, only some 57,000 ha had so far been planted as <strong>of</strong> 2005 and many <strong>of</strong> the<br />

plantations were replanted several times because <strong>of</strong> management failure and fire, and most performed<br />

poorly in strict economic terms (World Bank (2005). This has however, recently improved with large<br />

foreign investments in jatropa and teak plantations, as well as rubber, c<strong>of</strong>fee and banana plantations<br />

within the agriculture sector.<br />

1 Dr Majella Clarke, Department <strong>of</strong> Forestry, Vientiane, Lao PDR. Email: majella.clarke@indufor.fi.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 356<br />

However, expanding shifting cultivation areas with reduced fallow periods, hydroelectric, energy and<br />

mining projects, unsustainable timber extraction techniques and forest land conversion are some <strong>of</strong> the<br />

potential and recognized driving causes <strong>of</strong> deforestation and forest degradation in Lao PDR. In addition,<br />

little research and capacity development has addressed drivers <strong>of</strong> deforestation in the past in Lao PDR<br />

and therefore monitoring such drivers, and establishing historical trends is already creating a great<br />

challenge within the REDD R-PLAN preparation.<br />

FOREST ASSESSMENT IN LAO PDR<br />

A brief review <strong>of</strong> the history <strong>of</strong> the forest inventory illustrates why Lao PDR has accumulated a variety <strong>of</strong><br />

methods through a number <strong>of</strong> institutional arrangements over the past century for monitoring its forests.<br />

The first forest assessments were conducted between 1909 and 1943 by the French colonists but it was<br />

almost 30 years before another forest assessment would be attempted by the Mekong River Commission<br />

in 1970. In 1974, Cooperation between the Governments <strong>of</strong> Lao PDR and Canada provided an assessment<br />

<strong>of</strong> forest cover but only in some provinces. It was not until 1982 that the first national reconnaissance<br />

survey was conducted on forests with the assistance <strong>of</strong> the USSR. Between 1989 and 1990 a national<br />

level forest inventory at the provincial level was completed. In 1990, the Mekong River Commission did<br />

a second assessment <strong>of</strong> forest cover. This was followed shortly by the second national reconnaissance<br />

survey with the assistance <strong>of</strong> the Swedish International Development Agency (SIDA). During 1995-<br />

2000, the project FORMACOP did a number <strong>of</strong> provincial inventories in production forests supported by<br />

the Finnish Government and the World Bank. In 2002, the first digital forest assessment was produced<br />

with the assistance <strong>of</strong> SIDA and the Japanese International Cooperation Agency (JICA), in which an<br />

assessment <strong>of</strong> forest cover and land use for 1982-1992-2002 was produced and published in 2005.<br />

Finally, the Sustainable Forestry and Rural Development Project (SUFORD) supported by the Finnish<br />

Government and the World Bank, produced a number <strong>of</strong> inventories for production forests between 2004<br />

and 2008, with ongoing support for an expansion <strong>of</strong> further inventories planned for 2009-2012<br />

(Chanhsomone, 2008).<br />

Such efforts have led to a fragmentation in the collection, storage and publication <strong>of</strong> forest inventory data.<br />

The most reliable source <strong>of</strong> forestry data and which the majority <strong>of</strong> governmental decisions are based<br />

upon is the 1982-1992-2002 Assessment <strong>of</strong> Forest Cover and Land Use published by the Department <strong>of</strong><br />

Forestry (DoF) in 2005. The preliminary reference scenarios within this paper are also based on the<br />

assessment, as it is the most recent national level assessment <strong>of</strong> forest cover and land use and has also<br />

been digitalized.<br />

The current forest definition in Lao PDR is in line with the definition parameters used to report to FAO<br />

Global Forest Resource Assessment, and meets the UNFCCC COP decision no. 19/CP9 (A/R CDM).<br />

Based on the national forest resource assessment carried out between 1992 and 2002 the tables below<br />

illustrate the forest types and their respective areas in Lao PDR.<br />

Table 1. Forest Vegetation types and Distribution in Lao PDR in 2002<br />

“Current Forest” Vegetation Type Area (1000 ha) % <strong>of</strong> land area<br />

Dry Dipterocarp 1,317.2 5.5<br />

Lower Dry Evergreen 65.0 0.2<br />

Upper Dry Evergreen 1,387.9 5.9<br />

Lower Mixed Deciduous 881.0 3.7<br />

Upper Mixed Deciduous 5,499.5 23.2<br />

Gallery Forest 28.2 0.1<br />

Coniferous 89.1 0.4<br />

Mixed coniferous and broadleaved 525.8 2.2<br />

Wood Plantation 40.0 0.2<br />

Total 9,824.7 41.5<br />

Source: DoF (2005)<br />

“Current Forests” include natural forests and forest plantations and is applicable to land with tree canopy<br />

density <strong>of</strong> more than 20%, on an area <strong>of</strong> more than 0.5 ha, in which the trees should be able to reach a<br />

minimum height <strong>of</strong> 5 metres. A change in the canopy density parameter to 10% has implications and will


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 357<br />

be reviewed later in this article. Table 1 presents the forest vegetation types and respective areas included<br />

in the definition <strong>of</strong> current forests.<br />

Potential forests refer to previous forest areas where the canopy density has been reduced below 20% for<br />

some reason (logging or shifting cultivation), and also include bamboo, old shifting cultivation areas<br />

(young secondary forests) and temporary unstocked areas. Table 2 presents the forest vegetation types<br />

and respective areas included in the definition <strong>of</strong> potential forests.<br />

Table 2. Potential forest vegetation types and distribution in Lao PDR 2002<br />

Potential Forest Vegetation Type Area (1000 ha) % <strong>of</strong> land area<br />

Bamboo 539.0 2.3<br />

Unstocked 10,096.3 42.6<br />

Old shifting cultivation areas 516.9 2.2<br />

Total 11,152.2 47.1<br />

Source: DoF (2005)<br />

Other Wooded Areas are areas with trees where site conditions are so poor that the canopy density can<br />

never be expected to exceed 20%. This includes savannah forests and heath, and stunted or scrub forests.<br />

(Table 3)<br />

Table 3. “Other Wooded Areas” and distribution in Lao PDR 2002<br />

Other Wooded Area Vegetation Types Area (1000 ha) % <strong>of</strong> land area<br />

Savanah/Open Wood Lands 94.4 0.4<br />

Heath, Scrub Forests 192.1 0.8<br />

Total 286.5 1.2<br />

While current and potential forest areas may look optimistic at first glance compared with other countries<br />

within the South-East Asian region, deforestation and forest degradation have reduced the forest cover<br />

from about 16-17 million ha (70% <strong>of</strong> land area) in 1940 , to 9.8 million ha by the completion <strong>of</strong> the latest<br />

forest assessment in 2002. One study put forest clearing in 2001 at a rate <strong>of</strong> 300,000 ha per annum<br />

(AusAID 1996). The government attributed the deforestation rate <strong>of</strong> about 200,000 ha per annum to<br />

shifting cultivation and a further 100,000 ha per annum to activities other than shifting cultivation (STEA<br />

2000). The following section attempts to give a brief overview <strong>of</strong> the potential and recognised drivers <strong>of</strong><br />

deforestation and forest degradation in Lao PDR.<br />

POTENTIAL AND RECOGNISED DRIVERS OF DEFORESTATION IN LAO PDR<br />

Hydropower damn construction<br />

Hydropower is the most abundant and cost effective energy source in the Greater Mekong River Basin<br />

with theoretical hydroelectric potential <strong>of</strong> about 18,000 MW in Lao PDR (AusAID, 1996). However, as<br />

<strong>of</strong> 2001, less than 5 % (624 MW) <strong>of</strong> the country’s potential for hydroelectric power had been developed<br />

(DANIDA, 1998). As <strong>of</strong> 2007, Lao PDR had 10 hydropower plants operational, with an additional 12<br />

sites planned to be ready to export energy to Thailand by 2015. According to the latest forest assessment,<br />

31 forest areas for hydropower dam construction have been identified as potential sites in line with the<br />

government policy to respond to regional demand, amounting to 140,635 ha altogether, (DOF, 2005). The<br />

hydropower sites are drivers <strong>of</strong> deforestation as areas designated for development are financially<br />

attractive logging sites since no regeneration <strong>of</strong> forests is required after logging, and no other area to<br />

replant and <strong>of</strong>fset what is lost is required to be regenerated.<br />

.Mining and Mineral Exploration<br />

To date, there are no studies which analyse the relationship between mining and deforestation in Lao<br />

PDR. In 2000, although Lao PDR has good mining potential, mining activities only accounted for about<br />

1% <strong>of</strong> GDP in 2000 (STEA,(2000) although recent figures now suggest that the mining sector is a major<br />

contributor to GPD. So far, Laos has approved 181 mining projects by a total <strong>of</strong> 118 companies, 74 <strong>of</strong><br />

which are foreign owned. The Lao government reportedly plans to approve more than ten projects<br />

between 2008 and 2010. The majority <strong>of</strong> those projects will be bauxite and aluminium exploration and<br />

mining in Champasak and Attapeu provinces (Bird, 2008). Given the rich unexploited mineral resources


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 358<br />

in Lao PDR, including tin, coal, iron, copper, gold, gypsum, zinc, sapphire and other minerals, and the<br />

role mining activities have played in deforestation elsewhere, mining activities, both legal and illegal,<br />

could be a potential driver <strong>of</strong> deforestation in Lao PDR, and which should also be further researched.<br />

Slash and Burn for Land Conversion<br />

It is generally agreed that slash and burn systems can be sustainable with long fallow periods when the<br />

population densities are low (Roder, 2001). For Lao PDR, slash and burn agriculture is <strong>of</strong> particular<br />

importance as it is a major land use practice which involve more than 150,000 households, or 25% <strong>of</strong> the<br />

rural population (Lao PDR, 1999). Productivity <strong>of</strong> land used in slash and burn agriculture is extremely<br />

low, and even lower if it takes into consideration fallow land. Even with its low productivity, slash and<br />

burn is still practiced widely in Northern Laos today. With many areas still not easily accessible by roads<br />

and with no communication infrastructure in most rural areas, providing alternatives for income<br />

generation and food security remains a challenge. Moreover, the practice <strong>of</strong> slash and burn could be<br />

difficult to eradicate because <strong>of</strong> its significance to the traditions and culture <strong>of</strong> the Northern upland<br />

people, even though it is part <strong>of</strong> the environmental strategy 2020 to phase out all slash and burn practices<br />

by 2010.<br />

However, due to the potential emissions which can be reduced from phasing out slash and burn, it has<br />

recently become the focus <strong>of</strong> a number <strong>of</strong> REDD pilots and projects. To understand the scale <strong>of</strong> slash and<br />

burn, whether for shifting cultivation or illegal land conversion, a map produced by the University <strong>of</strong><br />

Maryland, in Thomas (2008) using MODIS satellite fire data for Laos 2007 is shown below, illustrating<br />

some 200,000 fire detections in 2007. The darker the dot, the higher the fire intensity and number <strong>of</strong><br />

burning days detected.<br />

Source: University <strong>of</strong> Maryland MODIS Fire Data in Thomas (2008)<br />

There are other drivers <strong>of</strong> deforestation in Lao PDR, including conversion to rubber plantations,<br />

resettlement, legal and illegal mining, unsustainable timber harvesting, illegal logging and illegal land<br />

conversion. The lack <strong>of</strong> studies on the impact <strong>of</strong>, these drivers means that accurate up to date statistics<br />

are hard to find, and such drivers are given only a general mention in most <strong>of</strong> the relevant government<br />

documents (DOF,2005, STEA, 2000, World Bank, 2005).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 359<br />

DRIVERS OF FOREST DEGRADATION<br />

Shifting Cultivation with Reduced Fallow Cycles<br />

Shifting cultivation practices and forest fires are still the main cause <strong>of</strong> forest degradation, particularly in<br />

the north (DOF 2005). There, the lack <strong>of</strong> flat land for permanent and stable agriculture contributes to the<br />

traditional practice <strong>of</strong> shifting cultivation. Generally, yields tend to be low, increasing pressure on<br />

cultivation areas for rice self-sufficiency. While the State <strong>of</strong> the Environment Report 2001 identified<br />

600,000 ha <strong>of</strong> area as being under shifting cultivation, the government attributed deforestation and<br />

degradation due to shifting cultivation at about 200,000 ha per annum in 2001 (STEA 2000). Of more<br />

concern is the declining trend in the percentage <strong>of</strong> “current forest” in the North <strong>of</strong> Lao over the past 20<br />

years, while there has been an increasing trend in the “potential forest” areas (which incorporate<br />

unstocked forests caused by shifting cultivation (Figure1).<br />

%<br />

Figure 1. Forest and land distribution in Northern Lao PDR<br />

Source: Department <strong>of</strong> Forestry (2005)<br />

A recent concern for shifting cultivation practices in Lao PDR, and particularly in the northern region, is<br />

that traditionally a 12-16 year cultivation cycle is required for the recovery <strong>of</strong> soil fertility. However, in<br />

recent years the fallow period has been shortening and slash and burn agricultural systems with fallow<br />

periods <strong>of</strong> only 1-3 years are now common in Luang Prabang province, where upland rice farming is the<br />

main livelihood. (Kiyono et al. 2007). Some <strong>of</strong> the reasons for the shortening fallow and increased land<br />

pressure stem from Government policies for forest management which mandated the conversion <strong>of</strong> some<br />

fallow slash and burn areas to conservation forests, which in turn reduced the area under the slash and<br />

burn system. Teak plantations established in the 1990s also decreased the amount <strong>of</strong> arable land in the<br />

northern region (Roder et al. 1995).<br />

Unsustainable exploitation <strong>of</strong> timber and Non-Timber Forest Products is another cause <strong>of</strong> forest<br />

degradation. For example, approximately 80% <strong>of</strong> domestic energy consumption for cooking is based on<br />

fuel wood. The estimated volume <strong>of</strong> annual fuel wood consumed by local communities is 4-5 million<br />

m 3 /yr leading to excessive gathering <strong>of</strong> fuel wood, tree felling, and further pressure on remaining forests<br />

(Phanovong 1997). Excessive timber harvesting occurs most seriously in the central and southern regions<br />

<strong>of</strong> Lao PDR with many forest concession areas being exploited at a rate greater than 15 m 3 /ha, which has<br />

been set as the sustainable rate <strong>of</strong> timber extraction (STEA 2000). This has led to forest fragmentation<br />

and increased area classified as temporary unstocked forests.<br />

CURRENT REDD ACTIVITIES IN LAO PDR<br />

In 2007, the Government <strong>of</strong> Lao PDR submitted an expression <strong>of</strong> interest to the World Bank’s Forest<br />

Carbon Partnership Facility (FCPF) to receive support for developing the Readiness Project Idea Note (R-<br />

PIN) and subsequently Readiness Plan (R-PLAN) for REDD preparation. Activities supported under the<br />

R-PLAN include stakeholder consultations, developing a national REDD strategy, designing a REDD


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 360<br />

implementation framework, developing the required reference scenarios, and establishing a monitoring,<br />

reporting and verifications system at the national level for changes in forest cover and carbon stocks.<br />

With respect to any REDD project implementation, forest carbon monitoring can be the most critical<br />

exercise in assessing the amount <strong>of</strong> carbon and biomass that forests store. Forest carbon monitoring<br />

requires remote sensing capacity, and the activities used to monitor general forest parameters are fairly<br />

new to Lao PDR, with the first digital forest assessment accomplished in 2002. In addition to the need for<br />

a high level <strong>of</strong> expertise, remote sensing activities for forest carbon assessments require expensive inputs,<br />

such as satellite data, field verification, aerial surveys and sometimes the establishment <strong>of</strong> permanent<br />

sample plots, altogether.<br />

THE IMPORTANCE OF THE CANOPY DENSITY THRESHOLD<br />

AS A PARAMETER IN THE FOREST DEFINITION<br />

The canopy density threshold is the most important parameter in the forest definition when accounting for<br />

carbon stocks, as it defines the boundaries <strong>of</strong> forest cover. Changes in this parameter are currently under<br />

review in Laos. The proposed change is from a 20% canopy density threshold, which now defines<br />

“current forests”, to 10%, which includes “current forests” and “potential forests”. This would be a<br />

significant change for Lao PDR.. On one hand, a change from 20% to 10% would greatly enlarge the area<br />

eligible for REDD emission reduction credits from about 9 million ha to 15.5 million ha, taking into<br />

account an average linear deforestation trend from 2002 forest assessment estimates. This change would<br />

also minimize the risk <strong>of</strong> degradation emissions; for example, if a forest with 40% canopy cover is<br />

degraded to 15% canopy cover, under a lower threshold, it is not clear that any penalties would apply.<br />

However, under the 20% threshold penalties would apply, as it would be classified as deforestation. A<br />

definition for degradation accepted within the international climate change framework has still to be<br />

agreed upon.<br />

Figure 2 illustrates the difference in forest cover area between a 10% canopy density threshold and a 20%<br />

canopy density threshold, and gives an idea <strong>of</strong> what future reference scenarios required under REDD may<br />

look like. The deforestation rates are also different, and under the current REDD and reference scenario<br />

models used in the voluntary market projections with high potential deforestation rates against a baseline<br />

generate more emission reduction credits.<br />

Million ha<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

1982<br />

1984<br />

Figure 2. Forest cover loss average linear trend<br />

1986<br />

1988<br />

1990<br />

1992<br />

1994<br />

1996<br />

1998<br />

2000<br />

2002<br />

2004<br />

2006<br />

2008<br />

2010<br />

2012<br />

2014<br />

2016<br />

2018<br />

2020<br />

2022<br />

2024<br />

Forest Cover Loss Average Linear Trend<br />

Av 0,003%/yr<br />

Av 1,26%/yr<br />

2026<br />

2028<br />

2030<br />

2032<br />

2034<br />

2036<br />

CD 20%<br />

CD 10%<br />

2038<br />

2040<br />

2042<br />

This is illustrated in Figure 3 which shows that under the assumption that deforestation for the relevant<br />

canopy threshold can be stabilized from 2012, the threshold which has the highest deforestation rate also<br />

receives more emission reduction credits. This has been one <strong>of</strong> the scrutinizing factors within the REDD


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 361<br />

framework, as it is perceived as a mechanism which rewards the “bad guys” who have historically high<br />

rates <strong>of</strong> deforestation, and gives little reward to nations which have had generally low historical<br />

deforestation rates. The following two diagrams are based only on the data from the Forest Cover and<br />

Land Use Assessment between 1982-2002, DoF (2005).<br />

So while a decrease in the canopy density threshold from 20% to 10% will greatly enlarge the area<br />

available for REDD credits, it could potentially reduce the emission reduction credits allocated based on<br />

the current reference scenarios because the 10% canopy density threshold has a much lower historical<br />

deforestation rate than the 20% canopy density threshold, and this is illustrated by the area under the<br />

baseline, above the projected forest loss between 2012-2042. Note that the baseline assumes that<br />

deforestation rate goes to 0% in 2012 and forest cover is maintained throughout the reference period. This<br />

is particularly optimistic baseline was designed to show the maximum possible benefits <strong>of</strong> REDD. More<br />

realistic and up to date reference scenarios and baselines will be developed within the R-PLAN process<br />

and will most certainly have a deforestation rate for canopy density <strong>of</strong> 20%, above 0% deforestation,<br />

while the canopy density threshold <strong>of</strong> 10% would most likely be very close to 0% deforestation.<br />

Million ha<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

1982<br />

1984<br />

1986<br />

1988<br />

1990<br />

1992<br />

1994<br />

1996<br />

1998<br />

2000<br />

2002<br />

2004<br />

2006<br />

2008<br />

2010<br />

CD 20% CD 10%<br />

2012<br />

2014<br />

2016<br />

2018<br />

2020<br />

2022<br />

2024<br />

2026<br />

2028<br />

2030<br />

2032<br />

2034<br />

2036<br />

2038<br />

2040<br />

2042<br />

Figure 3. Comparison <strong>of</strong> emission reduction credits available under assumed baseline<br />

The canopy density thresholds will also have an influence on the types <strong>of</strong> REDD strategies formulated, as<br />

we can see from the graphs and optimistic baseline scenarios, that the canopy density threshold <strong>of</strong> 20% is<br />

a more risk adverse option compared with the canopy density threshold <strong>of</strong> 10%. However, one <strong>of</strong> the<br />

most important areas <strong>of</strong> REDD is yet to be clarified – that is the inclusion <strong>of</strong> emissions from forest<br />

degradation. Most <strong>of</strong> the forest cover area between canopy density threshold 20% and 10% could<br />

potentially represent degraded forests.<br />

According to the latest forest cover assessment <strong>of</strong> 2002, about 10 million ha, or some 42% <strong>of</strong> the land<br />

area in Laos, is classified as unstocked forest, which is one <strong>of</strong> the three categories within the potential<br />

forest definition (more than 10%, less than 20% canopy threshold). The national forest inventory defines<br />

unstocked forest areas as:<br />

“Unstocked forest areas are previously forest areas in which crown density has been reduced to less<br />

than 20% because <strong>of</strong> logging, shifting cultivation or other heavy disturbance. If the area is left to grow<br />

again undisturbed, it can become a forest again. Abandoned ray (or shifting cultivation areas) and<br />

disturbed stands with a crown density <strong>of</strong> less than 20% should also be classified as unstocked forest<br />

areas.” (DOF, 2005)


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 362<br />

Given that 42% <strong>of</strong> the total land area <strong>of</strong> Lao PDR is classified as “unstocked forest area”, emissions from<br />

degradation could be quite significant and will be one <strong>of</strong> the focal areas in the REDD strategy,<br />

Nonetheless, it highlights the importance <strong>of</strong> reaching an internationally acceptable definition for forest<br />

degradation, so that such emissions can be accounted for.<br />

STRATEGIES FOR AVOIDING DEFORESTATION<br />

Deforestation and forest degradation are not new phenomena in Lao PDR and forest strategies have been<br />

drafted in the past to address such problems. This section will review the past forest strategies <strong>of</strong> Lao<br />

PDR noting the improvements made and challenges they face. Overall, the Government <strong>of</strong> Lao<br />

recognized the rapidly deteriorating forest resource situation and has set development targets for 2005,<br />

2010 and 2020. The targets include stabilizing shifting cultivation by 2005 and phasing it out completely<br />

by 2010. Tree plantations are also strongly promoted, and the classification, delineation and special<br />

management for protection forests, production forests and conservation forests has commenced (MAF<br />

2005).<br />

Establishing National Biodiversity Conservation Areas (NBCAs)<br />

Through Prime Ministerial Decree in 1993, 18 NBCAs were set aside specifically for conservation.<br />

Logging, collecting ,NTFPs, excavation or mining, expansion <strong>of</strong> shifting cultivation, exploitation <strong>of</strong><br />

cultural or historical assets, activities that degrade the environment such as the use <strong>of</strong> explosives,<br />

chemicals, poisons, and burning are prohibited on these areas. Since then, several more areas have been<br />

added so that there are currently 20 NBCAs and 2 green corridors, covering 5.3 million ha or about 22%<br />

<strong>of</strong> the total land area, under some degree <strong>of</strong> protection. However, management <strong>of</strong> the NBCAs is still in<br />

the initial stage, with many still lacking clear delineation <strong>of</strong> boundaries, and specific management plans<br />

for managing high conservation values (e.g. watershed and soil conservation values), and resource<br />

depletion continues with illegal harvesting and trade in wildlife and non-timber forest products occurring<br />

in such areas (MAF 2005).<br />

Reduction in Annual Harvest Quota <strong>of</strong> Logs by the Government<br />

There are now several Prime Ministerial Orders and Decrees controlling the harvesting and sales <strong>of</strong> forest<br />

products. Consequently, government annual log harvest quotas have fallen from 734,000m 3 in 1999 to<br />

150,000 m 3 in 2004/5. However, the operational capacity <strong>of</strong> sawmills is still far above the harvesting<br />

levels set by the Government and puts pressure on natural forests. Over the next few years the<br />

Government is implementing pilots for certification projects to examine the costs and benefits under<br />

existing policy and conditions; however, it is still unclear as to how cost effective certification is for<br />

sustainably managed production forests in Lao PDR (MAF 2005).<br />

Stabilizing Shifting Cultivation<br />

In 1996, a land and forest allocation program which aimed to encourage cash crop production, stabilize<br />

shifting cultivation and promote forest conservation was adopted. More than half the rural population<br />

(some 4 million) was allocated land for farming and tree planting, and forests for management by 2005.<br />

Consequently, the area used for shifting cultivation was reduced between 1996 and 2005. However, the<br />

program was implemented very quickly without a thorough stakeholder consultation with villagers, and<br />

other sectors, and village forest <strong>of</strong>ficers received little training. Thus the results <strong>of</strong> the total program were<br />

limited, with poor communication and lack <strong>of</strong> support being cited as reasons for not realizing the full<br />

program objectives. The forest strategy for 2020 will reintroduce the land and forest allocation system,<br />

working more closely with villages in the land use planning process. The strategy also states that a<br />

national land use policy will be formulated over the coming decade (MAF 2005).<br />

Tree Planting<br />

Tree planting and improving the existing forest areas have been identified as one <strong>of</strong> the major sector<br />

targets, which must be achieved to contribute to poverty reduction. The targets for 2020 are ambitious<br />

with natural regeneration occurring on up to 6 million ha and planting trees on up to 500,000 ha in<br />

temporary unstocked forests. In the past, tree planting increased by 1700 ha/yr in the early 1990s, to<br />

17000ha/yr in 2000s. Teak in particular, has started to bring benefits to farmers from this program,<br />

however, past challenges included an inadequate maintenance <strong>of</strong> young stands and a lack <strong>of</strong> attention to<br />

the thinning (rarely done) important for quality wood production. The surveys for planted trees prior to<br />

2005 were not well done with respect to the selection <strong>of</strong> sites and species. There is therefore a strong need


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 363<br />

to improve tree plantation pr<strong>of</strong>itability, technology and market research during the implementation <strong>of</strong> the<br />

2020 strategy.<br />

Improving Governance and the Regulatory Framework<br />

The legislative and regulatory framework are still being formulated and developed, and it is not<br />

uncommon to implement annual Prime Ministerial Degrees to cover gaps in the forest law when they are<br />

noticed (DOF 2005). Problems with forest law enforcement and governance relate mostly to harvesting<br />

and use <strong>of</strong> timber, and to NTFPs (MAF 2005). In response, a forest surveillance unit was established<br />

under the Ministry <strong>of</strong> Agriculture and Forestry in 2008, which aims to improve forest governance and<br />

decrease the rate <strong>of</strong> illegal logging.<br />

Overall, there have been some improvements in livelihoods due to improved forest health, and while<br />

progress appears to be slow in some areas, the REDD process provides an opportunity to renew such<br />

commitments. Some projects funded by donors are beginning to produce results which suggest that the<br />

participatory sustainable management <strong>of</strong> forests can lead to poverty reduction and increased earning<br />

opportunities.<br />

CONCLUSION<br />

REDD policies and mechanisms which support REDD strategies can impact rural forest livelihoods.<br />

However, policies and mechanisms from other sectors, such as energy, hydropower, mining and<br />

agriculture can have a greater impact on rural livelihoods, and potentially even on REDD activities,<br />

particularly in a country like Lao PDR where all land and forests are state owned.<br />

However, carefully planned community based REDD activities could, on the other hand, provide an<br />

additional source <strong>of</strong> income and work to rural communities in Lao PDR. Programs that focus on<br />

sustainable forest management and conservation, eco-tourism and protected areas, providing alternatives<br />

to slash and burn, afforestation and land rehabilitation, as well as community based forest surveillance<br />

can all have positive impacts on rural livelihoods, in addition to providing income flows. This has already<br />

been found with a few existing avoided deforestation/REDD projects in Africa and Latin America.<br />

In the past, forest strategies in Lao PDR to avoid deforestation and degradation and increase forest cover<br />

have met with mixed success and can provide important lessons for future REDD strategies. Expanding<br />

shifting cultivation areas with reduced fallow periods, hydroelectric, energy and mining projects, slash<br />

and burn, unsustainable timber extraction techniques and forest land conversion are some <strong>of</strong> the drivers <strong>of</strong><br />

deforestation and forest degradation in Lao PDR. Of particular concern, are the potential emissions from<br />

forest degradation caused by shifting cultivation, usually preceded by slash and burn activities. One <strong>of</strong> the<br />

key challenges in meeting the information requirements for strategy formulation associated with REDD<br />

activities in Lao PDR is the lack <strong>of</strong> available information on the direct and indirect drivers <strong>of</strong><br />

deforestation and forest degradation, and the lack <strong>of</strong> necessary data infrastructure to support forest carbon<br />

monitoring requirements.<br />

REFERENCES<br />

AusAID (1996) Natural Resource Management in the Mekong River Basin: Perspective for <strong>Australia</strong>n Development<br />

Cooperation, Final Overview Report, <strong>Australia</strong>n Agency for International Development, Online Document:<br />

http://usyd.edu.au/su/geography/hirsch/5/5/htm<br />

Bird, J. (2008) Intergrated Water Resources Management in the Rapidly Growing Private Sector Development<br />

Context: the Mekong Basin. Presentation at the World Water Week, 19 th August 2008, Stockholm, Sweden.<br />

Chanhsomone, P., Somchai, S., and Chittana, P. (2008). Current Status <strong>of</strong> Forest Cover in Lao PDR. Presentation to<br />

the 2 nd event <strong>of</strong> the Symposium on Preparing for Mitigation <strong>of</strong> Climate Change in the Mekong Region and the<br />

Workshop for preparing the REDD programme. Hanoi, 3-5 November, 2008, Vietnam.<br />

DANIDA (1998). Environmental Problems <strong>of</strong> the Energy Sector, Presentation to Danida Natural Resource and<br />

Environment Programme, DANIDA, Vientiane.<br />

Department <strong>of</strong> Forestry (DOF) (2005). Report on The Assessment <strong>of</strong> Forest Cover and Land Use During 1992-2002.<br />

Ministry <strong>of</strong> Agriculture and Forestry, Vientiane, Lao PDR.<br />

Foppes, J. and S. Ketphanh, (1997). “The use <strong>of</strong> Non Timber Forest Products in Lao PDR, paper presented at the<br />

International Workshop on Sustainable Management <strong>of</strong> Non-Wood Forest Products, UPM, Serdang, Selangor,<br />

Malaysia, 14-17 October 1997.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 364<br />

Ingles, A. (2006). Non-timber Forest Products and Rural Livelihoods in Lao PDR: reducing poverty through forest<br />

development and conservation interventions. Greater Mekong Sub-region BCI Symposium April 27 th 2006.<br />

Kiyono, Y., Ochiai, Y., Chiba, Y., Asia, H., Saito, K., Shiraiwa, T., Horie, T., Songnoukhai, V., Navongxaui, V.<br />

And Inoue, Y. (2007). Predicting Chronosequential Changes in Carbon Stocks <strong>of</strong> Pachymarph Bamboo<br />

Communities in Slash and Burn Agricultural Fallow, Northern Lao People’s Democratic Republic. Journal <strong>of</strong><br />

Forest Research 12: 371-383.<br />

Lao PDR (1999). The Government’s Strategic Vision for The Agriculture Sector, Ministry <strong>of</strong> Agriculture and<br />

Forestry, Vientiane, Lao PDR.<br />

MAF (2005). Forest Strategy to the Year 2020 <strong>of</strong> the Lao PDR. Ministry <strong>of</strong> Agriculture and Forestry, Department <strong>of</strong><br />

Forestry, Vientiane, Lao PDR.<br />

Phanovong (1997). Energy Demand and Supply in Lao PDR in World Bank (2005) Lao PDR Environment Monitor.<br />

World Bank, USA.<br />

Robinson, D,M. & McKean, S.T. (1992). Shifting Cultivation and Alternatives: An Annotated Biography 1972-<br />

1989. Wallingford UK, Centro Internacial de Agriculura Tropical and CAB International.<br />

Roder,W. (2001). Slash and Burn Rice Systems in the Hills <strong>of</strong> Northern Lao PDR: Description, challenges and<br />

opportunities. International Rice Research <strong>Institute</strong>, Manila, Philippines.<br />

Roder, W., Keoboualapha, B., Manivanh, V. (1995). Teak (Tectona grandis), fruit trees and other perennials used by<br />

hill farmers <strong>of</strong> Northern Laos. Agro forestry Systems 29:47-60.<br />

Roder, W., Manivong, V, Soukaphone, H., and Lealock, W. (1992). Fanning Systems Research In the Uplands <strong>of</strong><br />

Laos: In Proceedings <strong>of</strong> the upland Rice-based farming systems research planning meeting, Chiang Mai,<br />

Thailand, p.39-54.<br />

STEA (2000). National Environmental Action Plan 2000, Science Technology and Environment Agency, Vientiane,<br />

Lao PDR.<br />

Thomas, I. (2008). How to Map Land Cover and Forest Change in Lao PDR. Presentation to the Lao REDD Task<br />

Force, Department <strong>of</strong> Forestry, November, 2008, Vientiane, Lao PDR.<br />

World Bank (2005). Lao PDR Environment Monitor. World Bank, USA.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 365<br />

REDD IN ASIA PACIFIC:<br />

CHALLENGES AND IMPLICATIONS<br />

Andrew Morton 1 and Blair Freeman 1<br />

ABSTRACT<br />

The scope for Reducing Emissions from Deforestation and Forest Degradation (REDD) to<br />

address climate change is significant. This paper outlines some <strong>of</strong> the challenges that need<br />

to be addressed for effective REDD projects to be put in place. Key design challenges<br />

include the further development <strong>of</strong> national governance and registry systems; cost–effective<br />

baseline assessment methodologies; and practical systems for equitable and efficient<br />

distributions <strong>of</strong> carbon payments to stakeholders. Emerging carbon standards will play an<br />

important role in supporting ongoing investment in REDD projects. Recent cost<br />

assessments indicate the full costs <strong>of</strong> REDD projects will tend to be higher than opportunity<br />

costs based on regional assessments. Furthermore, project proponents will seek returns<br />

commensurate with the high level <strong>of</strong> risks around REDD during this formative stage <strong>of</strong><br />

development. Notwithstanding these challenges, REDD does <strong>of</strong>fer enormous potential to<br />

address climate change and can support the broader objectives <strong>of</strong> sustainable forest<br />

management. If REDD is to succeed in the longer term, integrated approaches to forest<br />

management will be required to tackle the underlying drivers for deforestation and<br />

degradation, some <strong>of</strong> which comes from beyond the traditional forest sector.<br />

THE EMERGENCE OF REDD<br />

There is a relatively new forest-based solution now attracting significant attention worldwide: the<br />

scope for Reducing Emissions from Deforestation and Forest Degradation (REDD) is widely<br />

considered to have great potential to assist global efforts to curb climate change.<br />

Avoided deforestation was initially excluded by the United Nations Convention on Climate Change<br />

(UNFCC) from eligible activities within the land use, land-use change and forestry sector. However<br />

REDD could be incorporated in the post-2012 framework for the successor to the Kyoto Protocol.<br />

The premise <strong>of</strong> REDD is simple: Deforestation and forest degradation in tropical forests account for a<br />

substantial proportion <strong>of</strong> global carbon emissions. Emissions from tropical deforestation in the 1990s<br />

were estimated to be approximately 1.6 billion tonnes <strong>of</strong> carbon per year, equating to approximately<br />

20% <strong>of</strong> global carbon emissions (IPCC 2007).<br />

REDD projects <strong>of</strong>fer the opportunity to utilise funding from developed countries to reduce<br />

deforestation and forest degradation in developing countries. REDD projects incorporate interventions<br />

to stopping or reducing the release <strong>of</strong> CO2 from forest lands. For example:<br />

• A forest that would have been cleared for oil palm is protected and therefore the CO2 is not<br />

released;<br />

• A forest that would have been logged conventionally is logged using Reduced Impact Logging<br />

(RIL) systems, thereby reducing damage and therefore CO2 emissions; and<br />

• An area that is excessively burnt is protected, thereby reducing forest degradation and carbon<br />

emissions.<br />

REDD is a compelling mechanism as it represents avoided emissions, rather than post-emission<br />

sequestration and <strong>of</strong>fsets. It also <strong>of</strong>fers scope for substantial reductions in global greenhouse gas<br />

1 URS Forestry, Level 6, 1 Southbank Boulevard, Southbank, VIC 3006, <strong>Australia</strong>. Email: blair_freeman@urscorp.com


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 366<br />

emissions, potentially at relatively low cost (as much <strong>of</strong> the project creation will be undertaken in<br />

developing countries) and with additional environmental and social benefits.<br />

In 2005, Papua New Guinea and Costa Rica, supported by eight other Parties to the United Nations<br />

Framework Convention on Climate Change (UNFCC), proposed a mechanism for Reducing<br />

Emissions from Deforestation in Developing Countries. The proposal received wide support from<br />

Parties and the convention established a contact group and thereafter began a two year process to<br />

explore options for REDD.<br />

The scope for REDD to be included in a post-2012 UNFCCC framework was discussed further in Bali<br />

in 2007. Under the Bali Action Plan (or Roadmap), the Conference <strong>of</strong> the Parties agreed that if REDD<br />

is to be incorporated beyond 2012, a decision about what a REDD mechanism will look like and what<br />

it will include needs to be agreed by next meeting <strong>of</strong> the Parties in Copenhagen in December 2009.<br />

A REGIONAL CONTEXT<br />

REDD is proposed as part <strong>of</strong> a global solution to climate change, and may be developed for projects<br />

based in tropical countries including Central and South America, Africa and Southeast Asia.<br />

From an <strong>Australia</strong>n perspective, there is particular interest in opportunities in Southeast Asia,<br />

specifically Indonesia and also Malaysia and Papua New Guinea. According to FAO reports,<br />

deforestation in Indonesia totaled approximately 1.8 million hectares per year through the 1990s, and<br />

the rate has increased from 1.7% to 2% per year since then (FAO 2006).<br />

A key feature <strong>of</strong> Indonesia’s capacity for REDD is its extensive peat swamp forests, which hold vast<br />

stores <strong>of</strong> carbon. Peat can extend to 20 metres in depth and can contain upwards <strong>of</strong> 20,000 tonnes <strong>of</strong><br />

CO2 per hectare. Recent estimates indicate that in the order <strong>of</strong> 42 - 55 gigatonnes <strong>of</strong> carbon are stored<br />

in Indonesia's peatlands (Jaenicke et al, 2008; Hooijer et al, 2009).<br />

However, deforestation, draining and burning <strong>of</strong> peatland is a major contributor to Indonesia’s carbon<br />

emissions. According to Hooijer et al (2009), global CO2 emission caused by decomposition <strong>of</strong><br />

drained peatlands was between 355 and 855 million tonnes per year in 2006, <strong>of</strong> which 82% came from<br />

Indonesia, largely Sumatra and Kalimantan. Based on these estimates, CO2 emission from peatland<br />

drainage in Southeast Asia is contributing the equivalent <strong>of</strong> 1.3 to 3.1% <strong>of</strong> current global CO2<br />

emissions from the combustion <strong>of</strong> fossil fuel.<br />

The opportunity to reduce this conversion through REDD projects is demanding attention due to the<br />

scale <strong>of</strong> emission reductions that could be achieved in targeted areas. Malaysia and Papua New<br />

Guinea have a range <strong>of</strong> adjacent forested lands and similar pressures on forests as in Indonesia.<br />

DESIGN CHALLENGES<br />

While the development <strong>of</strong> REDD has been rapid, it is still emerging as a policy option and there are<br />

numerous design issues and challenges that confront its proponents across a range <strong>of</strong> regions.<br />

Several key design challenges are outlined below, providing some perspectives on the complexity <strong>of</strong><br />

REDD project development and implementation. These challenges include:<br />

a) Development <strong>of</strong> national governance and registry systems;<br />

b) Baseline determinations, incorporating additionality, leakage and permanence; and<br />

c) Equitable and effective distribution <strong>of</strong> payments to stakeholders.<br />

Governance and registry systems<br />

REDD has been prioritised in international negotiations about the climate change policy framework<br />

beyond 2012 and incorporated in consideration <strong>of</strong> the capacity <strong>of</strong> individual countries to meet<br />

emission reduction targets. REDD credits would accrue nationally or sub-nationally, rather than the<br />

smaller scale project-based approach fostered under the Clean Development Mechanism.<br />

National sovereignty will play a major role in the shaping <strong>of</strong> REDD projects in different countries.<br />

Possible government roles would appear to be as: a seller; a buyer from a sub-national devolved<br />

payment system; or regulator and/or broker. Mayers et al 2008 note that high levels <strong>of</strong> central


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 367<br />

coordination will be important – strong and fair rules and institutions, macroeconomic and agricultural<br />

policies aligned with forest policies, effective monitoring – and issues <strong>of</strong> tenure at local level will be<br />

critical.<br />

National and sub-national governments need to resolve the role/s that they will fulfil and the processes<br />

and systems in which they will conduct these roles. In some respects these design issues are well<br />

developed, while in other respects they remain unclear.<br />

In Indonesia for example, there has been substantial pilot project activity at the provincial and<br />

kabupaten level, on the basis <strong>of</strong> direct support and facilitation from provincial governments and local<br />

stakeholders. However, the Central Government has more recently moved to establish its ultimate<br />

authority on REDD implementation. In May 2009, the Government enacted a Regulation <strong>of</strong> the<br />

Minister <strong>of</strong> Forestry on Procedures for Reducing Emission from Deforestation and Forest<br />

Degradation (P.30/Menhut-II/2009), which:<br />

• Recognises the establishment <strong>of</strong> a REDD Commission to manage REDD implementation;<br />

• Requires REDD proponents to submit a proposal to the Minister, incorporating requirements<br />

outlined in the Decree that include: (i) clear designation <strong>of</strong> the land tenure for the project area; (ii)<br />

recommendation for REDD implementation from the local government; (iii) detailed<br />

specifications in respect <strong>of</strong> location and baseline criteria and indicators for REDD implementation;<br />

and (iv) a REDD implementation plan.<br />

• Recognises the Minister will request the REDD Commission to assess the REDD proposal, and<br />

within 14 days after receiving the assessment <strong>of</strong> the REDD Commission, the Minister can decline<br />

the proposal or approve the proposal by issuing a licence;<br />

• Recognises the REDD Commission will request an Independent Appraiser Institution to conduct<br />

verification <strong>of</strong> monitoring reports, with costs borne by the proponent; In the event that all<br />

requirements are fulfilled, the Commission will issue a Carbon Emission Reduction Certificate,<br />

which can be traded by the proponent.<br />

These government guidelines have been established in advance <strong>of</strong> the decision by the UNFCC on the<br />

mechanism <strong>of</strong> REDD implementation at the international level, for the purpose <strong>of</strong> REDD<br />

demonstration activity capacity building, and transfer <strong>of</strong> technology and voluntary carbon trading. In<br />

this climate <strong>of</strong> change, further government regulations may follow shortly.<br />

In addition to the development <strong>of</strong> these guidelines, a World Bank workshop on developing REDD in<br />

Indonesia (2008) noted that key government challenges are to effectively implement the Forest<br />

Resource Information System and the National Carbon Accounting System, which will support central<br />

registry systems in which REDD projects can be reconciled against actual and project emission levels.<br />

Baseline determinations<br />

REDD project baseline determinations incorporate three key elements. These are:<br />

1. Current carbon stocks, which is effectively the carbon stored within the REDD project area at<br />

project inception, based on measurements and determinations <strong>of</strong> the carbon pools comprising<br />

above ground and potentially below ground biomass;<br />

2. Future carbon stocks without the proposed REDD intervention, taking into account deforestation<br />

or forest degradation that is expected to occur; and<br />

3. Future carbon stocks with the proposed REDD intervention, taking into account the impacts <strong>of</strong> the<br />

intervention within the REDD project area, the buffer area and any leakage impacts.<br />

Figure 1 illustrates the nature <strong>of</strong> a carbon baseline, and how it can change over time. For example, in<br />

this schematic, the REDD intervention may result in a short term reduction in emissions that is<br />

substantially larger than the longer term annual emission reductions.<br />

Current carbon stocks can to a large extent be readily determined with established forest inventory<br />

assessments, which include remote sensing technologies that provide scope for rapid appraisals across<br />

large forested areas. This is particularly the case for above ground biomass.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 368<br />

In the case <strong>of</strong> below ground biomass and soil carbon however, cost effective methodologies for<br />

measuring carbon stocks are not as well developed (Gibbs et al 2007). For forests on peat soils in<br />

particular, the absence <strong>of</strong> standardised, cost effective and rigorous methodologies is a significant<br />

technical challenge for the development <strong>of</strong> reliable baseline estimates to underpin REDD projects.<br />

CO2 stock<br />

(t CO2/ha)<br />

Figure 1. - Establishing a carbon baseline for REDD projects<br />

A<br />

Source: URS Forestry<br />

B<br />

Time<br />

Project<br />

eg: forest conservation, reduced<br />

impact logging and improved<br />

peat management<br />

Baseline<br />

eg: conversion to agriculture and<br />

degradation<br />

In respect to developing projections <strong>of</strong> future carbon stocks, there is a critical need for a sound<br />

understanding <strong>of</strong> the underlying drivers for deforestation and degradation. This will generally require<br />

technical forestry expertise and social research expertise to assess forestry utilisation under different<br />

baseline scenarios, within the REDD project area and the broader region.<br />

The success <strong>of</strong> the planned intervention will depend to a large extent on the nature <strong>of</strong> the underlying<br />

drivers, local stakeholders’ attitudes towards land rights and forest stewardship, and the way in which<br />

the planned intervention will address these drivers. For example, deforestation and degradation may<br />

be caused by (i) commercial interests in conversion to agricultural crops, (ii) markets for high value<br />

timbers within the project area, or (iii) land right disputes. Unless the underlying cause <strong>of</strong><br />

deforestation and degradation is addressed, there is a threat to the permanence <strong>of</strong> the intervention or a<br />

threat <strong>of</strong> leakage to surrounding areas.<br />

The assessment <strong>of</strong> forestry utilisation and the risk <strong>of</strong> deforestation or degradation is also central to the<br />

claims that a project can make on ‘additionality’. Additionality refers to the basis on which a project<br />

can show that the carbon benefit (from avoided emissions or emission <strong>of</strong>fsets) would not have<br />

happened but for the intervention. Furthermore, if the risk <strong>of</strong> deforestation or degradation is imminent,<br />

this can generally provide greater assurance <strong>of</strong> additionality. This can lead to some perverse outcomes<br />

in terms <strong>of</strong> process or in on-ground activity. For example, it could result in efforts to bring forward or<br />

otherwise enhance the threat <strong>of</strong> deforestation / degradation to substantiate additionality claims.<br />

