28.12.2012 Views

biology join - Coweeta LTER - University of Georgia

biology join - Coweeta LTER - University of Georgia

biology join - Coweeta LTER - University of Georgia

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Articles<br />

Long-Term Ecological Research<br />

in a Human-Dominated World<br />

G. PhiliP RobeRtson, scott l. collins, DaviD R. FosteR, nicholas bRokaw, huGh w. Ducklow,<br />

teD l. GRaGson, coRinna GRies, stePhen k. hamilton, a. DaviD mcGuiRe, John c. mooRe,<br />

emily h. stanley, RobeRt b. waiDe, anD maRk w. williams<br />

The US Long Term Ecological Research (<strong>LTER</strong>) Network enters its fourth decade with a distinguished record <strong>of</strong> achievement in ecological science.<br />

The value <strong>of</strong> long-term observations and experiments has never been more important for testing ecological theory and for addressing<br />

today’s most difficult environmental challenges. The network’s potential for tackling emergent continent-scale questions such as cryosphere loss<br />

and landscape change is becoming increasingly apparent on the basis <strong>of</strong> a capacity to combine long-term observations and experimental results<br />

with new observatory-based measurements, to study socioecological systems, to advance the use <strong>of</strong> environmental cyberinfrastructure, to promote<br />

environmental science literacy, and to engage with decisionmakers in framing major directions for research. The long-term context <strong>of</strong> network<br />

science, from understanding the past to forecasting the future, provides a valuable perspective for helping to solve many <strong>of</strong> the crucial environmental<br />

problems facing society today.<br />

Keywords: coupled natural–human systems, cyberinfrastructure, environmental observatories, environmental education, socioecological systems<br />

The US Long Term Ecological Research (<strong>LTER</strong>) Network<br />

was started in 1980 to provide sites for ecologists to<br />

address questions that require long periods <strong>of</strong> study in<br />

order to be resolved. Hypothesis-driven research conducted<br />

over an extended period is a hallmark <strong>of</strong> the <strong>LTER</strong> Network<br />

today and the foundation <strong>of</strong> its scientific contributions<br />

(see Callahan 1984, Franklin et al. 1990, Hobbie et al. 2003).<br />

At 26 sites (figure 1), ecologists conduct synthetic and crosssite<br />

research that builds on site-based data, experiments, and<br />

models across diverse regions.<br />

Historically, studies at <strong>LTER</strong> Network sites have addressed<br />

long-term questions not easily addressed in short-term<br />

funding cycles: How do populations change in response to<br />

long-term environmental forcings such as landscape and<br />

climate change? How do these changes affect biodiversity<br />

and trophic interactions and, in turn, primary productivity,<br />

element cycles, and other ecosystem processes? What are<br />

the lags in ecosystem responses to and the legacies <strong>of</strong> past<br />

human and natural disturbances? What precipitates ecological<br />

tipping points, and are such changes predictable?<br />

These questions are broadly applicable to all ecosystems,<br />

and as the value <strong>of</strong> addressing them became clear during<br />

the first 20 years <strong>of</strong> the program, the <strong>LTER</strong> Network grew to<br />

include additional biomes and ecosystem types, to encompass<br />

broader regional scales <strong>of</strong> inquiry, and to incorporate<br />

human-dominated systems in its research. Today’s network<br />

<strong>of</strong> forest, grassland, desert, freshwater, coastal, and other<br />

ecosystems spans a broad geographic range <strong>of</strong> both climate<br />

and human impact. Climates within the network range from<br />

polar to tropical and from maritime to continental, with<br />

correspondingly diverse biotic assemblages. Human influences<br />

among sites range from no intentional disturbance to<br />

intensive management for agricultural, rangeland, forestry,<br />

and urban outcomes.<br />

Creation <strong>of</strong> the network thus substantially altered the<br />

range <strong>of</strong> research sites used by US ecologists. Although<br />

sites in national parks, national forests, agricultural experiment<br />

stations, and biological field stations have historically<br />

provided a rich context for asking long-term ecological<br />

questions, most questions have been addressed in an ad hoc<br />

manner. The <strong>LTER</strong> Network provides an explicit opportunity<br />

to document ecological changes and to simultaneously<br />

address long-term questions across a broad array <strong>of</strong> ecosystems.<br />

Documenting these changes provides an important<br />

opportunity to ask mechanistic questions about the causes<br />

and consequences <strong>of</strong> change, an additional hallmark <strong>of</strong><br />

<strong>LTER</strong>: place-based long-term experimentation (Knapp et al.<br />

2012 [in this issue]).<br />

The ability to link results at one site to findings at another<br />

allows the exploration <strong>of</strong> questions at broader geographic<br />

BioScience 62: 342–353. ISSN 0006-3568, electronic ISSN 1525-3244. © 2012 by American Institute <strong>of</strong> Biological Sciences. All rights reserved. Request<br />

permission to photocopy or reproduce article content at the <strong>University</strong> <strong>of</strong> California Press’s Rights and Permissions Web site at www.ucpressjournals.com/<br />

reprintinfo.asp. doi:10.1525/bio.2012.62.4.6<br />

342 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


SBC<br />

CCE<br />

BNZ<br />

0 10 km<br />

ARC<br />

MCR<br />

AND<br />

CAP<br />

NWT<br />

JRN<br />

SEV<br />

SGS<br />

CDR<br />

KNZ<br />

NTL<br />

KBS<br />

CWT<br />

BES<br />

FCE<br />

scales in order to explore both regional patterns and controls,<br />

as well as to explore the degree <strong>of</strong> connectivity among<br />

disparate parts <strong>of</strong> regional and continental landscapes (sensu<br />

Peters et al. 2008). By the end <strong>of</strong> the network’s third decade,<br />

the number <strong>of</strong> cross-site studies had ballooned (figure 2;<br />

Johnson et al. 2010). The following five examples demonstrate<br />

this development:<br />

(1) In the Long-Term Intersite Decomposition Experiment,<br />

the decomposition rates <strong>of</strong> leaf litter and roots were<br />

measured in a 10-year reciprocal-transplant experiment<br />

among 21 long-term sites in seven biomes. The results<br />

showed that relatively simple models can predict decomposition<br />

rates on the basis <strong>of</strong> litter quality and regional<br />

climate (Gholz et al. 2000) but that the rate <strong>of</strong> nitrogen<br />

release from leaf litter is largely independent <strong>of</strong> climate<br />

(Parton et al. 2007). Nitrogen release instead depends on<br />

the initial tissue nitrogen concentrations and mass, except<br />

in arid environments in which exposure to large amounts<br />

<strong>of</strong> ultraviolet radiation overrides the influence <strong>of</strong> nitrogen<br />

content.<br />

(2) In the Lotic Intersite Nitrogen Experiment (LINX), a set<br />

<strong>of</strong> comparative studies <strong>of</strong> nitrogen dynamics were conducted<br />

HBR<br />

PIE<br />

HFR<br />

GCE<br />

Articles<br />

in 70 headwater streams from<br />

across North America on the basis<br />

<strong>of</strong> collaborations that grew out<br />

<strong>of</strong> the <strong>LTER</strong> Network and later<br />

included other sites. Using coordinated<br />

whole-stream nitrogen-15<br />

isotope-addition experiments,<br />

LINX studies demonstrated the<br />

importance <strong>of</strong> headwater streams<br />

for maintaining downstream<br />

water quality (Peterson et al.<br />

2001, Helton et al. 2011), quantified<br />

their sensitivity to excess<br />

nitrate loading (Mulholland et al.<br />

2008), and clarified their role as<br />

sources <strong>of</strong> the potent greenhouse<br />

gas nitrous oxide (Beaulieu et al.<br />

2011). LINX research has now<br />

expanded to include nitrogen<br />

cycling in large rivers and wetlands<br />

in addition to streams.<br />

(3) In a study <strong>of</strong> very longterm<br />

records <strong>of</strong> lake ice initiated<br />

at <strong>LTER</strong> Network sites and then<br />

expanded to other Northern<br />

Hemisphere locations, Magnuson<br />

and colleagues (2000) exposed a<br />

trend <strong>of</strong> shorter and more variable<br />

durations <strong>of</strong> ice cover over<br />

the past century. These trends<br />

<strong>of</strong> reduced ice cover <strong>of</strong>fered an<br />

integrated, long-term indication<br />

<strong>of</strong> a warming climate over broad<br />

geographic regions.<br />

(4) A working group convened at the National Center<br />

for Ecological Analysis and Synthesis to examine the relationship<br />

between plant productivity and diversity at <strong>LTER</strong><br />

Network and other sites (Waide et al. 1999) led to a<br />

meta-analysis <strong>of</strong> over 170 studies <strong>of</strong> species richness and<br />

productivity (Mittelbach et al. 2001), which changed the<br />

prevailing view that species richness peaks at intermediate<br />

productivities. Building on this result, a more recent multisite<br />

international experiment that included <strong>LTER</strong> Network<br />

sites showed that species richness per se cannot be used to<br />

predict productivity, except in reconstructed communities<br />

(Adler et al. 2011).<br />

(5) An analysis <strong>of</strong> more than 900 species responses from<br />

34 nitrogen-fertilization experiments across long-term sites<br />

(Cleland et al. 2008) showed that trait-neutral and traitbased<br />

mechanisms operated simultaneously to influence<br />

diversity loss as net primary production increased with<br />

fertilization (Suding et al. 2005). Although soil-buffering<br />

capacity modulated some responses (Clark et al. 2007),<br />

low abundance was consistently an important driver <strong>of</strong><br />

species loss across ecosystems, and both trait-based and<br />

species-specific responses were also evident (Pennings et al.<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 343<br />

VCR<br />

PAL<br />

0 2000 km<br />

MCM<br />

Figure 1. Map <strong>of</strong> the 26 Long Term Ecological Research Network sites on an ecoregion<br />

map from Olson and colleagues (2001). Descriptions <strong>of</strong> the sites can be found at<br />

www.lternet.edu. Abbreviation: km, kilometers. For the site-name abbreviations,<br />

see Knapp and colleagues’ (2012) table 1 (in this issue, on p. 379).<br />

LUQ


Articles<br />

1992<br />

BNZ<br />

AND<br />

HBR<br />

2003<br />

AND<br />

PAL<br />

SGS<br />

FCE<br />

HFR<br />

ARC<br />

VCR LUO<br />

VCR<br />

CWT<br />

NTL<br />

KNZ<br />

HFR<br />

LUQ<br />

CDR<br />

JRN<br />

HBR<br />

PAL<br />

CWT<br />

ARC<br />

SEV<br />

JRN<br />

NTL<br />

NWT<br />

MCM<br />

SEV<br />

NWT<br />

KBS<br />

KNZ<br />

KBS<br />

SGS<br />

GCE<br />

BNZ<br />

PIE<br />

CAP<br />

BES<br />

Figure 2. Evolution <strong>of</strong> cross-site research in the Long Term<br />

Ecological Research Network in the decade prior to 2003,<br />

quantified by <strong>join</strong>t intersite publications (lines between<br />

sites), recalculated from the data in Johnson and colleagues<br />

(2010). For the site-name abbreviations, see Knapp and<br />

colleagues’ (2012) table 1 (in this issue, on p. 379).<br />

2005, Cleland et al. 2011). The results from these syntheses<br />

demonstrated that rarity, species identity, functional traits,<br />

and physical environment all contributed to changes in<br />

plant-community composition in response to soil nitrogen<br />

availability.<br />

Clearly, cross-site syntheses add substantial value to<br />

highly mechanistic site-based analyses and provide the<br />

opportunity to develop and test ecological theories that can<br />

lead to broader ecological knowledge.<br />

The new emergence <strong>of</strong> environmental observatories—<br />

including the National Ecological Observatory Network<br />

(NEON; Keller et al. 2008); the Oceans Observatory<br />

Initiative (Isern and Clark 2003); the Global Lake Ecological<br />

Observatory Network (Kratz et al. 2006); and others,<br />

including open-source networks such as NutNet (Adler<br />

et al. 2011)—underscores a growing scientific appreciation<br />

for the power <strong>of</strong> cross-site observations (Carpenter 2008,<br />

Robertson 2008). The <strong>LTER</strong> Network was not designed<br />

to be a single integrated observatory in which each<br />

site employs standardized instrumentation and capacity,<br />

as does NEON, nor are network sites optimally located to<br />

capture environmental trends at continental scales. Rather,<br />

the <strong>LTER</strong> Network was designed to provide key places for<br />

long-term, biome-specific observations and experimentation,<br />

where investigations can reveal the underlying causes<br />

and future consequences <strong>of</strong> patterns detected by distributed<br />

observatories that include <strong>LTER</strong> Network sites. And<br />

by increasingly engaging with diverse stakeholders—land<br />

managers, policymakers, and decisionmakers at all levels—<br />

<strong>LTER</strong> Network scientists can ensure that their inquiries<br />

are relevant to addressing societal concerns (Driscoll et al.<br />

2012 [in this issue]).<br />

<strong>LTER</strong> Network sites also play a unique role in science<br />

education at all levels. The sites are closely associated with<br />

institutions <strong>of</strong> higher learning, and graduate, undergraduate,<br />

and postdoctoral scholarship at these sites serves to<br />

advance ecology, as well as to introduce undergraduates<br />

to field research and graduate and postdoctoral scientists<br />

to the value <strong>of</strong> distributed research networks (e.g., Kane<br />

et al. 2008). Researchers at the sites are also actively engaged<br />

with local communities and with state and national agencies<br />

and boards, through which they can advance a range<br />

<strong>of</strong> informal education approaches, including pr<strong>of</strong>essional<br />

advancement and environmental training for community<br />

leaders. The resulting synergies have included contributions<br />

to kindergarten through twelfth grade (K–12) science education<br />

through pr<strong>of</strong>essional-development activities for science<br />

teachers (e.g., McKnight 2010), broadening and strengthening<br />

<strong>of</strong> local and state science curricula, and pedagogical contributions<br />

to the development <strong>of</strong> an environmental literacy<br />

movement, as is described below. These diverse educational<br />

efforts are increasingly providing new avenues for improving<br />

the quality and relevance <strong>of</strong> <strong>LTER</strong> science (Driscoll et al.<br />

2012).<br />

As the <strong>LTER</strong> Network enters its fourth decade, it is poised<br />

to contribute substantively to helping society respond to<br />

the ever-growing challenges <strong>of</strong> environmental sustainability,<br />

including climate-change mitigation and adaptation. How<br />

will the network help meet these challenges? In this article,<br />

we describe the vision <strong>of</strong> that future that emerged through<br />

a multiyear process <strong>of</strong> engagement across and beyond the<br />

entire <strong>LTER</strong> community, its National Science Foundation<br />

(NSF) associates, and colleagues from many other programs,<br />

observatories, and agencies.<br />

We first lay out the place <strong>of</strong> <strong>LTER</strong> in a world increasingly<br />

subject to human influence. We describe a common<br />

framework for addressing important questions and examples<br />

<strong>of</strong> three overarching research themes that can best be<br />

addressed with a network <strong>of</strong> sites: landscape vulnerability<br />

and resilience to global change, cryosphere loss, and coastalzone<br />

climate change. We then describe <strong>LTER</strong> contributions<br />

toward building environmental science literacy among K–12<br />

students, undergraduates, and those who work with broader<br />

audiences, as well as the cyberinfrastructure demands <strong>of</strong><br />

long-term ecological science and the network’s approach to<br />

meeting these needs.<br />

344 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


<strong>LTER</strong> in a human-dominated world<br />

Thirty years <strong>of</strong> <strong>LTER</strong> Network research have yielded valuable<br />

knowledge about ecosystem change in response to<br />

both natural and human influences. Changes ranging from<br />

climate alteration to species introductions and to land- and<br />

water-use decisions have far-reaching impacts on ecosystem<br />

function, community structure, and population and evolutionary<br />

dynamics, which in turn strongly affect the critical<br />

ecosystem services on which we all depend. Ecological<br />

research seeks to test theory and to provide the empirical<br />

knowledge needed to forecast change and to devise effective<br />

management and policy responses. And theory and<br />

knowledge increasingly cross the boundary between natural<br />

and human systems, effectively linking science with policy<br />

(Liu et al. 2008, Driscoll et al. 2012).<br />

A framework for exploring coupled natural–human systems over<br />

the long term. Several recent studies have shown how couplings<br />

between human and natural systems exhibit nonlinear<br />

dynamics across space, time, and organizational scales<br />

and have revealed complexities that cannot be disentangled<br />

by ecological or social research alone. The importance <strong>of</strong><br />

understanding these dynamics cannot be underestimated:<br />

Without understanding the couplings between natural and<br />

human systems, workable policy solutions to some <strong>of</strong> the<br />

most recalcitrant environmental problems <strong>of</strong> today, which<br />

range from degraded water quality to biodiversity loss to<br />

climate-change vulnerability, will remain difficult to design<br />

and even more difficult to achieve.<br />

At individual <strong>LTER</strong> sites, research on the couplings<br />

between natural and human systems has a rich history,<br />

ranging from inherently coupled working lands (row crop<br />

systems, timber plantations, grazing lands, coastal fisheries)<br />

to urban and exurban areas and sites in which direct<br />

human impact ceased decades ago but in which its legacies<br />

continue to condition ecosystem patterns and processes.<br />

In fact, no <strong>LTER</strong> Network site is uncoupled from human<br />

influence: The network’s most remote Arctic and Antarctic<br />

sites are also affected by human decisions and behaviors,<br />

although humans are far away and their effects mostly<br />

unintentional. As a whole, the <strong>LTER</strong> Network provides a<br />

broad range <strong>of</strong> sites with differing intensities <strong>of</strong> human<br />

influence, degrees <strong>of</strong> intent, and levels <strong>of</strong> connectedness<br />

(Peters et al. 2008).<br />

The integration <strong>of</strong> social and ecological research within<br />

the context <strong>of</strong> <strong>LTER</strong> Network sites and scientists (Redman<br />

et al. 2004), coupled with a rich ecological information base<br />

for the sites, is a promising new research frontier for <strong>LTER</strong>.<br />

