07.04.2013 Views

Dealing with salinity in Wheatbelt Valleys - Department of Water

Dealing with salinity in Wheatbelt Valleys - Department of Water

Dealing with salinity in Wheatbelt Valleys - Department of Water

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

non-f<strong>in</strong>ancial benefits other than for <strong>sal<strong>in</strong>ity</strong> need<br />

to be sufficient.)<br />

In the past, the <strong>in</strong>terdependence <strong>of</strong> farmers <strong>with</strong><strong>in</strong><br />

a large catchment has been emphasised. At least<br />

<strong>with</strong> respect to movements <strong>of</strong> groundwater,<br />

<strong>Wheatbelt</strong> farmers are now seen to be much less<br />

dependent on each other for groundwater<br />

management. Importantly, it is now known to be<br />

possible for groundwater management to be<br />

effective locally, at least temporarily, even <strong>with</strong>out<br />

cooperation from neighbours. Typically, soils <strong>in</strong><br />

the large wheatbelt valleys <strong>of</strong> Western Australia<br />

have low “transmissivity”, mean<strong>in</strong>g that little water<br />

passes through them. This, comb<strong>in</strong>ed <strong>with</strong> the<br />

very low slopes, means that lateral water<br />

movement is very slow <strong>in</strong>deed and transmission <strong>of</strong><br />

pressure is low. Even if equilibrium areas <strong>of</strong><br />

<strong>sal<strong>in</strong>ity</strong> are reduced (as <strong>in</strong> parts <strong>of</strong> Figure 1), the<br />

plant<strong>in</strong>g <strong>of</strong> large areas <strong>of</strong> perennials is likely to<br />

delay the process <strong>of</strong> reach<strong>in</strong>g that equilibrium by<br />

between 20 and 80 years (Campbell et al. 2000).<br />

This delay is worth money directly, and it may also<br />

buy time for better management options to<br />

become available.<br />

But the treatments still need to generate money<br />

directly. The availability <strong>of</strong> perennial options<br />

which can generate an <strong>in</strong>come will be critical for<br />

our future <strong>sal<strong>in</strong>ity</strong> management <strong>in</strong> the <strong>Wheatbelt</strong>.<br />

A big effort to prove (and improve) lucerne is<br />

underway. The other option on the horizon is oil<br />

mallees. Oil mallees appear likely to become<br />

pr<strong>of</strong>itable for farms located <strong>with</strong><strong>in</strong> the transport<br />

limits <strong>of</strong> process<strong>in</strong>g plants/power generators<br />

(Bartle 1999; Cooper 2000; Herbert 2000). A<br />

pilot plant is planned for the town <strong>of</strong> Narrog<strong>in</strong>.<br />

However it is clear that we will need many more<br />

options that these two. The new Cooperative<br />

Research Centre for Plant-Based Management <strong>of</strong><br />

Dryland Sal<strong>in</strong>ity has as its prime objective the<br />

development <strong>of</strong> new pr<strong>of</strong>itable plant based options<br />

for <strong>sal<strong>in</strong>ity</strong> management.<br />

Liv<strong>in</strong>g <strong>with</strong> <strong>sal<strong>in</strong>ity</strong><br />

Farmers <strong>with</strong> large areas <strong>of</strong> salt-affected land are<br />

already triall<strong>in</strong>g and implement<strong>in</strong>g farm<strong>in</strong>g systems<br />

based on salt-tolerant species (e.g. salt bush, blue<br />

bush). In addition, there is grow<strong>in</strong>g <strong>in</strong>terest <strong>in</strong><br />

economic uses for sal<strong>in</strong>e water (e.g. aquaculture,<br />

electricity generation, irrigation <strong>with</strong> brackish<br />

water, algae [e.g. for agar, β-carotene, pigments,<br />

fish food], seaweed) and the potential to extract<br />

valuable salts and m<strong>in</strong>erals (e.g. magnesium,<br />

brom<strong>in</strong>e, potassium chloride) (see<br />

– 5 –<br />

Pannell<br />

http://www.ndsp.gov.au/opus/menu.htm for the<br />

OPUS database, accessed 5 June 2001).<br />

There has been little recent economic analysis <strong>of</strong><br />

these practices (which seems an important<br />

oversight). However, there is reason to expect<br />

that they will become <strong>of</strong> considerable economic<br />

importance. Much <strong>of</strong> the forecast sal<strong>in</strong>isation <strong>of</strong><br />

land is not technically avoidable <strong>with</strong>out<br />

implausibly large changes <strong>in</strong> land use. The plantbased<br />

options for sal<strong>in</strong>e land have at least one<br />

advantage over perennials planted <strong>in</strong> recharge<br />

areas. Recharge areas are still productive and<br />

valuable for traditional production, so farmers will<br />

be reluctant to switch to other land uses unless<br />

they are about as pr<strong>of</strong>itable. Establish<strong>in</strong>g salttolerant<br />

species on salt-affected areas does not<br />

<strong>in</strong>volve the same sort <strong>of</strong> sacrifice. Therefore,<br />

provided that up-front establishment costs are low<br />

enough and/or adequate productivity can be<br />

demonstrated, the prospects for widespread<br />

adoption <strong>of</strong> new salt-tolerant plants for economic<br />

production on salt-affected land appear good.<br />

Another advantage is that feed availability from<br />

salt-tolerant pastures can be timed to reduce the<br />

autumn feed gap, and therefore to <strong>in</strong>crease yearround<br />

stock<strong>in</strong>g levels. This rema<strong>in</strong>s an advantage<br />

so long as the area <strong>of</strong> salt-tolerant pastures is not<br />

so great as to more than fill the feed gap.<br />

Repair<strong>in</strong>g <strong>sal<strong>in</strong>ity</strong><br />

Many farmers would prefer to repair sal<strong>in</strong>ised land<br />

and cont<strong>in</strong>ue <strong>with</strong> traditional agriculture, if that is<br />

possible. This requires eng<strong>in</strong>eer<strong>in</strong>g. Deep open<br />

dra<strong>in</strong>s have been <strong>in</strong>stalled by many farmers to<br />

enhance discharge. Measurements by researchers<br />

have found that they reduce groundwater levels<br />

<strong>with</strong><strong>in</strong> only a few metres <strong>of</strong> the dra<strong>in</strong> on high-clay<br />

soils and rarely more than 40 metres on favourable<br />

soils (George 1985; George & Nulsen 1985;<br />

Speed & Simons 1992; Ferdowsian et al. 1997).<br />

However farmers have observed positive effects <strong>of</strong><br />

deep dra<strong>in</strong>s on plant growth over greater distances<br />

than this. It is not clear whether these effects at a<br />

distance are due to reductions <strong>in</strong> groundwater<br />

levels or reductions <strong>in</strong> water logg<strong>in</strong>g (which could<br />

probably be achieved <strong>with</strong> much cheaper<br />

structures). An economic analysis <strong>of</strong> deep open<br />

dra<strong>in</strong>s on agricultural land by Ferdowsian et al.<br />

(1997) reached negative conclusions about their<br />

cost effectiveness, but given farmer observations<br />

and new evidence that is emerg<strong>in</strong>g about their<br />

effectiveness <strong>in</strong> some situations, further research<br />

and analysis is needed. A key issue to resolve <strong>with</strong><br />

deep dra<strong>in</strong>age is the cost-effective and<br />

environmentally safe disposal <strong>of</strong> discharged waters.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!