Countries with different forest covers and historic deforestation rates will hold different interests in the<br />

way the baseline reference levels are constructed, and involving countries with high forest covers and<br />

low historic deforestation rates will be necessary to reduce perverse incentives.<br />

Equitable and effective distribution <strong>of</strong> payments<br />

The concept <strong>of</strong> REDD is based on funding from developed countries to reduce deforestation and forest<br />

degradation in developing countries. How this funding is distributed (financial flows) among<br />

stakeholders is likely to be critical to the long term success <strong>of</strong> REDD projects. How can the benefits<br />

from REDD be distributed to forest communities in a just, equitable way that minimises rent seeking<br />

<strong>of</strong> the benefits by other stakeholders or local elites?<br />

REDD projects can potentially comprise many key stakeholders that would likely seek entitlement to a<br />

share <strong>of</strong> REDD benefits. These include:


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 369<br />

• project proponents, which may the project developers or carbon brokers, or both;<br />

• tiers <strong>of</strong> government – including national, provincial and district/local government authorities;<br />

• concession license holders;<br />

• village leaders and clan leaders;<br />

• land owners and farmers; and<br />

• forest based community people dependent on the use <strong>of</strong> land as a source <strong>of</strong> economic activity,<br />

permitted or not.<br />

In addition, there are project support services, including project development assessments (comprising<br />

technical forestry and social services as well as legal and administrative services) and ongoing<br />

monitoring and verification assessment services.<br />

The distribution <strong>of</strong> payments from the proponent to local communities in the REDD project area will<br />

generally be net <strong>of</strong> service fees and distributed through the tiers <strong>of</strong> government. How the government<br />

manages this will relate to the respective positions on REDD governance and registry systems.<br />

Clearly there is a threat that a cascading series <strong>of</strong> payments may substantially diminish the benefits<br />

payable to whose livelihoods may depend on the forests subject to REDD interventions. This is a key<br />

design challenge for project proponents and regional governments. In their assessment <strong>of</strong> tenure<br />

considerations for REDD, Cotula and Mayers (2009) highlighted the need to balance efficiency and<br />

fairness, noting “REDD simply will not work unless it is locally credible: it will be undermined and<br />

overthrown”.<br />

EMERGING STANDARDS<br />

Concurrent with REDD policy developments, carbon standards are emerging to support REDD<br />

projects through project inception and ongoing monitoring and validation. There are a range <strong>of</strong><br />

standards and investment guidelines that are relevant to development <strong>of</strong> REDD.<br />

For example, the Voluntary Carbon Standard (VCS) has emerged as one <strong>of</strong> the leading carbon<br />

standards worldwide and can be regarded as the most prominent <strong>of</strong> carbon standards considered for<br />

REDD forestry projects in the Asia Pacific region currently. The VCS program has launched a global<br />

registry system (comprising multiple registries) which will provide the chain <strong>of</strong> custody platform for<br />

trading <strong>of</strong> Voluntary Carbon Units.<br />

The Climate, Community & Biodiversity Alliance (CCB) Standard is also prominent in existing forest<br />

carbon projects in Asia Pacific region. In some cases, forest carbon projects have chosen to pursue<br />

certification under both the VCS and the CCB standards to either progressively attain certification to<br />

credible standards or to enhance the credibility <strong>of</strong> the project within target markets.<br />

The development <strong>of</strong> the VCS and CCB standards may converge with carbon finance support programs<br />

and associated standards that REDD proponents are considering as relevant and provide additional<br />

assurances to potential purchasers. Relevant examples <strong>of</strong> related programs include:<br />

• The International Finance Corporation (IFC) project selection criteria established by the Carbon<br />

Finance Unit. This IFC unit is set up to assist project sponsors in emerging markets to access the<br />

market for carbon credits. Eligibility for assistance under this program is determined against<br />

criteria that include the location (with the primary focus on emerging countries ratified to the<br />

Kyoto Protocol), the amount <strong>of</strong> credits, environmental and social impact assessments and<br />

independent verification; and<br />

• Third party sustainable forest management certification, such as certification under the Forest<br />

Stewardship Council (FSC) or Programme for Endorsement <strong>of</strong> Forest Certification (PEFC). Such<br />

schemes may be aligned with or complementary to the outcomes and risks for forest carbon<br />

purchaser requirements.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 370<br />

The further development <strong>of</strong> internationally recognised carbon standards will be important to<br />

supporting ongoing investment in forest carbon projects, particularly during the formative phases <strong>of</strong><br />

voluntary and compliance markets.<br />

MARKET DYNAMICS<br />

In the absence <strong>of</strong> regulatory demand for avoided deforestation and deforestation, voluntary carbon<br />

markets are leading the establishment <strong>of</strong> the market dynamics and a carbon price.<br />

Forest carbon transactions and prices<br />

In their State <strong>of</strong> the Voluntary Carbon Markets 2009, Hamilton et al (2009) noted that between 2007<br />

and 2009, stakeholders seeking a means <strong>of</strong> halting deforestation had begun to aggressively influence<br />

policy and markets to create incentives for REDD. However, despite the positive sentiments towards<br />

REDD, the review found that REDD-based credits declined from 1.4MtCO2e in 2007 to 0.7MtCO2e in<br />

2008. This trend was attributed to the difficulties in developing projects due to a range <strong>of</strong> factors,<br />

including layers <strong>of</strong> complexity in relation to working with communities, mid-levels <strong>of</strong> government,<br />

national governments, the policies and regulations around carbon ownership, technical issues around<br />

measuring carbon.<br />

Comparing over the counter transaction prices between 2007 and 2008, Hamilton et al (2009) found an<br />

overall increase in credit prices for most project types, including avoided deforestation – up from a<br />

weighted average price <strong>of</strong> US$4.80/tCO2e in 2007 to US$6.30/tCO2e in 2008. However, transaction<br />

data for avoided deforestation credits is limited at present. This comparison <strong>of</strong> prices for avoided<br />

deforestation is based on 10 and 11 transactions in 2007 and 2008 respectively, principally in Africa<br />

and Latin America (none in Asia). Figure 2 shows the range <strong>of</strong> prices in 2008 varied between US$4.80<br />

– US$28/tCO2e.<br />

Figure 2. - Transaction prices in voluntary markets, 2008<br />

USD/tCO2<br />

$50<br />

$40<br />

$30<br />

$20<br />

$10<br />

$0<br />

Afforestation<br />

plantation<br />

Afforestation<br />

reforestation<br />

Avoided<br />

deforestation<br />

Source: Hamilton et al 2009, Ecosystem Marketplace, New Carbon Finance.<br />

Costs <strong>of</strong> REDD<br />

A range <strong>of</strong> recent studies have estimated the cost and potential <strong>of</strong> reducing emissions for deforestation.<br />

Kanounnik<strong>of</strong> (2008) noted that the economic concept <strong>of</strong> the supply curve suggests there is no single<br />

cost-value for REDD, but alternative levels <strong>of</strong> REDD supply are associated with different costs. It is<br />

important to recognise there are separate cost components <strong>of</strong> REDD. These include:<br />

o Opportunity costs – the foregone pr<strong>of</strong>its from alternative land uses such as cash or food crops and<br />

timber sales, for which the landowner or concessionaire will seek to be compensated;


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 371<br />

o Transaction costs – the costs associated with establishing the project or scheme, including the<br />

search for project areas and partners, project design, consultation with stakeholders, verification<br />

with standards and establishing contractual arrangements for all aspects <strong>of</strong> the project; and<br />

o Implementation costs – the operational costs associated with managing the project or scheme and<br />

the compliance costs associated with monitoring and evaluation for the life <strong>of</strong> REDD credits<br />

generated by the project or scheme.<br />

To date, opportunity costs have generally been recognised as the largest portion <strong>of</strong> REDD costs.<br />

Table 1 sets out a range <strong>of</strong> estimated opportunity costs <strong>of</strong> REDD in 2020, presented by the Union <strong>of</strong><br />

Concerned Scientists based on recent research covering a variety <strong>of</strong> regional and global studies.<br />

Table 1. - Estimated opportunity costs <strong>of</strong> REDD in 2020* (US$/CO2e)<br />

Approach Average Range<br />

Regional $3.51 $1.84 – 5.18<br />

Stern review $6.52 $3.76 – 9.28<br />

Global models $12.26 $7.77 - $18.86<br />

Source: Boucher, 2008; * Cost (in 2005 dollars) to reduce 1tCO 2e if overall there is a 46%<br />

reduction in global deforestation.<br />

Recent CIFOR research has also found global models have yielded far higher REDD prices than<br />

empirical models, including the Stern estimate (Kanounnik<strong>of</strong>f, 2009). One explanation is that global<br />

simulation models not only consider the opportunity costs, but also the costs arising from<br />

interrelations with other sectors and from the fact that for practical reasons land users are likely to be<br />

paid a uniform price, not differentiated according to their opportunity costs (Eliasch 2008).<br />

Other related work includes research on the opportunity costs associated with conversion <strong>of</strong> natural<br />

forests in Kalimantan to oil palm plantations. Based on a financial analysis <strong>of</strong> conversion scenarios,<br />

the University <strong>of</strong> Queensland estimated that halting oil palm conversion in Kalimantan would cost<br />

between US$10 and US$33 per tonne <strong>of</strong> CO2 (Venter et al 2008). These opportunity costs are higher<br />

than current average prices in voluntary markets (Figure 4), but lower than prices in the order <strong>of</strong><br />

US$30 per tonne in Kyoto compliance markets.<br />

Venter et al (ibid) also noted that opportunity costs can be considerably lower on peat soils, due to the<br />

substantially higher carbon stores in peat forests than in mineral soils. By targeting only peat areas for<br />

REDD, the opportunity cost was estimated to drop to a range <strong>of</strong> US$1.63 – $4.66 per tonne <strong>of</strong> CO2.<br />

This is <strong>of</strong> particular significance to the further development <strong>of</strong> REDD across Indonesia.<br />

However, transaction costs do need to be taken into account, particularly for the early movers, who<br />

face considerable risk at this early stage <strong>of</strong> REDD development. The full costs <strong>of</strong> establishing a project<br />

will incorporate substantial investment in project design documentation, a range <strong>of</strong> technical and<br />

social assessments, consultation with a broad range <strong>of</strong> stakeholders, and obtaining all necessary<br />

approvals at various tiers <strong>of</strong> government. Transaction costs could potentially be as high or exceed<br />

opportunity costs in some settings.<br />

In addition, there is the return on risk, or pr<strong>of</strong>it motive, that project developers will require to meet<br />

investment requirements. The risks associated with investing in REDD projects are high at present,<br />

and it is expected that investors will require a relatively high return on their investment, above the full<br />

range <strong>of</strong> costs associated with developing and implementing their projects.<br />

Global investment<br />

The success <strong>of</strong> REDD will ultimately be determined to a considerable extent by the market value <strong>of</strong><br />

REDD credits relative to the opportunity cost <strong>of</strong> other carbon emissions and alternative land uses.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 372<br />

Based on its review <strong>of</strong> costs <strong>of</strong> REDD, the Union <strong>of</strong> Concerned Scientists estimated that for<br />

US$5 billion a year, REDD could be applied to nearly 20% <strong>of</strong> the tropical forests in danger <strong>of</strong><br />

deforestation, and US$20 billion a year could cover about half <strong>of</strong> the tropical forests at risk (Boucher,<br />

2008). This study concluded that REDD can greatly reduce tropical deforestation and greenhouse gas<br />

emission with modest funding.<br />

REDD AND SUSTAINABLE FOREST MANAGEMENT<br />

From a forestry perspective, REDD must be considered in the context <strong>of</strong> sustainable forest<br />

management objectives and frameworks. Broadly, the objective <strong>of</strong> REDD is to reduce to greenhouse<br />

gas emissions by reducing deforestation and degradation in tropical countries. To some extent these<br />

are two distinct objectives, and proponents may be motivated by one or other, or both.<br />

The objectives <strong>of</strong> sustainable forest management, as set out under the Montreal Process, are broader.<br />

These objectives include maintenance <strong>of</strong> forest contribution to global carbon cycles, in addition,<br />

maintenance <strong>of</strong> the productive capacity <strong>of</strong> forest ecosystems and the maintenance and enhancement <strong>of</strong><br />

long-term multiple socio-economic benefits to meet the needs <strong>of</strong> society.<br />

If REDD is focussed primarily on excluding activity within a forested region - potentially a<br />

preservation regime - there is a risk that it foregoes sustainable forestry and livelihoods. Sustainable<br />

forest management is based on comprehensive forest planning to identify the range <strong>of</strong> values within<br />

forests and the means <strong>of</strong> managing for multiple long-term benefits. Contula and Mayers (2009) noted<br />

the ‘old worry’ among foresters was that the forest sector is so complex that it will not figure in<br />

climate change regimes; the new worry is almost the opposite: that forest carbon finance is coming<br />

forward so quickly that it will not support sustainable forestry and livelihoods.<br />

If REDD is to succeed in the longer term it will need to be nested within more integrated approaches<br />

to forest management. These approaches need to tackle the underlying drivers <strong>of</strong> deforestation and<br />

degradation, some <strong>of</strong> which comes from beyond the traditional forest sector. Sustainable forest<br />

management also needs to incorporate all <strong>of</strong> the various aspects <strong>of</strong> forest-based mitigation – such as<br />

forest restoration and reforestation - as well as adaptation to climate change.<br />

LOSING SIGHT OF THE FOREST FOR THE CARBON?<br />

Clearly the international interest in the scope for REDD is growing rapidly. The World Bank’s<br />

Learning Workshop in Indonesia noted that the COP13 Bali Action Plan led to burgeoning interest in<br />

implementing REDD pilot and demonstration projects, and created Indonesia’s “carbon rush” (not<br />

unlike the California gold rush in the mid 19 th century), with its touted high financial returns. Further,<br />

the potential earnings from avoided deforestation carbon credit sales in Indonesia have been estimated<br />

in a range from US$500 million to $2 billion per annum in today’s voluntary market (World Bank<br />

Indonesia, 2009).<br />

Amid this carbon rush, there is an important if not vital role for foresters to play in assessing and<br />

reviewing the key elements <strong>of</strong> the project.<br />

For example, recent technical project experience has found some projects that have been developed to<br />

a considerable extent, at considerable expense, without any field reconnaissance <strong>of</strong> the proposed<br />

project area by experienced forestry personnel – which would have identified some key issues or<br />

material impacts on carbon stock assessments.<br />

Similarly, deforestation is commonly the last act in a forest following a series <strong>of</strong> forest-degrading<br />

activities over time. Successful REDD initiatives will recognise and support sustainable forest<br />

management as a key and critical intervention.<br />

This experience highlights the important role that foresters can play in the development <strong>of</strong> REDD as<br />

an effective mechanism for reducing greenhouse gas emissions, utilising core forestry disciplines in<br />

the context <strong>of</strong> broader sustainable forest management objectives.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 373<br />

CONCLUDING REMARKS<br />

In this climate <strong>of</strong> change, REDD could play a major role in shaping the future management <strong>of</strong> tropical<br />

forests within Asia Pacific. This paper highlights the compelling nature <strong>of</strong> REDD, and the challenges<br />

that need to be addressed for effective REDD projects to be put in place.<br />

Design challenges include the further development <strong>of</strong> national governance and registry systems; cost–<br />

effective baseline assessment methodologies; and practical systems for equitable and efficient<br />

distributions <strong>of</strong> carbon payments to stakeholders, including local communities.<br />

Emerging carbon standards will play an important role in supporting ongoing investment in REDD<br />

projects.<br />

Recent cost assessments indicate that the full costs <strong>of</strong> REDD projects will tend to be higher than<br />

opportunity costs based on regional assessments. Furthermore, project proponents will seek returns<br />

commensurate with the relatively high level <strong>of</strong> risks around REDD during this formative stage <strong>of</strong><br />

development.<br />

Notwithstanding these challenges, REDD does <strong>of</strong>fer enormous potential to assist in curbing climate<br />

change and can support the broader objectives <strong>of</strong> sustainable forest management. If REDD is to<br />

succeed in the longer term, integrated approaches to forest management will be required to tackle the<br />

underlying drivers for deforestation and degradation, some <strong>of</strong> which comes from beyond the<br />

traditional forest sector.<br />

REFERENCES<br />

Boucher, D. 2008. 2008. Estimating the cost and potential <strong>of</strong> reducing emissions from deforestation.<br />

Union <strong>of</strong> Concerned Scientists. Available online at: www.ucsusa.org/REDD.html.<br />

Cotula, L. and Mayers, J. 2009. Tenure in REDD: Start-point or afterthought? Natural Resource<br />

Issues No. 15, IIED, London, UK.<br />

Eliasch, J. 2008. The Eliasch Review – Climate Change: Financing Global Forests. UK Office <strong>of</strong><br />

Climate Change. Available online at: http://www.occ.gov.uk/activities/eliasch.htm.<br />

FAO, (2006), Global forest resource assessment 2005: progress towards sustainable forest<br />

management. Forestry Paper 147.<br />

Gibbs, H., Brown, S., O Niles, J. and Foley, J. 2007. Monitoring and estimating tropical forest carbon<br />

stocks: making REDD a reality. Environmental Research Letters, 2 (2007) 045023.<br />

Hamilton K, Sjardin, M., Shapiro, A. and Marcello T. 2009. Fortifying the Foundation: State <strong>of</strong> the<br />

Voluntary Carbon Market 2009. Ecosystem Marketplace, New Carbon Finance: USA.<br />

Hooijer, A., S. Page, J. G. Canadell, M. Silvius, J. Kwadijk, H. Wösten, and J. Jauhiainen, 2009.<br />

Current and future CO2 emissions from drained peatlands in Southeast Asia, Biogeosciences<br />

Discuss., 6, 7207–7230, 2009, www.biogeosciences-discuss.net/6/7207/2009/.<br />

Intergovernmental Panel on Climate Change (IPCC) 2007 Climate Change 2007: The Physical<br />

Science Basis: Summary for Policymakers.<br />

Jaenicke, J et al. 2008. Determination <strong>of</strong> the amount <strong>of</strong> carbon stored in Indonesian peatlands.<br />

Geoderma. Vol. 147. pp. 151-158.<br />

Mayers, J., Bass, S. Bigg, T., Bond, I., Bradstock, A. and Grieg-Gran, M. 2008. Forest-based climate<br />

strategies: making REDD and other initiatives work. IIED, London, UK.<br />

Parker, C, Mitchell, A, Trivedi, M. and Mardas, N, 2008. The Little REDD Book, Global Canopy<br />

Programme, Oxford, UK.<br />

Richards, M. 2008. REDD, the last chance for tropical forests? Policy brief, FRR, Bristol, UK.<br />

World Bank, 2009. Convenient solutions to an Inconvenient Truth: Ecosystem-based Approaches to<br />

Climate Change. Environment Department, World Bank.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 374<br />

ABSTRACT<br />

A CHARACTERISATION OF THE CURRENT MARKET FOR<br />

REDUCING EMISSIONS FROM DEFORESTATION AND<br />

DEGRADATION (REDD) AND FOREST CARBON OFFSETS<br />

Majella Clarke 1<br />

Forest carbon <strong>of</strong>fsets from REDD and forestry projects are providing increasing<br />

opportunities in the voluntary carbon market. Currently, forest carbon <strong>of</strong>fsets from<br />

REDD and most forestry projects are only able to be traded in the voluntary carbon<br />

market on a project by project basis, and on a deal by deal basis through the Over-The-<br />

Counter market. The paper argues that because <strong>of</strong> the global nature and flexibility <strong>of</strong> the<br />

voluntary market, project based crediting mechanisms have been set up under a range <strong>of</strong><br />

different standards for accounting, monitoring, verifying and distributing carbon credits<br />

from forest carbon projects, leading to growing product complexity, difficult<br />

marketability and little overall transparency. The paper concludes that market and<br />

liquidity fragmentation have implications on achieving a common price signal and that<br />

inefficient price formulation is, at this stage, an unavoidable characteristic <strong>of</strong> the<br />

voluntary market for forest based carbon credits.<br />

INTRODUCTION<br />

The objective <strong>of</strong> the paper is to present a characterisation <strong>of</strong> the current market for forest carbon<br />

<strong>of</strong>fsets with a particular focus on Reducing Emissions from Deforestation and Degradation (REDD). It<br />

builds upon the State <strong>of</strong> the Voluntary Market Report 2008 (Hamilton et al. 2008) but diverges to<br />

focus specifically on carbon <strong>of</strong>fsets from forest projects and REDD to give a market characterisation.<br />

It distinguishes itself by reviewing a number <strong>of</strong> REDD projects generating <strong>of</strong>fsets on the voluntary<br />

market through a desk review <strong>of</strong> Project Design Documents (PDD), project websites, verification<br />

reports, newsletters, communications with project implementers and specific project presentations<br />

were reviewed for 12 REDD projects currently selling emission reductions as a result <strong>of</strong> the project<br />

activities on the voluntary carbon market. Also, carbon <strong>of</strong>fset prices were collected from 79 vendors<br />

selling carbon <strong>of</strong>fsets online with forestry projects in their portfolio.<br />

The first section <strong>of</strong> this paper will summarise the general trends in the trade <strong>of</strong> forest based carbon<br />

<strong>of</strong>fsets and briefly reviews current REDD projects, project types, their mechanisms and the spatial<br />

distribution. The second section looks into the voluntary market structure for forest carbon based<br />

<strong>of</strong>fsets. It argues that because <strong>of</strong> the global nature and flexibility <strong>of</strong> the voluntary market, project based<br />

crediting mechanisms have been set up under a range <strong>of</strong> different standards for accounting,<br />

monitoring, verifying and distributing carbon credits from forest carbon projects, leading to growing<br />

product complexity, difficult marketability and little overall transparency. The third section <strong>of</strong> this<br />

paper argues that market and liquidity fragmentation have implications on achieving a common price<br />

signal and that inefficient price formulation is, at this stage, an unavoidable characteristic <strong>of</strong> the<br />

voluntary market for forest based carbon credits.<br />

General Trends<br />

The market structure for REDD is, at this stage, to a large degree undefined as it is still not clear what<br />

“product functions” will be included within the international framework. Will deforestation and<br />

degradation be considered under the same mechanism? Or will it just be deforestation? Will there be<br />

multiple mechanisms? The most obvious and important market structure determinants for REDD will<br />

be what is <strong>of</strong>fered and how it is regulated, as these determinants will affect the supply and demand <strong>of</strong><br />

emission reduction <strong>of</strong>fsets directly. However, it may also follow the market development <strong>of</strong> other<br />

forest carbon <strong>of</strong>fsets, in which the Over-The-Counter (OTC) voluntary market has shown potential for<br />

filling compliance market gaps.<br />

1 Dr Majella Clarke, Department <strong>of</strong> Forestry, Vientiane, Lao PDR. Email: majella.clarke@indufor.fi


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 375<br />

According to Hamilton et al. (2008), OTC transactions for avoided deforestation credits grew from 3%<br />

to 5% <strong>of</strong> total OTC <strong>of</strong>fset credits in 2006 and 2007 respectively. Because <strong>of</strong> the nature <strong>of</strong> avoided<br />

deforestation <strong>of</strong>fset credits being traded solely OTC, prices for such credits had large variations<br />

ranging from about USD 2.50 - 30 tCO2e in 2007, with 11 data points used in calculating the volume<br />

weighted average <strong>of</strong> USD 4.80 /tCO2e for avoided deforestation <strong>of</strong>fsets. One <strong>of</strong> the key reasons for<br />

the divergence in price formulation is because avoided deforestation <strong>of</strong>fsets are sold on a project by<br />

project basis, and each project is unique.<br />

Types <strong>of</strong> Projects and Mechanisms to Reduce Emissions from Deforestation and Degradation<br />

Of the 12 REDD projects reviewed, at total <strong>of</strong> 2,572,670 ha <strong>of</strong> forest area was being used to<br />

implement REDD projects. one project was reviewed from <strong>Australia</strong>, two projects were reviewed from<br />

the USA, six projects from Latin America, two from Africa, both in Madagascar and finally, one<br />

project from Asia in Indonesia. Project areas ranged from 756 ha in the USA’s Fred Van Eck Forest<br />

Foundation registered on the California Climate Action Registry, to 750,000 ha <strong>of</strong> the Indonesian Ulu<br />

Masen REDD project, in which VERs have been sold to private investors in cooperation with donors<br />

and governments. So far, Latin America has the largest area and the most REDD projects within the<br />

sample reviewed. Within these REDD projects, a total <strong>of</strong> 362,315,886 tCO2e are expected to be <strong>of</strong>fset<br />

over the course <strong>of</strong> the project. Project based emission reduction durations ranged from 20 years to 100<br />

years. So far, the Latin American region has contributed to generating the most emission reductions.<br />

All 12 projects reviewed had the central objective <strong>of</strong> avoiding deforestation and reducing emissions.<br />

However, their emission reduction strategies included objectives like reducing emissions through<br />

avoided slash and burn <strong>of</strong> forests, avoided land conversion, avoided forest fragmentation, avoided<br />

logging concessions, etc. Figure 1 presents the different objectives in projects in addition to avoided<br />

deforestation.<br />

N. <strong>of</strong> Projects<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

Avoided<br />

Deforestation<br />

Avoided slash<br />

and burn<br />

Avoided land<br />

conversion<br />

Avoided illegal<br />

activities that<br />

lead to<br />

deforestation<br />

Avoided forest<br />

fragmentation<br />

and<br />

encroachment<br />

Avoided<br />

logging<br />

conessions<br />

Avoided land<br />

degradation<br />

Focus 12 5 5 4 2 4 1<br />

Figure 1. A Comparison <strong>of</strong> different project objectives to Reducing Emissions from<br />

Deforestation and Degradation (REDD)<br />

Avoided slash and burn, and avoided land conversion featured prominently throughout most REDD<br />

projects currently in implementation. One possible explanation for this is that avoiding slash and burn<br />

and avoiding land conversion (which <strong>of</strong>ten involves slash and burn in developing countries) can<br />

generate more emission reduction credits tCO2e/ha, than avoiding logging concessions without<br />

burning. A prime example <strong>of</strong> this is the Noel Kempff project in Bolivia, in which there are two<br />

components:


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 376<br />

Component 1 Stopping Industrial Timber Harvesting on an area <strong>of</strong> 524,000 ha generates 791,444<br />

tCO2 <strong>of</strong> certified <strong>of</strong>fsets, or between 1997-2005, an average <strong>of</strong> 1.51 tCO2/ha<br />

Component 2 Avoided Slash and Burn Agriculture on an area <strong>of</strong> 756 ha generates 371,650 tCO2<br />

certified <strong>of</strong>fsets, or between 1997-2005, an average <strong>of</strong> 491.6 tCO2/ha (Seifert-<br />

Granzin, 2007)<br />

This project illustrates that different REDD objectives yield different magnitudes <strong>of</strong> emission<br />

reduction credits, and may shed light on why an increasing number <strong>of</strong> REDD projects, pilots and<br />

initiatives are including additional components addressing slash and burn agriculture and land<br />

conversion. In response to the different emission reduction strategies presented above, a number <strong>of</strong><br />

standard project mechanisms for REDD are presented in Figure 2.<br />

No. <strong>of</strong> REDD Projects<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Figure 2. Project mechanisms for the implementation <strong>of</strong> REDD<br />

In general, two types <strong>of</strong> project mechanisms featured prominently within the 12 REDD projects<br />

reviewed.<br />

• Protected Areas and Corridors were the focus <strong>of</strong> REDD projects in Latin America and Africa<br />

and include:<br />

o Establishing Protected Areas<br />

o Enlarging Protected Areas<br />

o Creating/protecting Green Corridors<br />

o Protected Area Management<br />

• Community Development, Agreements and Land Use Planning were the focus <strong>of</strong> most REDD<br />

projects in developing countries. This includes activities like finding alternative incomes and<br />

employment for activities which deforest, land use planning and zoning at community and<br />

village levels so that livelihoods are improved, and agreements with surrounding communities<br />

on land and forest use.<br />

Other project mechanisms common in REDD projects include agr<strong>of</strong>orestry, Forest Law Enforcement<br />

and Governance initiatives, sustainable forest management with conservation, indigenous rights and<br />

ecotourism. Within the brief overview <strong>of</strong> on-going REDD activities which have received or are<br />

applying to obtain verified <strong>of</strong>fsets within the voluntary carbon market, there are several general<br />

conclusions which can be drawn from this so far:


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 377<br />

• There are different methods for reducing emissions from deforestation and degradation.<br />

• There are many different standards through which an REDD project can be verified and<br />

implemented upon.<br />

• Some activities are much more efficient (tCO2/ha) in yielding emission reduction credits than<br />

others (e.g. the Noel Kempff example).<br />

• There are a number <strong>of</strong> mechanisms which can be used to implement an REDD project, which<br />

require completely different financing requirements.<br />

• REDD activities and their emission reduction credits are not limited only to developing<br />

countries, like the CDM A/R, but are also able to be implemented in developed countries.<br />

VOLUNTARY MARKET FOR REDD AND FOREST CARBON OFFSETS<br />

Market Structure Determinants and the OTC Market<br />

Following Willet (1931) and Senn (2002), Over-The-Counter (OTC) markets can be characterised by<br />

some <strong>of</strong> the following features:<br />

1. Small capitalization.<br />

2. Limited distribution.<br />

3. Lack <strong>of</strong> speculative interest.<br />

4. High price.<br />

5. Trading is not centralised<br />

6. Dealers are normally market makers<br />

7. Desirability for portfolios <strong>of</strong> institutional investors (banks, insurance companies, etc.) who<br />

wish to negotiate the purchase or sale <strong>of</strong> a large block at one price.<br />

These characteristics are highly reflective <strong>of</strong> the REDD and forest carbon <strong>of</strong>fsets market. Often <strong>of</strong>fsets<br />

are purchased by individuals or businesses, on a deal by deal basis to <strong>of</strong>fset household emissions,<br />

flights or events. There is no specific or central location for the sale and purchase and there is an<br />

increasing number <strong>of</strong> internet vendors (for example, see www.carboncatalogue.com and<br />

www.carbon<strong>of</strong>fsetguide.com.au). A single purchase <strong>of</strong> <strong>of</strong>fsets is usually very small compared with the<br />

<strong>of</strong>fering (small capitalisation). Because the individual purchase <strong>of</strong> forest carbon <strong>of</strong>fsets have the sole<br />

purpose <strong>of</strong> <strong>of</strong>fsetting emissions, it is void <strong>of</strong> speculation.<br />

In general, there are several market structure determinants that are common and applicable to buying<br />

and selling forest carbon <strong>of</strong>fsets on the OTC market:<br />

1. Product complexity and marketability<br />

2. Data and pricing transparency<br />

3. Liquidity<br />

Product Complexity and Marketability<br />

Product complexity refers to the increasing number <strong>of</strong> product options an entity has to <strong>of</strong>fer, and can<br />

affect the marketability and adoption <strong>of</strong> the product (AMD (003, UGS PLM Solutions 2004). The<br />

rationality behind this is that the more simple a product is, the easier it is to sell. This is a particularly<br />

relevant market structure determinant to the selling <strong>of</strong> forest carbon <strong>of</strong>fsets due to the methodological<br />

complexities in measuring and monitoring forest carbon stocks and changes, multiple project options<br />

and strategies which can be used to implement an REDD or forest project, and the different forest<br />

values which can be <strong>of</strong>fered in addition to forest carbon credits, e.g. biodiversity conservation,<br />

community development etc. In all, it can be difficult for consumers <strong>of</strong> forest carbon <strong>of</strong>fsets to know<br />

what they are buying, and therefore, an evolving list <strong>of</strong> standards have emerged to verify a certain<br />

quality and standard are met by forest carbon <strong>of</strong>fset projects.<br />

According to the recent study by Hamilton et al. (2008), over the past 2 years, the number <strong>of</strong><br />

independent verification standards has risen in line with market trends which require the independent<br />

verification <strong>of</strong> projects which reduce GHG emissions. There are now more than a dozen standards for<br />

verification in existence to prove project activities reduce emissions. Some have been developed


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 378<br />

specifically for REDD and avoided deforestation while others have been developed for general<br />

emission reductions verification. In the State <strong>of</strong> the Voluntary Market Report for 2008, the emergence<br />

<strong>of</strong> carbon standards and registries was one <strong>of</strong> the most noticeable trends in 2007, in response to<br />

proving the legitimacy and quality <strong>of</strong> carbon credits. Table 1 illustrates the growing number <strong>of</strong><br />

standards and the different types <strong>of</strong> projects for forest based <strong>of</strong>fsets in the voluntary market.<br />

Table 1. Standards on the voluntary market for forest <strong>of</strong>fsets<br />

Standard Project types<br />

Climate, Community & Biodiversity Climate, Community & Biodiversity Alliance (CCBA) verifies all<br />

(CCB) Standard<br />

land based project types<br />

Climate, Community & Biodiversity<br />

(CCB) Standard, 2 nd In addition, it takes into account indigenous rights and high<br />

Edition biodiversity values<br />

Voluntary Carbon Standard (VCS) Afforestation, reforestation and revegetation, agricultural land<br />

Agriculture, Forestry and Other Land management, improved forest management, REDD<br />

Uses (AFOLU)<br />

Carbon Fix Standard Conversion <strong>of</strong> non-forest land to forest land, conservation forests,<br />

planted sustainably managed forests, protected areas leading to land<br />

use change <strong>of</strong> non forest to forest. Not for REDD.<br />

Carbon Fix Standard 2.1 As above, also includes agro-forestry projects<br />

Plan Vivo Afforestation, reforestation, agro-forestry, restoration, conservation,<br />

improved forest management and REDD.<br />

The Gold Standard Only REDD through fuel efficiency.<br />

VER+ Land Use, Land Use Change & Forestry (LULUCF) projects,<br />

including REDD, are accepted if implemented with a buffer approach<br />

to address the risk <strong>of</strong> potential non-permanence<br />

Green e Agriculture, Forestry and Other Land Uses (AFOLU) projects under<br />

Voluntary Carbon Standard (VCS) 2007, are eligible as long as the<br />

seller provides pro<strong>of</strong> that the native species requirements under the<br />

Green-e Climate Standard are met.<br />

ISO 14064 Special importance for the emerging voluntary approaches to<br />

greenhouse gas declarations by companies. Does not specify or<br />

exclude any project types.<br />

Chicago Climate Exchange (CCX) Protocol for sustainably managed forests, standard for afforestation<br />

and tree planting.<br />

Greenhouse Friendly Special guidelines for forest sink abatement projects.<br />

The Carbon Neutral Protocol Forestry and land use, manage forests to ensure their continued health<br />

and carbon storage capacity, make effective use <strong>of</strong> forest and farm<br />

residues.<br />

California Climate Action Registry Includes Forest Sector Protocols.<br />

Protocols<br />

Social Carbon The methodology is based on the Sustainable Livelihood Approach,<br />

and considers six basic resources: Social, Human, Financial, Natural,<br />

Biodiversity and Carbon.<br />

Voluntary Offset Standard (VOS) Only accepts Verified Emission Reductions (VERs) from projects<br />

implemented using CDM methodologies and Gold Standard <strong>of</strong>fsets.<br />

United National Framework for Only Afforestation and Reforestation, does not include REDD yet,<br />

Climate Change Convention though there is a proposal for the inclusion <strong>of</strong> REDD under the CDM<br />

(UNFCCC) Clean Development to be negotiated within the United Nations framework for climate<br />

Mechanisms (CDM) Afforestation change.<br />

/Reforestation (A/R)<br />

Sources: See websites for standards in references, in addition to Merger (2008), Kollmuss (2008) Ecosystem Marketplace, New Carbon<br />

Finance.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 379<br />

Data and Pricing Transparency<br />

Data and pricing transparency is an important aspect <strong>of</strong> pricing and valuation. This is backed up by<br />

Redmond (1989), Leach and Madhaven (1993) and García-Alonso et al. (1997). Price transparency is<br />

<strong>of</strong>ten limited by the private nature <strong>of</strong> the trades in the OTC market, and there are studies to conclude<br />

that prices on an exchange can better reflect information than prices in the over the counter market<br />

(Bessimbinder 1999). In order to deal with the lack <strong>of</strong> transparency <strong>of</strong> the OTC market, in addition to<br />

the standards, a number <strong>of</strong> carbon credit accounting registries have evolved, and are like public<br />

databases which provide transparency over ownership claims concerning emission reductions.<br />

Table 2. Registries and transparency<br />

Registry Host Transparency Forest<br />

Offsets<br />

American Carbon Registry USA Projects and Account Information Public No<br />

Blue Registry Germany Projects and Account Information Public No<br />

Chicago Climate Exchange USA CCX verified and registered <strong>of</strong>fset projects, list <strong>of</strong> Yes<br />

(CCX) Registry<br />

<strong>of</strong>fset providers and projects, list <strong>of</strong> members, but<br />

linkages between members, providers and projects is<br />

undisclosed.<br />

Asia Carbon Registry<br />

Singapore List <strong>of</strong> projects, place, Certified Emissions<br />

Reductions (CERs) and Verified Emission<br />

Reductions (VERs). Account info not public<br />

Yes<br />

<strong>Australia</strong> Climate Exchange <strong>Australia</strong> Linked with Greenhouse Friendly Abatement Yes<br />

Registry<br />

Register. Account info undisclosed<br />

<strong>Australia</strong>n Carbon Traders<br />

Registry<br />

<strong>Australia</strong> Project and account info public Yes<br />

New South Wales Greenhouse<br />

Gas Reduction Scheme (GGAS)<br />

<strong>Australia</strong> Project and account info public Yes<br />

Greenhouse Gas (GHG) Clean<br />

Projects Registry<br />

Canada Account info and projects are public Yes<br />

Canadian Greenhouse Gas (GHG) Canada Project info is public, but transaction info is not Yes<br />

Reductions Registry<br />

recorded on this registry<br />

California Climate Action USA Tracks and registers voluntary projects that reduce Yes<br />

Registry (CCAR)<br />

emissions <strong>of</strong> GHGs. Publically available.<br />

Triodos Climate Clearing House Netherlands Undisclosed Yes<br />

Globe Carbon Registry Canada The GLOBE Carbon Registry will ensure market<br />

integrity by issuing a unique serial number to each<br />

posted <strong>of</strong>fset, avoiding double counting and allowing<br />

chain <strong>of</strong> custody control from creation to retirement.<br />

Unclear<br />

Gold Standard Registry for Switzerland The registry’s infrastructure prevents double- Possible<br />

Verified Emission Reductions<br />

counting <strong>of</strong> VERs and provides numerous public<br />

(VERs)<br />

reports and a full audit <strong>of</strong> all transactions to ensure<br />

the integrity <strong>of</strong> the Gold Standard VERs.<br />

TZ1 Registry New<br />

Zealand<br />

Full account and project info available to public. Yes<br />

Carbon Offset Project Registry UK Project and credit info available, account info<br />

undisclosed to public<br />

Yes<br />

APX Voluntary Carbon Standard USA The Registry will <strong>of</strong>fer publicly accessible reports Yes<br />

(VCS) Registry<br />

showing the following: a directory <strong>of</strong> organizations Monitored,<br />

using the Registry, a list <strong>of</strong> Voluntary carbon Units measured,<br />

(VCU) registered projects, issued and retired VCUs. & verified<br />

only.<br />

While the registries aid in transparency over the transactions <strong>of</strong> credits and trace transactions and<br />

ownership, pricing transparency is still a problem as the nature <strong>of</strong> the OTC is a deal-by-deal basis. It is<br />

up to the buyer to look around and find the right price <strong>of</strong>fering. There is some independent online<br />

information which lists carbon <strong>of</strong>fset providers (www.carboncatelog.org), their projects and their<br />

prices, and based on a review <strong>of</strong> carbon price data collected from 79 vendors selling forest carbon<br />

<strong>of</strong>fsets within their portfolio (note that several <strong>of</strong> the vendors reviewed included forestry <strong>of</strong>fsets mixed<br />

with other non-forest emission reduction projects), Figure 3 illustrates the variance in the price for<br />

forest carbon <strong>of</strong>fsets (t/CO2).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 380<br />

USD/CO2<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

27.43<br />

13.15<br />

17.36<br />

Europe (8) USA (14) UK (16) <strong>Australia</strong> (34) Canada (6) New Zealand (1)<br />

13.76<br />

14.72<br />

16.54<br />

Series2<br />

Series1<br />

Source: Project reviews and price data were collected from www.carboncatalogue.com, March 19th 2009,<br />

www.carbon<strong>of</strong>fsetguide.com.au, May 22 nd 2009 . Numbers in parentheses indicate the number <strong>of</strong> carbon <strong>of</strong>fset sellers.<br />

Figure 3. Average price <strong>of</strong> forest carbon <strong>of</strong>fsets Q1, Q2 2009<br />

The range <strong>of</strong> price in which forest carbon credits are traded over the counter (i.e. not in auctions or on<br />

exchanges) in different regions <strong>of</strong> the developed world in the first quarter <strong>of</strong> 2009 shows that Europe<br />

(excluding the UK) had by far the largest carbon price range and the highest average price <strong>of</strong> carbon<br />

per ton for forest carbon <strong>of</strong>fsets, while the other regions had the average price <strong>of</strong> carbon between<br />

USD10-20. With such a variance in price for tCO2, it is clear that there is a strong need to improve the<br />

price discovery process for carbon <strong>of</strong>fsets traded on the over the counter market.<br />

Market and Liquidity Fragmentation in the Price Formulation Process<br />

Markets have two broad functions – liquidity and price discovery. The Stern Review, Stern (2006),<br />

states that where markets function, two conditions must hold to reduce GHG emissions efficiently.<br />

Firstly, the marginal social cost <strong>of</strong> carbon must equal the cost <strong>of</strong> abatement, and the second, more<br />

relevant to this analysis, is that to deliver emission reductions at least cost, a common price signal<br />

is required across countries and different sectors at a given point <strong>of</strong> time. The deeper and liquid a<br />

market, the harder it is for an individual trade to affect the overall price and the less volatile the market<br />

will be. Liquidity fragmentation, however, inhibits the efficient exchange interaction which would<br />

normally occur in an open integrated market, as each pool draws its own trades, market data becomes<br />

more difficult to collect and analyse, which can lead to imperfect information, volatility and inefficient<br />

price signalling and price formulation processes. This can already be seen from the lack <strong>of</strong> available<br />

pricing information available, different standards and registries being used across the board to verify<br />

and validate <strong>of</strong>fsets, and large variations in price ranges <strong>of</strong> carbon <strong>of</strong>fsets across countries (see Figure<br />

3). One <strong>of</strong> the key concerns <strong>of</strong> market fragmentation for REDD and forest carbon <strong>of</strong>fsets is that it<br />

could affect the efficient price formulation process via liquidity fragmentation. There is abundant<br />

literature which supports the link between market fragmentation and liquidity fragmentation 2 .<br />

2 Bennett and Wei (2003) showed that compared with a fragmented market structure, an exchange which requires all orders<br />

to interact and compete, produces higher quality price formulation, lower volatility and lower execution costs. Mendelson<br />

(1987) showed that a fragmented markets creates lower liquidity, and higher price volatility, while Madhaven (1995) showed<br />

that market fragmentation increases both the price variations and leads to price inefficiency. A number <strong>of</strong> papers have also<br />

shown that market fragmentation can reduce liquidity. Cohen, Maie, Schwartz and Whitcomb (1982) point out that <strong>of</strong>f-


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 381<br />

Markets fragment when instruments embodying the same risk trade in multiple locations or forms.<br />

Fragmentation may occur across multiple markets that trade identical instruments or across multiple<br />

instruments whose values depend on the same underlying risks (Harris and Mayhew 2005). This is<br />

already prevalent in the OTC dealings for forest carbon <strong>of</strong>fsets with different projects <strong>of</strong>fering carbon<br />

emission reduction <strong>of</strong>fsets using a range <strong>of</strong> implementation methods (sustainable forest management<br />

with conservation, enlargement <strong>of</strong> protected forest area, strategies to reduce emissions from slash and<br />

burn etc), different standards (ISO 16404, Gold Standard etc) and having the information and<br />

transaction information stored in an array <strong>of</strong> registries. There are different levels and strata <strong>of</strong> linkages<br />

between the standards, registries and exchanges. Some registries have linkages with other registries<br />

that have linkages with exchanges, other registries are linked through standards. For example, the<br />