The network has responded to this challenge by adopting as<br />

an organizing framework a common model that provides<br />

a standardized terminology and generalized structure to<br />

facilitate investigations <strong>of</strong> a wide variety <strong>of</strong> questions. The<br />

press–pulse dynamics (PPD) model (Collins et al. 2011)<br />

provides a comparative framework to integrate the biophysical<br />

and social sciences through an understanding <strong>of</strong> how<br />

human decisionmaking and behaviors interact with natural<br />

Articles<br />

processes to affect the structure, function, and dynamics <strong>of</strong><br />

ecosystems and the services they provide to people.<br />

The PPD model (see Collins et al. 2011, but cf. figure 3)<br />

is iterative, with linkages and feedbacks between biophysical<br />

and social domains (in figure 3, environmental and human<br />

systems, respectively). Model linkages are mediated by the<br />

biophysical system’s delivery <strong>of</strong> ecosystem services and by<br />

the perception <strong>of</strong> these services by the social system. Model<br />

feedbacks are mediated by how services change human<br />

outcomes, perceptions, and behaviors that in turn affect the<br />

biophysical systems and their capacity to deliver subsequent<br />

services. Behavioral changes in the PPD model range from<br />

shifts in consumer preferences to environmental and energy<br />

policies and reproduction and migration rates. Such changes<br />

deliver to the biophysical system short-term pulses, such as<br />

nutrient inputs, fires, and management interventions, as well<br />

as long-term presses, such as atmospheric carbon dioxide<br />

loading, climate change, nitrogen deposition, and sea-level<br />

rise. In time, presses, pulses, and pulse–press interactions<br />

affect community structure and ecosystem function (Smith<br />

et al. 2009), eventually changing the delivery <strong>of</strong> ecosystem<br />

services such as the provision <strong>of</strong> food and fiber, pest and disease<br />

suppression, soil fertility, greenhouse gas stabilization,<br />

and clean water.<br />

There are many potential socioecological questions that<br />

could be asked across a network <strong>of</strong> long-term sites. Building<br />

on a long history <strong>of</strong> prior research, <strong>LTER</strong> Network sites and<br />

scientists have identified several environmental challenges<br />

that represent critical issues for science and society that<br />

the network seems particularly well positioned to address<br />

today, including (a) landscape vulnerability and resilience<br />

to climate and land-use change, (b) the consequences <strong>of</strong><br />

cryosphere loss and changes in associated services that<br />

range from urban water supply to rural livelihoods, and<br />

(c) coastal-zone climate change as it interacts with rising<br />

sea levels and coastal population change. For each <strong>of</strong> these<br />

Exogenous<br />

forces<br />

Coupled human-ecosystem interactions<br />

Human<br />

systems<br />

Ecosystem<br />

services<br />

Actions/Policy<br />

Ecosystem boundary<br />

Environmental<br />

systems<br />

Exogenous<br />

forces<br />

Figure 3. A press–pulse dynamics (PPD) framework,<br />

simplified for use by K–12 learners. The complete PPD<br />

model with more comprehensive linkages and feedbacks is<br />

available in Collins and colleagues (2011).<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 345


Articles<br />

challenges, a comprehensive socioecological framework is<br />

required for them to be addressed effectively; each is best<br />

addressed with long-term observations and experiments in<br />

multiple locations; and for each, a subset <strong>of</strong> network sites<br />

in partnership with other networks and observatories could<br />

provide a core set <strong>of</strong> locations at which questions could be<br />

effectively addressed.<br />

Future scenarios: Examining landscape vulnerability and resilience<br />

to global change. Science to help us understand, anticipate,<br />

and adapt to global change, including land-use and climate<br />

change, is becoming an ever more pressing need. How will<br />

global change alter the future <strong>of</strong> regional socioecological<br />

systems, and how and why do regional systems differ in<br />

vulnerability, resilience, and adaptability to change? These<br />

questions cannot be addressed by discipline-bound thinking<br />

but, rather, require new approaches that also incorporate<br />

broad-scale comparative investigations <strong>of</strong> diverse systems.<br />

One such approach is that <strong>of</strong> scenario studies (e.g., Baker<br />

et al. 2004, Thompson et al. 2012 [in this issue]), which provide<br />

a framework for addressing socioecological questions<br />

by crafting and evaluating suites <strong>of</strong> plausible scenarios that<br />

follow from current and historical trajectories. By examining<br />

multiple visions <strong>of</strong> the future that reflect a range <strong>of</strong> assumptions<br />

about land and water use, the burden <strong>of</strong> prediction is<br />

lifted, and comparisons among contrasting scenarios can be<br />

used to understand the dynamics <strong>of</strong> complex systems. New<br />

insights come from the examination <strong>of</strong> the perceived bounds<br />

<strong>of</strong> plausibility and from the discovery <strong>of</strong> commonalities<br />

across scenarios. Indeed, intrinsic vulnerabilities and robust<br />

management strategies are <strong>of</strong>ten identified when patterns<br />

recur across disparate scenarios.<br />

Depictions <strong>of</strong> future scenarios are <strong>of</strong>ten articulated by<br />

regional stakeholders, including residents, policymakers,<br />

and social and ecological scientists, in order to illustrate<br />

major strategic choices (Hoag et al. 2005). These qualitative<br />

scenarios can be an end in themselves, or they may<br />

lead to quantitative simulations <strong>of</strong> future landscape change.<br />

This is frequently an iterative process whereby the narratives<br />

inform and are in turn informed by integrated spatial<br />

models <strong>of</strong> socioecological change that might include, for<br />

example, agent-based models that link land-use change,<br />

econometric, and ecosystem process models (Evans and<br />

Kelly 2004). At its best, fundamental site-based science<br />

underpins the development <strong>of</strong> the scenario-to-simulation<br />

framework, the creation <strong>of</strong> which is itself a form <strong>of</strong> scientific<br />

synthesis. This approach for coupling qualitative and<br />

quantitative scenarios has informed prescient planning and<br />

policy decisions and has generated a rich set <strong>of</strong> fundamental<br />

research questions.<br />

For example, researchers at the Harvard Forest <strong>LTER</strong> site<br />

have begun a statewide scenario-studies project to examine<br />

the future <strong>of</strong> Massachusetts’s forests. Their work began with a<br />

landscape-simulation study to examine the relative influence<br />

<strong>of</strong> 50 more years <strong>of</strong> the current trends in forest conversion,<br />

timber harvest, and climate change in terms <strong>of</strong> their effects<br />

on forest carbon storage and tree species composition<br />

(Thompson et al. 2011). This work was rooted in 20 years<br />

<strong>of</strong> ecological research at Harvard Forest. Researchers then<br />

convened a group <strong>of</strong> around 12 stakeholders, including<br />

natural-resource managers and decisionmakers from state<br />

government, representatives from conservation nongovernmental<br />

organizations, and academics from multiple<br />

disciplines. They asked this group to chart three alternative<br />

futures <strong>of</strong> their choosing to compare with the current<br />

trends that had already been modeled. Through spirited<br />

discussion, the group settled on (a) a “free-market future,”<br />

characterized by a rollback in environmental regulations and<br />

incentives for new business; (b) a “resource-limited future,”<br />

characterized by high energy prices, a resurgence <strong>of</strong> agriculture,<br />

and a strong demand for woody-biomass energy; and<br />

(c) a “green-investment future,” characterized by government<br />

incentives for conservation, green energy, and land-use<br />

planning. Through an iterative process with stakeholders,<br />

the researchers were able to describe the types, distribution,<br />

and intensity <strong>of</strong> land uses under each <strong>of</strong> the scenarios. Each<br />

<strong>of</strong> the scenarios is now being integrated into a simulation<br />

framework, which will superimpose the land-use scenarios<br />

onto a common template <strong>of</strong> climate-change and ecological<br />

dynamics. The goal is to examine the aggregate and interactive<br />

effects <strong>of</strong> land use within each scenario, as well as to<br />

make comparisons across the scenarios. Clearly, none <strong>of</strong><br />

the scenarios will manifest exactly as they were described;<br />

nonetheless, by examining multiple potential pathways, the<br />

effort should reveal characteristics <strong>of</strong> the Massachusetts<br />

landscape that are particularly vulnerable or resilient to the<br />

interactive effects <strong>of</strong> land-use and climate change.<br />

Cryosphere loss. The Earth’s cryosphere, which includes sea,<br />

lake, and river ice; glaciers; seasonal snow; and ice-rich permafrost,<br />

harbors over 80% <strong>of</strong> the freshwater on the planet.<br />

The cryosphere cools the planet through its albedo; regulates<br />

the global sea level; stores substantial stocks <strong>of</strong> carbon;<br />

insulates soil from subfreezing air temperatures; and serves<br />

as a seasonally refreshed water supply for human consumption,<br />

irrigation, nutrient transport, and waste disposal. The<br />

prospect <strong>of</strong> accelerated cryosphere loss under a warming climate<br />

portends great ecological change and poses enormous<br />

threats to these ecosystem services, with attendant social and<br />

economic costs. A 1-meter sea-level rise, now thought to be<br />

unavoidable with a 600–1000 parts per million atmospheric<br />

carbon dioxide peak over the coming century (Solomon<br />

et al. 2009), alone represents an estimated economic impact<br />

<strong>of</strong> $1 trillion that will be borne disproportionately by North<br />

America (Anth<strong>of</strong>f et al. 2010).<br />

The extent and rates <strong>of</strong> cryosphere loss are increasingly<br />

well monitored, and our ability to project the future rates<br />

<strong>of</strong> cryosphere decline is improving. However, the ecological<br />

consequences—and especially the nature and extent <strong>of</strong> and<br />

economic impacts on human society and institutions—<br />

are still poorly understood. Cryosphere loss represents an<br />

inadvertent press event caused by human decisions—driven<br />

346 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


y policies and markets—to extract energy from fossil fuel<br />

and to clear forests and other carbon-storing ecosystems<br />

for economic development. Changes in wintertime temperatures<br />

and snowfall will dramatically affect community<br />

structure and ecosystem processes in high-latitude and<br />

alpine ecosystems, but the effects will be felt even in arid,<br />

low-latitude ecosystems that depend on mountain meltwater<br />

for seasonal water supplies—riverine, floodplain,<br />

agricultural, and urban ecosystems in particular. Many <strong>of</strong><br />

these effects will be social, since some <strong>of</strong> the most populous<br />

cities and productive farmland in North America depend<br />

on these water supplies.<br />

Examples <strong>of</strong> cryosphere loss and its effects abound across<br />

the <strong>LTER</strong> Network, from Arctic sites undergoing long-term<br />

permafrost melt to Antarctic and northern lake sites losing<br />

ice cover and terrestrial sites experiencing shorter periods <strong>of</strong><br />

snow cover and more-frequent freeze–thaw events. Alpine<br />

communities, such as that at the Niwot Ridge <strong>LTER</strong> site,<br />

illustrate the degree <strong>of</strong> subregional connectivity involved<br />

(figure 4): Hydrological connectivity is driven by the duration<br />

and timing <strong>of</strong> the seasonal snowpack and snowmelt,<br />

and under a warming climate, increasing windborne dust<br />

will accelerate snowpack and glacial melt, which will result<br />

in the snowline’s moving to a higher elevation, which will in<br />

turn decrease hydrologic connectivity. With elevated nitrogen<br />

inputs from windborne dust and Denver air pollution,<br />

plant species diversity will decrease as alpine areas shrink,<br />

shrubland will expand, and the landscape will become more<br />

homogeneous. Exacerbating these trends is the regional<br />

outbreak <strong>of</strong> mountain pine beetles that will remove a large<br />

portion <strong>of</strong> the subalpine forest.<br />

Cryosphere change and its consequences are played out<br />

as long-term trends that in many places will be difficult to<br />

discern from short-term variability in climate and other<br />

environmental factors and in social dynamics, such as population<br />

and economic change. Long-term sites provide the<br />

perspective necessary to detect trends that would otherwise<br />

not be visible against this variability. And networked sites<br />

provide the potential for comparative tests <strong>of</strong> hypotheses<br />

that link cryosphere loss, ecosystem services, and human<br />

decisions, using, for example, the PPD model (figure 3).<br />

Key research questions (Fountain et al. 2012 [in this<br />

issue]) include (a) how climate regulation is affected by<br />

feedbacks from thawing permafrost and sea ice, especially<br />

because <strong>of</strong> the release <strong>of</strong> vast stores <strong>of</strong> carbon and changes in<br />

albedo, and what the implications are for regional and global<br />

economies and policies, including sovereignty; (b) what the<br />

economic implications <strong>of</strong> snow and ice loss are, including<br />

the future <strong>of</strong> winter recreation and related cultural activities;<br />

(c) how changing snowpack—the amount and timing<br />

<strong>of</strong> water storage and delivery—in the western United States<br />

will influence the economies <strong>of</strong> this region, and whether<br />

the impacts will be disproportionately imposed on disadvantaged<br />

groups; and (d) what the cultural mechanisms<br />

are by which cryosphere loss influences public opinion and<br />

what policies and legal instruments are most effective for<br />

Articles<br />

Figure 4. Expected changes in hydrologic connectivity<br />

related to cryosphere loss at the Niwot Ridge Long Term<br />

Ecological Research Network site. As windborne dust<br />

deposition increases in a warming climate, snowpack<br />

and glacial melt will be accelerated, which will result<br />

in a higher snowline, a shrunken alpine area, and the<br />

expansion <strong>of</strong> shrubland, exacerbated by a climate-induced<br />

mountain pine beetle outbreak that is now decimating the<br />

subalpine forest. This figure was created by Eric Parish.<br />

environmental protection, impact mitigation, and adaptation<br />

in the face <strong>of</strong> climate change. The PPD model provides<br />

an effective means for linking cryosphere change with the<br />

ecosystem dynamics that lead to altered ecosystem services<br />

and then with subsequent human activities that may—<br />

through policies, behaviors, and markets—either slow or<br />

hasten cryosphere loss.<br />

Coastal-zone climate change. Because they are at the interface<br />

<strong>of</strong> continental and oceanic realms, coastal systems are<br />

expected to be especially affected by climate change and<br />

to experience effects from both land and sea. With more<br />

than 50% <strong>of</strong> the US population living in coastal counties,<br />

many changes will play out in human communities and<br />

economies. Coastal-zone research sites, including nine<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 347


Articles<br />

<strong>LTER</strong> Network and many additional partner sites, differ<br />

in their biophysical vulnerability to the coastal impacts<br />

<strong>of</strong> climate change. Some ecosystems along the US eastern<br />

seaboard will be more affected by sea-level rise and<br />

storm-surge severities, and others will be more affected<br />

by ocean acidification (e.g., coral reef communities in the<br />

south Pacific), the loss <strong>of</strong> sea ice (Antarctica), or changes in<br />

water temperature and freshwater inflows. Human vulnerabilities<br />

will also differ among regions, which arises from<br />

differences in coastal population density and demographic<br />

composition and from the location and resilience <strong>of</strong> the<br />

regions’ built infrastructure, which ranges from cities to<br />

drilling platforms. All <strong>of</strong> these effects and vulnerabilities<br />

need to be considered in concert in order to provide a comprehensive<br />

understanding <strong>of</strong> coastal-zone climate change<br />

and the potentials for future adaptation.<br />

The need to understand and anticipate the effects <strong>of</strong><br />

climate change, assess the vulnerabilities <strong>of</strong> natural and<br />

human elements <strong>of</strong> coastal systems, and adapt to or mitigate<br />

the effects <strong>of</strong> changes is prompting new efforts for integration<br />

across academic disciplines and creation <strong>of</strong> partnerships<br />

among academic, public, and governmental constituents.<br />

As for cryosphere loss, long-term studies are needed in<br />

order to document patterns and consequences <strong>of</strong> coastalzone<br />

change. Unlike cryosphere loss, however, coastal-zone<br />

change is likely to be strongly episodic in response to storm<br />

events that are projected to be increasingly severe and whose<br />

inland consequences will, in any case, be magnified because<br />

<strong>of</strong> sea-level rise. Posing questions relevant to networked sites<br />

in the context <strong>of</strong> a common model allows for a fundamental<br />

understanding more difficult to gain from shorter-term or<br />

more geographically discrete research.<br />

Key questions include (a) how the presses and pulses<br />

associated with coastal climate change—altered water temperature,<br />

precipitation, run<strong>of</strong>f, sea level, solar radiation, wind<br />

and wave climates, pH, and salinity—affect the structure and<br />

function <strong>of</strong> coastal ecosystems and what attributes affect the<br />

vulnerability <strong>of</strong> those ecosystems; (b) how climate-induced<br />

changes in coastal systems affect critical ecosystem services<br />

such as carbon sequestration, wildlife habitat, food-web<br />

support, and storm protection; (c) what attributes <strong>of</strong> human<br />

systems such as built infrastructure; land use; governance<br />

structures; and population demographics, including wealth<br />

and ethnicity, interact to influence human vulnerabilities to<br />

coastal climate change and how these interact with changes<br />

in ecosystem services to prompt responses <strong>of</strong> adaptation and<br />

mitigation; and (d) how mitigation and adaptation strategies,<br />

such as coastal engineering and reductions in greenhouse<br />

gases, would feed back to affect climate drivers and the structure<br />

and function <strong>of</strong> coastal systems. Effectively addressing<br />

these questions requires an approach that acknowledges and<br />

deeply explores the linked socioecological processes that<br />

underlie the delivery <strong>of</strong> almost all <strong>of</strong> the ecosystem services<br />

provided by coastal-zone environments.<br />

Many <strong>LTER</strong> Network sites are located in coastal zones<br />

at different latitudes along the eastern and western US<br />

seaboards, as well as in the South Pacific and Antarctic<br />

Oceans, and provide a diversity <strong>of</strong> geomorphologies and<br />

degrees <strong>of</strong> human influence, ranging from urban to exurban,<br />

rural, and natural. They are therefore well positioned<br />

to address a subset <strong>of</strong> these key questions, most <strong>of</strong> which<br />

will require a combination <strong>of</strong> long-term baseline data<br />

and experiments designed to predict the consequences <strong>of</strong><br />

sea-level rise for the ecology <strong>of</strong> coastal communities.<br />

Toward an environmentally literate populace<br />

Society’s ability to understand and act on the coupled<br />

natural and human systems on which we depend, built<br />

on a foundation <strong>of</strong> complex scientific inquiry, is key to a<br />

sustainable future. And it is the public—decisionmakers at<br />

all levels, from landowners and local <strong>of</strong>ficials to national<br />

leaders—who must act. The importance <strong>of</strong> an environmentally<br />

literate public is hard to overstate: From the grocery<br />

store and car dealership to the voting booth and corporate<br />

boardroom, individuals make choices that have far-reaching,<br />

collective consequences. Education helps to ensure that<br />

those choices are based at least in part on evidence and reasoning<br />

underpinned by solid scientific research.<br />

The <strong>LTER</strong> approach to research, combined with an ability<br />

to implement long-term educational initiatives, has allowed<br />

for unique approaches to the training <strong>of</strong> future researchers<br />

and to the conveyance <strong>of</strong> ecological concepts and insights<br />

to a broad constituency. At individual network sites, educational<br />

activities range from K–12 students and teachers<br />

engaged in schoolyard ecology to undergraduates involved<br />

in field classes and research internships and to graduate<br />

students and postdoctoral scholars learning to frame questions<br />

and to conduct research in long-term and sometimes<br />

cross-site contexts. Public outreach in many forms reaches<br />

working pr<strong>of</strong>essionals, as well as the general public, <strong>of</strong>ten<br />

through the education <strong>of</strong> those best positioned to communicate<br />

with the general public. This outreach is increasingly<br />

placed in a socioecological context, which illustrates<br />

the natural–human system couplings that are central to<br />

addressing major environmental issues and that are <strong>of</strong>ten<br />

hidden to many.<br />

The vision for education in the <strong>LTER</strong> Network includes<br />

leveraging both long-term and cross-site perspectives to<br />

advance fundamental science learning by K–12, undergraduate,<br />

and graduate students and developing programs<br />

for key constituent and underrepresented groups. These<br />

groups include K–12 teachers, university students, education<br />

policymakers, and the pr<strong>of</strong>essional public, which includes<br />

policymakers, natural-resource managers, the working<br />

media, and others whose success depends on access to and<br />

imparting <strong>of</strong> sound ecological knowledge.<br />

A long-term framework for environmental science literacy. Environmental<br />

science literacy—the capacity to participate<br />

in and make decisions through evidence-based discussions<br />

<strong>of</strong> socioecological systems—is essential not only for<br />

many science careers but also for responsible citizenship.<br />

348 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


Environmental science literacy requires citizens to understand,<br />

evaluate, and respond to multiple sources <strong>of</strong> information.<br />

The development <strong>of</strong> an environmental science literacy<br />

framework is crucial for providing this capacity among K–12<br />

students, a key constituency that represents both future<br />

STEM (science, technology, engineering, and mathematics)<br />

pr<strong>of</strong>essionals and the 75% <strong>of</strong> the US population that will not<br />

earn a higher degree. It is also important for framing information<br />

provided to university students, STEM pr<strong>of</strong>essionals<br />

and the general public.<br />

The development <strong>of</strong> an environmental literacy framework<br />

requires that we understand stakeholders’ current<br />

state <strong>of</strong> knowledge in core areas and how math and science<br />

concepts are best used to provide a desired level <strong>of</strong> literacy.<br />

In this context, stakeholders range from K–12 students to<br />

the voting public. We know, in general, that for most stakeholders,<br />

the state <strong>of</strong> knowledge is low, which is reflected<br />

both in standardized K–12 test scores (Gonzales et al. 2008)<br />

and in college-level assessments <strong>of</strong> ecological concepts (e.g.,<br />

Hartley et al. 2011). We also know that there are troubling<br />

demographic disparities and, in particular, persistent gaps<br />

in science and mathematics achievement between white students<br />

and students <strong>of</strong> color (Vanneman et al. 2009). Building<br />

a capacity for principle-based environmental reasoning in<br />

all stakeholders and broadening the participation <strong>of</strong> underrepresented<br />

groups in environmental science careers should<br />

be important components <strong>of</strong> all science education efforts<br />

(George et al. 2001, ESA 2006).<br />

In K–12 education, the term learning progressions describes<br />

increasingly sophisticated ways <strong>of</strong> reasoning about an area<br />

<strong>of</strong> study, typically organized around a set <strong>of</strong> core topics<br />

that can be used to organize an integrated understanding<br />

<strong>of</strong> larger, complex issues (Duschl et al. 2007). A current<br />

effort within the network to build learning progressions<br />

into the K–12 science curriculum is being tested in 22<br />

school districts across the country with an <strong>LTER</strong> Network–<br />

associated NSF Math and Science Partnership (MSP) award.<br />

In districts around four network sites, MSP participants—<br />

scientists and science educators working with a diverse mix<br />

<strong>of</strong> K–12 science teachers and students—are developing<br />

learning progressions around key science strands. These<br />

strands include carbon, water, and biodiversity, plus a<br />

mathematical strand that addresses quantitative reasoning<br />

and the mathematics <strong>of</strong> modeling and a citizenship strand<br />

focused on the roles <strong>of</strong> culture and place. All <strong>of</strong> these strands<br />

are deeply embedded in state science and mathematics standards<br />

and are connected by the theme <strong>of</strong> education for citizenship:<br />

how students take on roles as consumers and voters<br />

using evidence-based reasoning about personal decisions<br />

that have environmental consequences. Multidisciplinary<br />

themes focused on the human impacts <strong>of</strong> land-use, ecosystem<br />

structure, and ecosystem services <strong>of</strong>fer rich experiences<br />

in STEM education that include atmospheric science, soil<br />

science, geology, agronomy, ecology, hydrology, computer<br />

science, and systems modeling. Placing these strands in<br />

a simplified PPD model (figure 3) provides an easily<br />

Articles<br />

understood context for showing coupled human–ecosystem<br />

interactions.<br />

Two observations have emerged thus far from our <strong>LTER</strong>based<br />

studies <strong>of</strong> learning progression. First, the PDD model<br />

embraced by the <strong>LTER</strong> Network (Collins et al. 2011) emphasizes<br />

that socioecological systems are organized as dynamic<br />

hierarchical systems. This basic tenet defines specific ways in<br />

which subjects or entities <strong>of</strong> interest are organized and interact<br />

with one another. Embedded in the concept is the notion<br />

<strong>of</strong> scale, wherein the boundaries between levels <strong>of</strong> interaction<br />

are defined by differences in the geographies and rates<br />

at which entities interact. Second, socioecological processes<br />

may include multiple principles that operate simultaneously<br />

in the social and ecological realms.<br />

Our research in student learning and understanding <strong>of</strong><br />

ecological systems indicates that students and teachers fail to<br />

adopt hierarchical reasoning when questioned about ecological<br />

systems and principles. The conservation <strong>of</strong> matter and<br />

energy is not understood. Processes operating at one scale<br />

are assumed to operate at the same scope and magnitude at<br />

other scales. The nature <strong>of</strong> the interconnectedness <strong>of</strong> systems<br />

is overstated (e.g., removal <strong>of</strong> one species leads to system<br />

collapse). Human social hierarchies and human agency are<br />

conflated with ecological hierarchies and processes and are<br />

applied to natural systems.<br />

<strong>LTER</strong> Network research on learning progressions provides<br />

insights into how to advance student understanding<br />

<strong>of</strong> socioecological systems and also provides the scaffolding<br />

<strong>of</strong> science and social-science principles that is needed for<br />

students’ environmental literacy—a key to understanding<br />

human agency and its application to the scale <strong>of</strong> action<br />

that the challenges demand and to engaging the broader<br />

public.<br />

Engaging the broader public. Extending this understanding to<br />

older stakeholders—both the voting public in general and<br />

working STEM pr<strong>of</strong>essionals, such as land managers, policy<br />

analysts, and public- and private-sector decisionmakers—<br />

presents a different set <strong>of</strong> challenges (Driscoll et al. 2012).<br />

Recent calls for a renewed effort by ecologists to engage in<br />

public outreach (e.g., Gr<strong>of</strong>fman et al. 2010) have noted that<br />

effective communication outside the classroom is influenced<br />

by learners’ interests, prior knowledge, social networks, and<br />

values and beliefs. This requires issues to be framed in ways<br />

that resonate with the public and messages to be delivered<br />

in ways that acknowledge the importance <strong>of</strong> emerging forms<br />

<strong>of</strong> media and informal learning environments (NRC 2009).<br />

Effective communication can also involve partnerships with<br />

boundary organizations that specialize in fostering the use<br />

<strong>of</strong> science knowledge in environmental policymaking and<br />

management (Osmond et al. 2010). These organizations<br />

range from university-based extension programs at landgrant<br />

universities to individual site-based efforts.<br />

One such site-based effort is the Science Links program<br />

at the Hubbard Brook <strong>LTER</strong> site. The program is explicitly<br />

aimed at communicating basic science findings at Hubbard<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 349


Articles<br />

Brook to interested audiences that range from the general<br />

public to congressional staffers and is a particularly apt<br />

example <strong>of</strong> a way to leverage limited funding to broaden the<br />

impact <strong>of</strong> ecological findings. Outreach is initiated early in<br />

a project and is informed by a group <strong>of</strong> policy and natural<br />

resource management advisors who help to craft a message<br />

that is most relevant to the audience at hand. Such efforts<br />

can help to shorten an otherwise distressing lag between<br />

recognizing and addressing important environmental problems,<br />

such as those associated with acid rain (Likens 2010),<br />

as well as to help an increasingly dubious public to gain confidence<br />

in their ability to understand—and ultimately act<br />

on—complex environmental issues. More broadly, efforts to<br />

shape the public’s perception <strong>of</strong> natural areas and to increase<br />

awareness <strong>of</strong> the linkages between human and natural systems<br />

(e.g., Foster 1999) are equally vital, and both targeted<br />

and more-generalized efforts to build public environmental<br />

literacy are important priorities for <strong>LTER</strong>.<br />

Information for the future<br />

Cyberinfrastructure describes the network <strong>of</strong> computing<br />

environments that support advanced data acquisition, storage,<br />

management, integration, mining, visualization, and<br />

other information-processing services (Atkins et al. 2003).<br />

When used for scientific purposes, cyberinfrastructure is a<br />

technical solution to the problem <strong>of</strong> efficiently connecting<br />

data, computers, and people.<br />

The development <strong>of</strong> cyberinfrastructure is integral to the<br />

success <strong>of</strong> all environmental networks, including the <strong>LTER</strong><br />

Network: Data for modeling and forecasting are essential<br />

for identifying the effects <strong>of</strong> accelerated and abrupt changes,<br />

and the explosion <strong>of</strong> real-time data availability calls for<br />

near-real-time analysis and distribution if those data are to<br />

be their most useful (AC-ERE 2009). Equally compelling<br />

is the need for data repositories and archives that allow<br />

the detection and synthesis <strong>of</strong> long-term trends and the<br />

effective integration <strong>of</strong> data from different networks and<br />

researchers.<br />

That less than 1% <strong>of</strong> ecological data is accessible after<br />

the publication <strong>of</strong> derived results (Reichman et al. 2011)<br />

reveals the social and technological challenges <strong>of</strong> curating<br />

that environmental data. Data dispersion, heterogeneity, and<br />

provenance (Jones MB et al. 2006), coupled with cultural<br />

norms that provide few rewards, make ecological information<br />

systems difficult to design, implement, and incentivize.<br />

Nevertheless, fueled by the emergence <strong>of</strong> environmental<br />

observatories charged with collecting and making openly<br />

available data from a variety <strong>of</strong> sensors (e.g., NRC 2004)<br />

and by the success <strong>of</strong> efforts to assemble and synthesize<br />

networked data toward broader-scale tests <strong>of</strong> ecological<br />

theory (e.g., Mittelbach et al. 2001, Suding et al. 2005), new<br />

efforts to develop centralized ecological information systems<br />

are under way. The <strong>LTER</strong> Network, with its 30 years <strong>of</strong><br />

experience in environmental information management, has<br />

been a pioneer in these efforts—a microcosm <strong>of</strong> the hard<br />

challenges and substantive benefits <strong>of</strong> networked ecological<br />

data—and will be among those linked by centralized efforts<br />

such as DataONE (Michener et al. 2011).<br />

The distributed data repositories <strong>of</strong> the 26 network sites<br />

reflect the vast diversity <strong>of</strong> ecological data and the breadth<br />

<strong>of</strong> approaches to environmental data-management systems<br />

as developed by field stations, museums, academic institutions,<br />

state and local governments, and individual scientists.<br />

Core <strong>LTER</strong> Network data conform to consistent metadata<br />

standards (Michener 2006) and are held by individual sites<br />

within Web-accessible catalogs. These data repositories,<br />

plus a network-wide policy <strong>of</strong> open data access, make data<br />

available to those wishing to assemble cross-site syntheses.<br />

Although open access has been crucial to the success <strong>of</strong><br />

cross-site studies such as those noted earlier (e.g., Mittelbach<br />

et al. 2001, Parton et al. 2007), the structure <strong>of</strong> site data <strong>of</strong>ten<br />

differs from catalog to catalog, which makes the discovery<br />

and subsequent integration <strong>of</strong> semantically similar data a<br />

task that is, at best, inconvenient. What is needed is a central<br />

repository that maintains the veracity and provenance <strong>of</strong> site<br />

data but allows single-portal access.<br />

Early examples include the climate and hydrology portals<br />

for <strong>LTER</strong> data. Centralized access to data from 26 sites provides<br />

an ability to detect and synthesize patterns and trends<br />

without the pain <strong>of</strong> querying 26 separate data catalogs with<br />

different keyword vocabularies and reporting units. Lowered<br />

transaction costs makes synthesis practical—and, in some<br />

cases, possible—when it had not been so previously, providing<br />

in this case a capacity to integrate multiple climate- and<br />

hydrologic-system components across disparate ecosystems<br />

and biomes to provide novel insights (Jones JA et al. 2012<br />

[in this issue]). What we need next is an ability to perform<br />

these syntheses for all system components. The <strong>LTER</strong><br />

Network Information System (NIS) is being developed to<br />

address this need.<br />

The <strong>LTER</strong> NIS will provide access to data from the<br />

26 network sites through a single point <strong>of</strong> access and at the<br />

same time ensure the long-term preservation <strong>of</strong> site data<br />

through centralized stewardship. Site data will continue to<br />

be curated at individual sites but will also be exposed to<br />

harvest by the NIS on a frequent, periodic basis. Further<br />

data processing will then provide common formats that<br />

can be easily queried by cross-network portals, such as<br />

EcoTrends (Peters et al. 2011), that are designed to create<br />

derived long-term data products and by storage systems<br />

such as DataONE that will provide a centralized facility<br />

for storing data from multiple sources, including the NIS.<br />

Importantly, the system will be scalable: Adding data from<br />

additional sites, whether they are existing field stations, sensor<br />

networks, future <strong>LTER</strong> sites, or individual field projects,<br />

will be straightforward.<br />

The nature <strong>of</strong> <strong>LTER</strong> data—and by extension, ecological<br />

data in general—makes a single rigid method for storing<br />

and accessing them impractical. By design, many long-term<br />

data sets are collected using common protocols at regular<br />

intervals from specified locations. As priorities, resources,<br />

and technologies shift, however, intervals change, protocols<br />

350 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


are improved, and locations sometimes become inappropriate<br />

or insufficient. Robust sampling programs will have<br />

precautions and methods in place to protect the veracity<br />

<strong>of</strong> long-term observations, including a data-management<br />

system sufficiently nimble to allow these changes. An additional<br />

challenge in ecological science, however, is archiving<br />

and exposing experimental data—data that may be collected<br />

over a short term, with additional protocols and different<br />

experimental treatments in a sampling matrix that may not<br />

correspond much with that <strong>of</strong> the long-term collection.<br />

This adds an additional important burden on information<br />

systems that aspire to address ecological science needs but<br />

also provides an invaluable opportunity for future users to<br />

query the full suite <strong>of</strong> observations available for a particular<br />

site or region.<br />

Informal users also need to be accommodated. The best<br />

information system will make derived products available<br />

to a variety <strong>of</strong> potential users—not just scientists but also<br />

educators, students, decisionmakers, and the public. So<br />

long as data within the system are fully exposed to all portal<br />

developers, this accommodation will be straightforward, as<br />

might be the accommodation <strong>of</strong> data from nontraditional,<br />

more-uncertain sources, such as citizen science networks<br />

(Cohn 2008).<br />

Conclusions<br />

The US <strong>LTER</strong> Network enters its fourth decade with a<br />

sound record <strong>of</strong> scientific achievement in the ecological sciences.<br />

At each <strong>of</strong> the network’s 26 sites, we have learned an<br />

extraordinary amount about the organisms and processes<br />

important at the biome it represents, about the way the<br />

site’s ecosystems respond to disturbance, and about human<br />

influences and long-term environmental change. Cross-site<br />

observations and experiments are increasingly revealing<br />

how key processes, organisms, and ecological attributes<br />

are organized and behave across major environmental<br />

gradients. In total, research in the <strong>LTER</strong> portfolio is contributing<br />

substantially to our basic knowledge <strong>of</strong> ecological<br />

interactions and to our ability to forecast change and test<br />

ecological theory.<br />

Against this backdrop, the <strong>LTER</strong> Network is undertaking a<br />

new kind <strong>of</strong> transdisciplinary science—one that ranges from<br />

local to global in scope, that blends ecological and social<br />

science theories, methods, and interpretations in order to<br />

better understand and forecast environmental change in an<br />

era when no ecosystem on Earth is free from human influence.<br />

Furthermore, the network is increasingly focused on<br />

conveying those results to an engaged audience <strong>of</strong> decisionmakers<br />

that can apply it. The <strong>LTER</strong> Network’s PPD model<br />

provides a unifying framework for better understanding<br />

coupled natural–human systems across regions and temporal<br />

scales and a means for examining feedbacks and testing<br />

hypotheses about, first, how humans perceive the critical<br />

services provided by ecosystems; second, how these perceptions<br />

change behaviors and institutions; and third, how<br />

these changes in turn feed back to affect ecosystem structure<br />

Articles<br />

and function and the ability <strong>of</strong> these systems to indefinitely<br />

sustain their delivery <strong>of</strong> services.<br />

Environmental literacy is an important ongoing legacy<br />

<strong>of</strong> <strong>LTER</strong> Network science and will remain so. Learners<br />

at all levels have benefited from <strong>LTER</strong> Network involvement:<br />

graduate and undergraduate students, K–12 students<br />

and educators, working pr<strong>of</strong>essionals involved in land<br />

and resource management, policymakers, and the general<br />

public. Future efforts will be directed toward ensuring<br />

that these groups understand the linkages and feedbacks<br />

between social and ecological systems to better inform their<br />

ability to make evidence-based environmental decisions at<br />

all levels.<br />

Advances in cyberinfrastructure are required in order to<br />

manage and organize the rapidly growing volume <strong>of</strong> ecological<br />

information and in order to enable integration and synthesis<br />

<strong>of</strong> that information over time. The <strong>LTER</strong> Network has<br />

led the ecological community in developing protocols and<br />

practices for documenting, curating, and sharing data, and<br />

it is now building the NIS, which will collect and curate data<br />

from <strong>LTER</strong> Network and other sites for storage in formats<br />

that can be queried by applications built to provide users<br />

with derived long-term data. Data in the system will thus be<br />

available to scientists, educators, students, decisionmakers,<br />

and the public for research, decision support, teaching, and<br />

informal education opportunities.<br />

The <strong>LTER</strong> Network’s primary mission is to use long-term<br />

observations and experiments to generate and test ecological<br />

theory at local to regional scales. Progress in solving environmental<br />

problems that today seem intractable depends<br />

on fundamental, long-term, integrated research that will<br />

generate a synthetic understanding <strong>of</strong> highly dynamic socioecological<br />

systems. Likewise, the early discovery <strong>of</strong> tomorrow’s<br />

surprises depends on long-term research that provides a<br />

capacity to detect new trends. Extending these capacities to<br />

continental scales will provide the necessary experimental<br />

context within which to address the causes and consequences<br />

<strong>of</strong> change documented both by the <strong>LTER</strong> Network and by<br />

the emerging constellation <strong>of</strong> environmental observatories.<br />