Chicago Climate Exchange (CCX) also hosts the CCX registry and has a board that administer CCX<br />

standards and protocols. Another example would be Triodos Climate Clearing House, which functions<br />

as an independent registry and trading platform in the transaction <strong>of</strong> carbon credits, yet standards<br />

accepted are based on discretion.<br />

One apparent observation from this exercise is that there are a number <strong>of</strong> voluntary market carbon<br />

instruments, which <strong>of</strong>ten carry the same type <strong>of</strong> risks (e.g. permanence) trading across multiple<br />

markets, and in different locations. Information on these trades in the OTC market, is generally<br />

registered to increase transparency, however, not only are the markets clearly fragmented, the<br />

information based in the registries has also become fragmented with the increasing number <strong>of</strong><br />

evolving standards and registries. As stated before, markets function by providing liquidity and a price<br />

discovery process, however, as illustrated, registries, standards and exchanges have linkages. Given<br />

the relatively small capital for voluntary market forest <strong>of</strong>fsets - 0.000527% <strong>of</strong> total global carbon<br />

market value (in 2006, a total <strong>of</strong> USD 21.06 million worth <strong>of</strong> forest <strong>of</strong>fsets was traded on the voluntary<br />

market, compared with a total global carbon market worth USD 40 billion, Hamilton et al. (2008))<br />

disbursed over a growing number <strong>of</strong> market instruments, liquidity fragmentation has become<br />

inevitable.<br />

What this means for REDD and forest <strong>of</strong>fsets is that transaction costs for each project are higher than<br />

exchange traded <strong>of</strong>fsets because it requires a broker to introduce new <strong>of</strong>fset buyers to new <strong>of</strong>fset<br />

sellers due to product and market complexity. When liquidity becomes highly fragmented for forest<br />

carbon <strong>of</strong>fsets, large projects may become difficult to finance and will also take time to finance as the<br />

broker/contributor looks for different sources to leverage funding.<br />

To counter this anomaly, some REDD projects which sell <strong>of</strong>fsets rely to some degree on donor funds<br />

to cover capacity building and monitoring costs. In addition, there are an increasing number <strong>of</strong> funds<br />

being established, and pledges made to tackle deforestation in tropical countries. Each fund is unique<br />

in terms <strong>of</strong> contributors, its aim and the amount it commits, as well as regional focus. The amounts <strong>of</strong><br />

money are actually quite significant – Norway pledged EUR 1.8 billion to REDD over 5 years, and<br />

Germany pledged EUR 500 million, with many other nations making large contributions(see Bellassen<br />

et al. (2008). At this stage it is unclear whether some <strong>of</strong> these funds will be directly linked to the<br />

carbon market to generate carbon <strong>of</strong>fsets, however, a number <strong>of</strong> proposals have suggested that there<br />

should be future efforts to link funds with a carbon market, see The Little REDD Book, for an outline<br />

<strong>of</strong> the 12 international proposals on approaches to market REDD credits (Parker et al. 2008).<br />

CONCLUSION<br />

Forest carbon <strong>of</strong>fsets from REDD and forestry projects are providing increasing opportunities in the<br />

voluntary carbon market. There are a number <strong>of</strong> options on how to reduce emissions from<br />

deforestation and forest degradation including avoiding slash and burn <strong>of</strong> forests, avoiding illegal and<br />

legal land conversion, avoiding illegal logging and encroachment, and the purchase <strong>of</strong> logging<br />

concession sites before harvest. There are a number <strong>of</strong> different instruments which can be used to<br />

exchange executions may benefit brokers but harm the market as a whole. Cohen, Conroy and Maier (1985) show that a<br />

fragmented market may result in a wider bid-ask spread because <strong>of</strong> decreased opportunity for order interaction. Mendelson<br />

(1987) finds that the fragmented market has less liquidity and increases price variances faced by investors. Madhavan (1995)<br />

demonstrates that fragmentation results in higher price volatility and violations <strong>of</strong> price efficiency. Empirically, Amihud,<br />

Lauterbach and Mendelson (2002) provide evidence that trading consolidation improves liquidity and adds value to investors.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 382<br />

achieve reduced emissions forest deforestation and forest degradation. Protected area enlargement,<br />

establishment and/or management, and community development and community level land use<br />

planning are the two most popular types <strong>of</strong> instruments for REDD projects, based on the review <strong>of</strong> 12<br />

REDD projects. Other instruments include sustainable forest management with conservation,<br />

improving forest law enforcement and governance, agr<strong>of</strong>orestry, ecotourism and working with<br />

indigenous rights.<br />

Currently, forest carbon <strong>of</strong>fsets from REDD and most forestry projects are only able to be traded in the<br />

voluntary carbon market on a project by project basis, and on a deal by deal basis through the Over-<br />

The Counter market. The market for REDD and forest based carbon credits can be characterised by<br />

the following trends:<br />

Growth in the number <strong>of</strong> projects <strong>of</strong>fering forest based carbon <strong>of</strong>fsets OTC<br />

Increasing number <strong>of</strong> standards for forest based carbon <strong>of</strong>fsets<br />

Increasing number <strong>of</strong> registries to improve transparency<br />

Increasing number <strong>of</strong> funds available for REDD activities<br />

Uncertainty <strong>of</strong> future linkages with the carbon market<br />

The market for REDD and forest based carbon credits can be characterised by the following market<br />

characteristics:<br />

Only available on the Voluntary Market OTC<br />

Market Fragmentation due to the nature <strong>of</strong> the OTC market<br />

Large degree <strong>of</strong> price variation USD/tCO2, particularly within the EU<br />

Lack <strong>of</strong> a common price signal across markets and sectors<br />

REDD and forest carbon <strong>of</strong>fsets are difficult to market<br />

Small capitalisation distributed over large geographical scope and across many markets has<br />

led to liquidity fragmentation<br />

High transaction costs<br />

Due to the nature <strong>of</strong> the OTC market, market fragmentation is unavoidable due to the product<br />

complexity, lack <strong>of</strong> data and price transparency, and divergent regulations to which different project<br />

regions are subjected. The lack <strong>of</strong> data and price transparency and product complexity means that<br />

forest carbon credits are traded on a deal by deal basis, with prices ranging from USD 2.50-68 per ton<br />

<strong>of</strong> CO2 in the first quarter 2009, based on price data collected from 79 vendors. As mentioned earlier,<br />

to reduce GHG emissions efficiently, one <strong>of</strong> the key market criteria which must hold is that a common<br />

price signal is required across countries and different sectors at a given point <strong>of</strong> time (The Stern<br />

Review). As this paper shows, this is far from the case and that inefficient price formulation is caused<br />

by market and liquidity fragmentation inherit in the over-the-counter market with small capitalisation.<br />

Given that reducing deforestation rates substantially will require significant levels <strong>of</strong> finance (USD<br />

17-30 billion per year to halve the emissions from the forest sector by 2030), there is an urgent need to<br />

consolidate standards, harmonise pricing and transparency mechanisms, to develop efficient and<br />

decisive methods to consolidate the required capital to reduce emissions from deforestation and<br />

degradation on a significant scale.<br />

REFERENCES<br />

American Carbon Registry - http://www.americancarbonregistry.org/carbon-registry [last viewed 19th March<br />

2009]<br />

AMD Global Consumer Advisory Board (2003) Technology Terminology and complexity Study. Online:<br />

http://www.amd.com/us-en/Weblets/0,,7832_8551,00.html<br />

APX VCS Registry http://vcsregistry.apx.com/ [last viewed 19th March 2009]<br />

Asia Carbon Global Registry Services - http://www.asiacarbon.com/Carbon_Registry_Services.html [last viewed<br />

19th March 2009]<br />

<strong>Australia</strong>n Climate Exchange Registry http://www.climateexchange.com.au/Content/rOperation.aspx [last<br />

viewed 19th March 2009]


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 383<br />

<strong>Australia</strong>n Carbon Traders Registry http://www.australiancarbontraders.com/carbonmart/public/register [last<br />

viewed 19th March 2009]<br />

Bellassen, V., Crassous, R., Dietzsch, L. & Schwartzman, S. (2008) Reducing Emissions from Deforestation and<br />

Degradation: What Contribution from Carbon Markets? Mission Climat <strong>of</strong> Caisse des Dépôts . Climate<br />

Report Issue 14, September, 2008, France.<br />

Bennett, P. & Wei, L. (2003) Market Structure, Fragmentation and Market Quality. Evidence from Recent<br />

Listing Switches. Draft Paper http://media.terry.uga.edu/documents/finance/bennett.pdf<br />

Bhattacharya, Mita and Bloch, Harry and Chen, Jong-rong. 2008. Determinants <strong>of</strong> domestic market structure: A<br />

comparative study between Japanese and Taiwanese manufacturing. Asia Pacific Journal <strong>of</strong> Economics and<br />

Business 11 (2): 25-38.<br />

Blue Registry - https://www.netinform.de/BlueRegistry/Projects.aspx [last viewed 19th March 2009]<br />

Bissembinder, H. (1999) “Trade Execution Costs On NASDAQ And The NYSE: A Post-<br />

Reform Comparison,” Journal <strong>of</strong> Financial and Quantitative Analysis, v34[3,Sep], 387-407.<br />

California Climate Action Registry - http://www.climateregistry.org [last viewed 19th March 2009]<br />

California Climate Action Registry Protocols - http://www.climateregistry.org/tools/protocols/projectprotocols.html<br />

[last viewed 19th March 2009]<br />

Cameron, P. (1995) Gas Regulation in Europe. From Monopoly to Competition. Fuel and Energy Abstracts,<br />

Volume 36, Number 4, July 1995, pp255.<br />

Canadian GHG Reductions Registry - http://www.ghgregistries.ca/reductions/index_e.cfm [last viewed 19th<br />

March 2009]<br />

Carbon Catalogue - http://www.carboncatalog.org/ [last viewed 19th March 2009]<br />

Carbon Fix Standard - http://www.carbonfix.info/ [last viewed 19th March 2009]<br />

Carbon Offset Guide – http://www.carbon<strong>of</strong>fsetguide.com.au/ [last viewed 10 th July 2009]<br />

Chicago Climate Exchange, Forestry Carbon Emission Offsets -<br />

http://www.chicagoclimatex.com/content.jsf?id=242 [last viewed 19th March 2009]<br />

Chicago Climate Exchange Registry - http://www.chicagoclimatex.com/content.jsf?id=582 [last viewed 19th<br />

March 2009]<br />

Climate, Community and Biodiversity Standard - http://www.climate-standards.org/ [last viewed 19th March<br />

2009]<br />

Crampton, P. & Kerr, S. (2002) Tradable Carbon Permit Auctions. How and Why to Auction Not Grandfather.<br />

Energy Policy 30, 333-345.<br />

EU_Commission (2008). Proposal for a Directive <strong>of</strong> the European Parliament and <strong>of</strong> the Council amending<br />

Directive 2003/87/EC so as to improve and extend the greenhouse gas emission allowance trading system<br />

<strong>of</strong> the Community (presented by the Commission). 2008/0013 (COD) COM(2008) 16 final.<br />

Federer, J.P. The rise <strong>of</strong> the Over the Counter Market. Department <strong>of</strong> Economics, Malacester College,<br />

USA.pp39.<br />

García-Alonso, M.D.C. & Levine, P. (1997) Strategic Procurement and Market Structure. International Journal<br />

<strong>of</strong> Industrial Organization, Volume 26, Issue 5, September 2008, Pages 1180-1190.<br />

GHG Clean Projects Registry http://www.ghgregistries.ca/cleanprojects/index_e.cfm last viewed 19th March<br />

2009]<br />

Green e Standard - http://www.green-e.org/getcert_ghg_standard.shtml [last viewed 19th March 2009]<br />

Greenhouse Friendly Guidelines - http://www.climatechange.gov.au/greenhousefriendly/publications/gfguidelines.html<br />

[last viewed 19th March 2009]<br />

Hamilton, K., Sjardin, M., Marcello, T., and Xu, G. (2008) Forging a Frontier: State <strong>of</strong> the Voluntary Carbon<br />

Markets 2008. A Report by Ecosystem Marketplace and New Carbon Finance. May 2008, USA.<br />

Harris, L.E. & MayhewS s. (2005) Market Fragmentation Metrics. Independent Paper. California, USA.<br />

IPCC (2007) Fourth Assessment Review. http://www1.ipcc.ch/ [last viewed 19th March 2009]<br />

ISO 14064 - http://www.climatechange.sgs.com/iso_14064_climatechange [last viewed 19th March 2009]<br />

Kollmuss, A., Zink, H., & Polycarp, C. (2008) Making Sense <strong>of</strong> the Voluntary Carbon Market. A Comparison <strong>of</strong><br />

Carbon Offset Standards. Stockholm Environment <strong>Institute</strong>. Tricorona, Sweden.<br />

Leach, J.C. & Madhaven, A.N. (1993) Price Experimentation and Security Market Structure. Review <strong>of</strong><br />

Financial Studies, Volume 6, Number 2, pp 375-404.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 384<br />

Madhaven A (1995) Consolidation, Fragmentation and the Disclosure <strong>of</strong> Trading Information. Review <strong>of</strong><br />

Financial Studies, 8, 579-603.<br />

Madhaven, A. (2000) Market Microstructure: A Survey, Journal <strong>of</strong> Financial Markets, 3, 205-258.<br />

Merger, E. (2008) Forestry Carbon Standards 2008 - A comparison <strong>of</strong> the leading standards in the voluntary<br />

carbon market and the state <strong>of</strong> climate forestation projects. Carbon Positive.<br />

Mendolson, H. (1987) Consolidation, Fragmentation and Market Performance. Journal <strong>of</strong> Financial and<br />

Quantitative Analysis, 22, 189-208.<br />

Parker, C., Mitchell, A., Trivedi, M., & Mardas, N. (2008) The Little REDD Book. Global Canopy Programs,<br />

Oxford, UK. pp 113.<br />

Nelson, J.P. Market Structure and Incomplete Information: Price Formation in a Real-World Repeated English<br />

Auction. Journal <strong>of</strong> Economic Behvaiour and Organisation , Volume 27 Number 3, August 1995, pp<br />

421.237.<br />

NERA Economic Consulting. (2007) Market Mechanism for Reducing GHG Emissions from Agriculture,<br />

Forestry and Land Management. Department for Environment, Food and Rural Affairs. United Kingdom.<br />

NSW GGAS Registry - http://www.ggas-registry.nsw.gov.au/ [last viewed 19th March 2009]<br />

Plan Vivo Standard - http://www.planvivo.org/fx.planvivo/scheme/news.aspx [last viewed 19th March 2009]<br />

Ravindranath, N.H., Ostwald, M. (2008) Carbon Inventory Methods: Handbook for Greenhouse Gas Inventory,<br />

Carbon Mitigation and Roundwood Production Projects, Springer, 304 pages.<br />

Redmond, W.H. (1989) The Effect <strong>of</strong> New Product Pricing on the Evolution <strong>of</strong> Market Structure. Journal <strong>of</strong><br />

Product Innovation Management. Volume 6, Number 2, June 1989, pp 99-108.<br />

Regi - http://www.regi.co.nz/regPublic/registry.mt_public.home [last viewed 19th March 2009]<br />

Senn, M. (2002) Supermontage in the American Securities Market Context. Journal <strong>of</strong> International Financial<br />

Markets. Volume 4. Issue 6. 2002.<br />

Seifert-Granzin, J. (2007) The Noel Kempff Climate Action Project: Challenges for Project Base Approaches in<br />

future REDD Schemes. Presentation to the Conservation Finance Workshop, Vlim, 29th July-Aug 3 rd ,<br />

2007.<br />

Social Carbon Standard - http://www.socialcarbon.org [last viewed 19th March 2009]<br />

Social Carbon registry - http://www.socialcarbon.org/Registry/ [last viewed 19th March 2009]<br />

Stern, N. (2006) The Economics <strong>of</strong> Climate Change. The Stern Review. Cabinet Office, HM Treasury,<br />

Cambridge Press, UK.<br />

The Carbon Neutral Protocol - http://www.carbonneutral.com/uploadedfiles/TCNC%20Protocol%202008.pdf<br />

[last viewed 19th March 2009]<br />

The Carbon Neutral Offset Project Registry - http://www.carbonneutral.com/pages/Offsetprojectregistry.asp [last<br />

viewed 19th March 2009]<br />

The Globe Carbon Registry - http://www.globecarbonregistry.com/ [last viewed 19th March 2009]<br />

The Gold Standard - http://www.cdmgoldstandard.org/ [last viewed 19th March 2009]<br />

Triodos climate Clearing House - http://www.triodos.com/com/climate/ [last viewed 19th March 2009]<br />

UGS PLM Solutions (2004) Reducing Complexity. NX Digital Product Development White Paper.<br />

UNFCCC (2007). COP 13. Revised draft decision -/CP.13 Ad Hoc Working Group on Long-term Cooperative<br />

Action under the Convention. FCCC/CP/2007/L.7/Rev.1 14 December 2007. Bali.<br />

VER+ Standard - http://www.tuev-sued.de/uploads/images/1179142340972697520616/Standard_VER_e.pdf<br />

[last viewed 19th March 2009]<br />

Voluntary Carbon Standard - http://www.v-c-s.org/afl.html [last viewed 19th March 2009]<br />

Voluntary Offset Standard - http://www.carboninvestors.org/documents [last viewed 19th March 2009]<br />

Willett, Meyer. “Over-the-Counter Possibilities.” Wall Street Journal (December 28, 1931): p11.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 385<br />

TASMANIA’S NEW FOREST INDUSTRY PLAN:<br />

RESHAPING THE <strong>FORESTRY</strong> AGENDA<br />

Aidan Flanagan<br />

ABSTRACT<br />

Tasmania is now past the half-way mark in relation to the Regional Forest Agreement.<br />

Over the last 10 years significant changes have occurred within the sector, reflecting<br />

community, environmental, commercial and political pressure. To ensure that<br />

Tasmania’s future incorporates a strong forest industry, the Forests and Forest Industry<br />

Council <strong>of</strong> Tasmania has facilitated the development <strong>of</strong> a New Forest Industry Plan<br />

(NFIP) which will fully utilise over 10 million tonnes <strong>of</strong> wood fibre annually by 2020.<br />

The NFIP contains commitments, strategies and actions which build on Tasmania’s<br />

competitive strengths and deliver investments which focus on: a) creating wealth by<br />

maximising the value <strong>of</strong> forest products and delivering positive economic benefits to<br />

regions; b) delivering rewarding careers for our children; c) creating stronger community<br />

engagement initiatives and responding to community concerns; d) generating healthier<br />

environments and protecting our natural values; and e) adapting to climate change and<br />

reducing the State’s carbon emissions.<br />

OVERVIEW<br />

Tasmania’s native forests have ancient origins that influence their unique characteristics, how they can<br />

be used and why they are highly valued by Tasmanians. Forests are an integral part <strong>of</strong> the landscape.<br />

These forests also produce some <strong>of</strong> the strongest and most beautiful and valuable timbers in the world<br />

which are renowned for their texture and character. In addition to the versatile Eucalypts (dominated<br />

by E. delegatensis, E. regnans and E. obliqua), other important species include Blackwood, Myrtle,<br />

Sassafras, and Celery Top, Huon and King Billy Pine.<br />

Figure 1. Tasmania’s land use<br />

In Tasmania, the forests and forest industry has a long history <strong>of</strong> effective innovation. It is an engine<br />

for advances in improved technology and management systems, and a driver <strong>of</strong> regional economies<br />

and establishing social bonds that make it a special feature <strong>of</strong> the Tasmanian way <strong>of</strong> life.<br />

Forests and Forest Industry Council <strong>of</strong> Tasmania, Level 5-2 Kirksway Place, Battery Point Tas 7004.<br />

Ph +613 62338221. Email aflanagan@ffic.com.au


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 386<br />

Forest operations around the world have gone through dramatic changes over the last 20 years and<br />

Tasmania is no exception. Gone are the days when a forest operation included a crew <strong>of</strong> men on<br />

chainsaws working in front <strong>of</strong> a relatively low cost tractor. Today’s forest industry is a high tech<br />

industry.<br />

Today, only 40% <strong>of</strong> Tasmania’s forests are potentially available for wood production (as indicated in<br />

Figure 1). However, the areas available for harvesting are likely to be less than the 40% identified as<br />

areas potentially available for timber production are <strong>of</strong>ten discounted. In Tasmania, these discounts<br />

include the Forest Practices Code and other conservation provisions, accessibility, economics, safety<br />

and operational constraints apply.<br />

The continuing success <strong>of</strong> the industry has been achieved by focusing on maximising value, improving<br />

productivity and pr<strong>of</strong>itability, committing to sustainable management practices, investing in new<br />

technologies and skilled career development, and responding to market demands. In summary,<br />

maximising value has been about adopting world’s best practice in a competitive market, and then<br />

adapting these to the Tasmanian forest environment.<br />

Its natural resources make Tasmania a State <strong>of</strong> opportunity which can build on its competitive<br />

strengths and deliver investments which focus on:<br />

creating wealth by maximising the value <strong>of</strong> forest products and delivering positive economic<br />

benefits to regions;<br />

delivering rewarding careers for our children;<br />

creating stronger community engagement initiatives and responding to community concerns;<br />

generating healthier environments and protecting out natural vales; and<br />

adapting to climate change and reducing the State’s carbon emissions.<br />

Tasmania’s forests and forest industry’s international competitive position is supported by the security<br />

<strong>of</strong> access to native forest resources provided under the Regional Forest Agreement (RFA); clear<br />

regulatory processes delivered through the Forest Practices Act; a growing hardwood pulpwood and<br />

sawlog plantation resource supported by Plantations <strong>Australia</strong>: the 2020 Vision; a diverse private<br />

contracting and processing sector; and highly skilled employees.<br />

In 1985, Tasmania became the first state in <strong>Australia</strong> to introduce a comprehensive planning system<br />

for forestry, encompassing forest practices legislation, policies and processes.<br />

The Forest Practices Act 1985 provides the world class framework under which balanced triple<br />

bottom-line principles are monitored and regulated “to achieve sustainable management <strong>of</strong> Crown and<br />

private forests with due care for the environment”. Tasmania remains a leader in forest regulatory and<br />

management practices. This system was independently assessed by researchers from Yale University<br />

who concluding in their 2008 report that:<br />

“Tasmania is unique among case study Organization for Economic Co-operation and<br />

Development jurisdictions in applying the same forest practice policies to both public and<br />

private lands in regards to riparian buffers, clearcut size limits, reforestation and road<br />

building 1 ”.<br />

Over the last decade, the Tasmanian RFA has supported a culture <strong>of</strong> continuous improvement and<br />

adaptive management which are embraced and driven by the forest and forest industry, and accredited<br />

under the <strong>Australia</strong>n Forestry Standard. Security provided under the RFA has underpinned<br />

investments in the industry which has diversified agricultural economies by injecting substantial<br />

capital into the State. Since 1997, the industry has generated over $14 billion in turnover, $2.5 billion<br />

in wages, and $2.3 billion in new forests, new technologies and new processing capacity. The RFA<br />

provides a balanced approach for managing natural and heritage conservation values in Tasmania’s<br />

forests. This was confirmed in 2008, when the joint UNESCO World Heritage<br />

1. McDermott C.L., Cashmore B. and Kanowski P, ‘A Global Comparison <strong>of</strong> Forest Practice Policies Using Tasmania as a<br />

Constant Case’, viewed at


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 387<br />

Centre/IUCN/ICOMOS mission report on the conservation <strong>of</strong> the Tasmanian Wilderness World<br />

Heritage Area and concluded that:<br />

“The area managed under the TWWHA management plan provides a good representation<br />

<strong>of</strong> well-managed tall Eucalyptus forest and there is similar forest outside the property<br />

which is also well-managed, but for both conservation and development objectives.”<br />

And<br />

“The threats to these forests from production forestry activities are well managed and<br />

there is no need for the boundary <strong>of</strong> the property to be changed to deal with such<br />

threats.”<br />

Ongoing support for the RFA and balanced regulatory policies will support an expansion in<br />

investment confidence by reducing sovereign risk and providing certainty for financial institutes to<br />

release the capital necessary for businesses to operate. It will also provide resource security and allow<br />

the industry to respond and contribute positively to the environmental challenges posed by climate<br />

change, by adopting new practices, investing in and developing new products and markets, and<br />

delivering community focused outcomes.<br />

Tasmania’s forests have also been a focus for active and robust community debate. This debate has<br />

reflected the variety <strong>of</strong> views and values held by people about the use and management <strong>of</strong> Tasmania's<br />

forests. These differences in views relate mainly to the value that individuals and groups place on<br />

forests and the wood products that flow from them. The espoused values are shaped by many factors,<br />

including economic dependency on forests. That is, the views held by tourist operators, contractors,<br />

apiarists, sawmillers, and craft wood users, differ from those with a non-commercial dependence,<br />

such as recreationists and tourists.<br />

Irrespective <strong>of</strong> these differences, it is clear that Tasmania has now achieved a balance between the<br />

needs <strong>of</strong> forest-based industries, the community, existing agricultural activities, and the environment.<br />

Research undertaken by FFIC (unpublished) indicates that the majority <strong>of</strong> Tasmanians support<br />

appropriate ‘triple bottom line’ outcomes from sustainable forestry activities. This research has shown<br />

that employment opportunities, retention <strong>of</strong> young people within the State, and wealth creation are<br />

recognised and valued by many within the wider community, along with forest-based tourism,<br />

indigenous participation, and environmental and cultural values.<br />

Ultimately, the success <strong>of</strong> the Tasmanian forests and forest industry will be determined by investment<br />

decisions, made outside <strong>of</strong> government, which are driven by innovation, financed by private<br />

investment, and responsive to market and community signals.<br />

It is now appropriate that the future forest industry focuses on new approaches and new opportunities.<br />

In order to support a future within which the Tasmanian forests and wood products industry has a lead<br />

role, the New Forest Industry Plan (NFIP) was developed to shape positive future government and<br />

industry programs and policy initiatives, and provide a sound basis for future planning and investment<br />

decisions which promote growth and innovation across Tasmania’s forest and forest industry supply<br />

logistics and value chains.<br />

The NFIP has been developed by applying three principles.<br />

1 There is only one forest industry in Tasmania and it is about highly skilled people and<br />

communities. Over 10,000 people and their families are directly dependent on the industry for<br />

employment and a sense <strong>of</strong> ‘place’.<br />

2 While the industry is made up <strong>of</strong> many businesses and many people, all are interdependent<br />

and reliant on the survival and prosperity <strong>of</strong> the supply, logistic and value chains as a whole.<br />

3 Good land management in Tasmania is no longer restricted by boundaries and there is no<br />

longer a need to extend reservations or further restrict access to forest resources.<br />

The NFIP builds on the success <strong>of</strong> the 1991 Secure Futures for Forests and People: the Forests and<br />

Forest Industry Strategy, and the 1997 Tasmanian RFA. It is also consistent with the National Forest


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 388<br />

Policy Statement and Plantations for <strong>Australia</strong>: The 2020 Vision. The NFIP is an industry plan, and<br />

was developed from input received through a series <strong>of</strong> industry workshops, and a number <strong>of</strong> broader<br />

symposia and strategic meetings held in early 2009. The NFIP provides strong commitments. These<br />

are supported by strategies, priority actions and general actions which promote investment and growth<br />

opportunities identified under five Elements: 1) Wealth Creation; 2) Industry <strong>of</strong> Choice; 3)<br />

Community Engagement; 4) Healthier Environments; 5) Climate Change.<br />

The NFIP provides the framework for industry, governments and the wider community to work<br />

constructively together to drive innovation and stimulate economic development essential to sustain<br />

services that support our aging population. It provides opportunities to create rewarding careers for our<br />

children, and create stronger links within the community so that Tasmanians can be confident that our<br />

forests are sustainably managed and our forest products meet international certification standards. It<br />

will create healthier environments by adopting new approaches which provide solutions to<br />

environmental challenges facing Tasmania, including climate change.<br />

The NFID also recognises that there is only one future for Tasmania’s forest industry. It is a future<br />

that incorporates native and plantation forests, and private and public forests. It includes saw and pulp<br />

logs; paper and specialty manufacturing; growing trees and processing wood; large companies and<br />

small businesses. It also consists <strong>of</strong> contractors and their employees; honey production and tourism;<br />

furniture and craft wood products; recreation and cultural heritage; and bioenergy and carbon storage.<br />

The NFIP recognises that the growth in the plantation estate, increased productivity from native<br />

forests, higher value processing and market development are creating opportunities for growth. It also<br />

sets forth commitments and strategies which will be progressed by supporting regulatory systems<br />

which promote open, competitive and non-discriminating markets.<br />

Tasmania’s extensive plantation and native forest resources are expected to provide over 10 million<br />

tonnes <strong>of</strong> wood products annually by 2020. This expanding resource creates opportunities for existing<br />

businesses to expand, and for new industries to develop. Over $2.1 billion <strong>of</strong> investment in new<br />

processing opportunities has been identified. These investments would benefit Tasmania by directly<br />

creating up to an additional 855 long term, highly technical career opportunities and directly generate<br />

an extra $1.17 billion annually in wealth to the state’s economy. Table 1 summarises income and<br />

employment impacts <strong>of</strong> new processing investments opportunities within Tasmania 2 .<br />

Table 1. Income and employment impacts <strong>of</strong> new processing investments<br />

Facility<br />

Minimum<br />

capital<br />

investment<br />

($ million)<br />

Annual direct<br />

income<br />

($ million)<br />

Long term<br />

employment in<br />

new facilities<br />

(No. <strong>of</strong> jobs)<br />

Input volume <strong>of</strong><br />

wood required<br />

(million m 3 or tpa)<br />

Hardwood pulp mill 1,450 660 290 4 Mt<br />

ESL plant 225 290 150 0.55 Mt<br />

Hardwood sawmilling<br />

- 3 reciprocated mills<br />

60 50 200 0.24 Mm 3<br />

Hardwood sawmilling<br />

65 50 65 0.25 Mm 3<br />

- 1 linear mill<br />

Hardwood plywood 15 20 50 0.25 Mm 3<br />

S<strong>of</strong>twood sawmilling 10 30 30 0.165 Mm 3<br />

Bioenergy – (based<br />

on 3 bio-electricity<br />

plants and 1 export<br />

wood pellet plant)<br />

370 120 165 1.15 Mm 3<br />

Total (reciprocated<br />

hardwood sawmills 2,130 1,170 885<br />

option)<br />

Total (linear hw<br />

sawmill option)<br />

2,135 1,170 750<br />

2 URS Forestry report, 2009. “Economic Impacts <strong>of</strong> Potential Forest Industry Developments in Tasmania’.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 389<br />

There are also opportunities to increase supply chain efficiencies, enhance forests management, reduce<br />

the threat <strong>of</strong> wildfire and increase the use <strong>of</strong> wood products. These opportunities can, in themselves,<br />

significantly improve competitiveness by increasing value recovery across the supply and logistic<br />

chains. Figure 2 displays the value recovery opportunities within the harvesting and haulage sector 3 .<br />

Sub optimal log<br />

making<br />

52%<br />

Felling Damage<br />

7%<br />

Thinning<br />

2%<br />

Sub optimal choice<br />

grades or stands<br />

30%<br />

Extraction breakage<br />

2%<br />

High stumps and<br />

butt damage<br />

5%<br />

Log Making<br />

Damage<br />

2%<br />

Figure 2. Value recovery opportunities within the harvesting and haulage sector<br />

These new opportunities, when added to the servicing <strong>of</strong> existing processing and manufacturing<br />

businesses, will provide secure opportunities for forest contractors. These opportunities will require<br />

progressive investments over 10 years in new harvesting machinery <strong>of</strong> around $230 million and<br />

transport investments <strong>of</strong> around $135 million. Investments in harvesting and haulage contracting<br />

capacity will directly support over 650 jobs. This level <strong>of</strong> investment will generate revenues <strong>of</strong> $240<br />

million annually, while delivering efficiencies and reducing delivery costs.<br />

However, the transition from native to plantation forest resources is reliant on the resolution <strong>of</strong> a<br />

number <strong>of</strong> technical and economical challenges, including loss in the value <strong>of</strong> dry output associated<br />

with creating high value appearance products from plantation grown Eucalyptus nitens. Currently E.<br />

nitens is unsuitable as a replacement for traditional native forest products, particularly within the solidwood<br />

‘appearance’ grade market. Further research is required to evaluate the increased suitability and<br />

utilisation <strong>of</strong> E. nitens through improved breeding and silviculture, harvest and transport handling<br />

systems, and sawing and drying methods.<br />

Maximising the value <strong>of</strong> Tasmania’s forests and forest products also will benefit the wider Tasmanian<br />

community. Applying standard forestry employment and income multipliers, additional support and<br />

service investments would generate at least an additional 1000 jobs and around $1 billion in direct and<br />

indirect income within the State. Service and support sectors which are likely to benefit from<br />

increased forest processing and manufacturing capacity include:<br />

• housing construction and accommodation – increased regional employment will stimulate<br />

demand<br />

• accounting, financial and administration services – there is likely to be demand for capital<br />

financing support, accounting and administration services<br />

• community services – increased regional employment will require community services, such<br />

as schooling, health services, and social and sporting facilities<br />

• communication services – efficient communication networks will be required to support sales<br />

and marketing systems<br />

• engineering, mechanical and technical services – these include computer maintenance,<br />

electrical, mechanical, civil and structural skills required during the design and construction<br />

phase, as well ongoing plant and equipment maintenance requirements<br />

• environmental services – construction and operations will require ongoing environmental<br />

assessment and monitoring services<br />

3 CRC for Forestry, ‘Maximizing Value Recovery along the Forest-to-Mill Supply Chain’, June 2009 workshop.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 390<br />

• labour services – ongoing career and skill development services will be required, and will<br />

drive demand for education and training support.<br />

• transport – increased movement <strong>of</strong> people (employees and their families) and materials will<br />

generate demand for efficient transport services within regions, and between regions and ports<br />

CREATING WEALTH<br />

Tasmanian forests produce around six million cubic metres <strong>of</strong> logs annually. These are used to make<br />

sawn timber and panels for homes and commercial buildings; industrial material for construction;<br />

unique and high quality flooring, joinery, designs in furniture and craft products; composite products<br />

such as Timbercrete (a brick which incorporates plantation timber waste with cement, sand and<br />

binders); and a range <strong>of</strong> high value products such as writing, packaging, tissue and other paper<br />

products. Forest and mill residues may also be used to produce landscaping material, firewood and<br />

bioenergy. Most <strong>of</strong> these products are sold within <strong>Australia</strong>n markets, although exports are an<br />

important part <strong>of</strong> the industry.<br />

This diversity in processing and manufacturing has helped to cushion our forest-reliant businesses,<br />

employees, their families and many rural communities from prolonged and deepening impacts <strong>of</strong><br />

market fluctuations experienced by many other sectors <strong>of</strong> Tasmania’s economy.<br />

Figure 3. Composition <strong>of</strong> the forests and forest industry employment in Tasmania<br />

Forestry and timber processing continues to contribute significantly to Tasmania’s economic<br />

prosperity and social wellbeing. The forests and forest industry:<br />

• is the second highest manufacturing contributor to Tasmania’s gross state product and<br />

contributes up to $1.6 billion to the State economy;<br />

• is an integral part <strong>of</strong> the Tasmanian community, and directly employs over 6000 people<br />

(Figure 2) and in excess <strong>of</strong> 10,000 in businesses providing support and services;<br />

• encompasses 1600 private forest growers (the majority being farm based); 500 businesses<br />

which includes large forest growers, contracting and transport providers, and processors;<br />

• is unique in that it is a net exporter <strong>of</strong> forest products; and<br />

• is a nationally important contributor to reducing our reliance on imported wood products.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 391<br />

If the forests and forest industry is not investing, developing and applying innovative solutions and<br />

approaches, and improving competitiveness, its ability to adapt and continue to provide ongoing<br />

benefits to the economy and community <strong>of</strong> Tasmania will be diminished.<br />

People within the forest and wood products industry have demonstrated a commitment and capacity to<br />

adapt to change by embracing technology, enhancing their skills and thus broadening their career<br />

opportunities. This commitment has resulted in a rationalisation <strong>of</strong> opportunities within certain sectors<br />

in the short to medium term, while providing new career opportunities in others.<br />

The relatively recent change in resource reliance from old-growth to smaller diameter regrowth and<br />

plantation sawlogs has resulted in significant adaptation along the supply and logistics chains, with<br />

innovation playing a major role in improving timber utilisation by focusing on maximising recovery.<br />

The growing and processing <strong>of</strong> smaller diameter logs is driving the adoption <strong>of</strong> new approaches to<br />

forest management, harvesting and transport, mill processing and drying, business management and<br />

market development.<br />

The increasing importance <strong>of</strong> carbon storage will require innovative solutions to achieve more<br />

accurate systems for measuring, accounting and reporting carbon sequestration within plantations and<br />

native forests, as well as assessing new products to reduce carbon emissions, such as bioenergy and<br />

bi<strong>of</strong>uels.<br />

The NFIP provides a range <strong>of</strong> options for what the industry could look like in 10+ years and highlights<br />

areas <strong>of</strong> innovation to:<br />

develop and adopt new processing and re-tooling technologies to deal with changing resource<br />

qualities and allow for more efficient utilisation <strong>of</strong> resources;<br />

respond to emerging markets for new products that may arise from global initiatives to combat<br />

climate change;<br />

drive efficiency and add value across the supply and logistics chains;<br />

adopt new designs and marketing strategies which strengthen consumer linkages;<br />

achieve the necessary scale required to remain internationally competitive;<br />

invest in people and create skilled careers;<br />

adopt strategic management techniques to target high value market opportunities with<br />

products that feature Tasmanian timber; and<br />

invest in sophisticated sales and distribution systems which respond to information and<br />

products sought by architects and specifiers.<br />

The NFIP identifies value adding opportunities for a range <strong>of</strong> new approaches in forest management,<br />

in supply and logistics, and in processing. These opportunities include the need to expand our existing<br />

pulp and paper production and to restructure existing hardwood sawmilling capacity to include greater<br />

utilisation <strong>of</strong> plantation resources. It also includes the establishment <strong>of</strong> engineered wood product<br />

production facilities, an ongoing need to focus on high value appearance type products for domestic<br />

and export markets, and the establishment <strong>of</strong> sustainable bioenergy production. These opportunities<br />

will require ongoing innovation, modernisation and development by industry partners, and should be<br />

viewed as evolutionary, reflecting a continuum in adding value rather than radical change.<br />

Tasmania’s extensive plantation and native forest resources are expected to provide up to 10.5 million<br />

tonnes <strong>of</strong> wood products annually (22.3% from native forests). This expanding resource creates<br />

processing investment opportunities totalling over $2.2 billion. These will benefit Tasmania by<br />

directly creating an additional 850+ long term, highly technical career opportunities and generate an<br />

extra $1.1 billion annually in wealth to the State’s economy.<br />

In addition, these new opportunities, will provide secure opportunities for forest harvesting contractors<br />

and transport providers,, and a wide range <strong>of</strong> service providers.<br />

Discriminatory or inefficient policies will increase sovereign risk and undermine investor confidence,<br />

jeopardise critical developments and place at risk Tasmania’s future economic and job creation<br />

prosperity. The NFIP recognises the need for governments to review polices to reduce sovereign risk


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 392<br />

and support economic growth. Strong, consistent and non-discriminatory regulatory systems will<br />

continue to support investments in Tasmania only where such a balance is achieved. The NFIP also<br />

identifies further scope to refine existing frameworks to ensure they deliver enhanced security and<br />

thereby contribute to attracting new investment.<br />

The extensive involvement <strong>of</strong> governments in promoting sustainable regional development means that<br />

they have an important role in developing a strong economy through a consistent approach to regional<br />

planning, and investing in integrated infrastructure networks which support efficiencies across supply<br />

chain networks and promote international competitiveness and market development.<br />

INDUSTRY OF CHOICE<br />

The NFIP recognises that people are the State’s most valuable natural resource. Tasmania’s success in<br />

today’s competitive, global economy depends on the pr<strong>of</strong>essionalism and skills <strong>of</strong> the people<br />

employed in businesses, and the daily interactions in workplaces. In effect, the success <strong>of</strong> the<br />

Tasmanian forests and forest industry relies on it being seen as ‘an industry <strong>of</strong> choice’ with attractive<br />

careers that <strong>of</strong>fer interesting work, safe and healthy work conditions, ready access to training and skill<br />

development, and remuneration <strong>of</strong> an acceptable level.<br />

The strength <strong>of</strong> this State’s economy relies on local businesses to provide career opportunities while<br />

maximising productivity and maintaining viable regional communities. However, Tasmania is<br />

entering ‘the age <strong>of</strong> the ageing’, and this demographic change is impacting on the industry’s ability to<br />

attract and retain skilled employees at all levels <strong>of</strong> the labour force. This is constraining industry<br />

development and investment by increasing costs and uncertainty for investors.<br />

Tasmania’s forest resources continue to change and the skills required in the areas <strong>of</strong> technology and<br />

environmental management will continue to evolve. Due to these changes, employees will require a<br />

broader and higher level <strong>of</strong> skills, and career development will require the support <strong>of</strong> industry decision<br />

makers and governments to meet these evolving requirements.<br />

Government policies must support stable employment, create opportunities for new careers, and<br />

ensure Tasmania’s economy remains strong during difficult times. The NFIP identifies the lead role<br />

that governments have in facilitating stronger links between industry, education and training<br />

organisations. We must also build on the strengths and synergies provided by vocational training and<br />

higher education structures while fostering the development <strong>of</strong> systems that support career<br />

development, not just job placement.<br />

The NFIP identifies future initiatives which will need to deliver targeted, relevant and flexible training<br />

systems, which attract and retain the people the forest industry requires if it is to become an industry<br />

<strong>of</strong> choice.<br />

Tasmania’s forest industry is committed to the continuing provision <strong>of</strong> career opportunities for<br />

indigenous people, and is proud to have the highest proportion <strong>of</strong> indigenous employees in the State.<br />

The industry supports initiatives, such as the National Indigenous Forestry Strategy. This provides a<br />

framework for industry to work with indigenous communities to build partnerships which focus on the<br />

creation <strong>of</strong> career opportunities across all sectors while protecting cultural values, and encouraging<br />

greater participation in land management.<br />

The NFIP identifies further employment and career development opportunities across many different<br />

disciplines and supports strong and stable communities. The industry is committed to working with<br />

governments and the community to develop a stronger and more resilient economy, through a<br />

consistent approach to regional and career project planning.<br />

COMMUNITY ENGAGEMENT<br />

Community-focused engagement is a central platform <strong>of</strong> the NFIP. The industry is committed to<br />

addressing community concerns by developing innovative solutions which are supported by<br />

scientifically based evidence, better practices and supportive policies.<br />

In Tasmania, rural populations continue to decline as families leave the regions, and farms are<br />

amalgamated into larger holdings. This has long been a global trend, but one which has been


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 393<br />

accelerated in Tasmania by the adoption <strong>of</strong> mechanisation which requires fewer workers, depressed or<br />

volatile prices for products, encouragement <strong>of</strong> youth to pursue alternative city-based careers, and<br />

family hardship associated with ongoing drought.<br />

During these times <strong>of</strong> change, improved management <strong>of</strong> private forests and expanded plantation<br />

development have helped to maintain rural populations. Thus, the expansion <strong>of</strong> plantations is one<br />

aspect <strong>of</strong> the continuum <strong>of</strong> change to rural landscapes, and helps maintain a viable traditional<br />

agricultural communities.<br />

It is therefore important that government and industry policy fosters innovative new investments in<br />

regional communities and alternative land uses as a means <strong>of</strong> creating employment, retaining young<br />

people, and maintaining regional economic, community and environmental values. The Tasmanian<br />

forests and forest industry brand will help to build and sustain community trust by promoting<br />

initiatives, such as the ‘Forestry Good Neighbour Charter’, and the uptake <strong>of</strong> forest certification<br />

systems. The industry is also committed to developing clear, and consistent approaches for<br />

communicating achievements in sustainability and regulatory compliance; and promoting public<br />

understanding <strong>of</strong> the environmental and carbon-friendly credentials <strong>of</strong> sustainable forestry.<br />

HEALTHIER ENVIRONMENTS<br />

Tasmania still has a higher percentage <strong>of</strong> native forest than any other State in <strong>Australia</strong>, and some 65%<br />

<strong>of</strong> this native forest is not available for wood production.<br />

Tasmania’s forests and forest industry is committed to improving the health <strong>of</strong> the urban and natural<br />

environments. It will achieve this by:<br />

promoting independent certification schemes;<br />

reducing forest impacts on the community’s health and lifestyle ;<br />

enhancing biodiversity values;<br />

conserving cultural heritage sites;<br />

reducing the threat <strong>of</strong> wildfire;<br />

improving the quality <strong>of</strong> water; and<br />

reducing our reliance on chemicals.<br />

Scientific research supports the view that conservation and biodiversity values are enhanced through<br />

the management practices being implemented by Tasmania’s forests and forest industry. Managed<br />

forests can redress many <strong>of</strong> the environmental and social problems facing rural and regional<br />

communities if they are strategically planned and located.<br />

Commercial forestry activities fund environmental and heritage initiatives which support publicly<br />

funded programs with over 23,500 hectares <strong>of</strong> formal reserves having been established under Forest<br />

Practices Plans lodged with the Forest Practices Authority, which administers the Forest Practices Act<br />

1985.<br />

Private forest owners manage more land for non-commercial returns than any other sector <strong>of</strong> the<br />

economy. In total, over 68,000 hectares <strong>of</strong> private plantation (19%) land are managed for noncommercial<br />

returns such as conservation, water quality enhancement and fire protection — all without<br />

public funding. In addition, over 600,000 (>50% <strong>of</strong> total) <strong>of</strong> State multi use forests are also managed<br />

for non-commercial returns.<br />

These commitments and achievements are being supported by the adoption <strong>of</strong> independent,<br />

international recognised forest management and sustainability standards. Over 1.83 million hectares<br />

<strong>of</strong> managed forests in Tasmania are accredited as being sustainably managed under the <strong>Australia</strong>n<br />

Forestry Standard (AFS), which is endorsed internationally by the Programme for the Endorsement <strong>of</strong><br />

Forest Certification — the world's largest forest certification body.<br />

The NFIP recognises the potential for organisations and governments to undermine the significant<br />

environmental and community benefits <strong>of</strong> the AFS certification processes through the adoption <strong>of</strong><br />

restrictive trade and discriminatory or punitive approaches to certification standards.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 394<br />