Acknowledgments<br />

The <strong>LTER</strong> Network owes its success to the several thousand<br />

scientists who have used its sites and data to conduct<br />

groundbreaking ecological research and to the support and<br />

leadership provided by the National Science Foundation and<br />

state and federal agency partners. The network’s principal<br />

partners include the US Forest Service, the Agricultural<br />

Research Service <strong>of</strong> the US Department <strong>of</strong> Agriculture, the<br />

US Fish and Wildlife Service’s Bureau <strong>of</strong> Land Management,<br />

and the US Geological Survey. We thank three anonymous<br />

reviewers for insightful comments on an earlier version <strong>of</strong><br />

this article.<br />

References cited<br />

[AC-ERE] US National Science Foundation Advisory Committee for<br />

Environmental Research and Education. 2009. Transitions and Tipping<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 351


Articles<br />

Points in Complex Environmental Systems. US National Science<br />

Foundation.<br />

Adler PB, et al. 2011. Productivity is a poor predictor <strong>of</strong> plant species<br />

richness. Science 333: 1750–1753.<br />

Anth<strong>of</strong>f D, Nicholls RJ, Tol RSJ. 2010. The economic impact <strong>of</strong> substantial<br />

sea-level rise. Mitigation and Adaptation Strategies for Global Change<br />

15: 321–335.<br />

Atkins DE, Droegemeier KK, Feldman SI, Garcia-Molina H, Klein<br />

ML, Messerschmitt DG, Messina P, Ostriker JP, Wright MH. 2003.<br />

Revolutionizing Science and Engineering through Cyberinfrastructure.<br />

National Science Foundation. Report no. cise051203.<br />

Baker JP, Hulse DW, Gregory SV, White D, Van Sickle J, Berger PA, Dole D,<br />

Schumaker NH. 2004. Alternative futures for the Willamette River<br />

basin, Oregon. Ecological Applications 14: 313–324.<br />

Beaulieu JJ, et al. 2011. Nitrous oxide emission from denitrification in<br />

stream and river networks. Proceedings <strong>of</strong> the National Academy <strong>of</strong><br />

Sciences 108: 214–219.<br />

Callahan JT. 1984. Long-term ecological research. BioScience 34:<br />

363–367.<br />

Carpenter SR. 2008. Emergence <strong>of</strong> ecological networks. Frontiers in Ecology<br />

and the Environment 6: 228.<br />

Clark CM, Cleland EE, Collins SL, Fargione JE, Gough L, Gross KL,<br />

Pennings SC, Suding KN, Grace JB. 2007. Environmental and plant<br />

community determinants <strong>of</strong> species loss following nitrogen enrichment.<br />

Ecology Letters 10: 596–607.<br />

Cleland EE, et al. 2008. Species responses to nitrogen fertilization in<br />

herbaceous plant communities, and associated species traits. Ecology<br />

89: 1175.<br />

Cleland EE, Clark CM, Collings SL, Fargione JE, Gough L, Gross KL,<br />

Pennings SC, Suding KN. 2011. Natural patterns <strong>of</strong> invasion in herbaceous<br />

plant communities are related to soil nitrogen and the functional<br />

similarity <strong>of</strong> native species. Journal <strong>of</strong> Ecology 99: 1327–1338.<br />

Cohn JP. 2008. Citizen science: Can volunteers do real research? BioScience<br />

58: 192–197.<br />

Collins SL, et al. 2011. An integrated conceptual framework for<br />

social–ecological research. Frontiers in Ecology and the Environment<br />

9: 351–357.<br />

Driscoll CT, Lambert KF, Chapin FS III, Nowak DJ, Spies TA, Swanson FJ,<br />

Kittredge DB Jr, Hart CM. 2012. Science and society: The role <strong>of</strong> longterm<br />

studies in environmental stewardship. BioScience 62: 354–366.<br />

Duschl RA, Schweingruber HA, Shouse AW, eds. 2007. Taking Science<br />

to School: Learning and Teaching Science in Grades K–8. National<br />

Academies Press.<br />

[ESA] Ecological Society <strong>of</strong> America. 2006. Women and Minorities in<br />

Ecology II (WAMIE II): Committee Report, March 2006. ESA.<br />

Evans TP, Kelley H. 2004. Multi-scale analysis <strong>of</strong> a household level<br />

agent-based model <strong>of</strong> landcover change. Journal <strong>of</strong> Environmental<br />

Management 72: 57–72.<br />

Foster DR. 1999. Thoreau’s Country: Journey through a Transformed<br />

Landscape. Harvard <strong>University</strong> Press.<br />

Fountain AG, Campbell JL, Schuur EAG, Stammerjohn SE, Williams MW,<br />

Ducklow HW. 2012. The disappearing cryosphere: Impacts and ecosystem<br />

responses to rapid cryosphere loss. BioScience 62: 405–415.<br />

Franklin JF, Bledsoe CS, Callahan JT. 1990. Contributions <strong>of</strong> the Long-Term<br />

Ecological Research Program. BioScience 40: 509–523.<br />

George YS, Neale DS, Van Horne V, Malcom SM. 2001. In Pursuit<br />

<strong>of</strong> a Diverse Science, Technology, Engineering, and Mathematics<br />

Workforce: Recommended Research Priorities to Enhance Participation<br />

by Underrepresented Minorities. American Association for the<br />

Advancement <strong>of</strong> Science.<br />

Gholz HL, Wedin DA, Smitherman SM, Harmon M, Parton WJ. 2000.<br />

Long-term dynamics <strong>of</strong> pine and hardwood litter in contrasting environments:<br />

Toward a global model <strong>of</strong> decomposition. Global Change<br />

Biology 6: 751–766.<br />

Gonzales P, Williams T, Jocelyn L, Roey S, Kastberg D, Brenwald S. 2008.<br />

Highlights from TIMSS 2007: Mathematics and Science Achievement<br />

<strong>of</strong> U.S. Fourth- and Eighth-Grade Students in an International Context.<br />

National Center for Education Statistics, US Department <strong>of</strong> Education.<br />

Report no. NCES 2009-001 Revised.<br />

Gr<strong>of</strong>fman PM, Stylinski C, Nisbet MC, Duarte CM, Jordan R, Burgin A,<br />

Previtali MA, Coloso J. 2010. Restarting the conversation: Challenges at<br />

the interface between ecology and society. Frontiers in Ecology and the<br />

Environment 8: 284–291.<br />

Hartley LM, Wilke BJ, Schramm JW, D’Avanzo C, Anderson CW. 2011.<br />

College students’ understanding <strong>of</strong> the carbon cycle: Contrasting<br />

principle-based and informal reasoning. BioScience 61: 65–75.<br />

Helton AM, et al. 2011. Thinking outside the channel: Modeling nitrogen<br />

cycling in networked river ecosystems. Frontiers in Ecology and the<br />

Environment 9: 229–238.<br />

Hoag DL, Keske-Handley C, Ascough II J, Koontz L. 2005. Decision making<br />

with environmental indices. Pages 159–182 in Burk AR, ed. New Trends<br />

in Ecology Research. Nova Science.<br />

Hobbie JE, Carpenter SR, Grimm NB, Gosz JR, Seastedt TR. 2003. The US<br />

Long Term Ecological Research Program. BioScience 53: 21–32.<br />

Isern AR, Clark HL. 2003. The ocean observatories initiative: A continued<br />

presence for interactive ocean research. Marine Technology Society<br />

Journal 37: 26–41.<br />

Johnson JC, Christian RR, Brunt JW, Hickman CR, Waide RB. 2010.<br />

Evolution <strong>of</strong> collaboration within the US Long Term Ecological<br />

Research Network. BioScience 60: 931–940.<br />

Jones JA, et al. 2012. Ecosystem processes and human influences regulate<br />

streamflow response to climate change at long-term ecological research<br />

sites. BioScience 62: 390–404.<br />

Jones MB, Schildhauer MP, Reichman OJ, Bowers S. 2006. The new bioinformatics:<br />

Integrating ecological data from the gene to the biosphere.<br />

Annual Review <strong>of</strong> Ecology, Evolution, and Systematics 37: 519–544.<br />

Kane ES, et al. 2008. Precipitation control over inorganic nitrogen import–<br />

export budgets across watersheds: A synthesis <strong>of</strong> long-term ecological<br />

research. Ecohydrology 1: 105–117.<br />

Keller M, Schimel DS, Hargrove WW, H<strong>of</strong>fman FM. 2008. A continental<br />

strategy for the National Ecological Observatory Network. Frontiers in<br />

Ecology and the Environment 6: 282–284.<br />

Knapp AK, et al. 2012. Past, present, and future roles <strong>of</strong> long-term<br />

experiments in the <strong>LTER</strong> network. BioScience 62: 377–389.<br />

Kratz TK, et al. 2006. Toward a global lake ecological observatory network.<br />

Publications <strong>of</strong> the Karelian Institute 145: 51–63.<br />

Likens GE. 2010. The role <strong>of</strong> science in decision making: Does evidencebased<br />

science drive environmental policy? Frontiers in Ecology and the<br />

Environment 8: e1–e9.<br />

Liu J, et al. 2008. Complexity <strong>of</strong> coupled human and natural systems.<br />

Science 317: 1513–1516.<br />

Magnuson JJ, et al. 2000. Historical trends in lake and river ice cover in the<br />

Northern Hemisphere. Science 289: 1743–1746.<br />

McKnight DM. 2010. Overcoming “ecophobia”: Fostering environmental<br />

empathy through narrative in children’s science literature. Frontiers in<br />

Ecology and the Environment 8: e10–e15.<br />

Michener WK. 2006. Meta-information concepts for ecological data<br />

management. Ecological Informatics 1: 3–7.<br />

Michener WK, et al. 2011. DataONE: Data Observation Network for<br />

Earth—Preserving data and enabling innovation in the biological and<br />

environmental sciences. D-Lib Magazine 2011, 17: 1–9. (17 January<br />

2012; www.dlib.org/dlib/january11/michener/01michener.html)<br />

Mittelbach GG, Steiner CF, Scheiner SM, Gross KL, Reynolds HL, Waide<br />

RB, Willig MR, Dodson SI, Gough L. 2001. What is the observed<br />

relationship between species richness and productivity? Ecology 82:<br />

2381–2396.<br />

Mulholland PJ, et al. 2008. Stream denitrification across biomes and its<br />

response to anthropogenic nitrate loading. Nature 452: 202–205.<br />

[NRC] National Research Council. 2004. NEON: Addressing the Nation’s<br />

Environmental Challenges. National Academies Press.<br />

———. 2009. Learning science in informal environments: People, places,<br />

and pursuits. National Academies Press.<br />

Olson DM, et al. 2001. Terrestrial ecoregions <strong>of</strong> the world: A new map <strong>of</strong> life<br />

on earth. BioScience 51: 933–938.<br />

352 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


Osmond DL, et al. 2010. The role <strong>of</strong> interface organizations in science<br />

communication and understanding. Frontiers in Ecology and the<br />

Environment 8: 306–313.<br />

Parton W, et al. 2007. Global-scale similarities in nitrogen release patterns<br />

during long-term decomposition. Science 315: 361–364.<br />

Pennings SC, Clark CM, Cleland EE, Collins SL, Gough L, Gross KL,<br />

Milchunas DG, Suding KM. 2005. Do individual plant species show<br />

predictable responses to nitrogen addition across multiple experiments?<br />

Oikos 110: 547–555.<br />

Peters DPC, Gr<strong>of</strong>fman PM, Nadelh<strong>of</strong>fer KJ, Grimm NB, C<strong>of</strong>fins SL,<br />

Michener WK, Huston MA. 2008. Living in an increasingly connected<br />

world: A framework for continental-scale environmental science.<br />

Frontiers in Ecology and the Environment 6: 229–237.<br />

Peters DPC, et al. 2011. Long-term Trends in Ecological Systems: A Basis<br />

for Understanding Responses to Global Change. US Department <strong>of</strong><br />

Agriculture, Agricultural Research Service.<br />

Peterson BJ, et al. 2001. Control <strong>of</strong> nitrogen export from watersheds by<br />

headwater streams. Science 292: 86–90.<br />

Redman CL, Grove JM, Kuby LH. 2004. Integrating social science into<br />

the Long Term Ecological Research (<strong>LTER</strong>) Network: Social dimensions<br />

<strong>of</strong> ecological change and ecological dimensions <strong>of</strong> social change.<br />

Ecosystems 7: 161–171.<br />

Reichman OJ, Jones MB, Schildhauer MP. 2011. Challenges and opportunities<br />

<strong>of</strong> open data in ecology. Science 331: 703–705.<br />

Robertson GP. 2008. Long-term ecological research: Re-inventing network<br />

science. Frontiers in Ecology and the Environment 6: 281.<br />

Smith MD, Knapp AK, Collins SL. 2009. A framework for assessing ecosystem<br />

dynamics in response to chronic resource alterations induced by<br />

global change. Ecology 90: 3279–3289.<br />

Solomon S, Plattner G-K, Knutti R, Friedlingstein P. 2009. Irreversible<br />

climate change due to carbon dioxide emissions. Proceedings <strong>of</strong> the<br />

National Academy <strong>of</strong> Sciences 106: 1704–1709.<br />

Suding KN, Collins SL, Gough L, Clark C, Cleland EE, Gross KL, Milchunas<br />

DG, Pennings S. 2005. Functional- and abundance-based mechanisms<br />

explain diversity loss due to N fertilization. Proceedings <strong>of</strong> the National<br />

Academy <strong>of</strong> Sciences 102: 4387–4392.<br />

Thompson JR, Foster DR, Scheller R, Kittredge D. 2011 The influence <strong>of</strong><br />

land use and climate change on forest biomass and composition in<br />

Massachusetts, USA. Ecological Applications 21: 2425–2444.<br />

Articles<br />

Thompson JR, Wiek A, Swanson FJ, Carpenter SR, Fresco N, Hollingsworth<br />

T, Spies TA, Foster DR. 2012. Scenario studies as a synthetic and integrative<br />

research activity for long-term ecological research. BioScience<br />

62: 367–376.<br />

Vanneman A, Hamilton L, Anderson JB. 2009. Achievement Gaps: How<br />

Black and White Students in Public Schools Perform in Mathematics<br />

and Reading on the National Assessment <strong>of</strong> Educational Progress.<br />

National Center for Education Statistics, US Department <strong>of</strong> Education.<br />

Report no. NCES 2009-455.<br />

Waide RB, Willig MR, Steiner CF, Mittelbach G, Gough L, Dodson GI, Juday<br />

GP, Parmenter R. 1999. The relationship between primary productivity<br />

and species richness. Annual Review <strong>of</strong> Ecology and Systematics 30:<br />

257–300.<br />

G. Philip Robertson (robertson@kbs.msu.edu) is a pr<strong>of</strong>essor with the W. K.<br />

Kellogg Biological Station and with the Department <strong>of</strong> Crop and Soil Sciences<br />

at Michigan State <strong>University</strong>, in Hickory Corners. Scott L. Collins is a pr<strong>of</strong>essor<br />

in the Department <strong>of</strong> Biology at the <strong>University</strong> <strong>of</strong> New Mexico, in<br />

Albuquerque. David R. Foster is a pr<strong>of</strong>essor in the Department <strong>of</strong> Organismic<br />

and Evolutionary Biology at Harvard <strong>University</strong> and director <strong>of</strong> the Harvard<br />

Forest in Petersham, Massachusetts. Nicholas Brokaw is a pr<strong>of</strong>essor with the<br />

Institute for Tropical Ecosystem Studies at the <strong>University</strong> <strong>of</strong> Puerto Rico, in San<br />

Juan. Hugh W. Ducklow is a senior scientist and director <strong>of</strong> The Ecosystems<br />

Center at the Marine Biological Laboratory in Woods Hole, Massachusetts.<br />

Ted L. Gragson is a pr<strong>of</strong>essor and chair <strong>of</strong> the Department <strong>of</strong> Anthropology<br />

at the <strong>University</strong> <strong>of</strong> <strong>Georgia</strong>, in Athens. Corinna Gries is a research scientist<br />

at the Center for Limnology at the <strong>University</strong> <strong>of</strong> Wisconsin, in Madison.<br />

Stephen K. Hamilton is a pr<strong>of</strong>essor in the Department <strong>of</strong> Zoology and the<br />

W. K. Kellogg Biological Station <strong>of</strong> Michigan State <strong>University</strong>, in Hickory<br />

Corners. A. David McGuire is a pr<strong>of</strong>essor in the Department <strong>of</strong> Biology<br />

and Wildlife at the <strong>University</strong> <strong>of</strong> Alaska, in Fairbanks, and is affiliated with<br />

the US Geological Survey. John C. Moore is a pr<strong>of</strong>essor and director <strong>of</strong> the<br />

Natural Resource Ecology Laboratory at Colorado State <strong>University</strong>, in Fort<br />

Collins. Emily H. Stanley is a pr<strong>of</strong>essor with the Center for Limnology at the<br />

<strong>University</strong> <strong>of</strong> Wisconsin, Madison. Robert B. Waide is pr<strong>of</strong>essor <strong>of</strong> <strong>biology</strong> at<br />

the <strong>University</strong> <strong>of</strong> New Mexico, in Albuquerque, and executive director <strong>of</strong> the<br />

Long Term Ecological Research Network. Mark W. Williams is a pr<strong>of</strong>essor in<br />

the Department <strong>of</strong> Geography at the <strong>University</strong> <strong>of</strong> Colorado, in Boulder.<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 353


Articles<br />

Science and Society: The Role <strong>of</strong><br />

Long-Term Studies in Environmental<br />

Stewardship<br />

Charles T. DrisColl, KaThleen F. lamberT, F. sTuarT Chapin iii, DaviD J. nowaK, Thomas a. spies,<br />

FreDeriCK J. swanson, DaviD b. KiTTreDge Jr., anD Clarisse m. harT<br />

Long-term research should play a crucial role in addressing grand challenges in environmental stewardship. We examine the efforts <strong>of</strong> five Long<br />