Consequently, the NFIP identifies the need for governments to adopt appropriate procurement<br />

policies, and build standards and associated regulations which recognise, , all internationally<br />

accredited, independent national forest certification schemes. Governments should act against the use<br />

<strong>of</strong> public funds to support organisations and businesses that restrict trade by discriminating against<br />

Tasmanian timber which meets third party certification.<br />

CLIMATE CHANGE<br />

The forestry and forest products industry helps to reduce the adverse impacts <strong>of</strong> climate change<br />

through reducing Tasmania’s reliance on imported coal-based electricity generation. From a national<br />

perspective, the total annual increase in carbon stored in managed native and plantation forests is 2.4<br />

times that lost through decay <strong>of</strong> forest wastes produced during harvesting. The industry’s ability to<br />

capture carbon within its forests and store it in wood products has reduced the State’s total carbon<br />

emissions by nearly 30% as indicated in Figure 4.<br />

million tonnes CO2-e<br />

3.0<br />

2.0<br />

1.0<br />

0.0<br />

-1.0<br />

-2.0<br />

-3.0<br />

Agriculture<br />

Figure 4. Tasmania's GHG emissions by sector 4<br />

Transport<br />

Energy:<br />

Manufacturing and<br />

construction<br />

Industrial<br />

processes<br />

Energy: Electricity<br />

.The NFIP has identified strategies and actions which will help the industry to reduce the State’s<br />

carbon emissions and increase carbon storage capacity by:<br />

capitalizing on carbon stored in forests through enhanced management practices;<br />

maximizing the use <strong>of</strong> wood products in residential and commercial constructions;<br />

expanding the capital value <strong>of</strong> farms by integrating forestry to <strong>of</strong>fset on-farm carbon<br />

emissions; and promoting renewable bioenergy opportunities.<br />

To achieve the benefits provided through these initiatives, the NFIP advocates a government review <strong>of</strong><br />

policies and the removal <strong>of</strong> impediments which distort markets by restricting the management and<br />

expansion <strong>of</strong> forests, and discriminate against the use <strong>of</strong> wood products in construction and the<br />

generation <strong>of</strong> power and heat.<br />

CONCLUSION<br />

Tasmania’s future economic and social growth relies on a strong forest industry embracing change by<br />

focusing on maximising value across the supply, logistics and value chains. The career development<br />

opportunities and benefits <strong>of</strong>fered by an innovative and pr<strong>of</strong>itable industry are significant. The NFIP<br />

will be achieved by developing partnerships between the industry and the community, and focusing on<br />

building trust and delivering on commitments. Tasmania’s forest industry can help in establishing a<br />

healthier Tasmania by providing solutions to environmental challenges, including climate change.<br />

The industry has a long history <strong>of</strong> innovation, and the NFIP represents a commitment by individuals<br />

and businesses to continue to adapt to change, and embrace solutions and technologies which meet the<br />

social, environmental and economic needs <strong>of</strong> the State.<br />

4 DCC (2008b). State and Territory Greenhouse Gas Inventories. Department <strong>of</strong> Climate Change. Available at: http://www.climatechange.gov.au/inventory/2006/index.html<br />

generation<br />

Energy: Other<br />

industries<br />

Waste<br />

Land use change<br />

Afforestation &<br />

Reforestation


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 395<br />

ABSTRACT<br />

RESHAPING THE <strong>FORESTRY</strong> AGENDA IN QUEENSLAND<br />

Sean Ryan 1<br />

Private Native Forestry will need to fill the coming void in hardwood production in<br />

Queensland. In 15 years with the closure <strong>of</strong> State Native forests and the production lag<br />

in the fledgling State-based nanoscale hardwood plantation program, there needs to be<br />

a seismic shift in forestry foresight and planning. It needs to be recognised that<br />

plantations with their significant initial carbon footprint, genetic pollution and cost<br />

factors are not the only solution to future production scenarios. In SE Queensland there<br />

are over 1.2 million hectares <strong>of</strong> remnant classified private native forests, mostly in a<br />

very unproductive condition due to lack <strong>of</strong> management. However, more importantly,<br />

from a future productivity point <strong>of</strong> view, there is in excess <strong>of</strong> 500 000ha <strong>of</strong> regrowth,<br />

non-remnant classified forest. This young regrowth has generally not had the negative<br />

impacts <strong>of</strong> high-grading and suppression evident in our mature forests, and is still in or<br />

near ‘free-growth’ mode. Thinning trials established by Private Forestry Southern<br />

Queensland in 10 different locations and forest types across south east Queensland, and<br />

the commencement <strong>of</strong> a thinning harvest <strong>of</strong> 1000ha <strong>of</strong> regrowth from a 20 year old<br />

clearfall operation, has demonstrated a cost effective, environmentally sound system to<br />

rapidly fill the hardwood production shortfall.<br />

The paper outlines the initial condition <strong>of</strong> the regrowth forests, the process involved<br />

including costs and growth data gathered to date and the improvements in forest health,<br />

ground cover and forest productivity achieved. It will also outline the potential extent<br />

<strong>of</strong> the resource, forecast product range and potential growth rates.<br />

INTRODUCTION<br />

<strong>Australia</strong> has supported a 2 billion dollar national forest product deficit for many years, which includes<br />

a sawn wood deficit <strong>of</strong> 450 000 m³. Over the last ten years the trade deficit (the extent to which<br />

imports exceed exports) has been running at a fairly constant level between $1.7 and $2.2 billion per<br />

year with an average <strong>of</strong> $1.9 billion (Stanton 2005).<br />

Queensland is a net importer <strong>of</strong> overseas forest products, with an overall trading deficit <strong>of</strong> about $447<br />

million in 2002/03, an increase <strong>of</strong> about $166 million, compared to a decade earlier. In 1991/92<br />

approximately 2 per cent <strong>of</strong> the total turnover <strong>of</strong> the forest industry was sold overseas. By the end <strong>of</strong><br />

the decade, in 1999/2000, this ratio had risen to 5.5 per cent.<br />

The import penetration <strong>of</strong> the Queensland market for forest products (that is, the share <strong>of</strong> imports in<br />

total estimated sales to the domestic market) rose from 12.6 per cent in 1991/92 to 17.5 per cent in<br />

1999/2000. An estimated 860 000 cubic metres <strong>of</strong> sawn timber products came out <strong>of</strong> the Queensland<br />

industry in 2002/03, <strong>of</strong> which 88 per cent was sold locally. Queensland consumes 1.1 million cubic<br />

metres <strong>of</strong> sawn timber product each year. The current hardwood sawn timber deficit <strong>of</strong> around 100<br />

000m³is expected to increase to around 200 000 m³ by 2015 (Fegely & Follas, 2006).<br />

About two-thirds <strong>of</strong> Queensland’s overseas imports <strong>of</strong> forest products are sourced from producers in<br />

four countries, New Zealand, Indonesia, Malaysia and China, with about 33% ($216 million) sourced<br />

from New Zealand (Marohasy 2005).<br />

1<br />

Executive Officer, Private Forestry Southern Queensland, 224 Mary Street Gympie 4570. Ph. 07 54836535.<br />

Email -pfsq@bigpond.com.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 396<br />

CURRENT PRODUCTION<br />

Forestry Plantations Queensland (FPQ) states in its 07/08 Annual Report that its business objectives<br />

are to operate as a responsible commercial plantation forest manager and maximise the market<br />

value <strong>of</strong> its assets consistent with:<br />

- best practice principles for sustainable plantation forest management;<br />

- the directions <strong>of</strong> its responsible Minister; repairs<br />

- the operations and financial parameters <strong>of</strong> its operational plan;<br />

- the Queensland government’s commitments for hardwood plantations.<br />

Although FPQ does not have as a specific business objective to increase the supply <strong>of</strong> timber to meet<br />

Queensland’s Timber deficit, it does produce 1.26 million m³ <strong>of</strong> exotic pine sawlogs and .42m m³ <strong>of</strong><br />

Hoop Pine sawlogs and purchased an additional 4,942ha <strong>of</strong> land suitable for exotic pine plantation<br />

development in the 07/08 financial year. (07/08 Annual report).<br />

POLICY PLANNING VOID<br />

Queensland does not have a development plan to address the trade deficit. It has been 15 years since<br />

the Federal and State Governments signed the National Forest Policy Statement (1992). This<br />

Statement outlined agreed objectives and policies for the future <strong>of</strong> <strong>Australia</strong>'s public and private<br />

forests.<br />

The strategy and its policy initiatives were to lay the foundation for forest management in <strong>Australia</strong><br />

into the next century. Some <strong>of</strong> the Policy Initiatives specific to Private Native Forests in the document<br />

include:<br />

• Sustainable management <strong>of</strong> private native forests will be encouraged through a combination <strong>of</strong><br />

measures that may include dissemination <strong>of</strong> information about, and technical support for, forest<br />

management, education programs, conservation incentives, land-clearing controls, harvesting<br />

controls, and codes <strong>of</strong> forest practice.<br />

• Encouraging private landowners to manage forests for long-term economic use by removing any<br />

unnecessary impediments or disincentives. The Governments will develop a range <strong>of</strong> incentives<br />

and programs to promote sustainable management <strong>of</strong> native forests on private land. These<br />

incentives and programs will be designed to ensure active management <strong>of</strong> private native forests<br />

for both ecologically sustainable wood production and nature conservation, so that the private<br />

native forest estate will remain a permanent resource.<br />

In his recent Statement to Federal Parliament ‘Preparing our Forest Industries for the Future’ Tony<br />

Burke, Minister for Agriculture, Fisheries and Forestry, stated that ‘The Government committed $20<br />

million to assist industry and support jobs, particularly in regional communities, through measures<br />

that:<br />

• invest in value-adding – through the Forest Industries Development Fund;<br />

• address long-term skills and training shortages – through the new Forest and Forest Products<br />

Industry Skills Council and database development;<br />

• deal with climate change – by addressing major knowledge gaps; and<br />

• work with our Asia-Pacific neighbours and industry to tackle illegal logging – by investing in<br />

capacity building, certification, improving governance and in developing a Regulatory Impact<br />

Statement’.<br />

In Queensland’s hardwood industry the problem is not the processing industry; it has always been the<br />

resource, or lack <strong>of</strong> it, that has resulted in seven hardwood mills closing down in SE Queensland in the<br />

last 12 months. Dr Gary Bacon in a recent interview for the in-wood magazine states ‘I am saddened<br />

by this because native forests management is a decentralised regionally-based activity. On one hand<br />

you hear this litany from government jobs, jobs, jobs, and yet they take away from these communities a<br />

sustainable production from native forests’.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 397<br />

The failed RFA process committed many millions <strong>of</strong> dollars to the mills with state hardwood<br />

allocation for retooling and value adding, but nothing to improving the productivity <strong>of</strong> the resource,<br />

where the problem clearly lay.<br />

Minister Burke’s statement went on to say ‘The Rudd government remains fully committed to RFAs as<br />

the primary mechanism to sustain jobs and support industry, to ensure high conservation values, and<br />

for the protection <strong>of</strong> biodiversity and threatened species.’ But <strong>of</strong> course Queensland does not have an<br />

RFA and there is no State-based management plan for the future <strong>of</strong> the hardwood industry other than a<br />

commitment to a total <strong>of</strong> 20,000ha <strong>of</strong> hardwood plantation to replace the 3 million hectares <strong>of</strong> Native<br />

hardwood resource.<br />

THE REGROWTH RESOURCE<br />

PFSQ’s mapping shows there is over 1.2 million<br />

ha <strong>of</strong> remnant classified private native forest in SE<br />

Queensland. Remnant classification was given to<br />

any vegetation considered to meet three criteria,<br />

namely:<br />

1. 50% <strong>of</strong> the original canopy cover<br />

2. 70% <strong>of</strong> the original canopy height and<br />

3. supporting the original species mix<br />

Surrounding and interspersed with the remnant<br />

forests is a similar area <strong>of</strong> non-remnant classified,<br />

regrowth forest. It is this resource that holds the<br />

key to a highly productive native forest resource.<br />

The majority <strong>of</strong> remnant forests have been subject<br />

to 100 years <strong>of</strong> opportunistic harvesting with little<br />

or no follow up management. Stocking rates <strong>of</strong><br />

800-1000 stems/ha over 10cm dbh is not<br />

Map 1. (Above) Aerial Photo showing<br />

remnant overlay and non-remnant<br />

d<br />

Map 2. (Left) Showing Plot location for:<br />

Spotted Gum remnant<br />

Non-remnant<br />

uncommon, and without a radical silvicultural<br />

reset these forests tend to demonstrate a very<br />

low productivity with diameter growth <strong>of</strong> 0.1-<br />

0.3cm per year. PFSQ’s research has<br />

demonstrated that even after a radical thin,<br />

existing stems take up to five years to regain a<br />

reasonable growth rate. Even then is at best 50% <strong>of</strong> that expected from a tree never subjected to<br />

excessive competition or suppression. It is not until the regeneration from the thin comes through that<br />

the productivity will reach acceptable levels.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 398<br />

PLOTS AND DEMONSTRATION SITES<br />

PFSQ has established 11 demonstration sites across SE Queensland over the last ten years (Map 2.),<br />

five in remnant forests and six in non-remnant forests. The plots are generally within the 900-1200mm<br />

rainfall zone dominated by Spotted Gum (7 <strong>of</strong> the 11 sites). Most sites have 3 replicated thinning<br />

regimes (nominally 200, 120, and 80 stems/ha) and a control. (Full site description and procedures<br />

have been written up as case studies for each site and appear at www.pfsq.net)<br />

The demonstration sites were implemented to ascertain the growth rates currently being achieved from<br />

our unmanaged private native forest resource and what management procedures need to be<br />

implemented to improve that productivity. There was little or no information on the dynamics <strong>of</strong> this<br />

significant resource.<br />

When early growth data was analysed in subsequent years it was clear that there had been a very slow<br />

response to the management intervention. It was then that we turned to the regrowth resource to<br />

investigate its productivity potential.<br />

REMNANT FORESTS<br />

Three locations were initially<br />

chosen for the Remnant forest<br />

demonstration sites, one in<br />

Nanango, Iron Pot and Miva. The<br />

pre-thinning stand data for the<br />

three remnant forests demonstrate<br />

an overstocked forest all <strong>of</strong> which<br />

had been subject to a medium<br />

harvest in the previous ten years.<br />

The plots were established in<br />

2000 (Miva NW Gympie), 2002<br />

(Iron Pot NW Kingaroy) and 2006<br />

Nanango. Trees selected for<br />

retention were generally in a<br />

dominant or co-dominant position<br />

with a Grimes Crown score in<br />

excess <strong>of</strong> 20 points. Any trees not<br />

marked for retention with a<br />

Photo 1. Miva plot 7 - thinned to 80 stems/ha<br />

merchantable product were<br />

harvested. Any trees left after the harvest that were unmarked were tree injected with 4:1 Tordon®.<br />

Table 1 represents the original stand stocking rates for each diameter class.<br />

Table 1. Original plot stocking rates for each diameter class<br />

DBH Classes 0 - 10 10 - 20 20 - 30 30 - 40 40 + Total<br />

Av Stems/ha - Iron Pot 149 213 111 14 10 497<br />

Av Stems/ha - Miva 130 91 94 70 - 385<br />

Av Stems/ha – Nanango 320 241 85 31 17 693<br />

Av Stems/ha - Rathdowney 63 118 14 6 8 209<br />

Av Stems/ha – Gin Gin 866 170 10 5 8 1059<br />

Eight years after the thinning operation annual diameter growth increments are still low to very low,<br />

ranging from .15 to .37 cm/yr dbh (see Table 2).<br />

NON-REMNANT FORESTS<br />

After completing a value adding case study in 2001 at a property owned by the Thompson’s at<br />

Gundiah, it became clear the management systems that they had introduced over the last 60 years,<br />

managing regeneration encroaching on to their open paddocks, were achieving growth rates well


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 399<br />

above the traditional State-based systems. It became evident that thinning Spotted Gum when it was<br />

young, and retaining only trees that never came under suppression or excessive competition, would<br />

triple productivity.<br />

Three locations were chosen as regrowth demonstration sites. The first was obviously at Thompson’s<br />

and were in good condition with appropriate stocking levels (130stems/ha). The other two sites Gin<br />

Gin, (200 km NW <strong>of</strong> Gympie) and Rathdowney (130 km south <strong>of</strong> Brisbane) were clearly regrowth.<br />

The Gin Gin stand had a significant regeneration response (866 stems


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 400<br />

Clearly the forest requires thinning from 1100 stems/ha to around 200, this can be achieved by<br />

A. Chemically treating the unwanted stems - at over 900 stems/ha to be injected it would be an<br />

expensive operation ($500-600/ha) and leave a very high volume <strong>of</strong> dead standing timber, a<br />

future work place and fire hazard, and leave little opportunity for Blackbutt to regenerate in the<br />

areas where woodsiana is dominating.<br />

B. Fall to waste, leaving up to 100 tonnes/ha fuel load<br />

C. Harvest the material as Masonite billets to supply the factory outside Ipswich (100km haul).<br />

This was considered the only viable option even though the mill gate price for the billets was<br />

only $42/tonne and it would cost $20/tonne haulage.<br />

To date 3 thinning trials have been undertaken, namely:<br />

1. Minor (1ha) hand falling operation with the product snug out with a 55HP 4wd tractor mounted<br />

with a hydraulic grapple.<br />

2. 10 ha mechanical harvest using a ‘Valmet 941’ (270 hp, 6 wheel drive with 11.5m reach).<br />

3. 7 ha mechanical harvest using a ‘Timbco 415’ (215hp, tracked drive with 6.5m reach).<br />

The trial’s objective was to test the viability <strong>of</strong> harvesting the non-productive/non-commercial portion<br />

<strong>of</strong> the stand to reduce the stand from around 1160 stems/ha, retaining only trees <strong>of</strong> a commercial<br />

species with a good quality bole and healthy crown. In simple terms the majority <strong>of</strong> the stems removed<br />

would be Angophora woodsiana a non-commercial species with only pulp values. The retained<br />

stocking rate varied with the species mix; in areas with a high proportion <strong>of</strong> woodsiana, all stems are<br />

removed creating a gap that is suitable for enrichment planting. In areas dominated by pilularis or<br />

microcorys the retained stocking levels would be around 200 stems/ha, if that quantity <strong>of</strong> straight<br />

healthy stems were available.<br />

The advantage <strong>of</strong> harvesting the material was to remove the bulk <strong>of</strong> the non-productive sector <strong>of</strong> the<br />

stand from the site instead <strong>of</strong> leaving it standing dead if chemically injected. It was considered that this<br />

may be able to be undertaken for the same price or cheaper than chemically treating the stand.<br />

The trial areas were paint marked for retention ahead <strong>of</strong> the harvest to ensure the best quality stems<br />

were retained and for ease <strong>of</strong> operation as the harvest operator did not have the skill set or the visual<br />

perspective from his machine to make a reasonable tree selection decision.<br />

MANUAL HARVEST OPERATION<br />

The hand falling operation was undertaken by two cutters using ‘660 Stihl’ chainsaws cutting on a<br />

face across the hectare. The 4wd tractor then moved through the stand, snigging the logs to a dump at<br />

either end <strong>of</strong> the stand. The cutters cut @ 3.1 tonnes/hr, and the tractor snug @ 6.8 tonnes/hr removing<br />

approximately 55 tonnes <strong>of</strong> product from the<br />

hectare. Net return = $ -280/ha.<br />

‘Valmet 941’ x 40 hr Trial<br />

The machine opened its own track between the<br />

trees and removed unmarked trees up to 11<br />

metres either side <strong>of</strong> the opened track. The<br />

harvester directional felled each tree, cut it to<br />

length and then stacked it for removal by the<br />

forwarder.<br />

The harvester proved very effective, harvesting<br />

stems with little or no damage to the residual<br />

stand. The large rubber tyres also rolled over the<br />

ground cover without exposing soil to possible<br />

erosion. The machine also broke the harvest<br />

residues into small pieces and left them flat on<br />

Photo 3. 55 hp 4wd tractor and grapple<br />

the ground reducing the potential fire risk and<br />

leaving a good layer <strong>of</strong> mulch.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 401<br />

The forwarder also reduced site impacts by lifting all harvest material and transporting it <strong>of</strong>f site by<br />

trailer in place <strong>of</strong> snigging the material along the ground, exposing soil in the process. Average rate <strong>of</strong><br />

production achieved was approximately 11.87 tonnes per hour completing 475 tonnes in the 40 hours.<br />

Net = return $-594/ha.<br />

Photo 4. Valmet 942 wheeled harvester -<br />

$150/hr<br />

Photo 5. Forwarder unloading logs at the dump<br />

$140/hr<br />

‘Timbco 415’ Trial<br />

The ‘Timbco 415’ (considerably older than the Valmet) proved to be very slow and resulted in a<br />

higher level <strong>of</strong> soil impact and damage to the residual stand. Harvesting at only 5.9 tonnes/hr. Net<br />

return = $ -1080/ha<br />

DISCUSSION<br />

The costs <strong>of</strong> the operation are significant and as such must be considered as a cost/return calculation<br />

per hectare. This varies considerably according to the density <strong>of</strong> the timber and its size class, ease <strong>of</strong><br />

access, type <strong>of</strong> harvesting machine etc and as such may vary considerably according to the site. The<br />

results <strong>of</strong> these small trials demonstrate similar costs per hectare to chemical treatment per hectare<br />

($450/ha), however the results are very different. The advantages <strong>of</strong> harvesting can be considered from<br />

two positions, namely:<br />

1. The stand is not left with a significant layer <strong>of</strong> dead trees (fuel load, visual and<br />

WH&S benefits).<br />

2. The stand is immediately opened up to allow for crown development in the retained<br />

trees and a new regeneration layer.<br />

3. The areas with large gaps are available for immediate enrichment planting.<br />

4. The stand is left with a significant mulch layer, immediately reducing soil erosion and<br />

increasing soil carbon.<br />

5. Product is utilised and the carbon locked into long term building products.<br />

6. Further differentiation <strong>of</strong> the product i.e. sorting fencing materials, minor saw logs<br />

should improve the returns in areas where durability one species and saw logs occur.<br />

Disadvantages <strong>of</strong> harvesting include:<br />

1. With variation in site quality a significant drop in returns may make the process<br />

unviable (difficult to ascertain from a small trial).<br />

2. As a result <strong>of</strong> analysing the cost <strong>of</strong> the operation, the owners <strong>of</strong> the machinery indicate<br />

the cost <strong>of</strong> the hourly rate <strong>of</strong> the machine for ongoing work would be considerably<br />

higher. However there are other available machines for a similar cost but performance<br />

will need to be tested.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 402<br />

3. Many <strong>of</strong> the stumps will coppice and there is also a layer <strong>of</strong> Angophora woodsiana<br />

regeneration that will need further treatment or it will again dominate the stand. This<br />

section <strong>of</strong> the trial will trial three chemicals, ‘Glyphosate’, ‘Grazon’, and ‘Garlon’ to<br />

establish which chemical is the most effective at killing the stump coppice as well as<br />

the cheapest. The harvested area will be broken into three units and all stump coppice<br />

that is not required for regeneration, as well as unwanted regeneration (mostly<br />

Angophora woodsiana) will be sprayed using a double reeled ‘Quickspray’ unit with<br />

remote automatic hose retraction capability.<br />

RECOMMENDATION AND WORK PROGRAM<br />

There are significant advantages in the thinning harvest scenario in areas that are viable for this type <strong>of</strong><br />

operation. However, the scope needs to be extended to an integrated harvest that includes removal <strong>of</strong><br />

fencing material, trees in decline where sawlog material is available and the removal <strong>of</strong> trees for pulp<br />

(bent or non-commercial) that are not required for habitat or species distribution.<br />

REFERENCES<br />

Bacon G (2009) ‘Time for Fresh’, Interviewed by Neilson T in In-Wood Magazine. http://www.inwoodmag.com<br />

Burke T (2009) ‘Preparing our Forest Industries for the Future’, Department <strong>of</strong> Agriculture, Fisheries and<br />

Forestry Ministerial Statement. Dated: Wednesday, 24 th June 2009.<br />

De Fegely R and Follas C (2006) ‘The Options for Increasing Harvest Security and Investment in Private<br />

Natural (Native) Forests in Queensland’, Report by Poyry Forest Industry for AgForests Queensland.<br />

Marohasy J (2005) ‘Terrorist or Freedom Fighter?’ The Politics and Environement Blog July 21, 2005,viewed<br />

18 June 2009, http://www.jennifermarohasy.com/blog/archives/000748.html<br />

Stanton R (2005) ‘Tracking the trade deficit is becoming harder’, <strong>Australia</strong>n Forest Grower. 28 (1): p. 10


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 403<br />

IMPROVING BUSINESS PERFORMANCE<br />

IN SOFTWOOD PLANTATIONS<br />

David Williams 1<br />

ABSTRACT<br />

<strong>Australia</strong>’s plantation estate is now approaching 2 million hectares and is composed <strong>of</strong><br />

similar areas <strong>of</strong> s<strong>of</strong>twood and hardwood species. There has been limited expansion <strong>of</strong><br />

the s<strong>of</strong>twood estate over the last 2 decades. As a consequence the current tight s<strong>of</strong>twood<br />

log supply will become critical as demand continues to grow while supply remains<br />

static. Investment in new plantings has been limited because <strong>of</strong> poor returns. Best case<br />

new plantation investment returns that could be expected currently would be around<br />

4%. This paper considers some initiatives to improve financial outcomes from existing<br />

businesses which are a pre-requisite for any investment in new plantings. The initiatives<br />

include improved log prices, win-win opportunities through growerprocessor<br />

partnerships and improved value recovery from implementing optimising technology<br />

into harvesting. These initiatives have potential to add a further 3-4% return on<br />

investment.<br />

INTRODUCTION<br />

The national plantation estate has been expanding at a healthy rate for the last four decades. The<br />

current national plantation estate <strong>of</strong> 1.97 million ha represents a nine fold expansion over the preexpansion<br />

program estate <strong>of</strong> the early 1960’s. The estate is now almost equal proportions <strong>of</strong><br />

s<strong>of</strong>twood and hardwood species. These plantations were established for different purposes with<br />

different end markets. Changing government policies, the reduction in native forest harvesting and<br />

Vision 2020 have created drivers for an expansion <strong>of</strong> hardwood over s<strong>of</strong>twood in the past 15 years.<br />

Traditionally s<strong>of</strong>twood plantations have been managed under longer rotations (~30 years) to<br />

maximise sawlog production for processing into structural timber. The more recently established<br />

hardwood plantations are managed under shorter rotations (10-15 years) primarily for woodchip<br />

production destined for export to Japanese (or North Asian) paper companies.<br />

Despite increasing demand for s<strong>of</strong>twood sawlogs, the s<strong>of</strong>twood plantation estate has remained almost<br />

static for two decades. Net new plantings have averaged less than 1,000 ha per annum since 1992.<br />

Given the time lag for maturing <strong>of</strong> long rotation plantations, the lack <strong>of</strong> new plantation investment is<br />

now becoming apparent. Ironically this gloomy outlook follows the golden decade <strong>of</strong> increased log<br />

supply and processing investment.<br />

This paper focuses on <strong>Australia</strong>’s s<strong>of</strong>twood plantations to show that whilst the aggregate picture <strong>of</strong><br />

the national plantation estate reveals a positive outlook, the future for the s<strong>of</strong>twood plantation sector<br />

is dire. A large expansion in new s<strong>of</strong>twood plantations is desperately required. S<strong>of</strong>twood plantation<br />

owners have various opportunities to improve the commercial outcomes from their existing<br />

businesses a pre-requisite for investing in new plantings. These initiatives are actions available to<br />

s<strong>of</strong>twood plantation owners and have been implemented by HVP.<br />

THE NATIONAL SOFTWOOD PLANTATION ESTATE<br />

There are three recognisable phases covering the national s<strong>of</strong>twood plantation estate, as presented in<br />

Figure 1 below.<br />

Phase 1 – Prolonged early period <strong>of</strong> modest plantings<br />

Plantings commenced in the 1870’s and it took 90 years to establish the first 200,000 ha. The driver<br />

for these plantations was the shortage <strong>of</strong> local sources <strong>of</strong> s<strong>of</strong>twood species.<br />

1 rd<br />

General Manager, Business Development, HVP Plantations, 3 Floor, 517 Flinders Lane, Melbourne, Victoria 3000.<br />

Ph: 03 9289 1400 Ph: 03 9289 1430. Email: dwilliams@hvp.com.au.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 404<br />

Phase 2 – The estate rapidly expands<br />

The national estate increased five-fold to almost 1 million ha over the period from the early 1960’s to<br />

the late 1980’s. The plantings were undertaken by State governments, with Federal government<br />

financial assistance. Government investment was aimed at rapidly expanding the estate consistent<br />

with post World War II government policy aimed at self sufficiency in strategic industries. It should<br />

be noted that many <strong>of</strong> these plantations were established on Crown land which was cleared <strong>of</strong> native<br />

vegetation; an action now banned across <strong>Australia</strong>.<br />

Phase 3 – Plantation expansion hits the wall<br />

Government investment ceased by the late 1980’s as <strong>Australia</strong> moved into recession. New plantings<br />

have almost stalled since government investment ceased. Community pressure to halt the clearing <strong>of</strong><br />

native vegetation resulted in many agencies purchasing and planting cleared agricultural land. A<br />

subsequent backlash and difficult economic climate saw these agencies halt plantation expansion.<br />

The net s<strong>of</strong>twood estate has expanded by less than 1,000 hectares per annum since the one millionth<br />

hectare was planted in 1992, following a period when expansion averaged more than 30,000 hectares<br />

per annum. Planting <strong>of</strong> first rotation areas by Willmott Forests and other s<strong>of</strong>twood MIS companies is<br />

<strong>of</strong>f-setting the area lost due to conversion to other species or land uses.<br />

'000 ha<br />

1,200<br />

1,000<br />

Figure 1 : Average rate <strong>of</strong> s<strong>of</strong>twood plantation establishment in<br />

<strong>Australia</strong><br />

800<br />

600<br />

400<br />

200<br />

0<br />

1870<br />

early modest<br />

plantings<br />

rapid expansion<br />

1965<br />

1992<br />

expansion<br />

hits wall<br />

THE SECTOR’S GOLDEN DECADE<br />

Log production increases as the estate matures<br />

Log production increased over the last decade as the national s<strong>of</strong>twood estate matured and the<br />

plantings from the 1970’s and 1980’s were harvested. These plantations are routinely re-planted. The<br />

current s<strong>of</strong>twood sawlog production <strong>of</strong> 14.4 million m 3 represents a 39% increase over the decade<br />

(BRS, 2009).<br />

Processors invest in expanding production capacity as log supply increases<br />

During the past decade, sawn timber production has increased 85% to 3.9 million m 3 (BRS, 2009).<br />

Progressively, the sawn timber processing sector invested in larger, world scale capacity mills using<br />

improved technology to process the high quality logs at lower unit costs.<br />

The other processing sectors <strong>of</strong> panels and paper products have also expanded production over the<br />

last decade by 27% and 32% respectively on the back <strong>of</strong> increased log availability.<br />

2008


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 405<br />

Disappointing prices during decade <strong>of</strong> expansion<br />

Despite increases in supply and efficiencies in processing and subsequent import replacement, timber<br />

prices have declined substantially in real terms over this period <strong>of</strong> unprecedented expansion in log<br />

supply and processing capacity. Sawlog and pulp log prices have declined in real terms by 22% and<br />

9% respectively (KPMG, 2009). Figure 2 below illustrates the decline in sawlog prices in real terms<br />

over the period from 2000-2008. Sawlog prices shown are the weighted average <strong>of</strong> the four log<br />

classes presented in the <strong>Australia</strong>n Pine Log Price Index (APLPI)(KPMG, 2009).<br />

Sawlog price $/m cubic metre<br />

$51.00<br />

$49.00<br />

$47.00<br />

$45.00<br />

$43.00<br />

$41.00<br />

$39.00<br />

$37.00<br />

$35.00<br />

Figure 2 : Average s<strong>of</strong>twood sawlog prices 2000-2008<br />

2000 2001 2002 2003 2004 2005 2006 2007 2008<br />

APLPI price - flat real $ APLPI price - actual real $<br />

STAGNANT LOG SUPPLY FOR DECADES AHEAD<br />

No net increase in plantation areas over the last 2 decades limit opportunities<br />

S<strong>of</strong>twood sawlog supply is currently fully committed, predominantly under long term contracts to a<br />

sector with processing capacity which exceeds the log supply. Log supply will remain static for the<br />

next 30 years even if major new plantings commenced now. While demand exceeds supply, the<br />

domestic demand results in a national trade imbalance in timber and paper products <strong>of</strong> $2 billion per<br />

annum. Advances in productivity will require increased supply which is not available in the short<br />

term, or rationalisation through mergers to maintain world scale efficiencies.<br />

For more than 50 years, the demand for timber products has been correlated with population growth.<br />

On average, <strong>Australia</strong>ns consume a little over one cubic metre <strong>of</strong> log equivalent volume <strong>of</strong> timber<br />

products per person per year (BRS 2009). Domestic consumption <strong>of</strong> timber products will therefore<br />

continue to grow over future decades in line with population growth. Future population growth is<br />

influenced by fertility and morbidity rates as well as government immigration and family policies. It<br />

is observed that population growth increased by an average <strong>of</strong> 1.5% per annum for the 5 years until<br />

June 2008 (www.census.abs.gov.au) This is a comparatively high growth rate historically. However,<br />

one can expect future population growth to remain positive at a rate <strong>of</strong> around 1% per annum.<br />

Assuming a 1% future growth rate, timber demand will increase by more than 200,000 cubic metres<br />

<strong>of</strong> log equivalent per year. By 2019 the supply-demand gap for s<strong>of</strong>twood log will have increased by a<br />

further 8 million cubic metres per annum. This will cause further deterioration in the trade imbalance<br />

in timber products.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 406<br />

Investment to maintain competitive mills will be challenging<br />

Processors have invested in new mills by securing increased log volumes from growers over past<br />

decades. This has resulted in more efficient mills which were world scale when initially constructed.<br />

Maintaining competitiveness requires ongoing investment in new milling technology and this will be<br />

more challenging for the processing sector in the future. Because <strong>of</strong> supply constraints, further<br />

upgrades will be difficult without rationalisation through amalgamations or merging <strong>of</strong> operations.<br />

This is a more expensive option than the past approach <strong>of</strong> securing increased volume directly from<br />

growers. Upgrades do not increase overall log supply or production capacity.<br />

Business expansion is limited for growers<br />

Competition and increasing prices for land limits the ability for the s<strong>of</strong>twood plantation businesses to<br />

expand their existing estates into the future. Supply expansion will depend on making the most <strong>of</strong> the<br />

current estate through improving operational efficiencies, genetics, nutrition and other plantation<br />

management activities. Government-supported past research provided a basis for substantial<br />

productivity gains in s<strong>of</strong>twood plantations. Cessation <strong>of</strong> this source <strong>of</strong> support will require other<br />

parties, including individual growers, to increases their investment in plantation research if ongoing<br />

productivity gains are to continue.<br />

There will also be opportunities for s<strong>of</strong>twood plantation owners to benefit through new<br />

environmental values, including carbon sequestration, bio-fuels, and other environmental values.<br />

THE CHALLENGE OF ATTRACTING NEW INVESTMENT<br />

Past players and future investors<br />

Governments and public companies largely financed the existing s<strong>of</strong>twood plantation estate. These<br />

traditional bodies are unlikely to be major investors in further new plantings.<br />

There has been a changing <strong>of</strong> the guard in the s<strong>of</strong>twood plantation sector. Timberland Investment<br />

Management Organisations (TIMO’s) and Managed Investment Schemes (MIS) companies are more<br />

recent entrants into the <strong>Australia</strong>n forestry sector and may continue as investors if comparable<br />

mainstream returns can be achieved from investing in new s<strong>of</strong>twood plantations. MIS companies are<br />

finding it more difficult to justify and secure new land for planting, opting in some cases to rent<br />

second rotation land from established plantation growers.<br />

The nature, structure and motivations <strong>of</strong> <strong>Australia</strong>n plantation investors varies considerably. There<br />

are a number <strong>of</strong> different investor categories that have contributed to the national plantation sector.<br />

In 1950-51 more than 90% <strong>of</strong> <strong>Australia</strong>’s total plantation estate was publicly owned. Public<br />

ownership has declined substantially since and new players, including TIMOs and MIS companies,<br />

now own almost half <strong>of</strong> the plantation estate (BRS, 2009) as shown in Figure 3.<br />

Figure 3 : <strong>Australia</strong>n plantation ownership 2008<br />

Other private<br />

growers<br />

Timber companies<br />

Super funds<br />

MIS<br />

Governments


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 407<br />

Governments – unlikely future investors<br />

Governments were the major drivers behind the establishment <strong>of</strong> the existing national estate but these<br />

sources <strong>of</strong> investment fell away sharply 2 decades ago. The previous investment motivation relied on<br />

government industry policies to create regional employment and infrastructure and to reduce<br />

dependence on imported wood and paper products. Once established, governments withdrew from<br />

their activity to be replaced by private enterprise and so the former policies are no longer relevant.<br />

Further, government ownership <strong>of</strong> plantation businesses has been progressively reduced through<br />

either the creation <strong>of</strong> commercial Crown corporations or the complete privatisation <strong>of</strong> plantations.<br />

Having withdrawn from the sector, Governments are unlikely to again be direct investors in new<br />

plantings. Those plantations operated by Crown corporations may well invest in new plantings but<br />

only based on business models which show reasonable returns on investment using commercial<br />

investment criteria similar to private investors.<br />

Public companies – unlikely future investors<br />

Vertically integrated plantation and processing companies made significant past contributions to the<br />

existing national estate. The trend over recent decades has been for integrated companies to divest<br />

their plantations and focus on their processing businesses. This move was precipitated by perceived<br />

poor returns from their timberlands businesses based on poor commercial models and internal pricing<br />

between enterprises. The original motivation for their plantation investment was resource security<br />

for their downstream processing businesses. Revision <strong>of</strong> business models has suggested that it is<br />

more cost efficient for corporations to concentrate on their processing activities, sell the plantations<br />

and retain long term supply contracts for the future volume from these plantations. Thus the existing<br />

public companies involved in processing are unlikely to be major direct investors in new plantings in<br />

future.<br />

Private companies – likely future investors<br />

Private companies are those with a limited number <strong>of</strong> shareholders and are unlisted on the <strong>Australia</strong>n<br />

Stock Exchange. These are <strong>of</strong>ten TIMOs which are active investors in private timberlands. They<br />

have been formed on the back <strong>of</strong> opportunities provided by public integrated companies and<br />

governments divesting their timberlands. TIMOs bring together interested investors <strong>of</strong>ten with<br />

considerable access to funds. They apply streamlined and efficient management with long term<br />

philosophy to what were poorly performing businesses.<br />

The first TIMO to enter the <strong>Australia</strong>n plantation sector was Hancock Natural Resources Group<br />

when it acquired the previously owned Victorian government plantation business, Victorian<br />

Plantation Corporation. The parent created Hancock Victorian Plantations (now renamed HVP<br />

Plantations) in 1998 to manage the Victorian plantations. Since this time a number <strong>of</strong> other TIMOs<br />

have become involved in plantation ownership and or management.<br />

MIS is a relatively new investment model for establishing new plantations. There were 10 major MIS<br />

companies <strong>of</strong>fering a variety <strong>of</strong> investment options for individuals seeking investment in plantations<br />

and other agricultural businesses. The sector has expanded at a rapid rate since the mid 1990’s. The<br />

overwhelming share <strong>of</strong> total investment has been in short rotation hardwood plantations and this has<br />

been the driver in the rapid expansion <strong>of</strong> the nation’s hardwood plantation estate – expansion <strong>of</strong> more<br />

than 800,000 ha since 1996 (BRS, 2009). The investor attraction to these hardwood plantations has<br />

been the relatively short investment period.<br />

There has also been some smaller scale investment by local and foreign investors in short rotation<br />

hardwood plantations, <strong>of</strong>ten by entities directly or indirectly associated with north Asian processors<br />

or trading houses.<br />

One MIS company, Willmott Forests, is an exception to this general picture. Willmott Forests invests<br />

in s<strong>of</strong>twood plantations on behalf <strong>of</strong> individual investors. Over the last 20 years Willmott has<br />

established an estate <strong>of</strong> more than 40,000 ha <strong>of</strong> s<strong>of</strong>twood plantations through MIS.<br />

The current outlook for MIS investment is unclear with the two largest MIS companies going into<br />

receivership. Time will reveal whether the apparent shortcomings <strong>of</strong> these companies will limit<br />

future investment via MIS companies. With longer term investment horizons it should possible for<br />

MIS to be a significant driver for new s<strong>of</strong>twood plantations if the shortcomings <strong>of</strong> the current failing


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 408<br />

companies have not soured an investment model for new plantings based on reasonable returns and<br />

established processing.<br />

Both TIMOs and MIS companies are potential investors in new s<strong>of</strong>twood plantations provided<br />

realistic investment returns can be achieved.<br />

WHAT’S NEEDED FOR PROFITABLE INVESTMENT<br />

Mainstream investment in new plantations will require normal returns on investment over the life <strong>of</strong><br />

the business. Expected returns on investment vary with the nature and risk <strong>of</strong> the business. A<br />

definitive acceptable rate <strong>of</strong> return on new plantation investment is debatable. However, let us<br />

assume, for the purposes <strong>of</strong> this discussion, double-digit annual return <strong>of</strong> at least 10% is a reasonable<br />

expectation for a mainstream investment in new plantations. The general term investment return or<br />

return on investment as used in this paper refers to economic internal rate <strong>of</strong> return (IRR).<br />

It is estimated that the investment returns for many existing s<strong>of</strong>twood plantation businesses range<br />

from 2-10% per annum (Industry Edge, 2008). Businesses in the upper quartile <strong>of</strong> returns, which<br />

includes HVP, are likely to be providing satisfactory returns to their investors. Returns for businesses<br />

in the lower quartile would not satisfy normal investors and are therefore probably disappointing<br />

their current owners.<br />

Why wouldn’t existing successful plantation businesses invest in new plantings?<br />

Returns from existing plantations are substantially greater than new plantations. The difference is the<br />

cash flow pr<strong>of</strong>iles. Having the ability to generate cash from older plantations is important to<br />

investment returns. Existing plantations usually have a range <strong>of</strong> age classes that provides an<br />

immediate and ongoing cash flow whereas cash flow from new plantation investment is delayed for<br />

many years.<br />

In addition, the valuation approaches used to assess investment returns for most existing plantation<br />

businesses either do not consider land value or apply conservative accounting treatments to land.<br />

Thus the cost <strong>of</strong> purchasing or leasing land for a full rotation or in perpetuity is the defining<br />

difference between existing plantations businesses and new plantings. Carrying the land cost at<br />

prevailing discount rate for around 30 years until the plantation is finally harvested is a burden for<br />

new plantation investment. Much <strong>of</strong> the existing national estate was established by governments<br />

clearing native forests on Crown land, so there were no land costs for these plantations.<br />

Analysis based on costs and future cash flows incorporating land value at a prevailing discount rate is<br />

an appropriate model for considering new plantings. A hypothetical investment scenario illustrates<br />

the substantially superior investment outcomes for existing plantations compared with new plantings.<br />

Consider a new plantation investment assuming medium growth rate (assume m.a.i. = 22 m 3 /ha/a),<br />

prevailing log prices (assume APLPI price – KPMG, 2009), average plantation and log production<br />

costs, and 2 thinnings with final harvest at 28 years. If land is available at no cost, an IRR <strong>of</strong> 8-9%<br />

would be a reasonable expectation. If land purchase cost $5,000/ha, or land lease at commercial rent<br />

<strong>of</strong> 5% per annum, then the expected IRR might be in the order <strong>of</strong> 4%.<br />

There is a myriad <strong>of</strong> variables that affect costs, growth rates, plantation regimes and sales returns,<br />

therefore determining the hypothetical optimal new planting option is problematic. An IRR <strong>of</strong> close<br />

to 4% is likely to represent the best outcome achievable for investing in new plantings assuming best<br />

case plantation variables and including full land cost. Such a best case investment option falls short<br />

<strong>of</strong> minimum expected market returns by around 6% per annum.<br />

INITIATIVES TO IMPROVE BUSINESS PERFORMNACE<br />

This section considers three initiatives that have the potential to improve returns from existing<br />

plantation businesses and therefore provides opportunities to contribute to bridging the gap for new<br />

plantation investor returns.<br />

Initiative 1. Log prices<br />

Log prices have been declining in real terms over the period when log supply has been progressively<br />

tightening until the current time where log supply is in deficit.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 409<br />

Normally when supply exceeds demand for any commodity in competitive markets, increasing prices<br />

reflecting the normal principles <strong>of</strong> supply and demand can be expected. This has not been observed<br />

in the <strong>Australia</strong>n s<strong>of</strong>twood log markets for a number <strong>of</strong> reasons including the fact that log sales are<br />

generally not exposed to competitive market pressures.<br />

A competitive market requires a number <strong>of</strong> willing sellers and buyers. <strong>Australia</strong>n s<strong>of</strong>twood log<br />

markets do not reflect normal competitive markets because in most cases there is only one or two<br />

major log sellers and buyers in each regional market. In some cases there is only one seller and one<br />

buyer in a market. Increasing log supply by importing logs from New Zealand, North America or<br />

Chile is not viable and has not occurred. Thus the lack <strong>of</strong> normal competitive markets excludes<br />

supply and demand pricing in the large majority <strong>of</strong> sales.<br />

HVP has experienced increased prices in all cases when it has had the opportunity to <strong>of</strong>fer logs<br />

through a competitive sale process, illustrating the potential <strong>of</strong> competitive markets. Anecdotal<br />

information indicates other plantation owners also achieve higher log prices whenever they are able<br />

to <strong>of</strong>fer logs for sale through competitive market processes.<br />

With expanding demand and tightening log supply, processors seek greater supply security which has<br />

additional value to the processor. Once the aggregate log supply is fully committed, long term supply<br />

contracts reduce opportunities for changes in the log volume shares between processors. Thus market<br />

share <strong>of</strong> processed timber is also locked in by log supply contracts. A processor cannot challenge<br />

another competitor's market without in fact withdrawing from some part <strong>of</strong> its own existing market.<br />

Fixed market shares provide a unique pricing environment for processors. There is less pressure for<br />

price discounting when market shares cannot be challenged. Thus long term supply contracts present<br />

unique advantages to processors providing opportunities for real price increases for timber.<br />

Real price increases for domestic sawn timber establishes the circumstance for real log price<br />

increases. The challenge is for processors to maintain their pr<strong>of</strong>itability through timber price<br />

increases and reducing processing costs during the future period <strong>of</strong> market-driven increasing log<br />

prices.<br />

Whilst the increasing deficit <strong>of</strong> domestic timber supports price increases, domestic timber also faces<br />

import competition which can constrain domestic prices. Other materials used in house building<br />

have achieved better price outcomes than domestic timber, notwithstanding tough competition within<br />

their own sectors.<br />

annual % increase<br />

8.0%<br />

7.0%<br />

6.0%<br />

5.0%<br />

4.0%<br />

3.0%<br />

2.0%<br />

1.0%<br />

0.0%<br />

Figure 4 : Price changes for some building materials<br />

used in house building<br />

All groups Timber, board,<br />

joinery<br />

Concrete, cement,<br />

& sand<br />

Steel products<br />

av increase - last 5 years av increase - last 40 years


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 410<br />

Figure 4 illustrates that the average annual price for concrete, cement and sand, and for steel products<br />

increased by 6.7% and 6.8% respectively over the last 5 years. Over the same period, timber, board<br />

and joinery managed an average annual price increase <strong>of</strong> only 1.8%. However, the trend over the last<br />

40 years has been comparatively flat around 6% per annum 2 . It would seem, therefore, that, during a<br />

period <strong>of</strong> major increase in log availability and expansion in timber production, timber prices have<br />

actually declined when compared with other house building materials.<br />

Scarcity provides opportunity for substantial real price increases for both logs and timber. A real<br />

price increase in log prices <strong>of</strong> 10-15% is realistic for the medium term, and this could add 1% to the<br />

IRR for new plantation investment.<br />

This thus constitutes another opportunity for growers seeking pr<strong>of</strong>itable new plantation investment.<br />

Initiative 2. Grower Processor partnerships<br />

Partnerships are defined here as grower and processor working together where business aspects<br />

overlap or interface. The benefits <strong>of</strong> the partnership are allocated between the parties as pr<strong>of</strong>it shares.<br />

Traditional growerprocessor relationships involve the grower supplying logs to the processor with<br />

very little business cross-over. Growers maintain their focus on growing, harvesting and delivering<br />

logs, while processors receive the logs and process them into timber products. The businesses remain<br />

separate with few additional benefits beyond those formal arrangements detailed at the time supply<br />

agreements are executed.<br />

This relationship can inhibit opportunities for mutual benefits that could flow from cooperative<br />

actions aimed at improved overall commercial outcomes. The past has seen opportunities go begging<br />

because <strong>of</strong> the inability to agree on how the costs and benefits <strong>of</strong> enhanced net outcomes can to be<br />

shared.<br />

For example, a change to an aspect <strong>of</strong> supply or log specifications could improve the value <strong>of</strong> the<br />

parcel <strong>of</strong> logs but this may incur some additional cost to the grower. The inability to agree on how<br />

the additional cost is to be <strong>of</strong>fset and how the net upside is to be shared may prevent the beneficial<br />

changes being adopted. Such situations represent wasted opportunities.<br />

The most prospective areas for partnerships will be in business activities that are more likely to be<br />

successful if carried out jointly. For example, where inputs are complementary, such as expertise,<br />

funds, materials, other resources and markets. The possibilities could include value adding<br />

opportunities where action is required by both parties to achieve the improved outcome. One<br />

example is the production and sale <strong>of</strong> chips. This is the basis <strong>of</strong> the successful partnership <strong>of</strong><br />

S<strong>of</strong>twood Plantation Exporters (SPE) (discussed further below). Other potential areas for<br />

collaboration or for joint involvement include:<br />

• Bio-fuel production including co-generation with fibre coming from the mill and the<br />

plantations.<br />

• Changes to enhance the value <strong>of</strong> a log parcel beyond expectations in the original supply<br />

agreement. Changing the parcel to add value for the processor could incur cost to the grower<br />

but the net result could still represent increased value overall.<br />

• Additional log volume provided preferentially to the processor.<br />

• Enhancing the value <strong>of</strong> supply contracts by amending conditions such as extending term. This<br />

is valuable to the processor as it locks in market share by excluding access by competitors to<br />

a fully committed national log market. There is a particular opportunity to realise this value at<br />

the time a processing business is sold. Amending supply contracts prior to sale is an example<br />

<strong>of</strong> where grower and processor could act together to add and share value which is<br />

immediately realisable in monetary terms.<br />

The formal arrangements for such collaboration can come in many forms. There is a full range <strong>of</strong><br />

options covering the format <strong>of</strong> growerprocessor partnerships, including formal partnerships, joint<br />

ventures (JV) and marketing arrangements. These partnerships do not require one party intruding into<br />

2 http://www.abs.gov.au/AUSSTATS/abs@nsf/Lookup/6427.0Main+Features1Junpercent202004?Open


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 411<br />

the core business <strong>of</strong> the other party. The assumption is that parties are expert and successful at their<br />

core businesses and will continue as separate businesses.<br />

Implementing successful partnerships is challenging as they need to exhibit some or all <strong>of</strong> the<br />

attributes <strong>of</strong> a high level <strong>of</strong> confidence and trust between the parties, a strong base <strong>of</strong> common<br />

objectives, complementary inputs, and shared business style and culture.<br />

S<strong>of</strong>twood Plantation Exporters (SPE) – an example <strong>of</strong> a successful partnership<br />

HVP and AKD S<strong>of</strong>twoods, one <strong>of</strong> HVP’s major sawlog customers, formed a JV partnership more<br />

than 10 years ago to export s<strong>of</strong>twood chips to Japanese paper customers.<br />

At the time, HVP’s plantation value was constrained by limited market outlets for pulp logs from<br />

thinning operations. This prevented full implementation <strong>of</strong> optimal thinning regimes. AKD was also<br />

adversely affected by limited volume outlet for its residual chips. The parties came together through<br />

necessity after each concluded there were limited opportunities for their stand alone pulp log and<br />

chip volumes. The larger total chip volume allowed consideration <strong>of</strong> new domestic processing or<br />

export options. Exporting chips to the Japanese paper market was the preferred option. The parties<br />

formed a JV and the export project was developed as a greenfield business with the parties sharing<br />

equally all aspects <strong>of</strong> the JV and selling their chips under joint arrangements.<br />

The business has been successful, providing improved financial returns for both parties as well as<br />

building a platform for expansion <strong>of</strong> the business under the SPE banner and providing an opportunity<br />

to approach aspects <strong>of</strong> the traditional log supply as a partnership. AKD and HVP both receive good<br />

prices for sawmill chips and plantation pulp logs as well as an export margin. The export chip<br />

business is a significant addition to the respective partners with an aggregate turn over <strong>of</strong> almost<br />

$300 million since 1997.<br />

Thus such partnerships are cost effective and provide potential to improve the financial performance<br />

for both parties. A plantation owner could reasonably aspire to increased returns sufficient to<br />

improve IRR by 1% or more.<br />

Initiative 3. Value Recovery from harvesting<br />

Value recovery (VR) refers to maximising the value <strong>of</strong> trees at harvest. A tree stem can be cut into<br />

various combinations <strong>of</strong> log quality (veneer, sawlog, pulp log), diameter and length. Maximising the<br />

aggregate value <strong>of</strong> the mix <strong>of</strong> log products provides improved financial returns from harvesting<br />

operations.<br />

Growers’ primary benefit derives from value recovery. Conventional mechanised harvesters have<br />

limited capacity to select the optimal mix <strong>of</strong> log products and therefore cannot maximise recovered<br />

value from harvesting operations. Optimisation technology is one method for determining the best<br />

log combinations.<br />

Invariably processors have a log product range they are seeking. Using machine-mounted computer<br />

systems and competent operators, the grower and the processor can select various log product<br />

combinations to maximise value. An effective system incorporates reconciliation processes to<br />

measure VR performance against pre-harvest inventory and forest model benchmarks, and timely<br />

auditing <strong>of</strong> operations to provide positive feedback to harvesting operators. A VR program needs to<br />

be supported by comprehensive procedures.<br />

An optimisation program with these attributes is being implemented by HVP across its estate.<br />

Procedures have been developed, supervisors and operators trained and optimising technology<br />

progressively fitted to harvesters. Two thirds <strong>of</strong> harvesters in HVP plantations are now fitted with<br />

optimising technology and most supervisors and machine operators have been trained in VR<br />

procedures. The majority <strong>of</strong> HVP’s logs are now produced by optimising harvesters with very<br />

positive results. Recovered value is consistently better than non-optimised operations and value<br />

recovery exceeds forest modelling forecasts. Three per cent overall value improvement from<br />

harvesting is a conservative achievable target if an effective VR program is in place. There will also<br />

be opportunities for further improvement as effective VR programs are developed beyond the initial<br />

phase.<br />

A 3% value recovery improvement could correspond to a 0.5% increase in IRR.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 412<br />

CONCLUSION<br />

<strong>Australia</strong> has a substantial trade imbalance in timber products including imported products processed<br />

from s<strong>of</strong>twood logs. Currently s<strong>of</strong>twood log demand exceeds a supply which will remain static for<br />

several decades as a result <strong>of</strong> lack <strong>of</strong> investment in new plantations. Historically timber demand has<br />

grown in line with population growth, so the current gap between supply and demand will further<br />

widen.<br />

New plantation investment is urgently required to provide opportunities for <strong>Australia</strong>n s<strong>of</strong>twood<br />

plantation growers and processors to redress the increasing national timber trade imbalance. The lack<br />

<strong>of</strong> investment is attributable to perceived low returns from new plantation investment and the<br />

diversion <strong>of</strong> potential funds to short rotation hardwood plantations.<br />

Improving returns from existing plantation businesses is a first step in creating an environment that<br />

may be attractive to private investors. There are a number <strong>of</strong> potential initiatives that existing<br />

plantation businesses could adopt. Some <strong>of</strong> these include increased log prices resulting from the<br />

increasing scarcity <strong>of</strong> logs and domestic timber, shared improved outcomes from win-win<br />

partnerships between growers and processors, and improved value recovery from harvesting by<br />

utilising optimising technology. These initiatives have been successfully incorporated into HVP’s<br />

business and all have added to overall business performance.<br />

These initiatives, and others, have the potential to move the currently low returns from investment in<br />

new plantation areas closer to the IRR expected by mainstream investors. Achieving these outcomes<br />

could make new plantation investment attractive and be a catalyst for a reinvigorated expansion <strong>of</strong><br />

s<strong>of</strong>twood plantations.<br />

REFERENCES<br />

BRS, Department <strong>of</strong> Agriculture Fisheries and Forestry, <strong>Australia</strong>n Government (2009). <strong>Australia</strong>’s forests at a<br />

glance 2009 Data to 2007-08.<br />

KPMG (2009). <strong>Australia</strong>n Pine Log Price Index Updated to December 2008.<br />

Industry Edge (2008). A Critical Analysis <strong>of</strong> the S<strong>of</strong>twood Industry in Eastern <strong>Australia</strong>.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 413<br />

ADDRESSING CLIMATE CHANGE IMPACTS ON THE BUSINESS OF<br />

PNG <strong>FORESTRY</strong>: IMPLICATIONS FOR AUSTRALIA AND PNG<br />

Dick McCarthy 1 and Gabriel Samol 2<br />

ABSTRACT<br />

This paper addresses the climate change impacts and adaptation challenges for forestry<br />

on the forests <strong>of</strong> PNG (Papua New Guinea) which are botanically related to the isolated<br />

pockets <strong>of</strong> surviving Gondwanan flora in <strong>Australia</strong> such as the wet tropical rainforests<br />

<strong>of</strong> northeastern Queensland.<br />

PNG, through large scale plantation development, has the potential and the capacity<br />

(land, soils, water, air and adaptable plantation species) to capture and store carbon in<br />

its biomass, soils and forest products in perpetuity. If, through the proper introduction <strong>of</strong><br />

an emissions trading scheme, increased value and returns to the forest owners and<br />

developers can be realised, PNG has the capacity to increase:<br />

• Five fold its commercial timber tree plantation estate to some 300,000 hectares.<br />

• 20 % increase in forest industry employment and flow on effects to other<br />

employment sectors<br />

• an increase in value <strong>of</strong> forest products production <strong>of</strong> more than 100 %<br />

INTRODUCTION<br />

The origins <strong>of</strong> <strong>Australia</strong>’s forests can be traced to the beginning <strong>of</strong> the Cretaceous period, when the<br />

super continent Gondwana began to fragment into Africa, South America, India, <strong>Australia</strong>/Antarctica<br />

and many smaller islands. Although this happened at least 135 million years ago, similarities can still<br />

be found in the flora and fauna <strong>of</strong> these now widely separated lands.<br />

Over the intervening millennia, <strong>Australia</strong> has evolved a new biota. About 38 million years ago,<br />

<strong>Australia</strong> broke away from Antarctica and shifted northwards, colliding with Asia about 13 million<br />

years ago. During <strong>Australia</strong>’s northward journey, the climate became progressively warmer and drier,<br />

and the vegetation adapted accordingly. Cool and warm rainforests were replaced by sclerophyllous<br />

genera such as Eucalyptus and Acacia.<br />

PNG is a land subject to continual catastrophe. The mountains are young and continuing to uplift as<br />

the <strong>Australia</strong>n plate subducts below the Pacific plate so earthquakes with associated landslips are<br />

frequent on the young steep slopes. There are numerous active volcanoes which create lava flows and<br />

mudflows and thick ash deposits. Strong destructive winds occasionally occur. In exceptionally dry<br />

years these forests, always slightly seasonal, become unusually dry and may catch fire. The big rivers<br />

which run on the coastal plains have unstable courses. Shifting cultivation and associated regrowth<br />

forest is also extensive.<br />

In reviewing PNG forest dynamics, it is no surprise that, in a listing <strong>of</strong> timber tree species for a tract <strong>of</strong><br />

lowland rainforest, there is usually a considerable proportion <strong>of</strong> pioneers, such as species <strong>of</strong> Albizia;<br />

Paraserianthes and Serianthes, and Eucalyptus deglupta (New Britain), and strongly light-demanding<br />

climax species, for example, Campnosperma; Pometia and Terminalia.<br />

As PNG has been <strong>Australia</strong>’s wood shed for over 100 years and politically continues to be a hardwood<br />

producing nation, this paper will cover, especially from a PNG perspective:<br />

• Climate change scenarios<br />

1 McCarthy & Associates (Forestry) Pty Ltd, formerly Executive Director, PNG Forest Industries Association<br />

2 Executive Officer PNG Forest Industries Association, Inaugural President Association <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> Papua New Guinea


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 414<br />

• Effects on PNG communities and on the full range <strong>of</strong> forest values<br />

• Management <strong>of</strong> risks and identification <strong>of</strong> opportunities<br />

The implications for <strong>Australia</strong> and PNG <strong>of</strong> political, economic and security development in terms <strong>of</strong><br />

PNG forestry is that there is a long standing and enormous wood trade between <strong>Australia</strong> and PNG<br />

encompassed within a sphere <strong>of</strong> similar forest types and forest products bound by many existing long<br />

term areas <strong>of</strong> mutual co-operation within the respective government and private sectors. Climate<br />

change scenarios should be no different<br />

PNG’s Vegetation<br />

PNG’s constant high temperature is characteristic <strong>of</strong> the tropical climate and hence creates tropical<br />

moist rainforest vegetation over the whole island.<br />

This tropical moist rainforest vegetation <strong>of</strong> PNG has considerable variation dependent on varying<br />

responses to rainfall, soil water; soils; elevation, and soil fertility which give rise to the following<br />

distinct forest formations:<br />

• Rainforests (where every month is wet (100 mm+ <strong>of</strong> rainfall per month)) and graded by<br />

elevation into: Lowland rainforest; Lower montane forest; Montane forest; Subalpine forest;<br />

Heath<br />

• Savannah forests (where there are several dry months <strong>of</strong> 60 mm or less <strong>of</strong> rainfall)<br />

• Soil water creating forest formations as mangrove forests; beach vegetation; peat swamp<br />

forests and freshwater periodic swamp forests.<br />

Nature <strong>of</strong> PNG Wood Business - Current Situation<br />

Currently PNG’s annual wood flow approaches 8 million m 3 (5 million m 3 firewood (FAO estimate);<br />

1.8 million m 3 log export; processed export 0.65 million m 3 , domestic industrial 0.2 million m 3 )<br />

• Forestry and forest products are PNG’s second largest industry. After 60 years, in terms <strong>of</strong><br />

revenue generation and forest management, only 3.7 million hectares are under active forest<br />

management regimes and there are only 65,000 hectares <strong>of</strong> plantations<br />

• Even so, if allowed to operate efficiently, this small forest base is capable <strong>of</strong> producing a<br />

sustainable cut <strong>of</strong> 3.9 million m 3 with a contribution <strong>of</strong> US$270 million to PNG’s GDP<br />

annually; with US$85 million in export taxes/levies, while landowners receive some US$20<br />

million in direct payments<br />

• The total value <strong>of</strong> manufactured forest products in 2002 was US$33 million, with log exports<br />

totalling US$100 million.<br />

• The forestry industry has various sectors: harvesting, sawn timber, plywood manufacture,<br />

veneer production, furniture making and forest plantation activities.<br />

• The wood business trade is based on price and individual consumer demand<br />

PNG Market Changes for Solid Wood Products – Export Markets<br />

• In recent years, PNG has seen a rapidly declining market for its forest produce in Japan, and<br />

now its China market is diminishing.<br />

• Yet China is becoming the top exporter <strong>of</strong> secondary wood processed products in the<br />

developing world and is expected soon to overtake Germany as the third largest exporter<br />

globally.<br />

• The PNG solid wood export products market is being taken over by panel products, wood<br />

composites and now even wood/plastic composites.<br />

• The <strong>Australia</strong>n sawn timber and panel product is now becoming PNG’s fastest growing<br />

market for PNG sawn timber and PNG plywood.<br />

• The <strong>Australia</strong>n market is the closest export market for PNG<br />

PNG Market Changes for Solid Wood – Domestic Markets<br />

• PNG has lost much <strong>of</strong> the domestic market for timber products to other non-wood materials,<br />

due mainly to failure <strong>of</strong> wood in use because <strong>of</strong> the lack <strong>of</strong> wood preservative treatments.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 415<br />

Property owners and developers, together with architects, engineers and builders have sought<br />

other materials because <strong>of</strong> the inherent dangers in using untreated wood.<br />

• This loss <strong>of</strong> domestic market share correlates directly with the cessation <strong>of</strong> proper monitoring<br />

with compliance checks <strong>of</strong> correct timber treatments and verification by the responsible<br />

government authorities; i.e. government agencies are not doing the job for which they are<br />

mandated..<br />

Future Development <strong>of</strong> the PNG Solid Wood Market<br />

• The existing annual sustainable wood harvest is 3.9 million m 3 per annum. Current production<br />

is averaging 2 million m 3 . Of this harvest approximately 30% is processed.<br />

• The 10 proposed new projects cleared for further development by the Moratorium/review <strong>of</strong><br />

the former government would add 1.3 million m 3 to the annual cut. This is well within the<br />

annual sustainable (allowable) cut.<br />

• A phased introduction <strong>of</strong> those 10 projects would provide a net gain in direct export tax <strong>of</strong><br />

K80 million per year, and an increase in cash and benefit payments to landowners <strong>of</strong> K24.5<br />

million per year.<br />

CLIMATE CHANGE SCENARIOS<br />

The Politics <strong>of</strong> “Ecology” or Climate Change<br />

In the 1960’s, “ecology” was the buzz word. Mass media bestsellers assured us that by the 1980’s we<br />

would all be starving or suffocated by our own wastes. What went wrong? Or what went right?<br />

Did the world’s problems go away in the intervening period? Of course not — but the difference is<br />

that we know a lot more now than we did in the 1960’s.The ozone layer and the oceans appear to be<br />

rather more resistant to pollution than we once thought. We are more conscious <strong>of</strong> the crises being<br />

caused by acid rain; deforestation; desertification and carbon dioxide build-up and population buildup.<br />

PNG is no different to the rest <strong>of</strong> the world with regard to the politics <strong>of</strong> population; the politics <strong>of</strong><br />

food; or the politics <strong>of</strong> forests and the politics <strong>of</strong> forest protesters. The politics <strong>of</strong> climate change is<br />

treated no differently.<br />

The politics <strong>of</strong> climate change. – “The developed world wants a carbon trading scheme where PNG’s<br />

clean environment, climate cleaning mechanisms and carbon storage facilities are utilised without<br />

having to pay for them”. Anonymous<br />

Effects <strong>of</strong> Change on Rainforest Communities<br />

Tim Whitmore, one <strong>of</strong> the world’s foremost rainforest experts, studied the world’s tropical rainforests<br />

for many years. His observations typify PNG’s forest development. These observations include:<br />

• In lands where tropical rainforests occur, man has lived in closest dependence on them since<br />

time immemorial.<br />

• Europeans became aware <strong>of</strong> them over two millennia ago.<br />

• Increased knowledge since the Renaissance with the voyages <strong>of</strong> discovery and then the<br />

colonial era revealed that there are in fact many different kinds <strong>of</strong> tropical rainforest. Plants<br />

exist in luxuriance and a diversity <strong>of</strong> bizarre forms undreamed-<strong>of</strong> in temperate latitudes.<br />

Animal life is also rich and diverse.<br />

• Tropical forests have waxed and waned since geological time, and the present patterns <strong>of</strong><br />

species distributions are a result <strong>of</strong> these historical events.<br />

• The former idea that these great forests have survived immutable “since the dawn <strong>of</strong> time” is a<br />

romantic fallacy, as investigations over the last three decades have shown.<br />

• The forests are continually changing at the other end <strong>of</strong> the time scale, the life span <strong>of</strong> an<br />

individual tree.<br />

• The elucidation <strong>of</strong> forest dynamics has been the other major breakthrough <strong>of</strong> recent years.<br />

Now we know a great deal about the ecology <strong>of</strong> individual tree species and the particular


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 416<br />

requirements for growth <strong>of</strong> their seedlings in canopy gaps formed by the deaths <strong>of</strong> big trees.<br />

Silviculture, the manipulation <strong>of</strong> forests by Man to favour tree species <strong>of</strong> his choice is<br />

dependent on understanding these innate characteristics.<br />

• Tropical rainforest nutrient cycles are now reasonably understood. The old idea <strong>of</strong> a closed<br />

cycle and with nearly all the nutrients in the plants has not survived.<br />

• Shifting agriculture in now well known to be a sustainable form <strong>of</strong> farming, suitable for<br />

infertile soils.<br />

• Sustainable human utilisation <strong>of</strong> forest lands for crops or trees depends, as with silviculture,<br />

on working within the natural limits <strong>of</strong> the nutrient cycle.<br />

Whitmore identified further major impacts on tropical societies as:<br />

• The colonial era, which had a pr<strong>of</strong>ound impact on tropical societies. Now largely disappeared<br />

it left tell-tale traces in forest structure or species composition. The spice trade shaped the<br />

history <strong>of</strong> the world. Useful plants were moved between continents and introduced to new<br />

regions via Botanic gardens. Today, staple foods and many fruits are pan-tropical as a result.<br />

• The industrial revolution increased demand for many forest products such as rubber and<br />

resins. Now trade is greater in tropical hardwood timbers.<br />

• Today, tropical woody vegetation is being altered at a rapid rate in one <strong>of</strong> two different ways..<br />

Some areas are converted to other land uses (agricultural pursuits). Other areas are logged but<br />

left to regenerate, but may be destroyed later by shifting agriculture as farmers gain readier<br />

access to land being opened up by logging roads.<br />

Climate Change Scenarios and their Impact on PNG’s Vegetation<br />

Patterns <strong>of</strong> distribution <strong>of</strong> plants, animals and vegetation have long fascinated biologists. Over the past<br />

few decades two major causes have been discovered for many <strong>of</strong> the present day patterns seen<br />

amongst rain forest plants and animals. These are continental drift and past fluctuations <strong>of</strong> climate.<br />

Similarities in flora between the three regions <strong>of</strong> tropical rainforest occur because all are part <strong>of</strong> old<br />

Gondwanaland, The major evolution <strong>of</strong> the flowering plants had occurred before Gondwanaland<br />

began to break-up and has continued on the different fragments.<br />

Climate is the principal factor which controls the growth <strong>of</strong> plants and constitutes the conditions which<br />

render a country suitable for the abode <strong>of</strong> man and animals. Of course other factors such as<br />

topography, altitude and soil type also influence plant growth.<br />

These changes in forest cover can thus induce feedback effects on the climate by modifying surface<br />

temperatures and by influencing carbon dioxide concentrations. Forests have a lower reflectivity than<br />

other ecosystems and through their extensive root systems have more access to soil water than other<br />

types <strong>of</strong> vegetation. In consequence, they absorb more solar energy, which can lead to heating and loss<br />

<strong>of</strong> more water through evaporation, which can lead to cooling. In tropical zones, evaporative processes<br />

tend to dominate and the net effect <strong>of</strong> forests is to cool and moisten the environment.<br />

Forests which have been managed primarily for timber production could also be managed for climate<br />

mitigation and other environmental values.<br />

The greatest amount <strong>of</strong> carbon is captured and stored when tree growth is most rapid which is in their<br />

infant stages. This paper explores how PNG can best exploit that attribute.<br />

MANAGEMENT OF RISK AND IDENTIFICATION OF OPPORTUNITIES<br />

The question <strong>of</strong> how to make the best use <strong>of</strong> PNG’s forest resources is a complex one and is affected<br />

by the laws <strong>of</strong> the market; supply, demand and competition, as well as by environmental concerns and<br />

even by history.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 417<br />

For PNG producers and exporters to be successful, they must be able to optimize their logistics chain<br />

to make their integrated timber harvesting/wood production processes an efficient and economical<br />

material flow and data flow. Without this, PNG producers will not be able to meet the increasing trend<br />

<strong>of</strong> globalization in creating a buyer’s market, which is forcing suppliers to produce customer tailored<br />

goods more cost effectively and more swiftly. Importers need to know volumes <strong>of</strong> product on <strong>of</strong>fer,<br />

schedules <strong>of</strong> delivery and, above all, consistency, quality, quantity and reliability <strong>of</strong> products being<br />

exported.<br />

The problem facing any successful PNG producer/exporter is to find markets for PNG’s residual logs<br />

and lower quality forest products.<br />

Risks Facing Forest Developers Include:<br />

1. Historical Risk<br />

The natural forests <strong>of</strong> PNG are not public forests – they are privately owned and unless they become<br />

part <strong>of</strong> a forest management ethic and practice on the part <strong>of</strong> resource owners, it can be questioned<br />

whether there will be any effective base for the long term management goals <strong>of</strong> forest policy.<br />

2. Erosion <strong>of</strong> Forest Investment Capital<br />

Unless PNG can halt the decline <strong>of</strong> the existing forest industrial base through the erosion <strong>of</strong><br />

investment capital — due to impediments such as unsustainable and discriminatory taxation burdens,<br />

lack <strong>of</strong> rural infrastructure development and maintenance, and a lack <strong>of</strong> reinvestment by PNG back<br />

into the forest sector — there will be enormous poverty in the rural areas.<br />

3. Impact <strong>of</strong> PNG Government Decisions<br />

The PNG forest industry is on the “user end” <strong>of</strong> government decisions and actions with regard to a<br />

stable investment climate, resource security, and consistent administration <strong>of</strong> rules and regulations,<br />

especially those relating to forest revenue systems in the context <strong>of</strong> fiscal stability. Without such<br />

government action, the sector will falter and PNG will fare poorly in terms <strong>of</strong> the sustainable<br />

development <strong>of</strong> its forest resources.<br />

4. PNG Forest Revenue System<br />

PNG needs to harness the potential <strong>of</strong> the forest industry to become partners in development by<br />

providing realistic operating and fiscal conditions, which would bring back confidence to the sector.<br />

The forest revenue system, including export tax, needs to be a tool for achieving sustainable forest<br />

management, not just a means <strong>of</strong> raising government revenue.<br />

5. Consistency <strong>of</strong> Long Term Policies<br />

With regard to policies and practices which would encourage industry development, it is not just a<br />

question <strong>of</strong> policies being conducive to attracting the initial investment. Rather it is the whole policy<br />

matrix, and how practices are formulated within that matrix, that will influence the investment<br />

outcome. To achieve an internationally competitive industry, policies must focus on international best<br />

practice to be successful. Accordingly, policies, which guarantee a secure resource base, in large<br />

enough volumes to support the investment, must be the starting point.<br />

6. PNG Forest Industry Association members are restricted by regulation from undertaking many<br />

development activities. For example:<br />

• Reforestation levies which are paid to government to plant more trees.<br />

• The government is responsible to maintain infrastructure in concession areas.<br />

• Roads in concession areas are constructed only to allow harvesting to be undertaken.<br />

7. Poor Land Use Planning Controls allowing the emergence <strong>of</strong> alternative economic<br />

development in designated forest development areas, which permanently remove forest cover, such as<br />

oil palm plantations, other cash cropping, and mining.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 418<br />

HOW CAN GOVERNMENT AND INDUSTRY COLLABORATE TO REMOVE OBSTACLES<br />

TO GROWTH IN THE FOREST SECTOR?<br />

• Together, industry and government must identify and then resolve issues which impede the<br />

development <strong>of</strong> a viable downstream processing industry in PNG.<br />

• The role <strong>of</strong> government is to act as a catalyst for change by providing support framework to<br />

help implement change and assist with specific actions.<br />

In the forest sector <strong>of</strong> PNG there is significant potential for new employment creation together with<br />

wealth creation and rural development for resource owners and the nation. This is <strong>of</strong> particular<br />

relevance when it is remembered that the forest sector <strong>of</strong> the economy operates in rural areas where<br />

job opportunities and development have been in decline over recent years.<br />

However, if PNG wishes to alleviate poverty by generating wealth and creating new employment it is<br />

trade and new trade opportunities that will be the catalyst, not foreign aid, .<br />

Success will be achieve if strong and competent operators are maintained, and attracted to invest more<br />

in PNG, under regulations which give a fair deal to all parties – resource owners, resource investors<br />

and GOPNG, both for current and future harvesting cycles from production forest.<br />

OPPORTUNITIES FOR PNG FOREST INDUSTRY<br />

The Capacity <strong>of</strong> PNG Species for Plantation Development and Carbon Capture<br />

Much work is required to bring rainforest under scientific management because, in a broad sense,<br />

rainforest varies in physiognomy and floristic composition with varying environmental factors which<br />

shape all plant communities. Any management techniques must ultimately rest upon a sound<br />

understanding <strong>of</strong> forest ecology.<br />

If PNG were to use the natural forests as carbon sinks, the management <strong>of</strong> these natural forests would<br />

pose difficulties since it requires such detailed ecological knowledge <strong>of</strong> a vast number <strong>of</strong> species. In<br />

the natural forest there is considerable variation in the growth rates <strong>of</strong> the valuable species besides an<br />

abundance <strong>of</strong> secondary species which have no solid wood value and may in fact be poor carbon fixers<br />

due to their poor growth rates.<br />

This is a gross oversimplification, as economic factors in particular frequently outweigh the ecological<br />

factors in significance. The basic premise remains, however, that economic conditions in this<br />

technological age can change with alarming rapidity, but the ecological behaviour <strong>of</strong> rainforest<br />

communities and <strong>of</strong> the species making up those communities is relatively immutable and can be<br />

disregarded only at great risk to the final success <strong>of</strong> management operations.<br />

However, for PNG, it has the notable feature that many <strong>of</strong> its native species behave well in plantation<br />

format and, with their notable feature <strong>of</strong> fast growth rates, demonstrate the capability to produce, in<br />

half the time or less <strong>of</strong> the northern and southern hemispheres, wood with acceptable properties<br />

suitable for all forms <strong>of</strong> value-adding processes, besides capturing and storing large amounts <strong>of</strong><br />

carbon.<br />

Through silvicultural plantation techniques, further refinement <strong>of</strong> plantation growth (and hence carbon<br />

capture) can be undertaken by:<br />

• Matching species to site conditions for optimum growth<br />

• Improving establishment, tending and maintenance regimes<br />

• Manipulating the environment and nutrients<br />

Potential PNG Plantation Species<br />

Once the end users have been clearly defined, the choice <strong>of</strong> species, whether indigenous or exotic, is<br />

dependent on the end use or the environmental role <strong>of</strong> the plantation, coupled with the matching <strong>of</strong><br />

specific species to specific sites. For PNG proven plantation species include:


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 419<br />

INDIGENOUS EXOTIC<br />

A. cunninghamii; Araucaria hunsteinii; Acacia<br />

mangium; other acacias; Eucalyptus deglupta;<br />

E.pellita; Terminalia sp; Pometia pinnata;<br />

Casuarina spp; Pterocartpus indicus;<br />

Octomeles spp<br />

Tectona grandis<br />

Pinus spp<br />

Ochroma lagopus<br />

Benefits <strong>of</strong> Plantation Development in PNG<br />

• It represents a major opportunity for growth in PNG’s long term wood supply as well as emerging<br />

emission trading scheme markets.<br />

• It would be a significant contributor to regional economic development by value adding locally to<br />

primary production and generating downstream processing jobs.<br />

• It provides a long term agricultural crop which can be managed to produce large volumes <strong>of</strong> wood<br />

per unit area.<br />

Action Plan for a properly constructed Emission Trading Scheme using PNG Forest Resources<br />

Such an action plan should include not just government policy measures and the supportive roles <strong>of</strong><br />

government agencies, but more importantly also indicate how forest plantation activities can be made<br />

an attractive business proposition to investors through various financing options.<br />

For the partnership in emissions trading to be sustainable and commercially successful, an action plan<br />

needs to ensure:<br />

• The optimum species are used for both wood quality and carbon capture<br />

• Verification <strong>of</strong> the mechanism for carbon capture<br />

• Development <strong>of</strong> core competencies in-country to accomplish the two dot points above<br />

• Establishment <strong>of</strong> a reputable carbon market and a sound regulatory framework<br />

• Development and adoption <strong>of</strong> a national financial incentive mechanism to fight deforestation<br />

• The proper use <strong>of</strong> forest products such that carbon is stored and not released to the atmosphere<br />

• Resolution <strong>of</strong> the conflict between the use <strong>of</strong> forest products as a source <strong>of</strong> biomass energy<br />

and its consequent release <strong>of</strong> CO2<br />

• Mobilization <strong>of</strong> the existing hardwood industries to undertake plantation development<br />

• Development <strong>of</strong> access to investment capital (for establishment and processing)<br />

• Encouragement <strong>of</strong> confidence in the economics <strong>of</strong> plantations<br />

• Mechanisms are in place to develop and implement a communication strategy to deal with:<br />

o The lack <strong>of</strong> silvicultural knowledge<br />

o Support for the implementation <strong>of</strong> plantation establishment<br />

o The need to influence landholder, industry and community attitudes<br />

• Identification <strong>of</strong> the land base and product priorities through planning initiatives and<br />

management over 5, 10, and 20 year horizons, in conjunction with regional development<br />

organizations under the auspices <strong>of</strong> the National Forest Plan<br />

The Need for Strategic Industry Status for PNG Industrial Forest Plantations<br />

PNG is at the crossroads <strong>of</strong> forestry development, with the painful experience (economically and<br />

socially) <strong>of</strong> developing its natural forest resources further, whilst at the same time greatly expanding<br />

its forest plantation resources. Of course, the key elements <strong>of</strong> structural change in the PNG forest<br />

industry, and especially plantation development, are related to a variety <strong>of</strong> issues concerning demand,<br />

supply and international trade in forest products. Above all, what products are going to be made and<br />

marketed from plantation grown resource?<br />

From industry experience, the challenge in the longer term if the industry wishes to survive is to<br />

establish its own plantations and to grow wood for a range <strong>of</strong> markets rather than a single product,<br />

removing the uncertainties <strong>of</strong> the single-product markets. A bigger challenge for the industry is that, if<br />

it is to create new forest resources to compete in the world markets, it must be competitive.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 420<br />

The importance <strong>of</strong> forest plantation development in PNG to augment the supplies <strong>of</strong> logs from the<br />

natural forest needs to be recognized by both the government and the private sector. Plantation forests<br />

are seen as complementary to natural forests, never able to replace all the values associated with the<br />

natural forests but, where appropriately developed, helping to divert some <strong>of</strong> the pressures away from<br />

them.<br />

The government needs to accord “strategic industry” status to the forest plantation sector. The<br />

response to date has not been encouraging (only some 60,000 hectares since 1975) and much more<br />

needs to be done to study the issues surrounding forest plantation development.<br />

There have been no programs to encourage small and large investors for many years. Yet, under<br />

previous programs <strong>of</strong> some 35 years ago, forest extension programs alone saw over one million trees<br />

being planted each year by small holders for fuel wood, shelter wood and to augment future sawlog<br />

supplies.<br />

The issues surrounding forest plantation development have been discussed at numerous meetings,<br />

seminars and conferences. Reasons for poor performance such as inadequate or ineffective incentives,<br />

poor support from financing institutions, resource security in terms <strong>of</strong> land base and social issues,<br />

markets and shortage <strong>of</strong> sizeable land areas have been cited. As forest plantation development is<br />

becoming increasingly critical for the sustainability <strong>of</strong> the PNG wood based industry, it is imperative<br />

that concrete steps be taken to address the issues regarding forest plantation development and a<br />

practical action plan that could be implemented as soon as possible be formulated.<br />

May be this action plan could be included in the next revision <strong>of</strong> the PNG National Forest Plan?<br />

CONCLUDING REMARKS<br />

It is hoped that this policy paper will create opportunities for the PNG forest sector to diversify into<br />

new markets, assist to set the scene; ensure proper market outlooks, allow for growth in the sector<br />

though privatisation and competition by attracting investment and strengthening PNG’s research and<br />

development base.<br />

PNG, through large scale plantation development, has the potential and the capacity (land, soils, water,<br />

air and adaptable plantation species) to capture and store carbon in its biomass, soils and forest<br />

products in perpetuity. If, through the proper introduction <strong>of</strong> an emissions trading scheme, increased<br />

value and returns to the forest owners and developers can be realised, PNG has the capacity to<br />

increase:<br />

• Five fold its commercial timber tree plantation estate to some 300,000 hectares.<br />

• 20 % increase in forest industry employment and flow on effects to other employment sectors<br />

• an increase in value <strong>of</strong> forest products production <strong>of</strong> more than 100 %.<br />

REFERENCES<br />

Appanah S & Weinland G 1993. Planting Quality Timber trees in Peninsular Malaysia –a review Malayan<br />

Forest Record # 38 Forest Research <strong>Institute</strong> Malaysia 1993.<br />

Baur George 1961/62. The ecological basis <strong>of</strong> rainforest management. FAO<br />

Hillary, Sir Edmund 1984. Ecology 2000. – The changing face <strong>of</strong> earth. Editor. Michael Joseph London.<br />

McCarthy R B 2003. The Business <strong>of</strong> PNG Forestry – Implications for <strong>Australia</strong> and PNG. PNG Forest<br />

Industries Presentation 20 th <strong>Australia</strong>n PNG Business Forum Cairns.<br />

Papua New Guinea Forest Industries Association, 2008 Submission to the Garnaut Review on the Impact <strong>of</strong> an<br />

ETS on commercial forestry and development outcomes in PNG. PNG Forest Industries Association.<br />

Silver, Jerry 2008. Global warming and climate change demystified. McGraw Hill.<br />

Tennesen Michael . The complete idiot’s guide to global warming.<br />

Whitmore T. C. 1990 An Introduction to Topical Rainforest. Oxford University Press.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 421<br />

2009 - 2059 PNGFIA PLANTATION VISION<br />

“Five fold increase in commercial timber plantation area to 300,000 hectares.”<br />

APPENDIX 1<br />

1. PNGFIA goal is to build an internationally competitive, market orientated industry driven by<br />

private sector investment<br />

2. PNG has the necessary land resources, rainfall, commercial tree growth productivities<br />

compared to rest <strong>of</strong> world<br />

3. However, PNG needs the correct investment climate and resource security in order to<br />

undertake large scale plantation investment<br />

4. Benefits <strong>of</strong> large scale plantation development in PNG is that there is a creation <strong>of</strong> separate &<br />

alternative income streams for rural dwellers; enhanced regional development opportunities;<br />

increased employment in rural areas and environmental benefits e.g. reduced soil erosion,<br />

improved water quality, carbon sinks under a properly constructed emissions trading scheme<br />

5. The achievement <strong>of</strong> 2059 vision is the responsibility <strong>of</strong> the plantation growing industry, i.e.<br />

industry will be planting the trees, not the government<br />

6. Plantation wood cannot replace all timber uses. PNG will require continuing use <strong>of</strong> native<br />

forests for wood production with plantations supplying an economically viable, reliable and<br />

high quality wood resource which complements wood supplied from native forests<br />

7. PNG one <strong>of</strong> only a few countries capable <strong>of</strong> increasing its plantation base in next 50 years<br />

8. PNG plantation species (native and exotic) have great potential particularly if properly dried<br />

and presented. However, any development must be internationally competitive.<br />

9. PNG needs to look at the use <strong>of</strong> effluent for fuel wood plantations within urban areas<br />

10. There needs to be development <strong>of</strong> PNG commercial tree growers/ use <strong>of</strong> SRPM model for<br />

commercial woodlot owners<br />

11. PNG needs to develop cabinet wood plantations<br />

12. PNG needs to develop a code <strong>of</strong> practice for afforestation companies


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 422<br />

MANAGING FOREST COUNTRY:<br />

ABORIGINAL AUSTRALIANS AND FORESTS<br />

Sue Feary 1 , Peter Kanowski, Richard Baker and Jon Altman<br />

ABSTRACT<br />

Aboriginal <strong>Australia</strong>ns have a diversity <strong>of</strong> interests in their forests, encompassing<br />

cultural, economic, environmental and social values. Historically, the agencies and<br />

industries comprising the forests sector have engaged with only some <strong>of</strong> these interests,<br />

and have typically done so in a segmented fashion. Our research with Aboriginal<br />

communities around <strong>Australia</strong> suggests a myriad <strong>of</strong> opportunities for a broadly defined<br />

forests sector, but this requires improved relationships between Aboriginal people and<br />

the dominant culture and much deeper understanding <strong>of</strong> Aboriginal aspirations at the<br />

local level. The National Indigenous Forestry Strategy promotes these aspirations, but<br />

requires a much stronger commitment from governments if it is to deliver them.<br />

INTRODUCTION<br />

Aboriginal people’s views and aspirations about the forest sector have not always been well<br />

understood by forest managers, policy makers, or the community at large. Forest debates <strong>of</strong> the 1970s<br />

in <strong>Australia</strong> and subsequent adoption <strong>of</strong> the principles <strong>of</strong> sustainable forest management (SFM) have<br />

enabled Aboriginal voices to be heard through several consultative processes. These have consistently<br />

demonstrated three important goals for Aboriginal people: employment and economic development in<br />

the forest sector; recognition <strong>of</strong> customary rights over forests; and protection <strong>of</strong> the spiritual and<br />

cultural values <strong>of</strong> forests.<br />

Relevant national policy initiatives, such as the National Indigenous Forestry Strategy, aim to achieve<br />

these multiple goals, but this can be difficult where they are seemingly at some odds with each other.<br />

One <strong>of</strong> the barriers is the sectoral approach that state and federal governments have adopted for the<br />

management <strong>of</strong> forests, with a dichotomy established between conservation <strong>of</strong> forest lands on the one<br />

hand and their commercial exploitation for timber on the other. Thus, for example, forestry operations<br />

provide employment, but logging and road construction can threaten significant cultural places.<br />

Alternatively, reservation <strong>of</strong> forests in conservation areas ensures protection <strong>of</strong> cultural values <strong>of</strong><br />

forests but tends to <strong>of</strong>fer limited opportunities for economic development and may restrict certain<br />

cultural activities. This creates something <strong>of</strong> a dilemma for Aboriginal people for whom traditional<br />

processes <strong>of</strong> ‘caring for country’ involved both resource exploitation and protection.<br />

At the local level, Aboriginal people engage with the forest sector in a variety <strong>of</strong> ways, reflecting both<br />

the diversity <strong>of</strong> the forest sector and the social heterogeneity <strong>of</strong> Aboriginal culture. This paper<br />

presents the results <strong>of</strong> recent doctoral research throughout <strong>Australia</strong>, demonstrating that a ‘cultural<br />

match’ between the type <strong>of</strong> forest-related activity and the type <strong>of</strong> Aboriginal social unit can result in<br />

many different forms <strong>of</strong> productive partnerships within the broad paradigm <strong>of</strong> ‘Aboriginal forestry’.<br />

This research demonstrates that, in some instances, the forestry agenda has already been reshaped in<br />

response to local community situations. The challenge <strong>of</strong> the National Indigenous Forestry Strategy is<br />

to scale up local successes for wider application.<br />

CREATING A SPACE FOR ENGAGEMENT<br />

Native forests are part <strong>of</strong> the traditional landscape <strong>of</strong> <strong>Australia</strong>’s Indigenous people who have<br />

managed them for their utilitarian and spiritual values for thousands <strong>of</strong> years. It has been argued that<br />

colonial history was a dual war – a war against nature and a war against the natives (Rose, 2002:2).<br />