Term Ecological Research Network sites to enhance policy, management, and conservation decisions for forest ecosystems. In these case studies,<br />

we explore the approaches used to inform policy on atmospheric deposition, public land management, land conservation, and urban forestry,<br />

including decisionmaker engagement and integration <strong>of</strong> local knowledge, application <strong>of</strong> models to analyze the potential consequences <strong>of</strong> policy<br />

and management decisions, and adaptive management to generate new knowledge and incorporate it into decisionmaking. Efforts to enhance<br />

the role <strong>of</strong> long-term research in informing major environmental challenges would benefit from the development <strong>of</strong> metrics to evaluate impact;<br />

stronger partnerships among research sites, pr<strong>of</strong>essional societies, decisionmakers, and journalists; and greater investment in efforts to develop,<br />

test, and expand practice-based experiments at the interface <strong>of</strong> science and society.<br />

Keywords: boundary spanning, environmental policy and management, Long Term Ecological Research Network, science communication<br />

The growing urgency and complexity <strong>of</strong> challenges to<br />

global sustainability demands new approaches for engaging<br />

the intellectual capital <strong>of</strong> expert communities worldwide.<br />

To meet this demand, the scientific and social-science communities<br />

must expand their capacity to work at the interfaces<br />

between ecological science and environmental policy,<br />

natural-resource management, and conservation. The need<br />

for stronger, more-reliable linkages between science and<br />

society is well documented in both popular media and the<br />

academic literature (e.g., Lubchenco 1998).<br />

The US National Science Foundation (NSF) recognized<br />

the importance <strong>of</strong> translating the benefits <strong>of</strong> research for<br />

society by establishing its “broader impacts” review criterion<br />

in 1997. In 2002, the NSF’s 20-year review <strong>of</strong> the Long Term<br />

Ecological Research (<strong>LTER</strong>) Network recommended that<br />

the <strong>LTER</strong> Network program assume a more powerful and<br />

pervasive role in informing environmental solutions at local,<br />

national, and international levels. In 2005, the Ecological<br />

Society <strong>of</strong> America (ESA) established the foundation for<br />

its Earth Stewardship initiative (Chapin et al. 2011) when it<br />

recommended that ecologists must play a greatly expanded<br />

role in communicating their research and influencing policies<br />

and decisions that affect the environment. It is critical<br />

for the <strong>LTER</strong> Network and for other ecosystem research<br />

programs to move beyond broad calls for action and to build<br />

deliberate and effective long-term relationships between<br />

ecological science and environmental decisionmaking. In<br />

this article, we illustrate, as do the other authors in this<br />

special section (e.g., Thompson et al. 2012 [in this issue]),<br />

the growing role that <strong>LTER</strong> is playing to enhance science<br />

engagement with local, regional, and national policy and<br />

management issues. To develop such programs, the scientific<br />

community needs to build experience and learn from<br />

practical examples <strong>of</strong> effective synthesis and integration <strong>of</strong><br />

<strong>LTER</strong> to meet the needs <strong>of</strong> society. In this article, we present<br />

and discuss five case studies <strong>of</strong> work at the interface <strong>of</strong> science,<br />

policy, and management from forested <strong>LTER</strong> Network<br />

sites across the United States. We distill a set <strong>of</strong> common<br />

strategies, lessons, and recommendations for improving and<br />

expanding interface efforts to improve the ability to meet the<br />

grand challenges in environmental science <strong>of</strong> our time.<br />

Effective science interface efforts<br />

Although the integration <strong>of</strong> science and society is <strong>of</strong>ten<br />

viewed as a relatively narrow issue <strong>of</strong> a need for more and<br />

better science communication, programs that build stronger<br />

interfaces between science and society require attention to<br />

the full range <strong>of</strong> boundary-spanning activities, such as public<br />

engagement, decision-relevant synthesis, distillation <strong>of</strong><br />

results, and science translation and dissemination, through<br />

BioScience 62: 354–366. ISSN 0006-3568, electronic ISSN 1525-3244. © 2012 by American Institute <strong>of</strong> Biological Sciences. All rights reserved. Request<br />

permission to photocopy or reproduce article content at the <strong>University</strong> <strong>of</strong> California Press’s Rights and Permissions Web site at www.ucpressjournals.com/<br />

reprintinfo.asp. doi:10.1525/bio.2012.62.4.7<br />

354 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


a variety <strong>of</strong> media to meet the needs <strong>of</strong> diverse audiences<br />

(Cash et al. 2003, Driscoll et al. 2011). Boundary spanning<br />

refers to “practices and processes that facilitate bringing<br />

science and society closer together in order to produce<br />

‘useful’ information—that is, information that is salient,<br />

credible, and legitimate” (McNie et al. 2008, p. 9). Building<br />

credibility, salience, and legitimacy with stakeholders helps<br />

to solidify long-term relationships and increases the influence<br />

<strong>of</strong> scientific research in the decisionmaking process<br />

over time (Cash et al. 2003).<br />

The role <strong>of</strong> the <strong>LTER</strong> Network<br />

After 30 years <strong>of</strong> coordinated research and education, the<br />

<strong>LTER</strong> Network is well positioned to facilitate the integration<br />

<strong>of</strong> science and society by using its highly credible, longterm<br />

science to support engagement with decisionmakers<br />

to frame relevant questions for research and synthesis that<br />

can inform environmental policy and conservation. The<br />

<strong>LTER</strong> Network consists <strong>of</strong> 26 research sites throughout the<br />

United States and a few outside the United States, some <strong>of</strong><br />

which have been operating for three decades or longer. The<br />

long-term ecosystem measurements and experiments that<br />

are a hallmark <strong>of</strong> the <strong>LTER</strong> Network address important<br />

environmental issues in coupled human–natural systems,<br />

including climate change, land use, pollution, and the loss<br />

<strong>of</strong> biodiversity (Knapp et al. 2012 [in this issue], Thompson<br />

et al. 2012 [in this issue]). Another distinguishing feature <strong>of</strong><br />

the <strong>LTER</strong> Network is its core <strong>of</strong> researchers at each site, who<br />

are attentive to the well-being and future <strong>of</strong> their respective<br />

bioregions.<br />

The <strong>LTER</strong> Network’s new Strategic Communication Plan<br />

(<strong>LTER</strong> Network 2010a) and Strategic and Implementation<br />

Plan (<strong>LTER</strong> Network 2010b) call for the network to reach<br />

out to decisionmakers at local, regional, national, and international<br />

levels. The communication plan specifically recommends<br />

engaging decisionmakers in the framing <strong>of</strong> cross-site<br />

synthesis and equipping these efforts with full-scale communication<br />

capacity, funding a supplement program for<br />

<strong>LTER</strong> Network sites to develop local and regional programs<br />

for public engagement and outreach, and partnering with<br />

existing <strong>LTER</strong> Network sites that have established science<br />

journalism programs to develop sustained outreach to the<br />

media.<br />

The value <strong>of</strong> long-term monitoring and research<br />

Environmental policy and management issues play out over<br />

decades or longer and benefit from the continuous advances<br />

in understanding that are derived from long-term research.<br />

Policy development is an iterative process that requires<br />

ongoing assessment, reevaluation, adaptive management,<br />

and consideration <strong>of</strong> future scenarios (Driscoll et al. 2010).<br />

For example, although the Clean Air Act was first passed<br />

by Congress in 1972, the development <strong>of</strong> amendments and<br />

rules to implement the act are ongoing and rely on quantitative<br />

information to evaluate the effectiveness <strong>of</strong> pollution-<br />

control measures and to guide program management (Lovett<br />

Articles<br />

et al. 2007). Long-term measurements that link decreases in<br />

emissions with changes in soil and water quality and the<br />

health <strong>of</strong> aquatic and terrestrial ecosystems are vital to<br />

assessing the extent to which air pollution regulations meet<br />

the intent <strong>of</strong> the act (Driscoll et al. 2001).<br />

Similarly, effective natural-resource management is adaptive<br />

and draws on lessons from past decisions and management<br />

experience distilled from the results <strong>of</strong> long-term<br />

measurements and experiments, regional surveys, and<br />

modeling (Spies et al. 2010). The practice <strong>of</strong> forestry in landscapes<br />

that support multiple uses must adapt to new knowledge<br />

regarding the nature and effects <strong>of</strong> climate change,<br />

forest management, land-use trends, intense storms, fire,<br />

and other disturbances. This understanding must include<br />

the impacts <strong>of</strong> these <strong>of</strong>ten interacting pulse and press stressors<br />

on management goals and ecosystem services, such<br />

as fiber production, biological diversity, carbon storage,<br />

trace-gas production and consumption, water quantity and<br />

quality, and recreation. Detailed, long-term measurements<br />

tied directly to management-relevant forest experiments<br />

have improved the scientific basis for forest management<br />

and policy. Important examples include the guiding principles<br />

for the conservation <strong>of</strong> old-growth forests (Franklin<br />

et al. 1981) and regional- and continental-scale carbon<br />

budgets important to climate-change mitigation (Lovett<br />

et al. 2007).<br />

The five case studies presented here represent examples <strong>of</strong><br />

outreach activities at selected forested <strong>LTER</strong> Network sites.<br />

We chose this suite <strong>of</strong> case studies because they have active<br />

programs for engaging decisionmakers, represent a range <strong>of</strong><br />

policy and management issues, and use different approaches<br />

to achieve their outreach goals for a common ecosystem type.<br />

Reviewing efforts across forest sites provides the opportunity<br />

to consider how audiences, management and policy<br />

issues, and communication approaches vary across diverse<br />

regional research sites and programs. Specifically, the case<br />

studies incorporate the impacts <strong>of</strong> atmospheric deposition<br />

on forested ecosystems (Hubbard Brook), land-use change<br />

and forest conservation in a predominantly private-lands<br />

landscape (Harvard Forest), endangered species and public<br />

lands management (Andrews), urban forestry in developed<br />

landscapes (Baltimore), and forest stewardship in<br />

the context <strong>of</strong> changing fire and climate regimes (Bonanza<br />

Creek). These case studies represent only some <strong>of</strong> the many<br />

science–policy integration efforts that exist across the <strong>LTER</strong><br />

Network (for other examples, see the Translating Science<br />

for Society brochure at http://intranet2.lternet.edu/sites/<br />

intranet2.l ternet.edu/files/documents/Network_Publications/<br />

Brochures/nsf0533.pdf ). In each <strong>of</strong> these cases, the ability<br />

to tap into core strengths <strong>of</strong> the <strong>LTER</strong> Network, such as<br />

long-term research that is relevant to policy and management<br />

issues, advanced information-management systems,<br />

and stores <strong>of</strong> long-term data, has proven essential to the<br />

synthesis and distillation <strong>of</strong> science for use in policy and<br />

management decisions related to coupled human–natural<br />

systems.<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 355


Articles<br />

Case studies in linking <strong>LTER</strong> science with policy,<br />

conservation, and management<br />

Below, we describe several case studies that link <strong>LTER</strong> science<br />

with policy, conservation, and management.<br />

Air pollution effects on ecosystems: The Hubbard Brook Research<br />

Foundation Science Links Program. Air pollution can have<br />

marked effects on the structure and function <strong>of</strong> ecosystems<br />

through elevated atmospheric deposition <strong>of</strong> sulfur, oxidized<br />

and reduced nitrogen compounds and mercury, and<br />

high concentrations <strong>of</strong> tropospheric ozone. Recent efforts<br />

to channel this knowledge into decisionmaking through<br />

organized outreach and communication have increased<br />

the influence <strong>of</strong> long-term research on air-quality management<br />

in the United States (Driscoll et al. 2010). The <strong>LTER</strong><br />

Network, through its long-term measurements and experiments<br />

(Driscoll et al. 2001), has been particularly effective<br />

in addressing policy issues concerning air pollution and<br />

atmospheric deposition effects on ecosystems.<br />

The effects <strong>of</strong> air pollution on forest and aquatic ecosystems<br />

have been a research focus since the inception <strong>of</strong> the<br />

Hubbard Brook Ecosystem Study and the Hubbard Brook<br />

<strong>LTER</strong> site. The value <strong>of</strong> long-term measurements <strong>of</strong> the<br />

chemistry <strong>of</strong> precipitation and streamwater at the Hubbard<br />

Brook Experimental Forest in documenting trends in acidic<br />

deposition and in assessing the effectiveness <strong>of</strong> the federal<br />

Clean Air Act represents an important example <strong>of</strong> the connections<br />

between long-term research and air-quality policy<br />

(figure 1). The Hubbard Brook Research Foundation (HBRF)<br />

launched Science Links in 1998 to build on this legacy and to<br />

develop new initiatives linking ecosystem science with public<br />

policy (http://hubbardbrookfoundation.org/12-2).<br />

Science Links projects are state-<strong>of</strong>-the-science synthesis<br />

efforts <strong>of</strong> an environmental issue in the context <strong>of</strong> current<br />

policy discussions. The first three Science Links projects<br />

addressed air pollution impacts on ecosystems, including the<br />

effects <strong>of</strong> acid, nitrogen, and mercury deposition (Driscoll<br />

et al. 2001, 2011). Science Links projects involve teams <strong>of</strong><br />

around 12 scientific experts, selected on the basis <strong>of</strong> their<br />

experience and disciplinary coverage, and a team <strong>of</strong> policy<br />

advisers. The science teams define the scope <strong>of</strong> the project,<br />

analyze relevant databases and conduct model calculations.<br />

The policy advisers are engaged in dialogue from the outset<br />

to frame policy-relevant questions, discuss the alternatives<br />

analyzed, and provide input on Science Links products.<br />

A communication and outreach plan is integral to the<br />

success <strong>of</strong> Science Links projects. The written plan provides<br />

a roadmap to facilitate an exchange between scientists and<br />

policy stakeholders as well as direct outreach to journalists.<br />

The centerpiece <strong>of</strong> any Science Links project is the translation<br />

report aimed at congressional and government-agency<br />

staff involved in policy development. These reports are<br />

structured to facilitate communication <strong>of</strong> the major findings,<br />

with the conclusions presented first in clear, straightforward<br />

terms, followed by supporting information with<br />

layered details. A proactive media strategy has been critical<br />

to the impact <strong>of</strong> Science Links projects. Accurate and widespread<br />

media coverage has brought attention to Science<br />

Links results and verified the societal importance <strong>of</strong> the<br />

findings for policymakers. The initial public release <strong>of</strong> a<br />

Science Links report is followed by additional interviews,<br />

seminars, and briefings for up to a year. Science Links<br />

projects have also been coupled with the Hubbard Brook<br />

<strong>LTER</strong> site educational activities through the development<br />

<strong>of</strong> supplemental teacher guides.<br />

There are several dimensions to quantifying the impact<br />

<strong>of</strong> Science Links projects (Driscoll et al. 2011). The scientific<br />

impact can be measured by the number <strong>of</strong> citations<br />

in the scientific literature; the six Science Links journal<br />

articles have been cited more than 1300 times in the peerreviewed<br />

literature. The media impact can be measured by<br />

the extent and quality <strong>of</strong> media cover. Science Links initiatives<br />

have been covered in more than 475 media stories<br />

and have appeared in major news outlets, including an<br />

opinion–editorial piece in The New York Times. The impacts<br />

on policy are more difficult to quantify. Moreover, they are<br />

356 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org<br />

a<br />

b<br />

Figure 1. Relationships between annual volume-weighted<br />

concentrations <strong>of</strong> sulfate (a) and nitrate (b) in precipitation<br />

at the Hubbard Brook Experimental Forest and emissions<br />

<strong>of</strong> sulfur dioxide and nitrogen oxides, respectively, in the<br />

emission source area <strong>of</strong> the northeastern United States.<br />

The emission source area used is specified in Driscoll and<br />

colleagues (2001). Abbreviations: NO x , nitrogen oxides;<br />

R 2 , regression coefficient; SO 2 , sulphur dioxide; Tg/year,<br />

teragrams per year; µeq/L, microequivalencies per liter.


<strong>of</strong>ten beyond the control <strong>of</strong> the scientists, regardless <strong>of</strong> the<br />

process used to link science to policy. Timing is everything<br />

in this dance between science and management. Forms <strong>of</strong><br />

evidence <strong>of</strong> policy uptake include reference to Science Links<br />

findings in proposed legislation (e.g., the Clean Power Act,<br />

the National Mercury Monitoring Act), legal briefs (e.g., the<br />

Northeast States New Source Review case against the US<br />

Environmental Protection Agency), and media accounts<br />

<strong>of</strong> major policy and court decisions (e.g., the Interstate<br />

Transfer Rule for nitrogen oxide emissions and the remand<br />

<strong>of</strong> the Clean Air Mercury Rule and its trading provisions).<br />

Beyond this evidence, policymakers and program managers<br />

routinely comment on the usefulness <strong>of</strong> Science Links in<br />

improving the scientific basis for decisionmaking because<br />

<strong>of</strong> its reliance on rigorous long-term research and effective<br />

translation.<br />

Uniting conservation science and policy: Examples from the<br />

Harvard Forest <strong>LTER</strong> site. The Harvard Forest has oriented its<br />

Articles<br />

long-term studies around forest management and conservation<br />

questions in New England since its inception in<br />

1907. When it was established by Harvard <strong>University</strong>, its<br />

objectives were to serve as a model forest to demonstrate<br />

the practice <strong>of</strong> forestry, an experiment station for research<br />

in forestry, and a field laboratory for students (Fisher 1921).<br />

Today, Harvard Forest scientists remain dedicated to the<br />

founding tenet <strong>of</strong> drawing on insights gained through the<br />

historical and retrospective study <strong>of</strong> forests, natural disturbance,<br />

and land use (Fisher 1933, Foster 2000), and the<br />

Harvard Forest serves as a central gathering spot for meetings<br />

and workshops among forest managers, conservationists,<br />

policymakers, and scientists in the Northeast.<br />

Long-term research at Harvard Forest has informed many<br />

important conservation efforts, as well as policy and management<br />

decisions in the region (figure 2; Foster et al.<br />

2010). For instance, a review <strong>of</strong> land-ownership history<br />

and conservation patterns led to the creation <strong>of</strong> the North<br />

Quabbin Partnership and to an increase <strong>of</strong> conservation<br />

Figure 2. Harvard Forest research and conservation linkages: primary research and synthesis examples related to the<br />

Wildlands and Woodlands Initiative.<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 357


Articles<br />

land in the region from 36.8% in 1993 to 45.1% in 2010<br />

(Golodetz and Foster 1997). Research on the potential for<br />

local constraints on forestry to displace harvesting pressures<br />

to other, more sensitive parts <strong>of</strong> the world has broadened<br />

public acceptance <strong>of</strong> forestry in the region (Berlik et al. 2002).<br />

Surveys have documented how underrepresented old-growth<br />

forests are in southern New England, which has aided in the<br />

preservation <strong>of</strong> the few remaining sites (Orwig et al. 2001,<br />