What was once a holistic landscape imbued with Indigenous meaning has become fragmented and its<br />

Aboriginal owners have, for the most part, become detached from it. Little attention was paid to<br />

1 The <strong>Australia</strong>n National University, Canberra. Contact author: suefeary@hotkey.net.au, Phone 02 44286312


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 423<br />

Aboriginal people’s views on forest management until after the 1970s, when timber production forests<br />

became the centre <strong>of</strong> debates about whether or not the forestry practices <strong>of</strong> the time were sustainable<br />

(Dargavel, 1995). The subsequent recognition <strong>of</strong> the multiple values <strong>of</strong> forests provided the first<br />

formal opportunities for Aboriginal people to articulate their views and interests at a national level.<br />

The principles <strong>of</strong> sustainable forest management have been important in creating a space for<br />

engagement between Aboriginal people and the forest sector in three main areas. Firstly, a greater<br />

recognition <strong>of</strong> environmental and socio-cultural values reflects a move from a forest management<br />

approach towards a forest ecosystem approach (Holling et. al., 1998). This is important from an<br />

Indigenous perspective because recognising trees as part <strong>of</strong> a functioning landscape <strong>of</strong> complex<br />

ecosystems and natural cycles resonates with indigenous worldviews <strong>of</strong> holistic, integrated ecosystems<br />

that also include humans (Rose, 1996). The second concerns the major objective <strong>of</strong> sustainability -—<br />

finding balances between social, cultural, environmental and economic factors. This opens the way for<br />

using participatory mechanisms to include Aboriginal people in negotiations, and recognises that they<br />

have interests in each <strong>of</strong> these arenas. Finally, both the principle and the practice <strong>of</strong> SFM specifically<br />

acknowledge the rights <strong>of</strong> indigenous peoples to have their concerns heard and to participate in<br />

decision-making over forests (Higman et al., 2005).<br />

Nationwide consultation with Aboriginal people about forests and forest management has occurred<br />

three times over the last three decades. The first was through the Resource Assessment Commission’s<br />

(RAC) Inquiry into the Forest and Timber Industry in the late 1980s, when anthropologist Dr Scott<br />

Cane conducted a review <strong>of</strong> Aboriginal attitudes to forests through a series or workshops and<br />

interviews (Cane 1990). Cane’s report highlighted a multiplicity <strong>of</strong> Aboriginal interests in the forest<br />

sector, including land ownership, site protection, consultation, land management and access for<br />

cultural purposes, such as hunting.<br />

The second consultative process was associated with the Regional Forest Agreements (RFA),<br />

emerging from the 1992 National Forest Policy Statement, which coincidentally was agreed in the<br />

same year as recognition <strong>of</strong> native title. A review <strong>of</strong> social assessments carried out during the<br />

Comprehensive Regional Assessments between 1996 and 2000 (Black et al., 2003; Brooks et al.,<br />

2001), demonstrated that Aboriginal people had reiterated their previous concerns. Indigenous<br />

people’s interests in the forest sector, arising from the RAC and CRA processes, are summarised<br />

below:<br />

¤ Consultation and participation in forest management decisions<br />

¤ Land management –soil and water quality<br />

¤ Access for cultural activities<br />

¤ Site management and protection<br />

¤ Cross-cultural awareness training for non-Aboriginal agency staff.<br />

¤ Land claims/native title./prior ownership recognition<br />

¤ Business development<br />

¤ Employment and training opportunities<br />

¤ Compensation/royalties<br />

¤ Formation <strong>of</strong> agreements and partnerships<br />

¤ Control <strong>of</strong> cultural information/ mechanisms for ensuring confidentiality for cultural<br />

information<br />

The most recent forest-related consultation with Aboriginal people was during development <strong>of</strong> the<br />

National Indigenous Forestry Strategy (NIFS) in 2003 (<strong>Australia</strong>n Government 2005). Aboriginal<br />

people’s views were canvassed at workshops on perceived opportunities and barriers to forest-based<br />

development in their local areas (BDO Consulting (SA) Pty Ltd, 2004). The focus <strong>of</strong> these<br />

consultations was on economic development, rather than on the broader matters <strong>of</strong> land management,<br />

rights and social justice, and drew attention to Indigenous interests in plantations, something which<br />

had not been documented previously.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 424<br />

Despite the intensely localised nature <strong>of</strong> the demands by Aboriginal people, three major themes,<br />

emerged from these consultations: economic development, recognition <strong>of</strong> rights, and protection <strong>of</strong><br />

cultural values. The fact that these aspirations have been articulated by Aboriginal people since the<br />

RAC inquiry in 1989 demonstrated that overcoming social and economic disadvantage is not just<br />

about employment and wealth generation, but must be situated in broader social justice agendas.<br />

It is important to acknowledge that the benefits <strong>of</strong> creating a space for engagement between<br />

Aboriginal people and the forest sector are not just one-sided. Aboriginal <strong>Australia</strong>ns now own 16% <strong>of</strong><br />

<strong>Australia</strong>’s land area and 14% <strong>of</strong> <strong>Australia</strong>’s forests (Montreal Process Implementation Group 2008).<br />

The National Indigenous Forestry Strategy noted that the forest sector could benefit from access to<br />

Aboriginal land, both for growing trees and for utilising Aboriginal owned forests and woodlands<br />

(<strong>Australia</strong>n Government 2005).<br />

Additionally, since 1993, all forests on crown land are potentially subject to native title. Managers <strong>of</strong><br />

both wood production and protected forests need to have dialogue with traditional owners to address<br />

uncertainty in regard to the impact <strong>of</strong> native title rights and interests on their activities. In most<br />

instances, these negotiations are opting for agreements and settlements outside <strong>of</strong> the lengthy court<br />

hearings, for example, Indigenous Land Use Agreements or joint management arrangements.<br />

Agreements <strong>of</strong> this kind go some way to addressing the three main aspirations <strong>of</strong> Aboriginal people<br />

while also providing some level <strong>of</strong> certainty to forest management authorities. The negotiation and<br />

engagement pathways <strong>of</strong> timber production and conservation agencies are discussed later in this paper.<br />

CONCEPTUALISING ENGAGEMENT<br />

The forest sector and Indigenous society are separate complex domains comprising multiple parts.<br />

Dovers (2003:26) points out that policy processes have not acknowledged that there is more than one<br />

forestry industry. Meanings ascribed to ‘forestry’ have broadened as a result <strong>of</strong> societal change and<br />

political processes over the last thirty years. Thus, ‘forestry’ is no longer just about growing tall,<br />

straight trees to cater for the needs <strong>of</strong> the timber industry, but about forest management catering for a<br />

range <strong>of</strong> forest based interests and values. Globally, the economic values <strong>of</strong> forest products have<br />

expanded to include non-timber forest products and, the commercial use <strong>of</strong> forest’s intrinsic values<br />

through ecotourism. Most recently, debate on climate change has focused attention on the<br />

environmental services <strong>of</strong> forests, and the role that indigenous people can play in environmental<br />

stewardship (Wunder, 2005).<br />

Turning to the Indigenous side <strong>of</strong> the equation, there is a prevalence in <strong>Australia</strong>n society to construct<br />

all Aboriginal people as united and part <strong>of</strong> a same-thinking, pan-Indigenous society, regardless <strong>of</strong><br />

whether its subjects live in inner-city Redfern or in the desert. Many non-Aboriginal people see the<br />

diversity <strong>of</strong> views and opinions expressed within an Aboriginal community group as signs <strong>of</strong><br />

dysfunction and barriers to resolving an issue. Consequently, rather than developing systems for<br />

responding to diversity, effort focuses on trying to get consensus (Sullivan, 2004). This is not confined<br />

to Aboriginal <strong>Australia</strong>ns; any minority group tends to be regarded by the dominant culture as<br />

homogeneous (Dyck, 1985).<br />

Conversely, recognising diversity and heterogeneity within both the forest and Indigenous domains<br />

creates a useful space for exploring connections between the fine-scaled elements <strong>of</strong> each within the<br />

respective bigger pictures. The conceptual tool used to bring these two complex systems together is a<br />

matrix, comprising categories <strong>of</strong> the forest sector along one axis and Indigenous social units along the<br />

other, as shown in Figure 2. Historical legacies and geographical location comprise third and fourth<br />

dimensions, in recognition <strong>of</strong> their influences on contemporary Aboriginal culture.<br />

The depiction <strong>of</strong> Indigenous social structures shown in Figure 2 does not purport to reflect any<br />

particular theoretical position in anthropology. Different kinds <strong>of</strong> Indigenous organisations and<br />

structures are formed in response to different legislative requirements or other external influences:<br />

some may be strongly influenced by customary factors, others less so. Rather, this model is schematic<br />

and illustrative <strong>of</strong> Aboriginal representation. Similarly, the forest sector axis is indicative rather than<br />

exhaustive.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 425<br />

Figure 2: A conceptual diversity matrix linking Aboriginal people and the forest<br />

sector.<br />

The matrix is a heuristic device for exploring the diversity <strong>of</strong> connections between Aboriginal society<br />

and a broadly defined forest sector, premised on finding the right ‘match’ between forestry sector<br />

category, and ‘type’ <strong>of</strong> Indigenous social structure. Examples <strong>of</strong> appropriate matches could be: a<br />

natural resource management project in a remote area involving a group <strong>of</strong> traditional owners on their<br />

country; or an apprenticeship in a timber yard for a young Aboriginal man living in a regional<br />

township. Examples <strong>of</strong> other potential matches are shown in Figure 2. An appropriate ‘match’ between<br />

the forest sector unit and the Indigenous unit may be a significant factor in creating an enabling<br />

opportunity for addressing disadvantage. The Harvard project in America is having considerable<br />

success in using a similar model for developing effective local governance structures in Indigenous<br />

American communities (Cornell and Begay, 2003), and is gaining popularity in <strong>Australia</strong> (Behrendt,<br />

2003). Application <strong>of</strong> this model in forms appropriate to <strong>Australia</strong> <strong>of</strong>fers good prospects for enabling<br />

mutually beneficial cooperation between Indigenous <strong>Australia</strong>ns and the forest sector.<br />

TESTING THE MODEL<br />

Case study research was used to conduct in-depth analysis <strong>of</strong> the relationships between the forest<br />

sector and Aboriginal people, and to test the matrix model described above. Four main themes guided<br />

case study selection – remoteness, management <strong>of</strong> Crown native forests, involvement with the<br />

plantation industry and forest-based economic development. Research was conducted in four forested<br />

case study locations – western Cape York in far north Queensland, southwest Western <strong>Australia</strong>, the<br />

Riverina where it straddles the NSW/Victorian border, and southeastern New South Wales (Feary,<br />

2007).<br />

Remoteness<br />

Forest based relationships in remote settings were explored through partnerships between the<br />

Queensland government and the small Aboriginal communities <strong>of</strong> Injinoo at the tip <strong>of</strong> Cape York and<br />

Napranum near Comalco’s mining town <strong>of</strong> Weipa. Over the last decade, government agencies have<br />

invested effort in establishing forestry enterprises to assist social and economic development in a


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 426<br />

region where opportunities for employment and business development are extremely limited and the<br />

majority <strong>of</strong> the Aboriginal population receives welfare support (Annandale and Feary, 2009).<br />

The Injinoo sawmilling project was a partnership between the Injinoo community and the Queensland<br />

Government, aimed at building a team <strong>of</strong> skilled local people capable <strong>of</strong> using their forest resource to<br />

construct houses and visitor facilities (Annandale et. al., 2002). From 2001, local Aboriginal men were<br />

trained and accredited in harvesting, transportation, milling and grading <strong>of</strong> timber (Annandale et. al.,<br />

2002). On returning to the area to evaluate progress in 2005, government staff were met with marked<br />

trees still unfelled and an idle sawmill (Feary, 2007). Although still enthusiastic, the community<br />

expressed little commitment to keeping the project running. Reasons included tensions between<br />

different community organisations and the need for ongoing mentoring by the government to build<br />

confidence and expertise among the trained men.<br />

The situation was different at Napranum in 2005, where bauxite mining, which displaced Aboriginal<br />

people from their traditional lands in the 1950s, is now assisting in a viable community-run sawmilling<br />

operation. Until recently, Darwin stringybark forests in the mining leases were cleared and burnt by<br />

Comalco to allow access to the bauxite layer beneath (Annandale and Taylor, 2007). Not only was this<br />

a waste <strong>of</strong> usable timber, burning <strong>of</strong> extensive areas <strong>of</strong> forest was contributing to greenhouse gases. In<br />

the early 1990s, a new Indigenous owned business called Nanum Tawap Ltd was established, to<br />

harvest useable timber from the lease area prior to mining. The business is run by members <strong>of</strong> the five<br />

clan groups whose traditional country is covered by Comalco’s mining lease. The sawmill sells to both<br />

a local and export market and currently has a non-Aboriginal manager and employs five local<br />

Aboriginal men on average.<br />

Superficially it could be a small sawmill operation anywhere in <strong>Australia</strong>, but conversations with<br />

sawmill workers gave insight into the connections between cultural behaviours, economic activity, and<br />

the capacity for adaptation. Both Aboriginal and non-Aboriginal people recognised barriers to<br />

mainstream employment arising from a history <strong>of</strong> lack <strong>of</strong> workforce participation, limited formal<br />

education and social problems characteristic <strong>of</strong> poor, marginalised peoples. The sawmill operation has<br />

some innovative and pragmatic adaptations for ensuring its economic viability. Firstly, in recognition<br />

that the workforce will vary in its numbers and skill levels from day to day, the make <strong>of</strong> sawmill,<br />

although relatively more expensive, can be operated by one man if necessary, i.e. if nobody turns up to<br />

work. Secondly, the issue <strong>of</strong> vouchers to buy sought after white goods at the local electrical goods<br />

store was an unorthodox productivity incentive.<br />

The Nanum Tawap sawmill is a mainstream forestry operation situated in a cultural setting. There are<br />

arguments both for and against separating culture from business in Aboriginal communities (Altman,<br />

2001) but the approach taken at Napranum has been to separate the business from the operations and<br />

activities <strong>of</strong> the rest <strong>of</strong> the community. The workplace is separated from the wider Napranum<br />

community by physical, administrative, and cultural boundaries. But, cultural traditions are recognised<br />

at higher levels; in the administrative and management structures <strong>of</strong> the business, where the Board has<br />

real power in decision-making and where traditional clan territories are recognised and respected.<br />

Evaluation <strong>of</strong> the timber resource by the Queensland government indicated that the sawmill’s capacity<br />

would need to be substantially increased to maximise economic returns from timber salvage<br />

operations (Annandale and Taylor, 2007). The Nanum Tawap Steering Committee considered these<br />

options and supported a business model that would maximize local employment by having three small<br />

sawmills, one in each community, rather than a single large mill requiring fewer people but needing<br />

higher levels <strong>of</strong> technical skills. This decision is significant for demonstrating that building social<br />

capital in local communities can take precedence over maximizing economic returns.<br />

Management <strong>of</strong> Crown native forests<br />

Naturally vegetated lands, including forests, are significant to many Aboriginal people. They are a<br />

reminder <strong>of</strong> a once holistic Indigenous-owned landscape from which they were alienated during<br />

colonisation, especially in ‘settled’ <strong>Australia</strong>. Since most native forests are owned by the crown<br />

(Dargavel 1995), Aboriginal connections with native forests occur mainly in the context <strong>of</strong><br />

government instrumentalities. All tiers <strong>of</strong> government recognise, to a greater or lesser extent, social<br />

justice and indigenous rights, together with state land rights and native title legislation, in their


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 427<br />

policies, vision statements, codes <strong>of</strong> practice and the like. Most include mechanisms for consulting<br />

with Aboriginal people with varying degrees <strong>of</strong> effectiveness.<br />

Mechanisms for involving Aboriginal people in forests on Crown land, both wood/fibre production<br />

and protected area forests, were explored through case study research in Western <strong>Australia</strong>, the<br />

Riverina and southeastern NSW (Feary, 2007). Interviews with Aboriginal people suggested that<br />

consultation processes were very important in defining their relationship with the resource extraction<br />

sector. In fact, Aboriginal people frequently judged forestry agencies as much on their consultative<br />

and participatory processes as on their skills and abilities as forest managers.<br />

Interviews with non-Aboriginal people revealed many <strong>of</strong> the complexities <strong>of</strong> consulting with<br />

Aboriginal people, especially in relation to representation and identifying those with the ‘right to<br />

speak’. The increasing numbers <strong>of</strong> Aboriginal community organisations with interests in land<br />

management, together with intra- and inter-community contestations over traditional connections with<br />

the land, <strong>of</strong>ten result in unsatisfactory outcomes or prolonged periods <strong>of</strong> negotiation. Despite the<br />

existence <strong>of</strong> agency policies for consultation, both sets <strong>of</strong> actors feel that the other side could do better.<br />

Case study research suggested that timber production agencies have applied the principles <strong>of</strong><br />

sustainable forest management quite narrowly; limiting it to consultation, protection <strong>of</strong> tangible<br />

cultural heritage, some limited employment and access for local communities to conduct cultural<br />

activities. At a global scale, this is something <strong>of</strong> a contrast to other settler societies (NZ and the<br />

Americas) and developing countries, where the broader rights and interests <strong>of</strong> indigenous people have<br />

been incorporated into SFM in these countries. The narrow focus is problematical because there is a<br />

danger <strong>of</strong> consultation being an end in itself, representing the beginning and end <strong>of</strong> Aboriginal<br />

involvement in a commercial forestry operation.<br />

Increasingly there are exceptions, including the recent Githabul Indigenous Land Use Agreement over<br />

forests in northern NSW and southern Queensland, and agreements for managing the river red gum<br />

forests in Victoria, over which the Yorta Yorta people lost their native title claim (Morgan et al 2006).<br />

On the NSW far south coast, an intensely negotiated Eden RFA delivered arrangements between the<br />

Eden Local Aboriginal Land Council and the local forestry <strong>of</strong>fice for contract employment and the<br />

transfer <strong>of</strong> ownership <strong>of</strong> some Crown forests to the Land Council.<br />

Consultation processes were not the main focus <strong>of</strong> Aboriginal interviewees’ interests in natural<br />

resource management and conservation. Instead, conversations moved beyond consultation to focus<br />

more on participatory processes and partnerships that integrated scientifically based land management<br />

and research with customary Aboriginal knowledge. The intrinsic synergies between nature<br />

conservation and Aboriginal worldviews about caring for the environment have resulted in relatively<br />

greater expectations for conservation agencies to play a major role in addressing Indigenous rights and<br />

social justice.<br />

However, the views <strong>of</strong> Aboriginal people reflected ambivalence towards protected area management.<br />

The challenge for both Aboriginal and non-Aboriginal people, is understanding the similarities and<br />

differences between nature conservation as practiced by western cultures and Aboriginal worldviews<br />

about land management, <strong>of</strong>ten labelled ‘caring for country’. There is a commonly held belief among<br />

non-Indigenous <strong>Australia</strong>ns that conservation <strong>of</strong> nature through protected area management is a<br />

modern equivalent <strong>of</strong> traditional practices <strong>of</strong> ‘caring for country’. It is true that many overlaps do exist<br />

and joint management <strong>of</strong> national parks across the country attests to the synergies. But, at the same<br />

time these cooperative arrangements mask the inevitable tensions between Aboriginal owners and<br />

national park agencies. From an Aboriginal perspective, ‘caring for country’ includes the exploitation<br />

<strong>of</strong> natural resources through hunting and collecting, which is at odds with the objectives <strong>of</strong><br />

contemporary nature conservation goals (Young 2001). Understandably, many Aboriginal people<br />

consider that conservation objectives <strong>of</strong> the western world are too restrictive and serve only to further<br />

separate humans from nature. At the same time, they also know that a national park (as opposed to<br />

development) <strong>of</strong>fers the best chance to protect the natural landscape and its associated cultural values.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 428<br />

Involvement in the plantation industry<br />

Plantation forestry <strong>of</strong>fers an alternative to logging native forests. Although the focus is on commercial<br />

wood and wood fibre production, the functions <strong>of</strong> plantations in reversing land degradation are<br />

becoming recognised (Varmola, 2005). Aboriginal people’s engagement with plantation forestry may<br />

occur as employees or as landowners, and both were examined through case study research in Western<br />

<strong>Australia</strong> and Cape York.<br />

Plantation forestry is a major industry in southern WA with thousands <strong>of</strong> hectares planted to<br />

Tasmanian blue gums and other species. An Esperance based company, Integrated Tree Cropping<br />

(ITC), contracted the Esperance Aboriginal Corporation (EAC) to undertake various projects at their<br />

plantations. ITC has organised training courses with a view to making Aboriginal workers ‘job ready’<br />

and is encouraging its Aboriginal workers to become pr<strong>of</strong>icient in harvesting techniques, to take<br />

advantage <strong>of</strong> new job opportunities when the current plantations are ready to harvest.<br />

Case study research in WA suggested that local Aboriginal people did not have a very good reputation<br />

in relation to work ethics, which, together with lack <strong>of</strong> training and education, has created barriers to<br />

Aboriginal employment in plantation forestry. However, ITC were making a considerable effort to<br />

engage Aboriginal people, while acknowledging that it required extra effort in providing training and<br />

may result in less efficiency and be less productive. The company gave Forest Stewardship Council<br />

(FSC) certification as a basis for their commitment to providing opportunities for Aboriginal<br />

employment. Recognising the rights and interests <strong>of</strong> indigenous peoples is an FSC criterion (Forest<br />

Stewardship Council, 2007).<br />

EAC members enjoyed the work because it was outside, enabled people to work together and had<br />

flexible working hours and to this end, demonstrated a strong work ethic. Other Indigenous people<br />

interviewed in WA supported sharefarm arrangements for plantations on their land because <strong>of</strong> long<br />

term and short economic returns, providing useful employment as well as helping to restore degraded<br />

agricultural land.<br />

Similar sentiments were expressed by an Indigenous man involved with sandalwood plantations on<br />

Cape York :<br />

‘The land needs to be in good condition when handed back after mining. That’s why we<br />

are growing sandalwood, to make the land healthy…‘We’ll just keep planting<br />

sandalwood so when Comalco does decide to shift out, traditional owners can still get<br />

money from the land by harvesting the sandalwood… Planting sandalwood will<br />

hopefully provide jobs and put money back into the community, reduce handouts from<br />

the government. That’s what I’d like to see’ (Feary, 2007).<br />

The need for land to grow trees, combined with positive responses by Aboriginal people to<br />

plantations, situates plantation forestry favourably for contributing to Indigenous people’s social and<br />

economic wellbeing. However, plantation forestry is highly commoditised, with engagement occurring<br />

in the mainstream market economy in a largely privatised industry. Aboriginal people are competing<br />

in a labour market for which they may not be ready or want to enter. There is little room for<br />

‘indigeneity’ in an investment driven industry and, without targeted education and training, neither<br />

individuals nor organisations will be competitive. The plantation industry is unlikely to enter into<br />

contractual arrangements if there is a real or perceived risk to business pr<strong>of</strong>its. Historically, the private<br />

sector has also been reluctant to reduce this risk by investing in training and development, tending to<br />

leave it to the public sector, although this is changing (Timber 2020). The Integrated Tree Cropping<br />

example indicates that forest certification may play an increasingly important role in building the<br />

capacity <strong>of</strong> Aboriginal people to be employed in plantation forestry.<br />

Forest-based economic development<br />

Arts and crafts and forest-based ecotourism: Outside <strong>of</strong> main stream employment, native forests <strong>of</strong>fer<br />

a broad range <strong>of</strong> potential economic opportunities at the local level. Small-scale enterprises such as<br />

arts and crafts and cultural tourism are <strong>of</strong>ten attractive to Aboriginal communities because they can be<br />

managed locally, do not require high levels <strong>of</strong> technology and provide local jobs. Most importantly,<br />

they are a mechanism for using customary knowledge and law for deriving an income in the market<br />

economy. In some remote areas it is one <strong>of</strong> the few ways for Indigenous people to gain some income


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 429<br />

from activity on their lands while maintaining social and cultural practices (Hall, 2007). However<br />

pressures <strong>of</strong> the market can undermine customary systems, as research into the use <strong>of</strong> the tree Bombax<br />

for making wooden artefacts has demonstrated (Koenig, et al 2005).<br />

Agr<strong>of</strong>orestry for the commercial production <strong>of</strong> bushfoods is another enterprise with potential for local<br />

Aboriginal communities and there has been recent research into the economic viability <strong>of</strong> small-scale<br />

plant harvest <strong>of</strong> a range <strong>of</strong> food species (Whitehead et. al., 2006). Non-wood forest products from<br />

<strong>Australia</strong>n forests, particularly bush foods and medicines are gaining popularity in <strong>Australia</strong> and<br />

overseas, and a considerable amount <strong>of</strong> research and development is directed towards the genetic<br />

development, domestication and finding sustainable levels <strong>of</strong> wild harvest <strong>of</strong> a wide range <strong>of</strong> native<br />

species (see www.rirdc.gov.au for examples). However, issues still surround the intellectual property<br />

rights <strong>of</strong> Aboriginal <strong>Australia</strong>ns (Janke, 1998) and although community based bush food enterprises<br />

seem an ideal, currently the industry is dominated by non-Indigenous companies.<br />

Cultural heritage management – managing and protecting sites – is an area where traditional<br />

knowledge forms the basis <strong>of</strong> viable economic ventures in Crown forests. All states have legislation<br />

for protecting Aboriginal heritage, together with policies requiring Aboriginal consultation. These<br />

have provided a springboard for employment and have greatly enhanced input by Aboriginal people<br />

into management <strong>of</strong> their cultural heritage. For example, Forest NSW engages the local Aboriginal<br />

Land Council to undertake heritage surveys prior to logging and the work has become a significant<br />

source <strong>of</strong> income for the Land Council.<br />

Western Arnhem Land Fire Abatement project: Research in west Arnhem Land has shown that<br />

strategic fire management in savannah landscapes can reduce greenhouse gas emissions from<br />

savannah fires. This research has led to a greenhouse gas <strong>of</strong>fset agreement between<br />

ConocoPhillips, the NT Government, and local Aboriginal groups. Indigenous fire managers<br />

will be paid around $1 million a year for 17 years to provide this fire management service.<br />

The project seeks to increase the proportion <strong>of</strong> controlled early dry season fires to create fire<br />

breaks and patchy mosaics <strong>of</strong> burnt and unburnt country to minimise the occurrence <strong>of</strong><br />

destructive late dry season wildfires, thereby reducing greenhouse gas emissions (NAILSMA<br />

2006).<br />

UNDERSTANDING ENGAGEMENT<br />

Analysis <strong>of</strong> case study data has demonstrated that engagement between the forest sector and<br />

Indigenous people took the form <strong>of</strong> individual and community based employment, various forms <strong>of</strong><br />

partnerships and agreements in economic and non-economic capacities, and a wide range <strong>of</strong><br />

consultative and participatory processes. The type <strong>of</strong> engagement was shown to be affected by the<br />

element <strong>of</strong> the forest sector under consideration, although this was tempered by physical setting and its<br />

associated historical legacies.<br />

Indigenous cultural factors were always present at the forest sector-Indigenous interface, but took<br />

many forms. Sometimes culture was expressed by the tangible and intangible values <strong>of</strong> the forest<br />

itself. Sometimes it was about the means for Aboriginal elders to pass on customary knowledge and<br />

for the next generation to practice and renew culture. In ‘settled’ <strong>Australia</strong>, many Aboriginal people<br />

wish to both participate in mainstream society and maintain cultural traditions. The Aboriginal owned<br />

land base in settled <strong>Australia</strong> is small, and naturally vegetated lands are mostly owned by the Crown.<br />

Uncertainty over native title and other processes affording Indigenous rights and interests over Crown<br />

native forests is a pragmatic reason for both state and federal governments to establish effective<br />

mechanisms for negotiating outcomes.<br />

Characterizing Aboriginal engagement with the forest sector prompts reflection about whether the<br />

stamp <strong>of</strong> ‘indigeneity’ can be disentangled from the influences <strong>of</strong> social and economic disadvantage.<br />

In his analysis <strong>of</strong> the relationship between identity and culture in the pastoral industry, Richard Davis<br />

notes that Aboriginal stockmen sometimes were distinctively Aboriginal while at other times their<br />

work was indistinguishable from that done by white stock workers and managers. He suggests further<br />

that Aboriginal stockmen also fashioned their own identity by positioning themselves in both worlds<br />

(Davis 2005).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 430<br />

An examination <strong>of</strong> Aboriginal engagement with the market components <strong>of</strong> the forest sector has<br />

demonstrated some parallels with the model developed by Davis for Aboriginal participation in the<br />

pastoral industry. Thus certain components <strong>of</strong> the forest sector, such as working in sawmills makes an<br />

Aboriginal person indistinguishable from a non-Aboriginal person, notwithstanding barriers arising<br />

from persistence <strong>of</strong> negative stereotypes and poor levels <strong>of</strong> education and training. Voices from case<br />

study research suggest that in these situations, ‘work is work’. That is, no special dispensation was<br />

given or sought for indigeneity and paid work was separated from social and cultural activities. Thus it<br />

is the similarities between cultures that take precedence, rather than the differences. For many parts <strong>of</strong><br />

<strong>Australia</strong> it is possibly little different from when Aboriginal people worked in local sawmills fifty<br />

years ago.<br />

In other examples, Aboriginal people undertake activities in Aboriginal ways. The forest heritage<br />

business, ecotourism enterprises and arts and crafts all rely on customary knowledge, and their<br />

expressions are distinctively Aboriginal. However, all operate within a modern market economy<br />

where Aboriginal people can be vulnerable to market pressure, leading to ecologically unsustainable<br />

practices, unequal partnerships and a possible concomitant reduction in the all important ‘Aboriginal<br />

flavour’.<br />

For remote areas such as Cape York, commerce and culture are intertwined in ways not easily<br />

explained by extant economic models. Nanum Tawap’s sawmill operation is firmly embedded in a<br />

cultural landscape but is separated from it physically and administratively. There are special<br />

adaptations such as the type <strong>of</strong> saw and use <strong>of</strong> white goods as work incentives, but essentially it<br />

operates as a mainstream workplace. Application <strong>of</strong> cultural practices occurs through traditional<br />

owners who are empowered through their positions on the Board <strong>of</strong> the company, where the important<br />

decisions are made. Culture is still evident in the management <strong>of</strong> sawmilling operations which is based<br />

on customary social boundaries and a desire to build social capital rather than maximise pr<strong>of</strong>it.<br />

Case studies dealing with commercial native forestry showed that sustainable forest management was<br />

focused largely on consultation processes and protection <strong>of</strong> tangible cultural heritage. Case studies <strong>of</strong><br />

conservation demonstrated that in addition to these aspects, a broader rights and social justice agenda<br />

operated, leading to active participation in land management and ownership <strong>of</strong> land, although with<br />

legislative constraints on customary activities such as hunting and gathering. Plantations and<br />

agr<strong>of</strong>orestry have the potential to <strong>of</strong>fer employment, business and revenue in the mainstream as was<br />

shown in the WA case study. But they can also contribute to ‘caring for country’ by repairing previous<br />

land damage or by providing food and raw materials for Aboriginal communities such as on Cape<br />

York.<br />

CONCLUSION<br />

This paper has presented the results <strong>of</strong> research that has demonstrated that ‘Aboriginal forestry’ is a<br />

spectrum <strong>of</strong> activities and values, ranging from purely economic to purely social. These reflect local<br />

responses by Indigenous people for balancing desires for maintaining cultural values and traditions<br />

with aspirations for social and economic development. Mainstream forest based employment and<br />

business development with no special recognition <strong>of</strong> indigeneity was found to be a preferred approach<br />

in several <strong>of</strong> the case studies. This is relevant for plantation forestry, which could capitalise on<br />

Aboriginal people’s support for plantations by developing policy to underpin effective and equal<br />

partnerships for growing trees that delivered economic, environmental and cultural benefits on<br />

Aboriginal land. In other cases, associations with forests were embedded in cultural responsibilities to<br />

care for the ancestral beings.<br />

As with all natural resource use by a nation state, the management and use <strong>of</strong> <strong>Australia</strong>’s forests is<br />

embedded in broader agenda <strong>of</strong> social justice for indigenous peoples. Experiences from mining have<br />

shown that royalties in return for resource exploitation do not overcome disadvantage on their own.<br />

Overcoming Indigenous social and economic disadvantage requires linkages between programmes for<br />

employment and business development and the broader social goals <strong>of</strong> health, education,<br />

empowerment and self-determination. Government policy must be pluralistic by having policies that<br />

allow Aboriginal people choices about the relative importance they want to place on the economic,<br />

cultural and environmental values <strong>of</strong> forests.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 431<br />

Ultimately, forests can be landscapes <strong>of</strong> reconciliation, through recognising and respecting cultural<br />

‘difference’: the sentient values <strong>of</strong> the forests to Aboriginal people, the cultural places they contain<br />

and their importance for maintaining cultural identity. Forests can also be landscapes for recognising<br />

‘sameness’ and overcoming discrimination between Aboriginal and non-Aboriginal workers in the<br />

mainstream market economy.<br />

REFERENCES<br />

Altman, J., 2001. Indigenous communities and business:Three perspectives, 1998-2000, Working Paper No.<br />

9/2001, Centre for Aboriginal Economic Policy Research, ANU.<br />

Annandale, M., McGavin, R. and Venn, T., 2002. Injinoo Sawmilling Project Phase 1: Small-scale Timber<br />

Processing., Queensland Forestry Research <strong>Institute</strong>.<br />

Annandale, M. and Feary, S., 2009. Consulting with Aboriginal and Torres Strait Islander communities:<br />

Agr<strong>of</strong>orestry in North Queensland, Canberra, Rural Industries Research and Development Corporation<br />

Publication 08-168.<br />

Annandale, M. and Taylor, D., 2007. Forest Futures - Indigenous Timber and Forestry Enterprises on Cape<br />

York., In Forestry for Indigenous Peoples. Proceedings <strong>of</strong> Technical Session 130, XXII IUFRO World<br />

Congress, 8-13 August 2005, Brisbane, <strong>Australia</strong>., (Ed.), Feary, S., ANU Fenner School <strong>of</strong> Environment<br />

and Society, Canberra, pp. 49-59.<br />

<strong>Australia</strong>n Government, 2005. The National Indigenous Forestry Strategy, Canberra, Commonwealth <strong>of</strong><br />

<strong>Australia</strong>.<br />

BDO Consulting (SA) Pty Ltd, 2004. Opportunities and Barriers for Greater Indigenous Involvement in<br />

<strong>Australia</strong>'s Forestry Industry, Commonwealth <strong>of</strong> <strong>Australia</strong>, Canberra.<br />

Behrendt, L., 2003. Achieving Social Justice, The Federation Press, Sydney.<br />

Black, A., Baines, P. and Bekle, H., 2003. Towards a Framework for Consultation and for Input <strong>of</strong> Indigenous<br />

Knowledge Pertinent to Ecologically Sustainable Forest Management at a Regional Level., Victoria,<br />

Forest and Wood Products Research and Development Corporation.<br />

Brooks, K., Kelson, S. and Tottenham, R., 2001. Summary <strong>of</strong> <strong>Australia</strong>n Regional Forest Agreement Social<br />

Assessments and a Recommended Assessment Methodology, Canberra, Forest Industry Branch, AFFA.<br />

Cane, S., 1990. Trees on the Land, Canberra, Report to the Resource Assessment Commission Forest and<br />

Timber Inquiry.<br />

Cornell, S. and Begay, M., 2003. What is Cultural Match and Why is it so Important? Lessons from 14 years <strong>of</strong><br />

the Harvard Project, Building Effective Indigenous Governance, Jabiru, NT. Available at:<br />

http://www.nt.gov.au/cdsa/indigenous_conference/web/html/papers.html (accessed 18/4/2005).<br />

Dargavel, J., 1995. Fashioning <strong>Australia</strong>'s Forests, Oxford University Press <strong>Australia</strong>, Melbourne.<br />

Davis, R., 2005. Identity and Economy in Aboriginal pastoralism, In The Power <strong>of</strong> Knowledge. The Resonance<br />

<strong>of</strong> Tradition, (Eds.), Taylor, L., Ward, G., Henderson, G., Davis, R. and Wallis, A., Aboriginal Studies<br />

Press, Canberra, pp. 49-60.<br />

Dovers, S., 2003. Are forests different as a policy challenge?, In Towards Forest Sustainability, (Eds.),<br />

Lindenmayer, D. and Franklin, J., CSIRO publishing, Collingwood, pp. 15-30.<br />

Dyck, N., 1985. Aboriginal Peoples and Nation States: An Introduction to the Analytical Issues, In Indigenous<br />

Peoples and the Nation State: Fourth World Politics in Canada, <strong>Australia</strong> and Norway., (Ed.), Dyck, N.,<br />

Canada, pp. 1-26.<br />

Feary, S., 2007. Chainsaw Dreaming: Indigenous <strong>Australia</strong>ns and the Forest Sector. PhD, School <strong>of</strong> Resources,<br />

Environment and Society, The <strong>Australia</strong>n National University: Canberra.<br />

Forest Stewardship Council, 2007. FSC <strong>Australia</strong> Update, Available at: www.fscaustralia.org/whats-new/fscaustralia-update/<br />

(accessed 7/10/2007).<br />

Hall, N., 2007. Telling the story: performance and narrative as reflective frameworks for Indigenous tourism,<br />

Sydney, CAUTHE Conference2007: Past Achievements, Future Challenges, 11-14 February, 2007,<br />

Higman, S., Mayers, J., Bass, S., Judd, N. and Nussbaum, R., 2005. The Sustainable Forestry Handbook. A<br />

practical guide for tropical forest managers on implementing new standards, Earthscan, London.<br />

Holling, C., Berkes, F. and Folke, C., 1998. Science, sustainability and resource management, In Linking social<br />

and ecological systems for resilience and stability, (Eds.), Berkes, F. and Folke, C., Cambridge University<br />

Press, Cambridge, pp. 342-362.<br />

Janke, T., 1998. Our culture our future: report on <strong>Australia</strong>n indigenous cultural and intellectual property<br />

rights, Canberra, <strong>Australia</strong>n <strong>Institute</strong> <strong>of</strong> Aboriginal and Torres Strait Islander Studies.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 432<br />

Koenig, J., Altman, J. and Griffiths, A. D., 2005. 'Too Many Trees' : Aboriginal Woodcarvers in <strong>Australia</strong>, In<br />

Carving out a Future, (Eds.), Cunningham, A., Campbell, B. and Belcher, B., Earthscan, pp. 136-146.<br />

Montreal Process Implememtation Group, 2008. <strong>Australia</strong>’s State <strong>of</strong> the Forests Report 2008. Bureau <strong>of</strong> Rural<br />

Sceinces: Canberra.<br />

Morgan, M., Strelein, L. and Weir, J., 2006. Authority, Knowledge and Values: Indigenous Nations Engagement<br />

in the Management <strong>of</strong> Natural Resources in the Murray-Darling Basin, In Settling with Indigenous<br />

People, (Eds.), Tehan, M., Palmer, L., Langton, M. and Mazel, O., The Federation Press, Sydney, pp.<br />

135-158.<br />

North <strong>Australia</strong> Indigenous Land and Sea Management Alliance (NAILSMA), 2006.<br />

http://www.nailsma.org.au/projects/indigenous_carbon_abatement.html (accessed 23/7/2009).<br />

Rose, D., 1996. Nourishing Terrains: <strong>Australia</strong>n Aboriginal Views <strong>of</strong> Landscape and Wilderness, <strong>Australia</strong>n<br />

Heritage Commission, Canberra, ACT.<br />

Rose, D., 2002. Love and Reconciliation in the Forest: A Study in Decolonisation., Magill, SA, Hawke <strong>Institute</strong><br />

Working Paper Series No. 19, Hawke <strong>Institute</strong>, University <strong>of</strong> South <strong>Australia</strong>.<br />

Sullivan, P., 2004. Indigenous Fragmentation, Minority Rights Scepticism and the Need for Both Re-integration<br />

and Co-integration, Canberra, Paper given at the AIATSIS Conference, Canberra, 22-25 November 2004.<br />

Timber2020 2009 Indigenous Forestry Great Southern Region, Western <strong>Australia</strong>. Strategy and Action Plan,<br />

Govt <strong>of</strong> Western <strong>Australia</strong> and Timber 2020.<br />

Whitehead, P., Gorman, J., Griffiths, A., Wightman, G., Massarella, H. and Altman, J., 2006. Feasibility <strong>of</strong><br />

Small Scale Commercial Native Plant Harvests by Indigenous Communities., Canberra, RIRDC<br />

Publication No. 04/149, Joint Venture Agr<strong>of</strong>orestry Program, RIRDC.<br />

Wunder, S., 2005. Payments for environmental services: some nuts and bolts, Jakarta, CIFOR Occasional Paper<br />

No. 42, Center for International Forestry Research.<br />

Varmola, M., 2005. Diversifying Functions <strong>of</strong> Planted Forests, In Forests in the Global Balance - Changing<br />

paradigms,(Eds.), Mery, G., Alfaro, R., Kanninen, M. and Lobovikov, M., IUFRO World Series Volume<br />

17, pp. 117-136.<br />

Young, E., 2001. Looking after country is men's and women's business: institutional support for Indigenous land<br />

management, Dialogue, 20(2): 28-32.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 433<br />

COMMUNITY CONSULTATION AND INDIGENOUS <strong>FORESTRY</strong><br />

OPPORTUNITIES IN CAPE YORK PENINSULA<br />

ABSTRACT<br />

Mark Annandale 1<br />

This paper explores community consultation with Indigenous people through the<br />

development <strong>of</strong> agr<strong>of</strong>orestry projects in north Queensland. Indigenous people have been<br />

practicing various forms <strong>of</strong> agr<strong>of</strong>orestry in <strong>Australia</strong> for thousands <strong>of</strong> years, as custodians<br />

<strong>of</strong> the natural environment. Traditional land management, <strong>of</strong>ten called ‘caring for<br />

country’, was practiced through strategic use <strong>of</strong> fire to encourage growth <strong>of</strong> certain<br />

species. Some species were planted and others were selectively cultivated to favour<br />

specific species or faunal habitats. Elements <strong>of</strong> these traditional systems <strong>of</strong> land<br />

management are contained in the modern practice <strong>of</strong> agr<strong>of</strong>orestry, yet there are few<br />

examples <strong>of</strong> commercial agr<strong>of</strong>orestry enterprises involving Indigenous <strong>Australia</strong>ns today.<br />

The case study resulted from an extensive consultation process that developed and tested a<br />

methodology for effective community engagement. This paper provides a review <strong>of</strong><br />

Queensland sandalwood and discusses options for growing sandalwood species to provide<br />

sustainable development opportunities in Cape York Peninsula (CYP) in far north<br />

Queensland. Strategies for further development <strong>of</strong> the sandalwood industry in CYP are<br />

discussed.<br />

INTRODUCTION<br />

The Sandalwood species that occurs naturally in Cape York Peninsula CYP is Santalum lanceolatum<br />

(R.Br), commonly known in northern <strong>Australia</strong> as Queensland sandalwood. Taxonomists have<br />

recently separated this species from the sandalwood which occurs south <strong>of</strong> 20 degrees latitude with<br />

suggested common names <strong>of</strong> northern and southern sandalwood respectively (Harbaugh, 2007). The<br />

first recorded sighting and collection <strong>of</strong> the plant by Joseph Banks and Daniel Solander occurred in<br />

1770 at the Endeavour River in south eastern CYP (Wharton, 2005). From around 1896 to the early<br />

1930s sandalwood cutters relied on Aboriginal knowledge <strong>of</strong> country and for labour to exploit the<br />

species in north Queensland (Pike, 1983; Wharton, 1985; Statham, 1990).<br />

Sandalwood is one <strong>of</strong> the most valuable timbers in the world (Baruah, 1999; Bragg et al., 2004). The<br />

high price commanded by sandalwood is a result <strong>of</strong> the increased commercial demand and<br />

international trade in sandalwood timber and its associated products, compounded by a decline in<br />

supply. New sandalwood plantations are required to address the overexploitation <strong>of</strong> the native forest<br />

resources experienced in most countries and to meet demand for sandalwood products. In the CYP<br />

min<strong>of</strong>orestry, the integration <strong>of</strong> forestry production systems into mine rehabilitated landscapes,<br />

provides an opportunity for Indigenous business, employment and a range <strong>of</strong> environmental benefits.<br />

The CYP is a diverse and ecologically important region <strong>of</strong> tropical <strong>Australia</strong>. It covers approximately<br />

13.72 M ha. The vegetation is diverse and includes extensive areas <strong>of</strong> dry eucalypt woodland<br />

dominated by Darwin Stringybark (Eucalyptus tetradonta), Melville Island Bloodwood (Corymbia<br />

nesophylla) and to a lesser extent Cooktown Ironwood (Erythrophleum chlorostachys), occurring<br />

mostly on deeply weathered plateaus (Neldner & Clarkson 1996). Areas <strong>of</strong> rainforest occur along the<br />

eastern coast, mostly in national parks, with extensive tropical grasslands, heathlands and mangroves<br />

are also present throughout the region.<br />

The current population <strong>of</strong> CYP is approximately 18,000 people, over half <strong>of</strong> whom are <strong>of</strong> Aboriginal<br />

and Torres Strait Islander origin and live in a number <strong>of</strong> Indigenous communities. Most are relatively<br />

small and isolated settlements which are self-governing, run by an elected council made up <strong>of</strong> local<br />

residents and funded by government, with the largest Indigenous community comprising about 1200<br />

1 Director. Environment Land Heritage Pty Ltd and PhD candidate at School <strong>of</strong> Natural and Rural Systems Management.<br />

University <strong>of</strong> Queensland St Lucia Qld, 4067 mark.annandale@landheritage.com.au


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 434<br />

people. The two regional centres include Cooktown, which serves as an administrative centre, and<br />

Weipa which is a large mining community, each comprising about 3000 people, including many<br />

Indigenous people from the region. There are well-recognised opportunities for Indigenous<br />

communities in CYP to develop, manage and operate viable forestry business enterprises based on<br />

sustainable management <strong>of</strong> native forests and plantations (Annandale, 2000; Annandale and Taylor,<br />

2000; Hopewell, 2001; Annandale et al., 2002; Annandale, 2003; Annandale and Bragg, 2004; Venn,<br />

2004). Several studies recognise the potential for commercial forest activities associated with pre- and<br />

post- mining land uses around the western CYP communities, including plantation-grown sandalwood<br />

and enterprises that can process timber cleared prior to mining activity (Annandale, 2003; Annandale<br />

and Bragg, 2004).<br />

Forestry-based business can play a lead role in regional development because <strong>of</strong> its potential to<br />

provide resources for value-adding and employment and encourage business. It will help foster<br />

regional economic development, including forestry-based businesses through development <strong>of</strong><br />

infrastructure, introduction <strong>of</strong> new technology and employment supported by training initiatives.<br />

Agr<strong>of</strong>orestry is the practice <strong>of</strong> combining trees, shrubs and forests with agricultural systems.<br />

Indigenous people in north Queensland have expressed a desire to develop sustainable commercial<br />

agr<strong>of</strong>orestry enterprises to provide local employment, enable people to work on country and to<br />

improve health and wellbeing.<br />

There is increasing recognition that the social justice agenda for Indigenous <strong>Australia</strong>ns incorporates<br />

the right to derive economic benefit from land-based enterprises and the expanding Indigenous-owned<br />

land base <strong>of</strong>fers opportunities for various agr<strong>of</strong>orestry enterprises.<br />

COMMUNITY CONSULTATION<br />

All non-Aboriginal people intending to work with Indigenous communities should undertake<br />

comprehensive induction and cultural awareness programs prior to entering communities. They must<br />

develop the interpersonal skills to freely exchange information and transfer skills that Indigenous<br />

people require to give a community greater autonomy. Additionally, skills and knowledge cannot be<br />

passed on without a basic level <strong>of</strong> understanding <strong>of</strong> the culture and society <strong>of</strong> community members.<br />

When working with Indigenous communities, one must first make steps towards understanding some<br />

<strong>of</strong> the background issues which have impacted on the communities and individuals in both recent and<br />

past times. Be aware that attitudes will be influenced by consultation processes undertaken in previous<br />

years, some <strong>of</strong> which may not have been adequate to meet the needs and aspirations <strong>of</strong> the community.<br />

The building <strong>of</strong> meaningful relationships is critical for the successful delivery <strong>of</strong> any project. This can<br />

<strong>of</strong>ten take a considerable amount <strong>of</strong> time, especially when many Indigenous people are involved. Lack<br />

<strong>of</strong> time to build relationships and a mutual understanding is the main reason that cross-cultural<br />

projects fail, so it is important to factor sufficient time into work programs. Once established, projects<br />

and enterprises evolve and mature. There is a vast amount <strong>of</strong> existing knowledge and experience in<br />

Indigenous communities, with people who have insights into what types <strong>of</strong> projects work. For<br />

successful project development there needs to be an exchange <strong>of</strong> information and experiences, which<br />

occurs over a long period <strong>of</strong> time.<br />

When discussing project ideas or opportunities both parties must be committed, have confidence in the<br />

process, be open, honest and provide reliable information. The project must be discussed at length and<br />

clearly outlined so that expectations remain realistic. Indigenous people are then in a better position to<br />

decide whether or not they wish to participate in projects.<br />

A holistic approach is required providing the communities with the tools to freely determine their own<br />

economic and social development, to empower people to control their destiny. The criteria for<br />

measuring the success <strong>of</strong> any business need to be identified and agreed to by all those involved. It<br />

should not be determined by government or any other external stakeholders. This will keep all<br />

expectations at a realistic level.<br />

Education and training needs to be adaptive and responsive to community needs. People need the<br />

opportunity to have inputs into curricula, including incorporation <strong>of</strong> cultural knowledge.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 435<br />

Community consultation does not have to be complicated, but it does need to be inclusive and follow<br />

some basic protocols. Protocol means following the customs or lores <strong>of</strong> the people with whom you are<br />

working.<br />

Effective consultation needs to be designed to meet the unique aspects <strong>of</strong> the situation and to identify<br />

and clearly define the issues. Enough time must be allocated for the consultation process and detailed,<br />

balanced and accurate information must be provided to enable informed discussion and negotiation.<br />

Once a proposed activity requiring involvement <strong>of</strong> Indigenous people has been identified, a<br />

consultation protocol should be followed to ensure that the appropriate Indigenous people are<br />

consulted in mutually agreed ways. The consultation process should begin as early as possible and be<br />

flexible.<br />

CASE STUDY<br />

Indigenous people in far north Queensland are interested in pursuing socially, culturally,<br />

environmentally and economically sustainable development on their own land. They have identified<br />

the need to become more self-sufficient and to engage with mainstream society on an equal footing.<br />

Indigenous people identify a common desire to return to country. They want to get back to country<br />

with a purpose, which includes educating younger people in traditional ways, looking after country<br />

and/or being able to make a living compatible with the lifestyle <strong>of</strong> the people involved. Therefore<br />

projects that involve people getting back onto country are attractive and <strong>of</strong>ten provide an opportunity<br />

for younger people to get to know their country in greater detail. Elders in several Cape York<br />

communities have talked about the time when many people lived on country, in outstations, as a time<br />

when many <strong>of</strong> the social problems, and dependency on welfare did not exist. There is a desire to find a<br />

balance between living in towns or communities and living on country. Agr<strong>of</strong>orestry business<br />

development may provide this opportunity.<br />

Currently, sustainable land use options available to Indigenous communities are limited. New<br />

industries involving some form <strong>of</strong> agr<strong>of</strong>orestry could <strong>of</strong>fer opportunities for ecologically sustainable<br />

development and provide long-term business and employment.<br />

People intending to work with Indigenous people should be adequately prepared. This may involve<br />

background research on the local Indigenous history and culture, and developing an awareness <strong>of</strong> local<br />

Indigenous political and governance structures.<br />

Consultation or other participatory processes are an essential component <strong>of</strong> any project. Inappropriate<br />

or inadequate consultation can cause a project to fail and result in ill feeling which jeopardises future<br />

projects.<br />

Sustainability is an essential part <strong>of</strong> any community development, achieved through a combination <strong>of</strong><br />

scientific and traditional knowledge. However, pressures <strong>of</strong> the market economy need to consider<br />

customary laws concerning resource use. To ensure that agr<strong>of</strong>orestry projects are addressing cultural<br />

and environmental sustainability, appropriate monitoring and evaluation mechanisms must be in place.<br />

These should include feedback from Traditional Owners on the impacts <strong>of</strong> the project on the natural<br />

environment and on cultural integrity.<br />

All issues associated with Indigenous Intellectual and Cultural Property Rights need to be<br />

acknowledged and respected prior to project implementation. To ensure appropriate and agreed use <strong>of</strong><br />

Indigenous peoples’ cultural knowledge and traditions.<br />

THE SANDALWOOD OPPORTUNITY<br />

The harvesting <strong>of</strong> native sandalwood in <strong>Australia</strong> has long been an important industry, with <strong>Australia</strong><br />

being one <strong>of</strong> the world’s two largest producers <strong>of</strong> native sandalwood (with India)(Statham, 1990;<br />

Radomiljac et al., 1999; Vernes and Robson, 2002; Bragg et al., 2004), supplying a global market<br />

estimated at 3000 tones per annum. Queensland and Western <strong>Australia</strong>n sandalwood, is exported to<br />

South-East Asia, primarily Taiwan, where it is powdered and mixed with various resins and other<br />

aromatics to make incense sticks (Jones, 2001). In 1988 <strong>Australia</strong> supplied 94% <strong>of</strong> the global incense<br />

stick sandalwood market (Statham, 1990).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 436<br />

Of more than 30 sandalwood species globally, about seven are economically important. Of the four<br />

Santalum species that occur in <strong>Australia</strong>, S. lanceolatum is the most widespread, occurring across<br />

northern <strong>Australia</strong> (Applegate et al., 1990b; Harbaugh, 2007), but Santalum album (L), Indian<br />

sandalwood, and Santalum spicatum (R.Br), Western <strong>Australia</strong>n sandalwood, are currently the two<br />

most commercially valuable species (Bragg et al., 2004).<br />

A recent study <strong>of</strong> natural sandalwood populations in CYP has identified several with outstanding oil<br />

qualities, and have high potential for domestication.<br />

In part because <strong>of</strong> its high-value, the exploitation <strong>of</strong> sandalwoods around the world has led to some<br />

uncertainty about the sustainability <strong>of</strong> the natural resources <strong>of</strong> most Santalum species. An attractive<br />

alternative or complement to a sandalwood industry based on <strong>Australia</strong>n native forest is the<br />

establishment <strong>of</strong> plantations which can be managed to provide a sustained yield (Applegate et al.,<br />

1990a; Baruah, 1999; Vernes and Robson, 2002).<br />

Sandalwoods have specific ecological and silvicultural requirements. They are all hemi-obligate root<br />

parasites, with inefficient root systems that extract a range <strong>of</strong> nutrients from the root system <strong>of</strong> their<br />

hosts. The oil content in the heartwood <strong>of</strong> sandalwood varies across their distribution (Baruah, 1999).<br />

Sandalwood products and markets<br />

The scented heartwood <strong>of</strong> sandalwood is a valuable commodity used historically by Indigenous people<br />

all over the world. In traditional medicine, sandalwood oil is used for a wide variety <strong>of</strong> conditions<br />

ranging from an antiseptic and astringent to the treatment <strong>of</strong> headache, stomach-ache, and urogenital<br />

disorders. The oil, with its distinctive and woody aroma, is a highly sought after fragrance.. It is also<br />

valuable for its binding qualities with other fragrances to create a characteristic bouquet for a branded<br />

perfume. In volume terms, its use in the perfumery industry far outweighs its medicinal usage.<br />

Sandalwood oil is produced commercially by steam distillation. Variation in oil content is attributable<br />

to tree age, site and position <strong>of</strong> heartwood within the tree.<br />

The heartwood is close-grained, finely and evenly textured, hard durable and renowned as a carving<br />

material. The timber seasons well when dried slowly. The wood can be worked to a smooth finish and<br />

takes on a satin-like polish. The heartwood was used for many centuries for carvings, prayer poles and<br />

other religious artifacts, valuable handicrafts, fuel for funeral pyres, c<strong>of</strong>fins and incense sticks<br />

(Statham, 1990; Baruah, 1999).<br />

Due to the high value <strong>of</strong> sandalwood species, illegal harvesting has become a management problem in<br />

some countries, to the extent that in some areas natural stands <strong>of</strong> sandalwood are threatened.<br />

Queensland sandalwood background<br />

The native forest sandalwood industry commenced in Queensland in the 1890s when trees were<br />

harvested from the northern part <strong>of</strong> CYP, and later near Normanton and the Mitchell River in the<br />

earlier parts <strong>of</strong> last century (Applegate et al., 1990a; Statham, 1990; Bristow et al., 2000). Reports<br />

dating back as early as 1902 detail sandalwood being exported to China from Somerset, at the top <strong>of</strong><br />

CYP (Statham, 1990). There is information referring to tonnes <strong>of</strong> sandalwood being cut and exported<br />

from CYP, adjacent to several Aboriginal missions in the late 19 th century (Wharton, 1985, 2005).<br />

Sandalwood cutters relied on Aboriginal labour and knowledge <strong>of</strong> country to exploit the species. The<br />

Aboriginal labourers located the trees, cut <strong>of</strong>f the bark and sapwood and helped transport the billets <strong>of</strong><br />

heartwood to the coast for export (Wharton, 1985).<br />

The sandalwood cutters first harvested sandalwood from the east coast <strong>of</strong> CYP, using the services <strong>of</strong><br />

up to two hundred Aboriginal people, including women and children, to dig up the valuable stumps<br />

and roots. When that resource became scarce, the sandalwood cutters spread to other areas <strong>of</strong> CYP. In<br />

the Aurukun area over 100 aborigines were employed to cut sandalwood and paid in tobacco and basic<br />

food rations, exporting 300 tonnes up to 1920 (Pike, 1983). The sandalwood industry in Queensland<br />

declined in the 1930s and ceased following the outbreak <strong>of</strong> civil war in China in 1946, and the<br />

collapse <strong>of</strong> the sandalwood market (Kealley, 1989). Some cutting resumed from 1983 (Applegate et<br />

al., 1990b; Statham, 1990).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 437<br />

Northern sandalwood is commonly harvested by pulling the entire tree and stump from the ground,<br />

because a large amount <strong>of</strong> the oil (in some species most <strong>of</strong> the oil) and a higher proportion <strong>of</strong> fragrant<br />

heartwood are found in the root ball (Doran et al., 2005).<br />

The natural regeneration <strong>of</strong> sandalwood species occurs through seed dispersal and vegetatively,<br />

including root suckering. In <strong>Australia</strong>, the critical factors for sandalwood establishment, growth and<br />

production appear to be the regularity <strong>of</strong> rainfall, the length <strong>of</strong> the growing season, grazing <strong>of</strong> the<br />

palatable leaves by animals, fire, and the regularity <strong>of</strong> frosts (Bristow et al., 2000; Bragg et al., 2004).<br />

Community development opportunities in the Cape York Peninsula: A case study<br />

Historically, the rate <strong>of</strong> commercial development in Cape York Peninsula has been slow, with most<br />

development activity occurring after World War II. Extensive cattle grazing and mining, together with<br />

some forestry operations, have been the main enterprises. A steady improvement in infrastructure<br />

(roads, transport, communication, towns) has accompanied the expansion <strong>of</strong> the mining, cattle and<br />

tourism sectors. Commercial forestry activities have been small-scale and low intensity harvesting <strong>of</strong><br />

native forest (Annandale and Crevatin, 2002).<br />

Commercial sandalwood plantings present an opportunity for the Indigenous communities <strong>of</strong> CYP to<br />

participate in sustainable business development. The unique ecology <strong>of</strong> the sandalwood, coupled with<br />

the high value <strong>of</strong> sandalwood products, allow for versatile silvicultural systems which in CYP could<br />

focus on sandalwood products and associated value-adding to create long-term business opportunities<br />

and alternative employment opportunities. In contrast, the plantation sandalwood industry in the Ord<br />

River region <strong>of</strong> WA, has adopted a highly mechanised, large-scale industrial forestry approach,.<br />

Specialised sandalwood plantations which target niche and high-value markets have the potential to<br />

meet both community expectations and high-value market opportunities. Sandalwood plantings in<br />

CYP could be designed to suit the ecological and socioeconomic conditions <strong>of</strong> the area through an<br />

adaptive management approach, including enrichment and min<strong>of</strong>orestry planting systems, utilising<br />

flexible labour systems that could include flexi-time, part-time, or seasonal work..<br />

Historically people in remote regions, are subject to decisions made by people <strong>of</strong>ten far removed from<br />

where the decisions have an impact, which are generally influenced by broader commercial directives<br />

and affected by government policies. These decisions may not suit the biophysical environments <strong>of</strong><br />

CYP; they don’t acknowledge the level <strong>of</strong> traditional forest knowledge resident in the communities<br />

and are not necessarily aligned with everyday life in this remote region. Industrial-scale forestry in<br />

<strong>Australia</strong> commonly relies on external investment together with mainstream economic systems and<br />

timetables that influence plantation management decisions. Also, plantation forestry companies<br />

already growing sandalwood typically require substantial financial and technological investments<br />

currently not available in CYP.<br />

Small-scale sandalwood planting systems provide an opportunity to maximise returns through<br />

integration <strong>of</strong> multiple goals to achieve both commercial and non-commercial benefits, tailored to suit<br />

other land uses, various land tenures, including freehold, community owned land and bauxite mine<br />

leases <strong>of</strong> the western CYP. Commercial benefits could be complemented by environmental and socioeconomic<br />

benefits. Such developments could be small in scale both spatially (e.g. planting, tending<br />

and harvesting small areas each year) and in terms <strong>of</strong> labour demands (e.g. providing jobs for small<br />

teams on a seasonal basis). In this way, sandalwood plantations could <strong>of</strong>fer opportunities for the<br />

Indigenous communities and individuals to determine scale, management regimes and other factors<br />

required for sustainable development in the socio-cultural environment <strong>of</strong> CYP.<br />

Strategies for further development <strong>of</strong> the sandalwood industry in CYP<br />

In Western CYP around 600,000 hectares <strong>of</strong> land is covered by bauxite mine leases near the mining<br />

town <strong>of</strong> Weipa, and around the Indigenous communities <strong>of</strong> the Mapoon, Napranum and Aurukun.<br />

Mining in this region has taken place for 50 years, and it is conservatively estimated that the bauxite<br />

resource will last for at least another 40 years (Rio Tinto Alcan newsletter 1, 2008). Bauxite mining<br />

involves clearing <strong>of</strong> vegetation, removal <strong>of</strong> the upper 0.3-1 m <strong>of</strong> soil, digging up <strong>of</strong> the bauxite<br />

through open-cut mining and progressive mine site rehabilitation. Mine rehabilitation can be<br />

developed to include plantation sandalwood, complemented by other commercial plantings, species


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 438<br />

for conservation outcomes and value-adding enterprises (Annandale 2003; Annandale and Bragg<br />

2004). On the completion <strong>of</strong> mining the land will be rehabilitated and eventually returned to the<br />

Traditional Owners.<br />

There is questionable value in regenerating native vegetation for marginal environmental benefit, as<br />

this provides no economic, cultural or recreational services to the community (Annandale, 2003) nor<br />

provides a long-term and sustainable business opportunity when the bauxite resource has been mined.<br />

A range <strong>of</strong> min<strong>of</strong>orestry options have developed (Bragg et al., 2004), including sandalwood<br />

establishment and management in either plantations or enrichment planting designs. Early enrichment<br />

planting trials in Weipa have established sandalwood seedlings in either native or mine rehabilitated<br />

forests to enrich the value <strong>of</strong> the forest. Presently both designs are being trialed on a small scale, but<br />

opportunities exist for broadacre systems. This integrated management approach, if developed to<br />

appropriate scale and commercialized, will increase the long-term sustainability <strong>of</strong> communities.<br />

Min<strong>of</strong>orestry — an integrated approach<br />

Min<strong>of</strong>orestry, the integration <strong>of</strong> forestry production systems into rehabilitated mine landscapes,<br />

provides an opportunity to include a commercial return on rehabilitation investments (miners have to<br />

pay for rehabilitation as part <strong>of</strong> the licence to operate), such as Indigenous business employment and a<br />

range <strong>of</strong> environmental benefits. There are many design options for min<strong>of</strong>orestry, some <strong>of</strong> which have<br />

demonstrated potential for mine rehabilitation and some commercial returns. These include:<br />

integration <strong>of</strong> seed <strong>of</strong> high-value species into standard rehabilitation seed mixes; enrichment planting<br />

into established mine rehabilitation areas; block plantings <strong>of</strong> selected commercial species; plantations<br />

as windbreaks and shelter for stock in pasture areas; and other inclusion <strong>of</strong> native species for food,<br />

fibre and medicinal purposes. In addition to economic and environmental benefits, some species may<br />

provide community health benefits i.e., traditional foods.<br />

DISCUSSION<br />

Consultation with Indigenous people invariably involves consultation, <strong>of</strong>ten involving large numbers<br />

<strong>of</strong> people from widely dispersed communities. The way consultation is undertaken will determine the<br />

success or otherwise <strong>of</strong> the process and ultimately the project itself. A large number <strong>of</strong> matters must<br />

be taken into consideration, some <strong>of</strong> which apply to all consultations and some that are unique to<br />

circumstances where the Indigenous people live.<br />

Sandalwood growing, harvesting, and research have a long history in north Queensland. Studies<br />

undertaken by Traditional Owners, James Cook University researchers and collaborators such as<br />

Harbaugh (2007) have provided evidence that sandalwood in CYP is able to meet international<br />

standards for sandalwood oil and compete with Indian sandalwood on the world market. The outcomes<br />

<strong>of</strong> this work provide support for further development <strong>of</strong> sandalwood plantations in the region through<br />

the selection <strong>of</strong> superior trees for the development <strong>of</strong> cultivars for plantation development.<br />

Experimental forestry plantations established over the last 40 years at Weipa have identified forest<br />

species suitable for mined land. Plantation on mined land can be developed to establish a viable<br />

forestry industry and associated value-adding business opportunities to create long-term and<br />

alternative business and employment opportunities for the Indigenous communities <strong>of</strong> the Western<br />

CYP. Specialised sandalwood plantations that target niche and high-value markets have the potential<br />

to meet both community expectations and high value market opportunity.<br />

In the CYP Indigenous community involvement in forestry is evolving. Community aspirations and<br />

expectations for mined land acceptable for relinquishment back to Traditional Owners include the<br />

need for industry diversification to support Indigenous business and employment opportunities when<br />

mining operations have ceased.<br />

Opportunities for growing sandalwood on Cape York Peninsula include plantation and forest<br />

management, min<strong>of</strong>orestry, or mixed agr<strong>of</strong>orestry systems on community lands. Conservation<br />

outcomes are further supported through growing sandalwood now only found in isolated and small<br />

populations, including preservation <strong>of</strong> in-situ trees.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 439<br />

ACKNOWLEDGMENTS<br />

This paper draws on a variety <strong>of</strong> work with Alan Bragg and Mila Bristow when they and the author<br />

worked with the Queensland Forestry Research <strong>Institute</strong> in north Queensland. and on work with Sue<br />

Feary <strong>of</strong> the <strong>Australia</strong>n National University on Indigenous community consultation, The inputs <strong>of</strong><br />

Alan, Mila and Sue were instrumental in developing some <strong>of</strong> the concepts discussed in this paper and<br />

are gratefully acknowledged.<br />

REFERENCES<br />

Annandale, M., 2000. Management and viability <strong>of</strong> forestry plantations at Weipa: Weipa Forestry Plantation<br />

Assessment. In. Queensland Forestry Research <strong>Institute</strong>, Department <strong>of</strong> Primary Industries, Atherton.<br />

Annandale, M., 2003. Economic development opportunities for indigenous communities: provided by<br />

regeneration <strong>of</strong> mined lands in the Western Cape York Peninsula. In, Land relinquishment criteria<br />

workshop, Weipa, Queensland.<br />

Annandale, M., Bragg, A., 2004. Min<strong>of</strong>orestry development at Weipa. In. Department <strong>of</strong> State Development and<br />

Innovation, Cairns, p. 31.<br />

Annandale, M., McGavin, R.L., Venn, T.J., 2002. Injinoo sawmill project phase 1:small-scale timber processing.<br />

In. Queensland Forestry Research <strong>Institute</strong>, Department <strong>of</strong> Primary Industries, Atherton.<br />

Annandale, M., Taylor, D., 2000. Developing a sustainable forest management system for the forests <strong>of</strong> Cape<br />

York Peninsula. In. Queensland Forestry Research <strong>Institute</strong>, Department <strong>of</strong> Primary Industries, Atherton.<br />

Applegate, G.B., Chamberlain, J., Daruhi, G., Feigelson, J.L., Hamilton, L., McKinnell, F.H., Neil, P.E., Rai,<br />

S.N., Rodehn, B., Statham, P.C., Stemmermann, L., 1990a. Sandalwood in the Pacific: A State-<strong>of</strong>-<br />

Knowledge Synthesis and Summary form the April 1990 Symposium. In: Hamilton, L., Conrad, C.E.<br />

(Eds.), Symposium on Sandalwood in the Pacific. USDA Forest Service, Honolulu, Hawaii, pp. 1-11.<br />

Baruah, A.D., 1999. The economics <strong>of</strong> sandal-oil. Indian Forester 125, 640-643.<br />

Bootle, K.R., 1983. Wood in <strong>Australia</strong>: types, properties and uses. McGraw-Hill Book Company, Sydney.<br />

Bragg, A., Annandale, M., Wapau, J., 2004. Min<strong>of</strong>orestry - Past, present and future commercial tree plantings at<br />

Weipa. In: Bevege, D.I., Bristow, M., Nikles, D.G., Skelton, D. (Eds.), Prospect for high-value hardwood<br />

timber plantations in the 'dry' tropics <strong>of</strong> northern <strong>Australia</strong>. Published as a CD ROM by Private Forestry<br />

North Queensland, Kairi, Queensland.<br />

Brand, J.E., 1993. Preliminary observations on ecotypic variation in Santalum spicatum 2. Genotypic variation.<br />

Mulga Research Centre Journal, Curtin University 11, 13-19.<br />

Brand, J.E., 1999. Conserving sandalwood (Santalum spicatum) in the rangelands, Western <strong>Australia</strong>. In.<br />

Department <strong>of</strong> Conservation and Land Management, Western <strong>Australia</strong>, pp. 1-4.<br />

Bulai, P., Nataniel, V., 2002. Research, development and extension <strong>of</strong> sandalwood in Fiji: A new beginning. In,<br />

Regional Workshop on Sandalwood Research, Development and Extension in the Pacific Islands and Asia.<br />

In Press, Noumea, New Caledonia.<br />

Bule, L., Daruhi, G., 1990. Status <strong>of</strong> Sandalwood Resources in Vanuatu. In: Hamilton, L., Conrad, C.E. (Eds.),<br />

Symposium on Sandalwood in the Pacific. USDA Forest Service, Honolulu, Hawaii, pp. 79-84.<br />

DPI Forestry, 1995. DPI Forestry yearbook 1994-95. In, Brisbane.Fox, J.E.D., Harbaugh, D.T., 2007. A<br />

taxonomic revision <strong>of</strong> <strong>Australia</strong>n northern sandalwood (Santalum lanceolatum, Santalaceae). <strong>Australia</strong>n<br />

Systematic Botany 20, 409-416. Harbaugh, D.T., Baldwin, B.G., 2007. Phylogeny and biogeography <strong>of</strong> the<br />

sandalwoods (Santalum, Santalaceae): Repeated dispersals throughout the Pacific. American Journal <strong>of</strong><br />

Botany 94, 1028-1040.<br />

Hopewell, G., 2001. Characteristics, utilisation and potential markets for Cape York Peninsula timbers. In.<br />

Forest Products program, Queensland Forestry Research <strong>Institute</strong>, Department <strong>of</strong> Primary Industries,<br />

Brisbane.<br />

Kealley, I.G., 1991. The management <strong>of</strong> sandalwood. In. Department <strong>of</strong> Conservation and Land Management,<br />

Perth.<br />

Pike, G., 1983. The Last Frontier. Pinevale Publications, Mareeba.<br />

Radomiljac, A.M., 1998. The influence <strong>of</strong> pot host species, seedling age and supplementary nursery nutrition on<br />

Santalum album Linn. (Indian sandalwood) plantation establishment within the Ord River irrigation Area,<br />

Western <strong>Australia</strong>. Forest Ecology and Management 102, 193-201.<br />

Rio Tinto Alcan newsletter 1, 2008. South <strong>of</strong> the Embley Project. Rio Tinto Alcan Brisbane <strong>Australia</strong>, June<br />

2008).


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 440<br />

Statham, P.C., 1990. The sandalwood industry in <strong>Australia</strong>: a history. In: Hamilton, L., Conrad, C.E. (Eds.),<br />

Symposium on Sandalwood in the Pacific. USDA Forest Service, Honolulu, Hawaii, pp. 26-38.Venn, T.J.,<br />

2004. Visions and realities for a Wik forestry industry on Cape York Peninsula, <strong>Australia</strong>. Small-scale<br />

Forest Economics, Management and Policy 3, 431-451.<br />

Vernes, T., Robson, K., 2002. Indian sandalwood industry in <strong>Australia</strong>. In, Sandalwood Research Newsletter, pp.<br />

1-4.<br />

Wharton, G., 1985. Antiquarians and sandalwood-getters: the establishment <strong>of</strong> the Cape York collection at<br />

Weipa. In, North <strong>Australia</strong>n Mine Rehabilitation Workshop No. 9, Weipa, Queensland.<br />

Wharton, G., 2005. Northern Sandalwood (Santalum lanceolatum) on Cape York Peninsula. A report on<br />

historical sources for the Queensland Department <strong>of</strong> State Development and Innovation. Prepared by Ge<strong>of</strong>f<br />

Wharton, Infotracker Historical and Information Research Services October 2005. Brisbane.<br />

Wilson, G. 2005. Traditional Owner Kanju elder Personal Communication.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 441<br />

A FORESTER’S GUIDE TO WORKING WITH<br />

INDIGENOUS AUSTRALIANS:<br />

AN INTRODUCTION<br />

Peter Shepherd 1<br />

ABSTRACT<br />

Indigenous issues have gained in importance in recent years. Forest managers are now<br />

required to work more closely with traditional land owners with the risk <strong>of</strong> possible<br />

litigation, long delays and high legal expenses if proposals are not supported by the<br />

local Indigenous community. <strong>Foresters</strong> will benefit from the broad introduction to<br />

Indigenous culture presented in this paper. Topics discussed include approach to land<br />

ownership, society structure, indigenous language groups and significance <strong>of</strong> flora to<br />

Indigenous people. The risk <strong>of</strong> treating the Indigenous community as a homogeneous<br />

society is recognised. Hence this paper is presented as a general guide and introduction<br />

to alert foresters about some topics they may need to consider when managing forests<br />

on land that is significant to Indigenous communities.<br />

INTRODUCTION<br />

Many foresters must work with Indigenous people who have an interest in the land; be they owners,<br />

traditional owners, neighbours or interested people. Over the last few decades, the need to work with<br />

Indigenous <strong>Australia</strong>ns has increased as their stake holding in land has increased.<br />

As Indigenous <strong>Australia</strong>ns are now expecting greater influence over decisions for land in which they<br />

are stakeholders, foresters need to be able to form positive relationships with Indigenous individuals<br />

and communities. This can only happen with respect, understanding and sensitivity. This paper will<br />

introduce some key information about Indigenous culture at a very basic level in order to help increase<br />

the general understanding <strong>of</strong> how to work constructively with Indigenous people.<br />

However, I am writing this paper with some trepidation since it is wrong to suggest Indigenous<br />

<strong>Australia</strong>ns are a homogenous population. Hence generalisations are dangerous. Further, I am not<br />

claiming to be an expert in Aboriginal communication and culture. This paper is presented as a very<br />

general and basic guide to inform and help those who are unfamiliar with working with Indigenous<br />

people to avoid mistakes and misunderstandings.<br />

I apologise in advance to any Indigenous person who feels I have made sweeping generalisations that<br />

do not apply to their culture.<br />

INDIGENOUS NATION STRUCTURE<br />

The Indigenous Nation is in effect a conglomerate <strong>of</strong> smaller cultural groups, like most countries.<br />

Language, culture and geography define these groups. It was assumed originally that there were over<br />

700 language groups, <strong>of</strong> which 250 have been recorded and over 200 different languages have been<br />

mapped (see http://www.abc.net.au/indigenous/map/ for an interactive map and Diagram 1). Extensive<br />

and strong relationships between language groups exist with respect for each “country” where the<br />

traditional owners are based. Respect is shown by acknowledging the owners <strong>of</strong> the land, the Elders<br />

and their ancestors, early in an interaction, event or entry to the territory.<br />

Each group has distinct cultural aspects which distinguish it from all others.<br />

1 Dr. Peter Shepherd, Manager, Business Development, Batchelor <strong>Institute</strong> <strong>of</strong> Indigenous Tertiary Education,<br />

ph 0419309281 (m) (08) 89397273; email: peter.shepherd@batchelor.edu.au


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 442<br />

Diagram 1: Map <strong>of</strong> Indigenous Language Groups<br />

(source: Tindale – revised 1974)<br />

LAND OWNERSHIP<br />

Indigenous people do not view land ownership in the same way that western cultures do. The concepts<br />

<strong>of</strong> freehold and land title are alien. In remote communities in north and central <strong>Australia</strong>, housing is<br />

provided by government (much <strong>of</strong> it well below standard). Many Indigenous people in their traditional<br />

setting do not understand how any person can own land or hold “title” to a house; hence there is no<br />

concept <strong>of</strong> mortgage over land or property. Urban Indigenous people in larger towns and cities do<br />

have a better understanding <strong>of</strong> home “ownership” but in many cases live in government-provided<br />

housing.<br />

Indigenous people have a very different relationship with land. Many believe they are culturally<br />

responsible to their “country”, which is part <strong>of</strong> their Dreaming, and they must act as custodians for that<br />

land. This is a spiritual obligation to their forefathers and traditional leaders.<br />

KINSHIP AND FAMILY<br />

Indigenous family and kinship relations are complex and difficult for untutored westerners to<br />

understand. I will start with family and work to the community. Sometimes a few related new mothers<br />

combine to raise a group <strong>of</strong> youngsters. Hence it is possible for an indigenous person to claim 2, 3 or<br />

more “mothers” all treated with similar respect.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 443<br />

An Aboriginal person will belong to an inherited skin group which is descended from a particular<br />

ancestral creator spirit. The skin group can be inherited from a person’s mother or father depending on<br />

the clan. There are very strict rules based on skin groups. The skin group has a strong ruling on a<br />

person’s selection <strong>of</strong> a marriage partner, land they will inherit and access to resources.<br />

Marriage between people <strong>of</strong> the same skin group is strictly forbidden. The women elders <strong>of</strong> a clan<br />

carefully considered all marriage options and arranged marriages to ensure correct skin group lineage.<br />

Each skin group has at least one opposing group with whom they cannot make contact,; especially<br />

people <strong>of</strong> the opposite sex. This includes every day contact such as talking, eye contact and even<br />

passing items to a forbidden person. It is understood these rules are in place to avoid relationships that<br />

are genetically close, for example, to forbid first, second or third cousins marrying.<br />

Because <strong>of</strong> the complex spiritual beliefs and the intertwining <strong>of</strong> clans through marriage, the estate<br />

boundaries would not always have been clearly definable as boundary lines on a map (such as<br />

Diagram 1), but none-the-less are clearly perceived and understood by individuals knowing<br />

geographic markers on their traditional land and passing this on to younger clan members as part <strong>of</strong><br />

Indigenous knowledge.. Through intermarriage and other alliances people were able to access land and<br />

resources far beyond their own estates. Access to land and resources were negotiated through<br />

discussion, marriage, ceremony and adherence to Aboriginal lore.<br />

Clans were united in alliances based on a shared language, spirituality, seasonal hunting and gathering<br />

events and lore. These alliances are called language-culture groups.<br />

It needs to be remembered that a young indigenous child will be exposed to a number <strong>of</strong> languages.<br />

The child’s mother will speak in many cases a few <strong>of</strong> the common dialects in the area. The child’s<br />

father will speak the same languages and several others needed for cultural interaction purposes. Some<br />

speak 7 languages or more to a high level <strong>of</strong> pr<strong>of</strong>iciency. English may well be the 4 th (or more)<br />

language spoken in their house, only at school, and rarely in the community.<br />

DREAMING AND TIME CONCEPT<br />

Western people <strong>of</strong>ten equate dream time to the creation, or some time in the past. This is true in one<br />

sense, but a mistake in that it is not limited to the past. Dreaming is also today. Dreaming is infinite<br />

and, for Aboriginal people, links the past and the present to the future. Dreaming can explain how<br />

landscapes and environments were shaped. But an Indigenous person will also have a “Dreaming”; a<br />

special inheritance and link to the land. This may be represented in his/her totem. Dreaming stories are<br />

the truth from the past which becomes the code <strong>of</strong> law for today. Often the story is only a little part <strong>of</strong><br />

a much greater story which provides insight into behaviour and future being. Some dreaming stories<br />

are very sacred when told in their country and can only be told in special places under specific<br />

circumstances.<br />

In most stories <strong>of</strong> the Dreaming, the Ancestor Spirits came to the earth in human form and as they<br />

moved through the land, they created the animals, plants, rocks and other forms <strong>of</strong> the land. They also<br />

created the relationships between groups and individuals to the land, the animals and other people.<br />

Once the ancestor spirits had made the world, they did not disappear but changed into trees, the stars,<br />

rocks, watering holes or other objects. These are the sacred places and organisms <strong>of</strong> Aboriginal culture<br />

and have special properties. Because the ancestors did not disappear at the end <strong>of</strong> the Dreaming, but<br />

remained in these sacred sites, the Dreaming is never-ending, linking the past and the present, the<br />

people and the land. Hence an Indigenous person has his/her “Dreaming” which is their relationship<br />

to the land and ancestors and a special creation story <strong>of</strong>ten through a plant species, landscape<br />

formation or animal.<br />

I have a very special boomerang made by an aboriginal elder who made a special effort to put his<br />

Dreaming on to it. It is a special item <strong>of</strong> great significance to me. It only takes one look to feel and<br />

know the difference between this and mass produced tourist items.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 444<br />

The protocols for social behavior and consequences, including punishments and disciplines learnt, are<br />

evident in Dreaming stories. Virtue in Aboriginal religion lies in the obligation to follow ancestral<br />

precedent, which involves keeping the Dreaming stories alive. This takes the forms <strong>of</strong> painting, song,<br />

dancing or ceremony - all <strong>of</strong> which are therefore necessarily inextricably linked. This is part <strong>of</strong> a living<br />

tradition based on ritual practices. Traditions and practices also merge with economic and ecological<br />

responsibilities for 'looking after country'. Looking after country means to continue to express these<br />

ritual forms <strong>of</strong> the Dreaming. Clan groups have the right to use the land regarded as their 'territory,<br />

estate or country, and any <strong>of</strong> its products, based on their duties to tend the land through the<br />

performance <strong>of</strong> ceremonies.<br />

Smith (2005, pp.4-5) outlines a mistake made by some Park Managers who were explaining the<br />

origins <strong>of</strong> fruit bats in North Queensland to an Aboriginal community when they suggested they had<br />

migrated from a colony further away. The local Indigenous people were amused because local<br />

knowledge pertaining to flying foxes was that they came from the mouth <strong>of</strong> the Rainbow Serpent, a<br />

powerful creator that was still present and active in the region’s landscape. Smith’s account is valuable<br />

because it discusses how difficult it was for park management pr<strong>of</strong>essionals to accommodate<br />

Indigenous knowledge into their thinking without being patronising. The result is a reluctance for<br />

Aboriginals to share and demonstrate a public dissent with non-aboriginal perspectives. Smith<br />

continues to outline how constructive relationships with local Aboriginals around land management<br />

can be created and is recommended reading for those considering working in this area.<br />

It only takes a moment to ponder the vast differences between the spiritual and ancestral values and<br />

the Western system <strong>of</strong> valuing land and forest products.<br />

TOTEMS<br />

A totem is an object or thing in nature that is adopted as a family or clan emblem. Different clans are<br />

assigned different totems and in some cases individuals are given personal totems at birth. In some<br />

areas personal pendants are worn and these pendants are mostly carved out <strong>of</strong> wood, turtle shell or<br />

shells and <strong>of</strong>ten represent the person's totem. There are well established rules as to when the pendants<br />

can be worn, <strong>of</strong>ten only allowed during ceremonies or rituals.<br />

Some Aboriginal people can be identified by their totems, which can be birds, reptiles (like turtles),<br />

plants, trees, sharks, crocodiles and fish. They are an important part <strong>of</strong> their cultural identity and are<br />

especially significant in song, dance and music as names and on cultural implements. Some clans<br />

forbid their individuals from eating the animal that is their totem, while other tribes make exceptions<br />

for special occasions such as ceremonies. People can collect seeds, fruits and other forest products<br />

from trees and plants as they have been provided by the Ancestor to the Clan. But the special Totem<br />

Trees cannot be harvested as they are the Dreaming <strong>of</strong> ancestors.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 445<br />

SORRY TIME AND FESTIVALS<br />

“Sorry Time” is the mourning period after a person has died. This can last up to a month and it is<br />

mandatory for clan members to be present. Employees will not be able to attend work for this period.<br />

Further, few meetings will be attended or decisions made. People will travel huge distances to attend a<br />

funeral. The dead person’s name cannot be mentioned in the community.<br />

There is usually an annual festival in the clan’s social calendar. This is generally to celebrate the end<br />

<strong>of</strong> a period <strong>of</strong> difficult weather, so it is usually held at the end <strong>of</strong> the Wet Season in the north or in<br />

spring in the south. The celebrations can last for days or weeks. Most Indigenous people will want to<br />

attend part or all <strong>of</strong> their community’s festival to meet clan and family obligations.<br />

LEARNING SYSTEMS<br />

As with many Pacific Island cultures, <strong>Australia</strong>n Aborigines are a “talking” culture. The need to take<br />

time to talk with visitors, to get to know them and establish a relationship before they will have<br />

meaningful talks is paramount.<br />

Indigenous learning is obtained from narrative leading to understanding and insight. Learning and<br />

understanding come from stories, songs and skills acquired in the field under the supervision <strong>of</strong> a more<br />

senior clan member. Learning only takes place within the context <strong>of</strong> a narrative, preferably by doing or<br />

observing a related fact or skill from the real world, Dreaming and Elder. Smith (2005) gives a good<br />

insight into how Indigenous knowledge is passed on.<br />

I cringe when western thinking managers and even teachers believe that they can further their<br />

objectives by preparing a PowerPoint based on our western experience and learning paradigm. Classic<br />

PowerPoint presentations are sure to quickly loose the interest <strong>of</strong> an Indigenous audience, fail to<br />

achieve the outcomes and do a disservice to the reputation <strong>of</strong> the person who delivers it in the eyes <strong>of</strong><br />

the clan.<br />

LEADERSHIP<br />

Finally a word about leadership. Defining leadership is a vexed task in both the business literature and<br />

reality <strong>of</strong> pr<strong>of</strong>essional life. I will leave discussions about definitions <strong>of</strong> leadership to others. The focus<br />

here is on the different approaches to leadership between Western and Indigenous cultures so that<br />

foresters can be alert to some traps that await them when working with Aboriginal people.<br />

Firstly, an Elder is an old and respected member <strong>of</strong> a clan. This position has been obtained by the fact<br />

<strong>of</strong> being a senior member. Their opinions are important. A leader is usually an Elder but does not<br />

necessarily have to be. Usually a Leader has a significant leadership role requested by the senior<br />

members <strong>of</strong> the clan and has a significant relationship with the clan such as being a traditional owner.<br />

Indigenous leadership is not about having a high pr<strong>of</strong>ile and being seen at the front. Working with an<br />

Elders’ Council may be seen as a slow and cumbersome process lacking in focus by a person<br />

expecting a western decision making experience.<br />

The leader may not be the person greeting you or chairing the meeting. The most influential people<br />

may be the Elder women sitting under a tree at the back <strong>of</strong> the meeting listening quietly and observing<br />

you carefully.<br />

Some country is Women’s Country. If this is the case, women will have a different position in<br />

discussions than usual. It is an advantage to have a female as part <strong>of</strong> the delegation who can meet with<br />

the women. However, consult the Elders first to ensure you are not making a mistake.<br />

Indigenous people are shy, charming, witty, charismatic and delightful to work with but they take time<br />

to develop relationships and respect for an outsider. Attempt a quick, in/out meeting with great caution<br />

as it would be better to spend more time, looking, listening while becoming familiar with the people<br />

and context.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 446<br />

Smith (2005, pp: 9-11) describes the work <strong>of</strong> David Claudie who compared Indigenous thinking (he<br />

called it “round thinking”) to western thinking (“square thinking”). Claudie describes bureaucrats who<br />

can’t listen and perceive themselves as “flash” and better than Aboriginal people. This may not be the<br />

case but is how they are perceived. A number <strong>of</strong> causes <strong>of</strong> this misunderstanding <strong>of</strong> people are <strong>of</strong>fered<br />

including a misalignment between Western and Indigenous people’s “way <strong>of</strong> knowing”.<br />

Noel Pearson (2009, pp: 251-253) speaks <strong>of</strong> a leadership continuum made up with realism and<br />

idealism on opposite ends <strong>of</strong> a two dimensional plane. Indigenous leadership models are at the<br />

idealism end while Western leadership is more pragmatic. Pearson suggests those who harbour ideals<br />

need to work with those who have real power to form a compromise.<br />

“It takes insight, skill and creativity, careful calculation as well as bold judgement, prudence<br />

as well as risk, perseverance as well as preparedness to alter course, belief as well as humility<br />

and great competence as well as ability to make good from mistakes to bring ideals closer to<br />

reality” (Pearson 2009, p. 252).<br />

Clearly the instruction to foresters here is to prepare well, gain insight, think carefully, listen and be<br />

flexible when dealing with Indigenous leaders and communities.<br />

CONCLUSION<br />

Working with other cultures is not rocket science. To be successful just takes a little time and thought.<br />

The web has vast resources available. A search for “working with Aboriginal Communities” will<br />

result in a vast collection <strong>of</strong> resource and guide books as well as bibliographies.<br />

It will take much more than this brief overview <strong>of</strong> some parts <strong>of</strong> Indigenous culture to ensure<br />

successful interactions for people who are not familiar with working with other cultures. There are a<br />

couple <strong>of</strong> key behaviours that need to be adopted that will help. Firstly listen and do more listening<br />

than talking. It is so important to listen to individuals, communities and read the press. If possible,<br />

attend community visits with someone who is experienced, known and respected by the people.<br />

Observe carefully how this person behaves. Develop expectations that are realistic for the Indigenous<br />

decision making model and not a western expectation. And finally think about what you have read and<br />

observed to create a deep behaviour model. Understanding will come with work and effort.<br />