D’Amato et al. 2006). Many <strong>of</strong> these linkages grew out <strong>of</strong><br />

strong informal ties between scientists and stakeholders built<br />

by serving on local, state, and regional committees.<br />

In 2005, the Harvard Forest launched its Wildlands and<br />

Woodlands (W&W) Initiative, which emphasizes decisionrelevant<br />

synthesis, communication, and stakeholder partnerships.<br />

The knowledge gained from dozens <strong>of</strong> studies at<br />

the Harvard Forest was synthesized into a series <strong>of</strong> W&W<br />

publications that were aimed at nonscientists and that called<br />

for stemming the loss <strong>of</strong> forest cover now occurring in all<br />

six New England states as large areas (e.g., in Maine) experience<br />

significant shifts in landownership. The publications<br />

call for balancing the preservation <strong>of</strong> wildlands with large<br />

areas <strong>of</strong> actively managed woodlands and for promoting<br />

civic engagement through landowner-conceived woodland<br />

councils (Foster et al. 2005, 2010).<br />

Since 2005, the W&W Initiative has produced two major<br />

reports, two update publications, and a Web site (www.<br />

wildlandsandwoodlands.org), with the purpose <strong>of</strong> raising<br />

awareness about the pace and consequences <strong>of</strong> land-cover<br />

change. Both W&W reports had extensive stakeholder input,<br />

and the second garnered comments from several hundred<br />

agency, nongovernmental-organization (NGO), landowner,<br />

and industry representatives. Harvard Forest has since<br />

teamed up with the nonpr<strong>of</strong>it organization Highstead to<br />

form a partnership with more than 60 participating groups<br />

to sustain stakeholder engagement and to help implement<br />

the vision <strong>of</strong> the W&W Initiative. The reports were accompanied<br />

by press releases; webinars; stakeholder briefings;<br />

and, in May 2010, a public event with Harvard <strong>University</strong>’s<br />

Kennedy School <strong>of</strong> Government.<br />

Assessing the societal impact <strong>of</strong> Harvard Forest research<br />

over the past 100 years is beyond the scope <strong>of</strong> this case<br />

study. However, we compiled information on the impact<br />

<strong>of</strong> W&W communication to shed light on the value <strong>of</strong> this<br />

coordinated outreach effort. In the two months following<br />

its release, the 2010 report generated 137 media and<br />

newsletter stories and 62 visits per day to the new W&W<br />

Web site, including visitors from 35 countries from five<br />

continents. By contrast, Harvard Forest garnered 21 non-<br />

W&W news stories between 2008 and 2010. W&W authors<br />

participated in 21 briefings, presentations, and workshops<br />

in the nine months since publication, which expanded the<br />

project’s influence and reach. These W&W synthesis and<br />

communication efforts have contributed to several notable<br />

policy and management advances, including the decision by<br />

the state <strong>of</strong> Massachusetts to establish permanent wildland<br />

reserves, the introduction <strong>of</strong> a conservation-finance bill in<br />

the Massachusetts General Assembly to accelerate the pace<br />

<strong>of</strong> conservation, and the launching <strong>of</strong> an innovative effort<br />

to aggregate multiple parcels into a single project with the<br />

goal <strong>of</strong> conserving approximately 10,000 acres <strong>of</strong> forest in<br />

western Massachusetts. The W&W efforts also fueled new<br />

research, including the establishment <strong>of</strong> new long-term<br />

study plots across sites with diverse histories, ownership, and<br />

management objectives; and a new cross-site <strong>LTER</strong> proposal<br />

on the Future Scenarios <strong>of</strong> Forest Change (see Thompson<br />

et al. 2012 [in this issue]).<br />

Sustained research–management partnerships at the Andrews<br />

Forest <strong>LTER</strong> site. The H. J. Andrews Experimental Forest and<br />

<strong>LTER</strong> site in the Oregon Cascade Range contains many <strong>of</strong><br />

the iconic and hotly debated elements <strong>of</strong> Pacific Northwest<br />

forests: old-growth trees; northern spotted owls; and cold,<br />

clear, fast streams. Societal conflicts over the future <strong>of</strong> the<br />

vast tracts <strong>of</strong> federal forestlands in the region have been<br />

pr<strong>of</strong>oundly affected by science findings from the Andrews<br />

Forest and, in turn, have strongly influenced the course <strong>of</strong><br />

science in the region and more broadly.<br />

The research history <strong>of</strong> the Andrews Forest, stretching<br />

back to its establishment in 1948, reflects a commitment<br />

to long-term ecological and watershed research by the US<br />

Forest Service and with NSF-funded programs under the<br />

International Biological Program in the 1970s, followed by<br />

<strong>LTER</strong> Network since 1980. These integrated science programs<br />

have produced high-quality studies and long-term<br />

records that underpin interpretations <strong>of</strong> ecosystem and<br />

environmental change and sustain an interdisciplinary cadre<br />

<strong>of</strong> scientists, all <strong>of</strong> whom are essential in investigating ecosystems<br />

that change abruptly and also gradually over time<br />

scales <strong>of</strong> decades and centuries. The context <strong>of</strong> extensive<br />

federal forestlands (e.g., US Forest Service, US Bureau <strong>of</strong><br />

Land Management) provides an audience <strong>of</strong> land managers<br />

who are required to guide management using current<br />

science. And, if they fail to do so, litigants and the courts<br />

remind them.<br />

A central feature <strong>of</strong> the Andrews Forest program is a<br />

research–management partnership that develops, tests, demonstrates,<br />

and critically evaluates alternative approaches to<br />

management so that when the policy window opens, new,<br />

scientifically and operationally credible approaches to management<br />

are ripe for broad adoption (http:// andrewsforest.<br />

oregonstate.edu/resmgt.cfm?topnav=35). This partnership<br />

involves the research community centered on the Andrews<br />

Forest <strong>LTER</strong> site and land managers <strong>of</strong> the Willamette<br />

National Forest. The partnership has made substantial<br />

impacts on forest management and policy on topics such<br />

as the characteristics <strong>of</strong> and conservation strategies for oldgrowth<br />

forest ecosystems (Franklin et al. 1981, Spies and<br />

Duncan 2009); the ecological roles and management implications<br />

<strong>of</strong> dead wood on land and in streams (Gregory et al.<br />

1991); the ecology and population dynamics <strong>of</strong> the northern<br />

spotted owl (Forsman et al. 1984); the effects <strong>of</strong> forest cutting<br />

and roads on streamflow, including floods (Jones 2000);<br />

358 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


Articles<br />

Figure 3. Andrews Experimental Forest research and links with management and policy for forestlands and watersheds.<br />

interactions <strong>of</strong> road and stream networks (Jones et al. 2000);<br />

and interactions <strong>of</strong> climate change with management and<br />

policy (Spies et al. 2010).<br />

With roots in the early 1950s and the assignment <strong>of</strong> the first<br />

scientist to the Andrews Forest, the research–management<br />

partnership has become a continuous, place-based learning<br />

program with balanced, reciprocal communication<br />

between the management and research communities and<br />

their respective cultures. To facilitate communication, the<br />

research–management interface is staffed with a research<br />

liaison position at the Willamette National Forest, which<br />

facilitates outreach to land managers and the public. The<br />

technical findings <strong>of</strong> research and management experience<br />

are communicated through diverse media, such as journal<br />

articles, including ones <strong>join</strong>tly composed by scientists and<br />

land managers (Cissel et al. 1999); publications prepared for<br />

land managers and the general public (e.g., in the Science<br />

Findings and Science Update series <strong>of</strong> the US Forest Service;<br />

www.fs.fed.us/pnw/publications/scifi.shtml, www.fs.fed.us/<br />

pnw/publications/sci-update.shtml); workshops; and field<br />

tours. In some cases, social scientists have examined the<br />

effectiveness <strong>of</strong> communications on challenging topics, such<br />

as the use <strong>of</strong> historic disturbance regimes to guide future<br />

land management (Shindler and Mallon 2009). The net<br />

effect <strong>of</strong> this communication program is a continuing public<br />

discussion <strong>of</strong> the future <strong>of</strong> forest and watershed management<br />

and policy in the region.<br />

The impacts <strong>of</strong> long-term research from the Andrews<br />

Forest and the research–management partnership are<br />

manifest in federal agencies’ management <strong>of</strong> forest stands<br />

and landscapes throughout the Pacific Northwest and more<br />

broadly (figure 3). In particular, the Northwest Forest<br />

Plan, which drew heavily from research from the Andrews<br />

Forest, ushered in a new era <strong>of</strong> ecosystem-based management<br />

on 10 million hectares <strong>of</strong> federal lands in northern<br />

California and western Oregon and Washington (FEMAT<br />

1993, USDA and USDI 1994). Andrews Forest–based science<br />

on old-growth forests, forest–stream interactions, aspects<br />

<strong>of</strong> biodiversity, and the roles <strong>of</strong> dead wood in forests and<br />

streams helped shape new federal land-management policies<br />

(FEMAT 1993). Several <strong>of</strong> the key publications have been<br />

cited in the scientific literature more than 1000 times each,<br />

which indicates the influence <strong>of</strong> the concepts in the environmental<br />

sciences. Since 1994, individual research themes have<br />

continued to influence management practices in the region<br />

( figure 3). Publications from the Andrews Forest–based work<br />

are widely cited in planning documents for timber sales and<br />

fuel-treatment projects on National Forests and US Bureau<br />

<strong>of</strong> Land Management districts across the Pacific Northwest.<br />

The impact <strong>of</strong> the research–management partnership has<br />

drawn social scientists to examine the dynamics, motivations,<br />

and public perception <strong>of</strong> these science–management–<br />

policy connections (Lach et al. 2003).<br />

Tools for assessing services and values to improve urban naturalresources<br />

stewardship: Baltimore Ecosystem Study long-term<br />

data. Information on natural resources in urban areas is<br />

<strong>of</strong>ten lacking and limits the ability <strong>of</strong> planners and managers<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 359


Articles<br />

to properly steward or incorporate natural-resources services<br />

within urban ecosystems. Long-term research is currently<br />

being conducted in the Baltimore area to foster a better<br />

understand how urbanization affects natural system processes<br />

(e.g., Pickett and Cadenasso 2006). Baltimore, through<br />

its participation in the <strong>LTER</strong> Network, was one <strong>of</strong> the first<br />

cities to have its entire forest and tree structure assessed,<br />

along with the concomitant ecosystem services and values<br />

(e.g., pollution removal, carbon storage and sequestration,<br />

effects on building energy use; see, e.g., Nowak et al. 2008).<br />

It is also the first city (along with Syracuse, New York) to<br />

establish (in 1999) permanent vegetation-monitoring plots<br />

to assess long-term vegetation changes (Nowak et al. 2004).<br />

These data provide critical information for better understanding<br />

<strong>of</strong> urban vegetation systems, their environmental<br />

effects, and how these ecosystems are changing. These data<br />

have also helped in the development and testing <strong>of</strong> publicdomain<br />

s<strong>of</strong>tware tools designed to aid managers and the<br />

general public in assessing urban trees and their associated<br />

ecosystem services and values. Data collected in Baltimore<br />

and other cities in the mid to late 1990s led to the development<br />

<strong>of</strong> s<strong>of</strong>tware to assess urban forest structure and functions:<br />

the UFORE (urban forest effects) model (Nowak and<br />

Crane 2000). Through time, a diverse collaboration developed<br />

among numerous partners to expand the development<br />

<strong>of</strong> this and other urban forest computer programs into a<br />

suite <strong>of</strong> free s<strong>of</strong>tware tools known as i-Tree (www.itreetools.<br />

org), which was released in 2006.<br />

The information provided by i-Tree s<strong>of</strong>tware has been<br />

used to inform management and policies throughout<br />

the world in relation to urban forestry. The influence <strong>of</strong><br />

i-Tree results from the use <strong>of</strong> the model and local data<br />

by consultants, managers, and local citizens to guide<br />

management and policies decisions related to issues such<br />

as emerald ash borer protection (Siyver 2009), building<br />

financial support for urban forestry programs (Society<br />

<strong>of</strong> Municipal Arborists 2008), linking local tree data<br />

with the US Conference <strong>of</strong> Mayors Climate Protection<br />

Agreement (Hyde 2009), public outreach campaigns (e.g.,<br />

billboards) on the benefits <strong>of</strong> trees (Siyver 2009), developing<br />

urban forest strategic management plans (McNeil<br />

and Vava 2006), and helping secure financing for tree<br />

planting and management (e.g., Ibrahim 2009). Most <strong>of</strong><br />

the data collected and analyzed through i-Tree are used to<br />

encourage municipal, county, and state leaders to establish<br />

or improve urban forestry programs, to recognize<br />

the role that trees play among urban natural resources,<br />

and to focus funds to improve stewardship. New tools<br />

were released in early 2010 (version 4.0) that include new<br />

approaches to help integrate science into local policy decisions<br />

related to streamflow, tree pests, local tree cover and<br />

effects, and related ecosystem services.<br />

Information and results from i-Tree, its analyses, and<br />

impacts are generally communicated by the research partners<br />

and users to others through public presentations,<br />

reports and articles, webinars, the i-Tree Web site, and word<br />

<strong>of</strong> mouth. To assist in communicating project results, i-Tree<br />

automatically produces a standard report with graphics that<br />

users can export and customize for their own use (figure 4).<br />

Users can also report ideas, questions, or problems back to<br />

the i-Tree team, which are then used to update or develop<br />

future versions.<br />

To date, more than 8200 unique users in 99 countries have<br />

downloaded the s<strong>of</strong>tware. Use <strong>of</strong> i-Tree has grown at about<br />

30% per year since its release in August 2006. i-Tree Web site<br />

traffic has increased about tenfold since the release <strong>of</strong> version<br />

3.0 in June 2009 and continues to increase. Currently,<br />

about 20,000 unique users access the Web site every three<br />

months. Focused surveys <strong>of</strong> users have been conducted to<br />

help determine the types <strong>of</strong> impacts. Between 50 and 100<br />

journal articles and reports have been published in which<br />

i-Tree was used, and the numbers have increased annually.<br />

New programs in development are focused on temporal and<br />

spatial modeling <strong>of</strong> forest effects, and the Baltimore longterm<br />

permanent field-plot data are critical to the development<br />

<strong>of</strong> these new tools. International urban forest data<br />

standards are also in development to aid in sharing and in<br />

the use <strong>of</strong> the programs among nations.<br />

Climate-change impacts on wildfire: Bonanza Creek engagement<br />

with fire managers and indigenous communities. Alaska is<br />

warming twice as quickly as the global average, with little<br />

change in precipitation (Chapin et al. 2006a). The resulting<br />

drying <strong>of</strong> the boreal forest has increased the annual area<br />

burned, primarily through increased frequency <strong>of</strong> dry years<br />

and larger wildfires, which have important consequences for<br />

changes in forest cover and the closely coupled human and<br />

ecological communities (figures 5 and 6; K<strong>of</strong>inas et al. 2010).<br />

Bonanza Creek scientists collaborate with fire managers and<br />

indigenous communities to share knowledge for predicting<br />

and adapting to changing fire regimes.<br />

Working with fire managers, Bonanza Creek <strong>LTER</strong> ecologists<br />

have developed predictive models that provide a scientific<br />

foundation for fire-management decisions. Spatially<br />

explicit models <strong>of</strong> climate and wildfire suggest that, by 2050,<br />

a “typical” fire year in interior Alaska will be similar to the<br />

most extreme fire years in the historical record (www.snap.<br />

uaf.edu). These models were developed through extensive<br />

input from climatologists, ecologists, and fire managers<br />

(Duffy et al. 2005).<br />

At the community level, village tribal councils have invited<br />

Bonanza Creek ecologists to collaborate in developing new<br />

ecosystem-management strategies to respond to increasing<br />

wildfire risk. These strategies include the sustainable harvest<br />

<strong>of</strong> flammable black spruce stands near communities to heat<br />

public buildings, create new jobs, and generate secondary<br />

successional habitat that favors moose—an important food<br />

source (Chapin et al. 2008, K<strong>of</strong>inas et al. 2010). Bonanza<br />

Creek social scientists and ecologists have also participated<br />

in federally mandated community wildfire protection planning<br />

by conducting interviews and focus groups among<br />

local residents and resource managers. These interviews<br />

360 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


Figure 4. Example <strong>of</strong> the user interface for i-Tree. The page shows the i-Tree canopy survey page for urban forests.<br />

demonstrated that local residents trusted managers to plan<br />

community-level wildfire protection but felt disenfranchised<br />

in regional wildfire planning for the surrounding<br />

lands, because their knowledge and concerns about future<br />

subsistence opportunities and places <strong>of</strong> cultural value were<br />

overlooked (Ray 2010).<br />

Fire-modeling results are communicated to fire managers<br />

and the public through participation in annual wildfire<br />

strategic-planning workshops, agency meetings with the<br />

public, <strong>join</strong>t agency <strong>LTER</strong> planning <strong>of</strong> prescribed burns, and<br />

production <strong>of</strong> site-specific 2-kilometer-resolution climate<br />

projections and fire-risk projections on request (www.snap.<br />

uaf.edu).<br />

Community workshops coorganized by tribal councils<br />

and Bonanza Creek ecologists allow an exchange <strong>of</strong> local,<br />

traditional, and scientific knowledge about wildfire ecology.<br />

This dialogue has enriched understanding by the <strong>LTER</strong> scientists<br />

<strong>of</strong> the ecological and societal consequences <strong>of</strong> climate<br />

change. The trust that develops through community partnerships<br />

enables Bonanza Creek researchers to learn from<br />

Articles<br />

and contribute to societal responses to a rapidly changing<br />

socioecological environment.<br />

The Bonanza Creek <strong>LTER</strong> site is forging new ground in<br />

identifying climate-change impacts that require immediate<br />

management and community action. The associated metrics<br />

<strong>of</strong> impact are therefore recent and qualitative. Judging from<br />

the ESA’s Sustainability Science Award for Chapin and colleagues’<br />

(2006b) article, in which they described the socioecological<br />

framework for this research, the Bonanza Creek<br />

<strong>LTER</strong> site is contributing to fundamental science and to new<br />

approaches for integrating community knowledge and concerns<br />

in socioecological research (figure 6). Bonanza Creek<br />

collaborations contributed to Alaska fire managers’ capacity<br />

to adapt federal guidelines on the basis <strong>of</strong> fire issues <strong>of</strong> the<br />

lower 48 US states to conditions and issues that are relevant<br />

to Alaska. Managers use the fire-risk model and routinely<br />

invite Bonanza Creek ecologists to participate in the training<br />

<strong>of</strong> wildfire managers, which indicates that they value the<br />

practical relevance <strong>of</strong> <strong>LTER</strong>. Bonanza Creek ecologists and<br />