Failure to do the work will result in problems.<br />

ACKNOWLEDGEMENTS<br />

Thank you to Patrick Anderson from Batchelor <strong>Institute</strong> for pro<strong>of</strong> reading this paper and providing to<br />

me many deep and insightful explanations <strong>of</strong> the mysteries <strong>of</strong> Indigenous culture.<br />

REFERENCES<br />

Pearson, N. (2009). “Up from the Mission”. 400 pp. Black Inc.<br />

Smith, B. R. (2005). “We got our own management”: local knowledge, government and development in Cape<br />

York Peninsula. <strong>Australia</strong>n Aboriginal Studies 2005/2 pp:4-15.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 447<br />

AN INTEGRATED APPROACH TO SPECIES REGIME SELECTION<br />

IN PNG: BEGINNING WITH LOCAL PEOPLE’S NEEDS<br />

ABSTRACT<br />

Braden Jenkin 1 , Michael Blyth 2 , Peter Kanowski 3 and Don Yakuma 4<br />

Selection and development <strong>of</strong> a tree species regime to deliver benefits to small-scale treegrowers<br />

must be driven by the needs <strong>of</strong> the individuals, families and local communities<br />

involved. A species regime must consider all elements <strong>of</strong> the production system from the<br />

tree species requirements through each stage <strong>of</strong> the supply chain to the point <strong>of</strong> sale. As<br />

part <strong>of</strong> an <strong>Australia</strong>n Centre for International Agricultural Research project (ACIAR<br />

FST/2004/050), a systematic approach to planted woodlot and individual tree species<br />

regime choice is being developed to assist local communities in three pilot locations<br />

(Madang area, Madang Province; Ramu Valley, Morobe Province; the Fly River corridor,<br />

Western Province) in Papua New Guinea. A parallel component <strong>of</strong> the project seeks to<br />

better understand the decision making process <strong>of</strong> why people would need or wish to plant<br />

trees.<br />

The key variables <strong>of</strong> a species regime selection framework are the physical and financial<br />

input requirements, the timing <strong>of</strong> returns, and inherent and exogenous risk and uncertainty<br />

factors. Individuals’ preferences for the timing <strong>of</strong> returns may be linked to significant<br />

household events. In some cases, a species regime may provide intergenerational returns<br />

with trees treated as a bank account. Other regimes may provide continuous cash flow<br />

after a startup period, with a final lump sum (e.g. rubber trees). A species regime that<br />

provides additional non-financial benefits may be <strong>of</strong> interest (e.g. nut production as part <strong>of</strong><br />

a food security).<br />

The ACIAR project has developed protocols to consider market and processing needs, to<br />

assist focus on least-risk options that are feasible under current conditions. While other<br />

options may become attractive in the future, a risk minimisation approach suggest such<br />

options should be regarded as the potential icing on a likely cake.<br />

INTRODUCTION<br />

There is an increasing need in Papua New Guinea (PNG) to investigate, develop and implement tree<br />

growing regimes which assist rural communities to generate income from planted trees and forests, in<br />

part to relieve pressures on natural forests and in part to enhance landowners’ livelihoods. Everywhere<br />

in PNG, tree growing and management <strong>of</strong> trees are incorporated into both traditional and modern<br />

farming systems. However, because there has been little incentive to focus on commercially-valuable<br />

forestry species, such species have seldom been adopted. Where a critical mass <strong>of</strong> resource can be<br />

established, commercial tree-growing appears a good prospect for landowners with limited incomegeneration<br />

alternatives (ACIAR, 2009). Smallholders’ (in the PNG context, usually described as<br />

‘landowners’) resource allocation decisions to bridge these two issues require support and guidance in<br />

a number <strong>of</strong> critical areas, in particular species choice.<br />

Kanowski et al., (2007) conducted a pilot assessment (the Scoping Study for ACIAR Project<br />

C2005/050) <strong>of</strong> the potential to value add to current PNG agr<strong>of</strong>orestry systems and facilitate<br />

smallholder uptake <strong>of</strong> high-value tree species. This project identified suitable candidate regions and<br />

1<br />

Managing Director, Sylva Systems Pty Ltd (contact point), PO Box 1175, Warragul, Vic, 3820 braden@latrobe.net.au.<br />

2<br />

Director, Four Scenes Pty Ltd., PO Box 50, Kippax, ACT 2615.<br />

3<br />

Pr<strong>of</strong>essor, The Fenner School <strong>of</strong> Environment & Society ANU Forestry Building 48, Linnaeus Way, The <strong>Australia</strong>n<br />

National University, Canberra ACT 0200, <strong>Australia</strong>.<br />

4<br />

Forestry Program Coordinator, Ok Tedi Development Foundation, Ok Tedi Mining Limited, P.O. Box 1, Tabubil, WP,<br />

Papua New Guinea.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 448<br />

partners, together with preliminary tree species and production systems (ACIAR, 2009), and resulted<br />

in a subsequent project (ACIAR project FST/2004/050). A multidisciplinary team was formed to<br />

research the adoption <strong>of</strong> commercial-scale high-value tree growing in PNG, developed through a<br />

relationship fostered between landowners and selected business partners (ACIAR, 2009), and to apply<br />

the results to benefit smallholders (see Kanowski et al., 2008).<br />

A systematic approach to tree species selection has evolved to assist local communities in three pilot<br />

locations (Madang area, Madang Province; Ramu Valley, Morobe Province; the Fly River corridor,<br />

Western Province). A parallel component <strong>of</strong> the project is to develop an understanding <strong>of</strong> smallholder<br />

motivations and decision making and is the subject <strong>of</strong> ACIAR-sponsored PhD research (see Mulung,<br />

2007, p.26).<br />

This paper describes the development <strong>of</strong> a decision-making framework to assist individual<br />

landholders, communities and their advisors through the critical step towards successful tree species<br />

selection for commercial production.<br />

KEY CONCEPTS<br />

Smallholder needs: An appropriate species regime must satisfy the needs <strong>of</strong> those growing the trees,<br />

and hence this project includes a foundation in social research. Chambers (2007, p.38-41) considered<br />

that the dimensions <strong>of</strong>ten omitted in social research are:<br />

• Tropical seasonality: disease, ground hardness, lack <strong>of</strong> food and the impact on ability to<br />

work;<br />

• Places <strong>of</strong> the poor: people can be isolated and lack access to infrastructure;<br />

• Poverty <strong>of</strong> time and energy: an imbalance between the desire to work and lack <strong>of</strong> work<br />

options;<br />

• The body: <strong>of</strong>ten the only resource available, which may be weakened by poor health and<br />

nutrition.<br />

The analysis will focus on the needs <strong>of</strong> the local people while considering these points <strong>of</strong> caution.<br />

High-value: High-value is a philosophical adjective, unless qualified by a statement <strong>of</strong> who is the<br />

beneficiary. In some cases, what is high-value to one party may be at the expense <strong>of</strong> another party; or<br />

there may be little connection between value and return to growers: e.g. does an increase in high-value<br />

exports contributing to gross domestic product necessarily result in higher pr<strong>of</strong>its for producers?<br />

Non-standard economics: Non-standard economics looks beyond simply financial costs and returns.<br />

Economic analyses relevant to PNG landowners’ decisions should include measures such as utility and<br />

returns on labour (both <strong>of</strong> which are standard economic parameters) to fully assess livelihood assets<br />

deployment options and outcomes.<br />

SMALLHOLDER AGRO<strong>FORESTRY</strong><br />

‘In most countries in Southeast Asia, government driven reforestation (under both traditional and socalled<br />

community-based reforestation approaches) have had limited success’ (Bertomeu et al., 2008,<br />

p.27). By contrast, agr<strong>of</strong>orestry systems implemented by smallholder farmers can re-establish tree<br />

cover, provide environmental services and enhance local livelihoods (Bertomeu et al., 2008, p.27).<br />

Agr<strong>of</strong>orestry is a dynamic, ecologically based natural resource management system that, through the<br />

integration <strong>of</strong> trees on farms and in the agricultural landscape, diversifies and sustains production for<br />

increased social, economic, and environmental benefits for land use at all levels (Leakey, 1998).<br />

Smallholder agr<strong>of</strong>orestry systems typically follow less intensive management regimes than industrialscale<br />

plantations. After a site is prepared and trees are planted, ongoing management <strong>of</strong> the tree crop<br />

such as application <strong>of</strong> fertilizer, thinning and pruning trees and weeding during the establishment<br />

phase are less likely to occur on a strict schedule than in an industrial plantation.<br />

However, where silvicultural activities directly benefit the agricultural crops inter-planted with the<br />

trees, they are practised more commonly. While small-scale agr<strong>of</strong>orestry systems can achieve


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 449<br />

pr<strong>of</strong>itability levels comparable to large-scale plantations, such performance levels may negatively<br />

affect the productivity <strong>of</strong> associated agricultural crops (see Tomich et al., 2001; Holding and<br />

Roshetko, 2003). Within an agr<strong>of</strong>orestry system, trade-<strong>of</strong>fs between the productivity <strong>of</strong> timber<br />

production and the productivity <strong>of</strong> the agricultural food crops have to be made, to varying degrees.<br />

Food production or food security is, necessarily, <strong>of</strong>ten the priority <strong>of</strong> smallholders.<br />

Given the situation <strong>of</strong> smallholders, their decisions to invest in tree planting cannot be expected to<br />

parallel those <strong>of</strong> large-scale industrial plantation investors. As Byron (2001, p.212) suggested ‘there<br />

may be fundamental differences in objectives, which inevitably lead to the choice <strong>of</strong> quite different<br />

processes and techniques’. He explained that where an industrial plantation operator may aim to<br />

maximise yield per hectare, a smallholder may choose to maximise the cash return per unit <strong>of</strong> labour.<br />

Understanding the motivation or purpose <strong>of</strong> smallholders to integrate trees into their agricultural<br />

production system is a critical first step in determining successful small-scale agr<strong>of</strong>orestry investment<br />

strategies.<br />

Successful decision-making for agr<strong>of</strong>orestry depends firstly on the purpose for which a landholder<br />

intends to plant trees. If the purpose is achievable, then success is possible. Smallholders plant trees<br />

for a variety <strong>of</strong> purposes, including: timber for various end uses, fuelwood, fruits, fodder, medicines,<br />

resins, shade for livestock, wind protection and for conservation <strong>of</strong> soil and water resources (Holding<br />

and Roshetko, 2003). Timber may be a secondary product, harvested after trees have served their<br />

primary purpose. In some cases, farmers plant trees as a form <strong>of</strong> savings for future family needs such<br />

as education expenses or as an intergenerational investment to help children and grandchildren. Where<br />

the objective is to produce roundwood for a third party on a commercial basis, other issues come into<br />

play, and the process <strong>of</strong> species selection becomes a marketing issue, rather than being based purely<br />

on utility.<br />

PNG SMALLHOLDERS<br />

In PNG, trees have been and continue to be grown and managed as an integral part <strong>of</strong> customary<br />

farming and land use systems. The diversity and dynamism <strong>of</strong> these systems reflect high levels <strong>of</strong><br />

innovation and adaptation in agriculture by land users (Kanowski et al., 2007, p.4). Many land users<br />

have experiences growing c<strong>of</strong>fee, cocoa, coconut, oil palm and a range <strong>of</strong> fruit trees. In some districts,<br />

such as in the Western Province smallholders have many years experience growing rubber. There are<br />

four broad customary agr<strong>of</strong>orestry systems described, based primarily on differences in geographic<br />

factors (Kanowski et al., 2007, p.8&9):<br />

• Highlands agr<strong>of</strong>orestry systems – characterised by sweet potato and casuarina (Casuarina<br />

oligodon);<br />

• Lowlands to mid-montane agr<strong>of</strong>orestry systems – characterised by combinations <strong>of</strong> annual and<br />

perennial food plants and trees including breadfruit (Artocarpus altilis), pandanus, betel nut<br />

palm, sago palm and bananas;<br />

• Coastal agr<strong>of</strong>orestry systems – characterised by fruit and nut tree species including breadfruit<br />

and Canarium species in combination with cassava, sweet potato and bananas;<br />

• Islands agr<strong>of</strong>orestry systems – characterised by fruit and nut tree species including Canarium,<br />

Barringtonia, breadfruit and coconut, in association with numerous species <strong>of</strong> annual and<br />

perennial food crops.<br />

DECISION MAKING<br />

The time lag between planting a tree and yielding a benefit must be considered in the planning: what<br />

by when? This involves a process beginning with the end in mind whether physical, financial or other<br />

outcomes. Formal analytical approaches are available: e.g. Mendoza et al., (1986) presented an<br />

application <strong>of</strong> multiple objective programming as applied to agr<strong>of</strong>orestry as a mechanism to consider<br />

multiple attributes inherent is such systems.<br />

Analysis should include an assessment <strong>of</strong> associated risk and uncertainty. Risk is defined as ‘a state in<br />

which the number <strong>of</strong> possible future events exceeds the number <strong>of</strong> events that will actually occur, and<br />

some measure <strong>of</strong> probability can be attached to them’ (Bannock et al., 1992, p. 374). Uncertainty is<br />

defined as ‘the state in which the number <strong>of</strong> possible outcomes exceeds the number outcomes and


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 450<br />

when no probabilities can be attached to each possible outcome’ (Bannock et al., 1992, p. 431).<br />

Bigsby (2001) presents a series <strong>of</strong> papers considering plantation investment risk, primarily in terms <strong>of</strong><br />

impacts on financial outcomes. Kirby et al., (1993) presented an appraisal <strong>of</strong> agr<strong>of</strong>orestry investment<br />

under uncertainty by applying a range <strong>of</strong> values to cash flow parameters resulting in net present values<br />

(NPV) at set discount rates. This level <strong>of</strong> sophistication is appropriate if the parameters and the likely<br />

range is understood. For some individuals and households in PNG, levels <strong>of</strong> risk and uncertainty may<br />

lead to catastrophic outcomes, if food security strategies are compromised or people cannot buy high<br />

protein / high fat foods deficient in local diets (Bourke and Allen, 2008).<br />

A DECISION MAKING FRAMEWORK<br />

Development <strong>of</strong> an understanding <strong>of</strong> smallholder motivations and decision making related to land use<br />

and tree growing is a primary objective <strong>of</strong> the ACIAR project (ACIAR FST/2004/050; ACIAR, 2009)<br />

under which the work reported in this paper is being undertaken. A species regime must fit within an<br />

individual’s or household’s livelihood framework and with the strategies developed to target<br />

individuals and household goals and aspirations. ‘A livelihood comprises the capabilities, assets<br />

(stores, resources, claims and access) and activities required for a means <strong>of</strong> living: a livelihood is<br />

sustainable which can cope with and recover from stresses and shocks, maintain or enhance its<br />

capabilities and assets, and provide sustainable livelihood opportunities for the next generation; and<br />

which contributes net benefits to other livelihoods at the local and global levels and in the short and<br />

long term’ (Chamber and Conway, 1991, p.6). A long-standing understanding is that Melanesians<br />

living in resource rich landscapes are <strong>of</strong>ten in a state <strong>of</strong> ‘subsistence affluence’. Hence when <strong>of</strong>fered<br />

commercial cropping opportunities to increase individual income and to boost national development,<br />

uptake is minimal or only to a level to meet their vital needs (Kennedy and Clarke, 2004, p.31). If a<br />

minimalist approach is taken, it is important to understand how such targets evolve with individuals’<br />

or families’ life stages, as they influence cash flow needs.<br />

A range <strong>of</strong> guides which assess livelihood strategies may assist in species regime selection (e.g.<br />

Messer and Townsley, 2003). Ashley and Carney (1999, p.47) presented an <strong>of</strong>ten used sustainable<br />

livelihood framework (SLF). It is a dynamic tool that links the deployment <strong>of</strong> social group livelihood<br />

assets to invest in strategies to achieve livelihood outcomes. Livelihood assets are classified as human,<br />

natural, financial, social and physical capital (see Messer and Townsley, 2003, p.8&9). The SLF<br />

defines the resources available and the formation <strong>of</strong> strategies to deploy such scarce resources.<br />

Although dynamic, the framework lacks prioritisation <strong>of</strong> scarce resource allocation and target<br />

outcomes. Individuals’ and families’ livelihood priorities are unlikely to be linear, and could be<br />

described as a series <strong>of</strong> buckets. The first priority is to fill bucket A to 25% capacity to reach a point<br />

considered to be sufficient, after which resources can be directed to buckets B and C, filling each up to<br />

a pre-determined level. Once this has been achieved, resources can be directed back to bucket A and it<br />

is filled to say 50%. The process will continue until the available resources are exploited. The order <strong>of</strong><br />

priority and levels <strong>of</strong> allocation will be an individual decision.<br />

In PNG, 87% <strong>of</strong> the population live outside <strong>of</strong> urban areas. About one million people live in severe<br />

poverty (Bourke and Allen, 2008), which takes a variety <strong>of</strong> forms in different locations (Hanson et al.,<br />

2001). For many, capital assets are limited. The average life expectancy for men in rural PNG is 53<br />

years (Bourke and Allen, 2008). Therefore, tree species regimes should complement the need for food<br />

security within the allocation <strong>of</strong> those livelihood assets necessary to achieve priority outcomes over<br />

one or more generations.<br />

A FRAMEWORK TO GUIDE DECISION MAKING FOR SMALLHOLDER<br />

AGRO<strong>FORESTRY</strong><br />

Several elements need to align or be aligned before successful agr<strong>of</strong>orestry can occur within a given<br />

area. For example, while a particular tree species may grow successfully on a smallholder’s land,<br />

sustained commercial viability within an agr<strong>of</strong>orestry system is complex, involving many factors.<br />

Before planting it is necessary to investigate and analyse market prospects for the products from<br />

particular species and assess the economic, institutional and technical performance <strong>of</strong> the supply chain<br />

including harvesting, transport, primary and secondary processing. The main elements <strong>of</strong> the<br />

agr<strong>of</strong>orestry decision support framework for smallholders, rural communities and their advisors in


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 451<br />

PNG are perceptions, attitudes and motivation <strong>of</strong> smallholders, institutional and legal factors, financial<br />

or economic factors, social and community factors as well as physical and biological factors.<br />

These have been discussed by Kanowski et al., (2007, p.24&25) and others (e.g. see: Mercer, 2004;<br />

Scherr, 2004; Pannell et al., 2006). In addition to these factors uncertainty has been identified as a<br />

particularly important factor in adoption <strong>of</strong> farming practices such as agr<strong>of</strong>orestry which are<br />

associated with land conservation. Consideration <strong>of</strong> uncertainty is critical in agr<strong>of</strong>orestry investment<br />

assessments. An example <strong>of</strong> institutional factors which may constrain commercial tree growing is the<br />

current export ban on roundwood <strong>of</strong> teak (Tectona grandis) (Teak); balsa (Ochroma pyramidale);<br />

Blackbean (Castanospermum australe); Cordia (Cordia dicotoma); Black ebony (Diospyros ferrea);<br />

rosewood (Pterocarpus indicus), and all coniferous species logs (Customs Tariff Act 1990: Schedule 2,<br />

Export Item 44.03).<br />

If, after initial assessment, it appears that each <strong>of</strong> the framework elements is satisfactory, or potentially<br />

so, then a prospective grower can move to the next step on their agr<strong>of</strong>orestry investment pathway. It is<br />

likely that some factors will provide more substantial constraints than others; these are the areas that<br />

require attention before an agr<strong>of</strong>orestry investment is likely to be successful. Some <strong>of</strong> these may be<br />

within the control or influence <strong>of</strong> an individual or community investor, while others may not. For<br />

example, improving inadequate or unreliable infrastructure is beyond the financial capacity <strong>of</strong> an<br />

individual or a community. Prospective growers will make judgments based on the best information<br />

available to them, reflecting their risk aversion and the quality <strong>of</strong> their analysis.<br />

Questions that persons advising smallholders should consider when deciding what tree species to<br />

recommend include (based on Holding and Roshetko, 2003):<br />

• Is there a domestic and/or export market demand for resources and products from planted<br />

forests? What product characteristics do markets or buyers require (e.g. for timber the market<br />

may specify characteristics such as form, length, volume and/or specific wood properties such<br />

as density; certification may be a requirement for some markets)?<br />

• Can harvested forest resources that meet market requirements be transported to processors and<br />

traders from the proposed production site efficiently and reliably? Can processing be<br />

conducted effectively – i.e. able to consistently meet market requirements?<br />

• Will sufficient volume <strong>of</strong> resources be produced in the area to attract and sustain<br />

establishment <strong>of</strong> processing facilities or are there existing processing facilities within<br />

economic reach <strong>of</strong> the growing area that can accommodate increased throughput?<br />

• Can quality germplasm <strong>of</strong> the right species or mix <strong>of</strong> species be sourced for the proposed site<br />

(where species are selected in accordance with growth potential for the site and the potential<br />

to deliver wood properties required by markets)?<br />

• Will the selected mix <strong>of</strong> species allow growers to balance market requirements, household<br />

needs, agricultural crop productivity and sustainability <strong>of</strong> the tree resource?<br />

• Do growers have access to technologies and practices to ensure optimal site productivity and<br />

resource quality?<br />

FACTORS INFLUENCING THE DECISIONS OF PNG LANDOWNERS<br />

Factors influencing adoption <strong>of</strong> agr<strong>of</strong>orestry by smallholders can be grouped under the following<br />

headings.<br />

Growing trees within garden systems: Application <strong>of</strong> the SLF outlined above highlights the broader<br />

issue <strong>of</strong> the impact <strong>of</strong> tree growing on livelihood outcomes. As the term agr<strong>of</strong>orestry implies, forestry<br />

is one component <strong>of</strong> a smallholder’s farming activities. This must be taken into account when<br />

measuring the benefits <strong>of</strong> planting trees. Assessing the returns from the trees is necessary but not<br />

sufficient to determine the benefits <strong>of</strong> an agr<strong>of</strong>orestry system. The net returns from the total<br />

agr<strong>of</strong>orestry operation must be assessed and compared with other land uses to establish whether or not<br />

agr<strong>of</strong>orestry is a better land use option. Assessment is further complicated where the decision to plant<br />

trees is not a purely economic or financial decision but includes social (family and community) and<br />

environmental considerations, as is <strong>of</strong>ten the case.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 452<br />

Financial focus: While an acceptable net present value (NPV) or internal rate <strong>of</strong> return (IRR)<br />

indicates an attractive investment option, a smallholder’s decision may consider other factors, such as<br />

the length <strong>of</strong> time before a positive cash flow occurs or the need to achieve a target income level by a<br />

particular year which aligns with anticipated education or other significant family expenses. Studies <strong>of</strong><br />

the motivations for tree planting by small-scale forest growers in <strong>Australia</strong> found that financial returns<br />

per se were a minor factor (Venn, 2005). Non-commercial benefits such as shade, shelter, land<br />

rehabilitation, water table management, wildlife habitat conservation and stock feed were more<br />

important factors. The influence <strong>of</strong> social and environmental factors on landholders’ decisions to adopt<br />

farm forestry suggests that it is a more than a purely financial decision for many landowners (Schirmer<br />

et al., 2000; Schirmer, 2004). For smallholder agr<strong>of</strong>orestry, conventional performance measures <strong>of</strong><br />

NPV, IRR, land expectation value (LEV), annual equivalent value (AEV) and return to labour will<br />

need to be supplemented with specific measures that reflect the particular needs and preferences <strong>of</strong><br />

smallholder investors. This involves understanding the processes that smallholders apply when making<br />

land use decisions.<br />

Attitudes to risk and uncertainty: A challenge for allocation <strong>of</strong> livelihood assets to forestry is the<br />

length <strong>of</strong> time before returns can be realised. Rotation lengths can range from four to six years for fastgrown<br />

pulpwood species (e.g. mangium: Acacia mangium), to 20 years for exotic sawnwood species<br />

(e.g. teak: Tectona grandis), and substantially longer for native hardwood species (e.g. kwila: Instia<br />

bijuga). Growers’ advisors and growers have to make judgments on how institutional, financial, social<br />

and physical conditions may change from establishment to harvest. If changes are likely, growers must<br />

consider how they might affect the fate <strong>of</strong> production. For some factors there may be a high degree <strong>of</strong><br />

certainty about likely outcomes (e.g., a government decision may have been made to develop roads or<br />

provide power supply that will benefit growers and downstream processors and enhance delivery to<br />

markets in the future; for other factors, there may be a high degree <strong>of</strong> uncertainty about their future<br />

state or level (e.g., will market requirements for timber resource characteristics be more or less<br />

specific in 20 to 30 years time?). Dealing with the uncertainty <strong>of</strong> future conditions is a major issue in<br />

forestry investment for small-scale and large-scale growers alike.<br />

Innovation, experience and adoption: There are many factors to consider before a tree is planted with<br />

confidence that it will be part <strong>of</strong> a successful agr<strong>of</strong>orestry production system. As a land use system,<br />

agr<strong>of</strong>orestry is more knowledge intensive than agricultural innovations such as a new crop variety or<br />

fertilizer regime. Consequently the uptake <strong>of</strong> agr<strong>of</strong>orestry is not a simple exercise in enterprise or<br />

input substitution but involves modification to the whole farming operation. Effective implementation<br />

requires substantial investment in education and training (Mercer, 2004) complemented by a good<br />

understanding <strong>of</strong> the capacity and motivation <strong>of</strong> prospective growers.<br />

There is added uncertainty for smallholders where agr<strong>of</strong>orestry requires establishing a new mix <strong>of</strong><br />

trees and annual food and fodder crops <strong>of</strong>ten with a new land management technique such as contour<br />

hedgerows, alley cropping or enriched fallows (Mercer, 2004). Successful outcomes depend on access<br />

to good knowledge and technology inputs and the capacity and patience <strong>of</strong> growers to apply and adapt<br />

that knowledge to their situation. This inherent uncertainty <strong>of</strong> agr<strong>of</strong>orestry production systems will<br />

require a slow initial adoption rates suggesting ‘that agr<strong>of</strong>orestry projects will require longer time<br />

periods before becoming self-sustaining and self diffusing than the earlier Green revolution<br />

innovations’ (Mercer, 2004, p.312). In the initial stages <strong>of</strong> implementation <strong>of</strong> agr<strong>of</strong>orestry systems, it<br />

will not be known what impacts tree crops will have on the productivity and/or quality <strong>of</strong> annual crops<br />

until various combinations have been tried and demonstrated, and estimates <strong>of</strong> the impacts <strong>of</strong> annual<br />

cropping regimes on the productivity and quality <strong>of</strong> the tree crops need to be made over a number <strong>of</strong><br />

years. Smallholder agr<strong>of</strong>orestry growers and their advisors will benefit from tools that can guide them<br />

to identify the best species to plant and how to manage inherent uncertainties and risks to best meet<br />

their objectives.<br />

ASSESSING THE RISKS – SPECIES SELECTION<br />

Where sufficient data and information exist, the decision framework should be sufficient to provide an<br />

indication <strong>of</strong> the feasibility <strong>of</strong> agr<strong>of</strong>orestry regimes in a given area. Where there are several gaps,


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 453<br />

alternative approaches may be needed to support decisions, allowing prospective investors to explore<br />

possible futures in a highly uncertainty environment.<br />

Markets: Markets exist for a range <strong>of</strong> natural forest and plantation species grown in PNG. The<br />

existence <strong>of</strong> a market is important to collect price information and roundwood specifications to help<br />

inform the grower’s allocation <strong>of</strong> livelihood assets. This can be combined with species experience to<br />

better understand likely uncertainty. Table 1 presents a matrix based on these two factors for a sample<br />

group <strong>of</strong> species. The least uncertain option is where a planted species has been grown and sold into a<br />

market, provided that the market continues. Thus, teak is a least uncertain species; in contrast,<br />

although a market exists for natural forest Intsia bijuga, it is not clear whether plantation grown<br />

material will be accepted, or if that market will still exist if there is a substantial gap between reduced<br />

supply from natural forests and wood from plantations coming on stream in perhaps 50 years time.<br />

Market<br />

Existing New<br />

New species Intsia bijuga (natural forest) Intsia bijuga (plantation)<br />

Existing species Acacia mangium, Tectona grandis,<br />

Eucalyptus deglupta, Araucaria hunsteinii<br />

Table 1. An example <strong>of</strong> an assessment matrix based on experience with the species in<br />

plantations and its present market status.<br />

Species experience: Where there is experience with the silviculture <strong>of</strong> a species, the projected<br />

outcomes <strong>of</strong> management should be more certain. In PNG there are few reliable data on indigenous<br />

species as planted forests on which to base commercial management decisions (Karmar et al., 2004,<br />

p.16). The level <strong>of</strong> domestication and breeding <strong>of</strong> a species is an important factor in reducing<br />

uncertainty. Domestication and breeding is a series <strong>of</strong> sequenced repeated selections and mating to<br />

change gene frequencies (Eldridge et al., 1997, p.1). Leakey (2006, p.4) presents a species<br />

commercialisation continuum for non-timber forest products, indicating that the process has a series <strong>of</strong><br />

stages. The development <strong>of</strong> trees for wood production would require a similar development cycle.<br />

Species that have already been grown to full operational rotations would represent the least level <strong>of</strong><br />

uncertainty, whereas wild seed <strong>of</strong> an untested species would have the greatest uncertainty. A<br />

smallholder formulating strategic plans for future cashflows, e.g. for education fees, is likely to prefer<br />

planting a species with more certain outcomes. The other dimension to risk and uncertainty<br />

management is rotation length as a proportion <strong>of</strong> a person’s life. The higher the proportion, the lower<br />

the chance <strong>of</strong> undertaking another crop and the greater the likelihood that the plantation will be an<br />

intergenerational crop.<br />

ASSESSING THE BENEFITS OF SMALLHOLDER AGRO<strong>FORESTRY</strong><br />

Fit with food security. The SLF proposes that livelihood assets are first allocated to ensure food<br />

security as a primary concern. Where subsistence agriculture is practiced, development <strong>of</strong> a species<br />

options must consider labour requirements for different components <strong>of</strong> the land use system. Figure 1<br />

presents the labour inputs to establish, maintain, and harvest to roadside on-truck mangium pulpwood<br />

in the Madang area. Mangium is <strong>of</strong> interest in part because its production interval complements that <strong>of</strong><br />

a garden rotation in the Madang region, which includes a 4 to 5 year fallow period. The initial crop<br />

establishment (including year 0 maintenance = 21.0% <strong>of</strong> labour) and the final harvest (39.5% <strong>of</strong><br />

labour) consume the greatest labour. Assuming that total labour available per year is 260 days (net <strong>of</strong><br />

community and church commitments), tree growing would be a significant proportion <strong>of</strong> labour,<br />

potentially at the expense <strong>of</strong> food production. However, site visits indicate that the local farmers were<br />

incorporating mangium into their garden system, suggesting that farmers were managing the time<br />

demands for both gardening and mangium growing.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 454<br />

Figure 1. The labour inputs to establish, manage and harvest a mangium pulpwood crop<br />

established at 1,100 stems/ha in the Madang area. The site was assumed to be 7<br />

years <strong>of</strong> regrowth forest on an ex-garden site.<br />

Returns on labour: agr<strong>of</strong>orestry can generate community benefits, provide employment opportunities<br />

for local communities and contribute to social stability within a community or region, especially<br />

where there is excess labour capacity. A smallholder must find a balance between allocation <strong>of</strong> their<br />

own and contract labour, which will depend on their broader community focus compared to their own<br />

interests. Taking caution from Chambers (2007, p.38-41), labour productivity will be impacted on by a<br />

range <strong>of</strong> factors. Returns on labour, (with cash crops) can be used to compare crop options by<br />

smallholders or the option to work for wages (K8/day is the base wage). Figure 2 presents returns on<br />

labour and total labour input for PNG cash crops (taken from Bourke and Harwood, in press) and a<br />

series <strong>of</strong> mangium management scenarios (using Jenkin, 2002). The aim was to test for drivers <strong>of</strong><br />

returns, while keeping the returns constant. The impact on garden output has been excluded as, at this<br />

stage <strong>of</strong> the project, this information is yet to be collected.<br />

Figure 2. Returns on labour and labour demands for PNG cash crops (Bourke and Harwood<br />

in press) and a series <strong>of</strong> mangium management scenarios. The tree crop was assumed<br />

to grow at a mean annual increment <strong>of</strong> 13.3 m 3 /ha/y at age 6 years. The percentage<br />

indicates the area covered e.g. 25% would indicate 1 in 4 rows are planted. ‘G’ indicates<br />

gardening applied for the number <strong>of</strong> years indicated, replacing maintenance inputs. ‘FG’<br />

indicates gardening throughout the rotation. ‘FM’ indicates full labour maintenance<br />

applied. ‘CH’ indicates that contract harvesting is used. Seedlings are assumed to be<br />

provided free to the smallholders.<br />

Smallholder intercropping results in frequent and intensive tending operations for the crops, improving<br />

planted tree survival and growth rates (Bertomeu et al., 2008, p.28). Increasing replacement <strong>of</strong><br />

maintenance labour by gardening improves returns on the marginal increase in labour for the planted<br />

tree crop e.g. from K8.88 / day when providing all labour in a woodlot to K217.33 / day with planting


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 455<br />

labour only and the maintenance by gardening throughout the rotation. The practicalities <strong>of</strong> each<br />

scenario will be determined in the field. The use <strong>of</strong> contract labour at K8.00 / day to harvest to<br />

roadside and load trucks is a key driver <strong>of</strong> smallholder returns.<br />

Further scenario analysis, assuming different levels and forms <strong>of</strong> labour inputs, <strong>of</strong> a mangium<br />

pulpwood regime was conducted (Figure 3). The analysis indicates the impact <strong>of</strong> increasing reliance<br />

on gardening to achieve tree inputs, and <strong>of</strong> the choice to contract harvest to roadside. The use <strong>of</strong><br />

contract harvest labour was observed in the field, reflecting local landowners’ assessment <strong>of</strong> the most<br />

efficient use <strong>of</strong> their time. The analysis indicates that, under the assumptions made, integration <strong>of</strong> tree<br />

crops as a marginal activity within a garden system provides an important cash returns on the marginal<br />

smallholder labour deployed.<br />

Figure 3. Returns on labour and labour demands for a series <strong>of</strong> mangium management<br />

scenarios. The tree crop was assumed to grow at a mean annual increment <strong>of</strong> 13.3<br />

m 3 /ha/y at age 6 years. The labour inputs are: ‘Full labour’ = 100% smallholder labour;<br />

‘M1’ = 3 rounds <strong>of</strong> 6days/ha/round <strong>of</strong> maintenance per year; ‘M2’ = 2 rounds <strong>of</strong><br />

6days/ha/round <strong>of</strong> maintenance per year; ‘M3’ = 1 rounds <strong>of</strong> 6days/ha/round <strong>of</strong><br />

maintenance per year; ‘G’ = gardening to achieve maintenance in the years indicated;<br />

‘P’ = smallholders planted the tree crop; ‘CH’ indicates that contract harvesting is used.<br />

Seedlings are assumed to be provided free to the smallholders.<br />

CONCLUDING COMMENTS<br />

In situations where risk and uncertainty can have catastrophic results, such as farming systems<br />

decisions in many parts <strong>of</strong> PNG, it is necessary ensure that any recommendations for changes to these<br />

systems helps minimise risk and uncertainty in decision making. This paper has reviewed a conceptual<br />

framework for appropriate decision making, and applied that approach to a sample <strong>of</strong> species and tree<br />

growing systems used in PNG. The next stage <strong>of</strong> the project will integrate this approach with the<br />

outcomes <strong>of</strong> research investigating landowners’ attitudes to tree growing, and more precise<br />

information about the basis for their decision making.<br />

ACKNOWLEDGMENTS<br />

The project is funded by the <strong>Australia</strong>n Centre for International Agricultural Research (ACIAR). We<br />

thank ACIAR and its Forestry Program Manager, Dr Russell Haines, for their support. The project<br />

team has been supported by many PNG institutions – including the PNG Forest Authority, PNG<br />

University <strong>of</strong> Technology, and Ok Tedi Development Foundation. We have been hosted by many<br />

people in PNG during field visits; their assistance is invaluable. The knowledge <strong>of</strong> the other team<br />

members has helped frame the concepts presented; in particular, we thank Dr Mike Bourke, whose<br />

PNG experience and passion has been a key ingredient, and acknowledge the contributions <strong>of</strong> Hartmut<br />

Holzknecht, Lastus Kuniata, Andrew McGregor, Kulala Mulung and Ruth Turia.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 456<br />

REFERENCES<br />

Ashley, C. & Carney, D., 1999. Sustainable livelihoods: lessons from early experience. London: Department for<br />

International Development.<br />

<strong>Australia</strong>n Centre for International Agricultural Research, 2009. Value-adding to PNG agr<strong>of</strong>orestry systems.<br />

Available at: http://www.aciar.gov.au/project/FST/2004/050 [accessed 04/03/09].<br />

Bank <strong>of</strong> Papua New Guinea, 2009. Table 8.7. Non-agricultural exports – quantities exported <strong>of</strong> major<br />

commodities. Available at http://www.bankpng.gov.pg [accessed 05/03/09].<br />

Bannock, G., Baxter, R.E & Davis, E., 1992. The Penguin dictionary <strong>of</strong> economics. 5 th ed. England: Penguin<br />

Press.<br />

Bertomeu, M., Roshetko, J.M. & Fay, C. 2008. Domesticating landscapes through farmer-led tree cultivation: an<br />

agr<strong>of</strong>orestry approach to reforestation in Southeast Asia, In: IUFRO International Symposium, Smallscale<br />

rural forest use and management: global policies versus local knowledge. Gerardner, France, 23-<br />

27 June 2008. Nancy: France.<br />

Bigsby, H.R. ed., 2001. Assessment and management <strong>of</strong> forest investment risk. Canterbury, New Zealand,<br />

Lincoln University Press.<br />

Bourke, M. & Allen, B., 2008. Poverty in rural Papua New Guinea. [Presentation at Divine World University,<br />

Madang] May 2008. Canberra: Land Management Group, The <strong>Australia</strong>n National University.<br />

Bourke, R.M. & Harwood, T. eds, (in press). Food and Agriculture in Papua New Guinea. Canberra: The<br />

<strong>Australia</strong>n National University.<br />

Byron, N., 2001. Keys to smallholder forestry in developing countries in the tropics. In: Harrision, S. R. and<br />

Herbohn, J. L. eds. 2001. Sustainable Farm Forestry in the Tropics: Social and Economic Analysis and<br />

Policy. UK: Edward Elgar Publishing. Part. IV<br />

Chambers, R., 2007. Poverty research: Methodologies, mindsets and multidimensionality. IDS Working Paper<br />

293. Brighton (UK): <strong>Institute</strong> <strong>of</strong> Development Studies, The University <strong>of</strong> Sussex.<br />

Chambers, R. & Conway, G.R., 1991. Sustainable rural livelihoods: practical concepts for the 21 st century. IDS<br />

Discussion Paper 296. Brighton (UK): <strong>Institute</strong> <strong>of</strong> Development Studies, The University <strong>of</strong> Sussex.<br />

Customs Tariff Act 1990 (Schedule 2: Rates <strong>of</strong> Export Duties). Port Moresby: Government <strong>of</strong> Papua New<br />

Guinea. Available at http://www.customs.gov.pg [accessed 05/03/09].<br />

Eldridge, K., Davidson, J., Harwood, C. & van Wyk, G., 1997. Eucalypt domestication and breeding. Oxford:<br />

Clarendon Press.<br />

Hanson, L.W., Allen, B.J., Bourke, R.M. & McCarthy, T.J., 2001. Papua New Guinea rural development<br />

handbook. Canberra: The <strong>Australia</strong>n National University.<br />

Holding, C. & Roshetko, J.M., 2003. Farm-level timber production: orienting farmers towards the market.<br />

Unasylva 212. (54), pp.48-56.<br />

Jenkin, B.M., 2002. The search for transparency in determining the share <strong>of</strong> residual woodchip stumpage returns<br />

to growers. The <strong>Australia</strong>n Forest Grower, 25(4), Special Liftout No. 62.<br />

Karmar, M., Spencer, D. & Aken, K., 2004. Report on the assessment <strong>of</strong> the Papua New Guinea Forest<br />

Authority’s (Forest Research <strong>Institute</strong>) planted trials <strong>of</strong> indigenous species. [Draft report] March 2004.<br />

Part <strong>of</strong> ACIAR Project FTS/1998/115 Domestication <strong>of</strong> Papua New Guinea’s indigenous forest species.<br />

Canberra: CSIRO Forestry and Forest Products.<br />

Kanowski, P., Holzknecht, H. & McGregor, A., 2007, Value adding to Papua New Guinea agr<strong>of</strong>orestry systems<br />

through the incorporation <strong>of</strong> high-value tree species. [Scoping Project Report] November 2007.<br />

ACIAR Project FST/2005/050. Canberra: The <strong>Australia</strong>n National University.<br />

Kanowski, P, Holzknecht, H, Mulung, Blyth, M, Jenkin, B & McGregor, A. 2008, Fostering commercial tree<br />

growing by small-scale landowners in Papua New Guinea from a base in traditional agr<strong>of</strong>orestry<br />

systems, In: IUFRO International Symposium, Small-scale rural forest use and management: global<br />

policies versus local knowledge. Gerardner, France, 23-27 June 2008. Nancy: France.<br />

Kennedy, J. & Clarke, W., 2004. Cultivated landscapes <strong>of</strong> the Southwest Pacific. Resource Management in Asia-<br />

Pacific Working Paper No.50. Canberra: Resource Management in Asia-Pacific, The School <strong>of</strong> Pacific<br />

and Asian Studies, The <strong>Australia</strong>n National University<br />

Kirby, C.M., Sinden, J.A. & Kaine, G.W., 1993. Appraisal <strong>of</strong> agr<strong>of</strong>orestry investment under uncertainty: A<br />

South <strong>Australia</strong>n case study. <strong>Australia</strong>n Forestry, 56 (2), pp.109-119.<br />

Leakey, R. 1998. Editorial. Agr<strong>of</strong>orestry Today. September.


Proceedings <strong>of</strong> the Biennial Conference <strong>of</strong> the<br />

<strong>Institute</strong> <strong>of</strong> <strong>Foresters</strong> <strong>of</strong> <strong>Australia</strong>, Caloundra, 2009 457<br />

Leakey, R.R.B., 2006. Domestication and commercialisation <strong>of</strong> agr<strong>of</strong>orestry trees for AFTPs. In:<br />

Agr<strong>of</strong>orestry.net, Trees for improving pr<strong>of</strong>itability, sustainability, and resource conservation on farms<br />

and ranches 2006: Pacific Island Agr<strong>of</strong>orestry Workshops and Field Visits 2006. Keauhou, Kona,<br />

Hawaii, 16 th to 19 th May, 2006. Available at: http://www.agr<strong>of</strong>orestry.net [accessed 13/03/09].<br />

Mendoza, G.A., Campbell, G.E. & Rolf, G.L., 1986. Multiple objective programming: An approach to planning<br />

and evaluation <strong>of</strong> agr<strong>of</strong>orestry systems – part 1: Model description and development. Agricultural<br />

Systems, 22, pp.243-253.<br />

Mercer, D.E., 2004. Adoption <strong>of</strong> agr<strong>of</strong>orestry innovations in the tropics: A review. Agr<strong>of</strong>orestry Systems, 61(1),<br />

pp.311-328.<br />

Messer, N. & Townsley, P., 2003. Local institutions and livelihoods: Guidelines for analysis. Rome: Rural<br />

Development Division, Food and Agricultural Organisation <strong>of</strong> the United Nations.<br />

Mulung, K, 2007, Land owner decision making processes in relation to commercial tree growing in Papua New<br />

Guinea (PNG). [Draft research proposal] October 2007. ACIAR Project FST/2004/050. Canberra:<br />

<strong>Australia</strong>n National University.<br />

Pannell, D. J., Marshall, G. R., Barr, N., Curtis, A., Vanclay, F. & Wilkinson, R., 2006. Understanding and<br />

promoting adoption <strong>of</strong> conservation practices by rural landholders. <strong>Australia</strong>n Journal <strong>of</strong> Experimental<br />

Agriculture. 46 (11), pp.1407-24.<br />

Scherr, S. J., 2004. Building opportunities for small-farm agr<strong>of</strong>orestry to supply domestic wood markets in<br />

developing countries. Agr<strong>of</strong>orestry systems. 61(1), pp.357-370.<br />

Schirmer, J., Kanowski, P. & Race, D., 2000. Factors affecting adoption <strong>of</strong> plantation forestry on farms:<br />

implications for farm forestry development in <strong>Australia</strong>. <strong>Australia</strong>n Forestry, 63 (1), pp.44-51.<br />

Schirmer, J., 2004. Review <strong>of</strong> social perceptions <strong>of</strong> plantation expansion in <strong>Australia</strong>. In: Woodhead, A.,<br />

Hyndman, D. & Dunn, T., eds 2004. Pratt Water: Timber Plantation Project. Social Assessment.<br />

Canberra: Bureau <strong>of</strong> Rural Resources, Department <strong>of</strong> Agriculture, Fisheries and Forestry.<br />

Tomich, T. P., Van Noordwijk, M., Budiddarseno, S., Gillison, A., Kusumanto, T., Murdiyarso, D., Stolle, F.<br />

and Fagi, A.M., 2001. Agricultural intensification, deforestation, and the environment: assessing<br />

trade<strong>of</strong>fs in Sumatera, Indonesia. In: D.R Lee and C.B. Barrett, eds. 2001. Trade<strong>of</strong>fs or synergies?<br />

Agricultural intensification, economic development and the environment. Wallingford UK: CAB-<br />

International, pp.221-24.<br />

Venn, T. J., 2005. Financial and economic performance <strong>of</strong> long-rotation hardwood plantation investments in<br />

Queensland, <strong>Australia</strong>. Forest Policy and Economics. 7(3), pp.437-454.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!