Alaskan indigenous leaders have formed the Working Group<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 361


Articles<br />

Ecoregions<br />

Subarctic<br />

tundra<br />

Arctic tundra<br />

Boreal forest<br />

Coastal rainforest<br />

on Rural Alaska Self-Reliance, a collaboration to implement<br />

community visions <strong>of</strong> adaptation to global change. This collaboration<br />

suggests that indigenous leaders value and trust<br />

their interactions with Bonanza Creek scientists.<br />

Discussion <strong>of</strong> the case studies. Boundary-spanning efforts can<br />

facilitate the bridging <strong>of</strong> science and society by producing<br />

information that is salient, credible, and legitimate (Cash<br />

et al. 2003, McNie et al. 2008), which ultimately enriches<br />

scientific research through stakeholder engagement, the<br />

expansion <strong>of</strong> public awareness, and the improvement <strong>of</strong><br />

the scientific basis for decisionmaking. The five <strong>LTER</strong> case<br />

studies presented here <strong>of</strong>fer experiences and lessons to help<br />

answer the question <strong>of</strong> what characterizes successful collaborative<br />

outreach efforts. The case studies suggest that<br />

efforts to build a stronger interface between science and<br />

society are shaped in part by three overarching attributes<br />

that pertain to all ecosystems but vary in detail among<br />

ecosystems: (1) Landscape and social context refers to the<br />

pattern <strong>of</strong> land ownership (e.g., private versus public) and<br />

the role <strong>of</strong> the different types <strong>of</strong> knowledge (e.g., local versus<br />

expert) that influence the framing <strong>of</strong> environmental issues,<br />

the management objectives, and the science used in the<br />

decisionmaking process. (2) Issue definition involves<br />

Indigenous Languages and Peoples<br />

Central Yup'ik<br />

Deg Xinag<br />

Eyak–Tlingit<br />

Tlingit<br />

determining the relevance <strong>of</strong> particular long-term research to<br />

policy and management issues at local, regional, or national<br />

scales (e.g., local fire- or fuel-management issues, regional<br />

air-quality concerns, federal forestland policy) and the extent<br />

to which individual actions or government actions are central<br />

to resolving the issues <strong>of</strong> concern. (3) Communication<br />

pathways entail understanding which communication approaches<br />

are most appropriate for specific decisionmakers,<br />

and the choice <strong>of</strong> pathway is determined in part by the<br />

context and issues addressed (e.g., direct briefings between<br />

scientists and policymakers; outreach to the media; working<br />

groups with managers; discussions with local communities,<br />

including tribes).<br />

In addition to these three overarching attributes that<br />

distinguish individual efforts, a set <strong>of</strong> common elements <strong>of</strong><br />

successful science communication efforts emerges from the<br />

case studies:<br />

In all <strong>of</strong> the case studies, boundary-spanning efforts were<br />

built on credible, multidecade, interdisciplinary science, and<br />

peer-reviewed publications. These efforts combine retrospective<br />

analysis; long-term measurements and experiments;<br />

quantitative modeling; and, increasingly, scenarios planning.<br />

For example, the ability <strong>of</strong> the HBRF Science Links projects<br />

to assess the impacts <strong>of</strong> air-quality regulations and the extent<br />

362 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org<br />

Ahtna<br />

Eyak<br />

Haida<br />

Tsimshian<br />

Eskimo–Aleut Athabascan<br />

Iñupiaq<br />

Gwich'in Holikachuk<br />

Unanga{ (Aleut)<br />

Sugpiaq (Alutiiq) Dena'ina Tanacross<br />

St. Lawrence Island Yupik<br />

Hän<br />

Upper Kuskokwim<br />

Upper Tanana<br />

Lower Tanana<br />

Denaakk'e<br />

(Koyukon)<br />

Map by C. West, M. Wilson and J. Kerr - ISER<br />

Figure 5. Cultural (linguistic) groups and ecoregions are closely coupled in the boreal forest region <strong>of</strong> Alaska (Chapin<br />

2009). The Bonanza Creek Long Term Ecological Research Network site uses understanding <strong>of</strong> socioecological responses<br />

to climate change as a platform for exploring and implementing adaptation options that rural Athabascan communities<br />

would find consistent with their history and current commitment to sustainable subsistence lifestyles. Source: Reprinted<br />

from F. Stuart Chapin III, “Managing ecosystems sustainably: The key role <strong>of</strong> resilience.” Pages 29–53 in Chapin FS III,<br />

K<strong>of</strong>inas GP, Folke C, eds. Principles <strong>of</strong> Ecosystem Stewardship: Resilience-Based Natural Resource Management in a<br />

Changing World (2009), with permission from Springer.


Community concern<br />

about global change<br />

Informal<br />

input<br />

Informal dialogue<br />

<strong>LTER</strong> or community<br />

request for collaboration<br />

Tribal council approval<br />

Focus groups, interviews<br />

and knowledge sharing<br />

based on shared concerns<br />

Revisions Initial results<br />

Community review<br />

<strong>LTER</strong> interest in<br />

Socioecological change<br />

Formal<br />

input<br />

Community actions Joint publications<br />

Figure 6. Processes <strong>of</strong> interaction between the Bonanza<br />

Creek Long Term Ecological Research (<strong>LTER</strong>) Network site’s<br />

scientists and indigenous communities in sharing ecological<br />

knowledge. Informal discussions between scientists and<br />

community members lead to a formal request to the village<br />

tribal council to explore specific questions (e.g., climatechange<br />

effects on fire regime). One or more rounds <strong>of</strong><br />

interaction involving interviews or focus groups, a review<br />

<strong>of</strong> the findings by the community, and a revision based<br />

on feedback. This leads to a formal report to the village<br />

council and <strong>join</strong>t publications by scientists and community<br />

members, as well as informal input to the village council,<br />

other community members, and <strong>LTER</strong> scientists.<br />

to which ecosystems have recovered was entirely dependent<br />

on the existence <strong>of</strong> long-term precipitation, soil and stream<br />

chemistry, and relevant biological measurements. These data<br />

enabled scientists to analyze changes in atmospheric deposition<br />

and associated chemical and biological responses, to<br />

establish impact thresholds, and to apply dynamic models<br />

to evaluate the extent to which future emissions reductions<br />

would achieve policy objectives.<br />

Among the most important activities is the collaboration<br />

<strong>of</strong> scientists and decisionmakers at the outset <strong>of</strong> and<br />

throughout a research effort. This interaction helps to define<br />

issues and questions salient to decisionmakers, to identify<br />

sources <strong>of</strong> knowledge beyond traditional scientific data sets,<br />

and to envision outputs that best meet user needs. This<br />

Articles<br />

process also enriches scientific research. For example, the<br />

Bonanza Creek research framework was expanded through<br />

interactions with community groups to larger temporal and<br />

spatial scales and integration <strong>of</strong> cultural dimensions. This<br />

led to the recognition <strong>of</strong> critical thresholds <strong>of</strong> the resistance<br />

<strong>of</strong> the boreal socioecological system to climatic and socioeconomic<br />

changes.<br />

Although scientists are accustomed to publishing focused<br />

research in peer-reviewed publications, these case studies<br />

point clearly to the need for policy- and managementrelevant<br />

synthesis and distillation to support the effective<br />

use <strong>of</strong> science in the policy and management processes. This<br />

problem-oriented synthesis is necessary but not sufficient<br />

for promoting knowledge sharing and should be accompanied<br />

by work to translate the key findings into compelling<br />

terms relevant to stakeholders. For example, by pulling<br />

together disparate findings from across dozens <strong>of</strong> articles<br />

produced over a decade or more, the Harvard Forest W&W<br />

publications have drawn public and stakeholder attention to<br />

that body <strong>of</strong> work and catalyzed conservation initiatives far<br />

beyond what any single study could accomplish.<br />

Successful outreach should not be an afterthought but<br />

a major and well-funded initiative with adequate staffing<br />

and supporting expertise ranging from traditional print<br />

publications to media, including Web-based, outreach.<br />

Innovative online tools that promote interaction and social<br />

networking and that are open source and easily accessible<br />

are increasingly important communication vehicles. For<br />

example, the i-Tree project built a program interface that<br />

is easy to use, open to all, supported, and free: The i-Tree<br />

partnership has built a platform to which others can contribute,<br />

and new peer-reviewed tools can be added and<br />

then supported through the existing i-Tree partnership and<br />

model structure.<br />

Partnerships are critical to sustaining reciprocal flow <strong>of</strong><br />

information among scientists, citizen leaders, managers, and<br />

policymakers; to applying scientific findings to policy and<br />

management through an adaptive process; and to fueling<br />

processes in which stakeholder experiences and knowledge<br />

inform research. For example, the research–management<br />

partnership developed by the Andrews <strong>LTER</strong> site and the<br />

US Forest Service provides a platform for sustained, placebased<br />

learning with substantial attention to communications<br />

with many audiences. Similarly, the Baltimore <strong>LTER</strong><br />

i-Tree project also functions as a partnership that meets<br />

regularly and has open discussions, working toward meeting<br />

the needs <strong>of</strong> the urban community. In both cases, the<br />

involvement <strong>of</strong> a public entity (the US Forest Service) has<br />

been instrumental in coordinating and managing the activities<br />

<strong>of</strong> the partnerships.<br />

In addition to these lessons, the case studies presented<br />

here demonstrate the need for stronger metrics to measure<br />

the impact <strong>of</strong> science communications and outreach to<br />

decisionmakers. In general, metrics for evaluating publicengagement<br />

outreach efforts can be divided into three<br />

categories: output, uptake, and impact. The five case studies<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 363


Articles<br />

present outputs such as the number <strong>of</strong> publications and<br />

presentations given to nonscientific audiences. They also<br />

provide strong evidence <strong>of</strong> uptake such as media coverage,<br />

scientific citations, and Web site visitation. Quantifying the<br />

impact on policy, conservation, and stewardship decisions<br />

remains elusive.<br />

Developing and applying meaningful metrics <strong>of</strong> impact<br />

is a common challenge. Under the auspices <strong>of</strong> the White<br />

House Office <strong>of</strong> Science and Technology Policy, the National<br />

Institutes <strong>of</strong> Health and the NSF are developing metrics <strong>of</strong><br />

impacts for science, called STAR METRICS (Science and<br />

Technology for America’s Reinvestment: Measuring the<br />

Effects <strong>of</strong> Research on Innovation, Competitiveness, and<br />

Science; Lane and Bertuzzi 2011). Several metrics have been<br />

proposed to measure the usefulness <strong>of</strong> scientific knowledge,<br />

many <strong>of</strong> which are applied in these case studies (e.g.,<br />

download or hit rates, media coverage, citations in federal<br />

or state regulations). But in the area <strong>of</strong> broader impacts or<br />

social outcomes, such as those in health, safety, and the environment,<br />

recommendations are under development by an<br />

interagency working group. Impact metrics for science are<br />

an important gap in understanding that should be remedied<br />

by the STAR METRICS program and other science-policy<br />

research efforts.<br />

Conclusions<br />

If science is to aid in the advance toward a more resilient and<br />

sustainable society, we must experiment with more effective<br />

means <strong>of</strong> integrating ecological research and decisionmaking.<br />

As is evidenced by the five case studies presented<br />

here for forest ecosystems and by many other examples,<br />

the <strong>LTER</strong> Network has an important and unique role to<br />

play in addressing the grand challenges in environmental<br />

and sustainability science. The <strong>LTER</strong> Network and associated<br />

research, with its long-term interdisciplinary focus,<br />

its focus on place-based study, its geographic distribution,<br />

its sophisticated information-management systems, and its<br />

public-outreach capabilities, are well suited to boundaryspanning<br />

initiatives that address emerging environmental<br />

issues related to changes in biogeochemistry, biological<br />

diversity, climate change, ecohydrology, infectious disease,<br />

and land use. Policy-relevant synthesis and science communication<br />

should be a focus <strong>of</strong> the <strong>LTER</strong> Network, and these<br />

activities, in turn, would probably promote cross-site and<br />

network-wide coordination <strong>of</strong> matters important to both<br />

science and society. This work could be enhanced by partnerships<br />

with established scientific societies that are dedicated<br />

to similar work. For example, the ESA is advancing a<br />

partnership among academic societies, agencies, and NGOs<br />

“to foster Earth Stewardship by (a) clarifying the science<br />

needs for understanding and shaping trajectories <strong>of</strong> change<br />

at local-to-global scales; (b) communicating the basis for<br />

Earth Stewardship to a broad range <strong>of</strong> audiences, including<br />

natural and social scientists, students, the general public,<br />

policymakers, and other practitioners; and (c) formulating<br />

pragmatic strategies that foster a more sustainable trajectory<br />

<strong>of</strong> planetary change by enhancing ecosystem resilience and<br />

human well-being” (Chapin et al. 2011, p. 45).<br />

Harnessing the power <strong>of</strong> long-term ecological studies<br />

to address the grand challenges in environmental science<br />

will require learning from and building on existing efforts<br />

to better integrate scientific research with societal concerns.<br />

The NSF can facilitate this process by expanding the<br />

bounds <strong>of</strong> informal education to include the engagement <strong>of</strong><br />

decisionmakers and journalists in order to provide the<br />

requisite research and learning needed to develop, test, and<br />

expand these critical experiments at the interface <strong>of</strong> science<br />

and society.<br />

Acknowledgments<br />

The effort and contributions <strong>of</strong> KFL were generously supported<br />

by the Bullard Fellowship program <strong>of</strong> Harvard<br />

<strong>University</strong>’s Harvard Forest and by a grant from Highstead.<br />

This is a contribution <strong>of</strong> the Long Term Ecological Research<br />

Network, which is supported by the US National Science<br />

Foundation and by the US Forest Service. We appreciate<br />

the efforts <strong>of</strong> David R. Foster in shepherding this series <strong>of</strong><br />

articles.<br />

References cited<br />

Berlik MM, Kittredge DB, Foster DR. 2002. The illusion <strong>of</strong> preservation:<br />

A global environmental argument for the local production <strong>of</strong> natural<br />

resources. Journal <strong>of</strong> Biogeography 29: 1557–1568.<br />

Cash DW, Clark WC, Alcock F, Dickson NM, Eckley N, Guston DH,<br />

Jäger J, Mitchell RB. 2003. Knowledge systems for sustainable development.<br />

Proceedings <strong>of</strong> the National Academy <strong>of</strong> Sciences 100:<br />

8086–8091.<br />

Chapin FS III. 2009. Managing ecosystems sustainably: The key role <strong>of</strong><br />

resilience. Pages 29–53 in Chapin FS III, K<strong>of</strong>inas GP, Folke C, eds.<br />

Principles <strong>of</strong> Ecosystem Stewardship: Resilience-Based Natural Resource<br />

Management in a Changing World. Springer.<br />

Chapin FS III, Oswood MW, Van Cleve K, Viereck LA, Verbyla DL, eds.<br />

2006a. Alaska’s Changing Boreal Forest. Oxford <strong>University</strong> Press.<br />

Chapin FS III, Lovecraft AL, Zavaleta ES, Nelson J, Robards MD, K<strong>of</strong>inas<br />

GP, Trainor SF, Peterson GD, Huntington HP, Naylor RL. 2006b. Policy<br />

strategies to address sustainability <strong>of</strong> Alaskan boreal forests in response<br />

to a directionally changing climate. Proceedings <strong>of</strong> the National<br />

Academy <strong>of</strong> Sciences 103: 16637–16643.<br />

Chapin FS III, et al. 2008. Increasing wildfire in Alaska’s boreal forest:<br />

Pathways to potential solutions <strong>of</strong> a wicked problem. BioScience 58:<br />

531–540.<br />

Chapin FS III, Pickett STA, Power ME, Jackson RB, Carter DM, Duke C.<br />

2011. Earth stewardship: A strategy for social-ecological transformation<br />

to reverse planetary degradation. Journal <strong>of</strong> Environmental Studies and<br />

Sciences 1: 44–53.<br />

Cissel JH, Swanson FJ, Weisberg PJ. 1999. Landscape management using<br />

historical fire regimes: Blue River, Oregon. Ecological Applications 9:<br />

1217–1231.<br />

D’Amato AW, Orwig DA, Foster DR. 2006. New estimates <strong>of</strong> Massachusetts<br />

old-growth forests: Useful data for regional conservation and forest<br />

reserve planning. Northeastern Naturalist 13: 495–506.<br />

Driscoll CT, Lawrence GB, Bulger AJ, Butler TJ, Cronan CS, Eagar C,<br />

Lambert KF, Likens GE, Stoddard JL, Weathers KC. 2001. Acidic deposition<br />

in the northeastern United States: Sources and inputs, ecosystem<br />

effects, and management strategies. BioScience 51: 180–198.<br />

Driscoll CT, Cowling EB, Grennfelt P, Galloway JM, Dennis RL. 2010.<br />

Integrated assessment <strong>of</strong> ecosystem effects <strong>of</strong> atmospheric deposition:<br />

Lessons available to be learned. EM Magazine November 2010: 6–13.<br />

364 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


Driscoll CT, Lambert KF, Weathers KC. 2011. Integrating science and policy:<br />

A case study <strong>of</strong> the Hubbard Brook Research Foundation Science Links<br />

program. BioScience 61: 791–801.<br />

Duffy PA, Walsh JE, Graham JM, Mann DH, Rupp TS. 2005. Impacts <strong>of</strong><br />

large-scale atmospheric–ocean variability on Alaskan fire season severity.<br />

Ecological Applications 15: 1317–1330.<br />

[FEMAT] Forest Ecosystem Management Assessment Team. 1993.<br />

Forest Ecosystem Management: An Ecological, Economic, and Social<br />

Assessment. US Department <strong>of</strong> Agriculture Forest Service.<br />

Fisher RT. 1921. The management <strong>of</strong> the Harvard Forest, 1909–1919.<br />

Harvard Forest Bulletin 1: 1–27. (18 January 2012; http://harvardforest.<br />

fas.harvard.edu/publications/pdfs/HFpubs/b1.pdf )<br />

———. 1933. New England forests: Biological factors. Pages 213–223 in<br />

Wood W, ed. New England’s Prospect. American Geographical Society.<br />

Special Publication no. 16.<br />

Forsman ED, Meslow EC, Wight HM. 1984. Distribution and <strong>biology</strong> <strong>of</strong> the<br />

spotted owl in Oregon. Wildlife Monographs no. 87: 3–64.<br />

Foster DR. 2000. From bobolinks to bears: Interjecting geographical history<br />

into ecological studies, environmental interpretation, and conservation<br />

planning. Journal <strong>of</strong> Biogeography 27: 27–30.<br />

Foster DR, Orwig DA. 2006. Preemptive and salvage harvesting <strong>of</strong> New<br />

England forests: When doing nothing is a viable alternative. Conservation<br />

Biology 20: 959–970.<br />

Foster DR, Aber JD, Melillo JM, Bowden RD, Bazazz FA. 1997. Forest<br />

response to disturbance and anthropogenic stress: Rethinking the 1938<br />

hurricane and the impact <strong>of</strong> physical disturbance vs. chemical and<br />

climate stress on forest ecosystems. BioScience 47: 437–445.<br />

Foster D[R], Kittredge D[B], Donahue B, Motzkin G, Orwig D[A], Ellison<br />

AM, Hall B, Colburn EA, D’Amato A. 2005. Wildlands and Woodlands:<br />

A Vision for the Forests <strong>of</strong> Massachusetts. Harvard <strong>University</strong>, Harvard<br />

Forest.<br />

Foster DR, Aber JD, Cogbill CV. 2010. Wildlands and Woodlands: A Vision<br />

for the New England Landscape. Harvard <strong>University</strong>, Harvard Forest.<br />

Franklin JF, Cromack K Jr, Denison W, McKee A, Maser C, Sedell J, Swanson<br />

F, Juday G. 1981. Ecological characteristics <strong>of</strong> old-growth Douglas-fir<br />

forests. US Department <strong>of</strong> Agriculture, Forest Service, Pacific Northwest<br />

Forest and Range Experiment Station. General Technical Report<br />

no. PNW-GTR-118.<br />

Golodetz AD, Foster DR. 1997. History and importance <strong>of</strong> land use and<br />

protection in the North Quabbin Region <strong>of</strong> Massachusetts (USA).<br />

Conservation Biology 11: 227–235.<br />

Gregory SV, Swanson FJ, Mckee WA, Cummins KW. 1991. An ecosystem<br />

perspective <strong>of</strong> riparian zones. BioScience 41: 540–551.<br />

Hall B, Motzkin G, Foster DR, Syfert M, Burk J. 2002. Three hundred<br />

years <strong>of</strong> forest and land-use change in Massachusetts, USA. Journal <strong>of</strong><br />

Biogeography 29: 1319–1335.<br />

Halpern CB, Haugo RD, Antos JA, Kaas SS, Kilanowski AL. 2012.<br />

Grassland restoration with and without fire: Evidence from a<br />

tree-removal experiment. Ecological Applications. Forthcoming.<br />

doi:10.1890/11-1061.1<br />

Harmon ME, et al. 1986. Ecology <strong>of</strong> coarse woody debris in temperate<br />

ecosystems. Pages 133–302 in MacFadyen A, Ford ED, eds. Advances in<br />

Ecological Research, vol. 15. Academic Press.<br />

Hyde G. 2009. President’s message. City Trees: Journal <strong>of</strong> the Society <strong>of</strong><br />

Municipal Arborists 45 (6): 5.<br />

Ibrahim M. 2009. US benefit analysis snared over $220m for trees.<br />

Horticulture Week 01 October 2009. (18 January 2012; www.hortweek.<br />

com/news/942310).<br />

Jones JA. 2000. Hydrologic processes and peak discharge response to forest<br />

removal, regrowth, and roads in 10 small experimental basins, western<br />

Cascades, Oregon. Water Resources Research 36: 2621–2642.<br />

Jones JA, Swanson FJ, Wemple BC, Snyder KU. 2000. Effects <strong>of</strong> roads on<br />

hydrology, geomorphology, and disturbance patches in stream networks.<br />

Conservation Biology 14: 76–85.<br />

Kittredge DB, Finley AO, Foster DR. 2003. Timber harvesting as ongoing<br />

disturbance in a landscape <strong>of</strong> diverse ownership. Forest Ecology and<br />

Management 180: 425–442.<br />

Articles<br />

Kizlinski ML, Orwig DA, Cobb RC, Foster DR. 2002. Direct and indirect<br />

ecosystem consequences <strong>of</strong> an invasive pest on forests dominated by<br />

eastern hemlock. Journal <strong>of</strong> Biogeography 29: 1489–1503.<br />

Knapp AK, et al. 2012. Past, present, and future roles <strong>of</strong> long-term experiments<br />

in the <strong>LTER</strong> network. BioScience 62: 377–389.<br />

K<strong>of</strong>inas GP, Chapin FS [III], BurnSilver S, Schmidt JI, Fresco NL, Kielland<br />

K, Martin S, Springsteen A, Rupp TS. 2010. Resilience <strong>of</strong> Athabascan<br />

subsistence systems to interior Alaska’s changing climate. Canadian<br />

Journal <strong>of</strong> Forest Research 40: 1347–1359.<br />

Lach D, List P, Steel B, Shindler B. 2003. Advocacy and credibility <strong>of</strong> ecological<br />

scientists in resource decisionmaking: A regional study. BioScience<br />

53: 170–178.<br />

Lane J, Bertuzzi S. 2011. Measuring the results <strong>of</strong> science investments.<br />

Science 331: 678–680.<br />

Levitt JN, Lambert KF. 2006. Report on the Woodlands and Wildlands<br />

Conservation Finance Roundtable. Harvard <strong>University</strong>, Harvard<br />

Forest.<br />

Lovett GM, Burns DA, Driscoll CT, Jenkins JC, Mitchell MJ, Rustad L,<br />

Shanley JB, Likens GE, Haeuber R. 2007. Who needs environmental<br />

monitoring? Frontiers in Ecology and the Environment 5:<br />

253–260.<br />

[<strong>LTER</strong> Network] Long Term Ecological Research Network. 2010a. <strong>LTER</strong><br />

Strategic Communication Plan: Bridging to Broader Audiences. <strong>LTER</strong><br />

Network. (8 February 2012; http://intranet2.lternet.edu/sites/intranet2.<br />

lternet.edu/files/documents/<strong>LTER</strong>%20History/Planning%20Documents/<br />

Final%20<strong>LTER</strong>%20Strategic%20Communication%20Plan%20-%20<br />

Nov%2011%202010.pdf )<br />

———. 2010b. Strategic and Implementation Plan. <strong>LTER</strong> Network.<br />

(18 January 2012; http://intranet2.lternet.edu/documents/lter-strategicand-implementation-plan)<br />

Lubchenco J. 1998. Entering the century <strong>of</strong> the environment: A new social<br />

contract for science. Science 279: 491–497.<br />

Mayer AL, Kauppi PE, Angelstam PK, Zhang Y, Tikka PM. 2005. Importing<br />

timber, exporting ecological impact. Science 308: 359–360.<br />

McDonald RI, Motzkin G, Foster DR. 2008. The effect <strong>of</strong> logging on<br />

vegetation composition in western Massachusetts. Forest Ecology and<br />

Management 255: 4021–4031.<br />

McNeil J, Vava C. 2006. Oakville’s urban forest: Our solution to our<br />

pollution. Town <strong>of</strong> Oakville, Parks and Open Space Department,<br />

Forestry Section. (8 February 2012; www.itreetools.org/resources/reports/<br />

Oakville%27s%20Urban%20Forest.pdf )<br />

McNie EC, van Noordwijk M, Clark WC, Dickson NM, Sakuntaladewi<br />

N, Suyanto JL, Joshi L, Leimona B, Hairiah K, Khususiyah N. 2008.<br />

Boundary Organizations, Objects and Agents: Linking Knowledge<br />

with Action in Agr<strong>of</strong>orestry Watersheds. Center for International<br />

Development at Harvard <strong>University</strong> and World Agr<strong>of</strong>orestry Centre.<br />

Working Paper no. 80.<br />

Nowak DJ, Crane DE. 2000. The Urban Forest Effects (UFORE) Model:<br />

Quantifying urban forest structure and functions. Pages 714–720<br />

in Hansen M, Burk T, eds. Integrated Tools for Natural Resources<br />

Inventories in the 21st Century. US Department <strong>of</strong> Agriculture, Forest<br />

Service, North Central Research Station. General Technical Report<br />

no. NC-212.<br />

Nowak DJ, Kurodo M, Crane DE. 2004. Urban tree mortality rates and tree<br />

population projections in Baltimore, Maryland, USA. Urban Forestry<br />

and Urban Greening 2: 139–147.<br />

Nowak DJ, Crane DE, Stevens JC, Hoehn RE, Walton JT, Bond J. 2008.<br />

A ground-based method <strong>of</strong> assessing urban forest structure and ecosystem<br />

services. Arboriculture and Urban Forestry 34: 347–358.<br />

Nudel M. 2003. Better conservation through partnerships. Exchange:<br />

Journal <strong>of</strong> the Land Trust Alliance 22: 17–21.<br />

Orwig DA, Cogbill CV, Foster DR, O’Keefe JF. 2001. Variations in oldgrowth<br />

structure and definitions: Forest dynamics on Wachusett<br />

Mountain, Massachusetts. Ecological Applications 11: 437–452.<br />

Pickett STA, Cadenasso ML. 2006. Advancing urban ecological studies:<br />

Frameworks, concepts, and results from the Baltimore Ecosystem Study.<br />

Austral Ecology 31: 114–125.<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 365


Articles<br />

Ray L. 2010. Can Community-Based Natural Resource Management<br />

Improve Wildfire Policy Planning in Interior Alaska? Addressing Value<br />

Differences, Ineffective Participatory Processes, and Conflicts over<br />

Traditional Ecological Knowledge. Doctoral Dissertation. Department<br />

<strong>of</strong> Geography. Clark <strong>University</strong>, Worcester, Massachusetts.<br />

Shindler B, Mallon AL. 2009. Public Acceptance <strong>of</strong> Disturbance-Based<br />

Forest Management: A Study <strong>of</strong> the Blue River Landscape Strategy<br />

in the Central Cascades Adaptive Management Area. US Department<br />

<strong>of</strong> Agriculture, Forest Service, Pacific Northwest Research Station.<br />

Research Paper no. PNW-RP-581.<br />

Siyver D. 2009. Milwaukee UFORE: Taking it to the streets. City Trees:<br />

Journal <strong>of</strong> the Society <strong>of</strong> Municipal Arborists 45 (5): 10–12.<br />

Society <strong>of</strong> Municipal Arborists. 2008. More than cows and cowboys: Urban<br />

forestry in Casper, Wyoming. City Trees: Journal <strong>of</strong> the Society <strong>of</strong><br />

Municipal Arborists Sept/Oct: 44(5): 6–10.<br />

Spies TA, Duncan SL, eds. 2009. Old Growth in a New World: A Pacific<br />

Northwest Icon Reexamined. Island Press.<br />

Spies TA, Giesen TW, Swanson FJ, Franklin JF, Lach D, Johnson KN. 2010.<br />

Climate change adaptation strategies for federal forests <strong>of</strong> the Pacific<br />

Northwest, USA: Ecological, policy, and socio-economic perspectives.<br />

Landscape Ecology 25: 1185–1199.<br />

BRINGING<br />

BIOLOGY<br />

JOIN<br />

AIBS<br />

www.aibs.org<br />

AIBS INDIVIDUAL MEMBERS<br />

RECEIVE PRINT AND/OR<br />

ELECTRONIC ACCESS<br />

TO BIOSCIENCE. AIBS<br />

ORGANIZATIONAL MEMBERS<br />

PARTICIPATE IN THE NATIONAL<br />

DIALOGUE ABOUT HOW TO<br />

ENSURE A BRIGHT FUTURE<br />

FOR BIOLOGY.<br />

Thompson JR, Wiek A, Swanson FJ, Carpenter SR, Fresco N, Hollingsworth<br />

T, Spies TA, Foster DR. 2012. Scenario studies as a synthetic and integrative<br />

research activity for long-term ecological research. BioScience<br />

62: 367–376.<br />

[USDA and USDI] US Department <strong>of</strong> Agriculture, US Department <strong>of</strong> the<br />

Interior. 1994. Record <strong>of</strong> Decision for Amendments to Forest Service<br />

and Bureau <strong>of</strong> Land Management Planning Documents within the<br />

Range <strong>of</strong> the Northern Spotted Owl. USDA Forest Service.<br />

Charles T. Driscoll (ctdrisco@syr.edu) is affiliated with the Department <strong>of</strong><br />

Civil and Environmental Engineering at Syracuse <strong>University</strong>, in New York.<br />

Kathleen F. Lambert and Clarisse M. Hart are affiliated with Harvard <strong>University</strong>’s<br />

Harvard Forest, in Petersham, Massachusetts. F. Stuart Chapin III is<br />

affiliated with the Institute <strong>of</strong> Arctic Biology at the <strong>University</strong> <strong>of</strong> Alaska, Fairbanks.<br />

David J. Nowak is affiliated with the US Department <strong>of</strong> Agriculture,<br />

Forest Service, Northern Research Station, in Syracuse, New York. Thomas A.<br />

Spies and Frederick J. Swanson are affiliated with the US Forest Service, Pacific<br />

Northwest Research Station, in Corvallis, Oregon. David B. Kittredge, Jr., is<br />

affiliated with the Department <strong>of</strong> Environmental Conservation at the <strong>University</strong><br />

<strong>of</strong> Massachusetts–Amherst.<br />

TO INFORMED<br />

DECISION<br />

MAKING<br />

JOIN THE<br />

COMMUNITY...<br />

As an umbrella organization<br />

AIBS works with biologists and<br />

their pr<strong>of</strong>essional organizations,<br />

to ensure that reliable<br />

information is used when<br />

decisions are made–in public<br />

policy, research funding, and<br />

the public forum.<br />

To learn more about our<br />

impact and to <strong>join</strong> AIBS, visit<br />

www.aibs.org.<br />

1900 CAMPUS COMMONS DR.<br />

STE 200<br />

RESTON, VA 20191<br />

WWW.AIBS.ORG<br />

366 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


Articles<br />

Scenario Studies as a Synthetic and<br />

Integrative Research Activity for<br />

Long-Term Ecological Research<br />

Jonathan R. thompson, aRnim Wiek, FRedeRick J. sWanson, stephen R. caRpenteR, nancy FResco,<br />

teResa hollingsWoRth, thomas a. spies, and david R. FosteR<br />

Scenario studies have emerged as a powerful approach for synthesizing diverse forms <strong>of</strong> research and for articulating and evaluating alternative<br />

socioecological futures. Unlike predictive modeling, scenarios do not attempt to forecast the precise or probable state <strong>of</strong> any variable at a given<br />

point in the future. Instead, comparisons among a set <strong>of</strong> contrasting scenarios are used to understand the systemic relationships and dynamics<br />

<strong>of</strong> complex socioecological systems and to define a range <strong>of</strong> possibilities and uncertainties in quantitative and qualitative terms. We describe five<br />

examples <strong>of</strong> scenario studies affiliated with the US Long Term Ecological Research (<strong>LTER</strong>) Network and evaluate them in terms <strong>of</strong> their ability<br />

to advance the <strong>LTER</strong> Network’s capacity for conducting science, promoting social and ecological science synthesis, and increasing the saliency <strong>of</strong><br />

research through sustained outreach activities. We conclude with an argument that scenario studies should be advanced programmatically within<br />

large socioecological research programs to encourage prescient thinking in an era <strong>of</strong> unprecedented global change.<br />

Keywords: socioecological systems, science synthesis, participatory engagement, futures<br />

Public investment in science increasingly comes with the<br />

expectation that research will address the complex<br />

challenges that society faces and will contribute to strategies<br />

that sustain the vitality and integrity <strong>of</strong> socioecological<br />

systems (Reid et al. 2010). During the past 30 years, as the<br />

US National Science Foundation (NSF)–sponsored Long<br />

Term Ecological Research (<strong>LTER</strong>) Network has grown from<br />

6 to 26 research sites, ranging from Alaska to Antarctica and<br />

from urban ecosystems to forests to coral reefs. The <strong>LTER</strong><br />

Network has increasingly used innovative approaches to<br />

connect science to society: from direct engagement with<br />

policymakers to educational programs in public schools<br />

and s<strong>of</strong>tware applications that inform the public about the<br />

structure, function, and state <strong>of</strong> socioecological systems<br />

(Bestelmeyer et al. 2005, Driscoll et al. 2012 [in this issue],<br />

Robertson et al. 2012 [in this issue]). Scenario studies are<br />

one such approach and are emerging as a powerful tool for<br />

synthesizing the results <strong>of</strong> <strong>LTER</strong> science and for engaging<br />

with stakeholders to consider socioecological futures. There<br />

is a rich history involving the use <strong>of</strong> scenarios to encourage<br />

prescient thinking, largely born out <strong>of</strong> military and corporate<br />

planning (Kahn 1962, Wack 1985). Scenarios <strong>of</strong>fer a<br />

framework for practitioners to integrate diverse modes <strong>of</strong><br />

knowledge and to explicitly recognize those components<br />

<strong>of</strong> complex systems that are understood and those that are<br />

uncertain. In this way, scenario studies can describe multiple<br />

ways in which our shared socioecological future may unfold,<br />

sometimes including a visioning process that is focused on<br />

the specific attributes <strong>of</strong> one preferred future condition.<br />

Although the first type <strong>of</strong> scenario studies yield impartial<br />

descriptions <strong>of</strong> a range <strong>of</strong> possible future states, visioning<br />

describes desirable future states (or visions)—for instance,<br />

according to sustainability principles or stakeholder agreement<br />

(Swart et al. 2004, Carpenter and Folke 2006, Weaver<br />

and Rotmans 2006). Both types <strong>of</strong> scenario are contrasted<br />

with probability-based approaches to forward-looking<br />

research below.<br />

We refer here to scenario studies in general as any strategy<br />

for describing plausible future conditions while explicitly<br />

incorporating relevant science, societal expectations, and<br />

internally consistent assumptions about major drivers, relationships,<br />

and constraints (Xiang and Clarke 2003, Iverson<br />

Nassauer and Corry 2004, Raskin et al. 2005, Bolte et al. 2007,<br />

Mahmoud et al. 2009). Unlike predictive modeling, scenario<br />

studies acknowledge the uncertainty inherent in socioecological<br />

systems and therefore do not attempt to forecast the<br />

precise or probable state <strong>of</strong> any variable at a given point in<br />

the future. Instead, comparisons among a set <strong>of</strong> contrasting<br />

scenarios are used to understand the systemic interrelation<br />

and dynamics <strong>of</strong> complex socioecological systems<br />

BioScience 62: 367–376. ISSN 0006-3568, electronic ISSN 1525-3244. © 2012 by American Institute <strong>of</strong> Biological Sciences. All rights reserved. Request<br />

permission to photocopy or reproduce article content at the <strong>University</strong> <strong>of</strong> California Press’s Rights and Permissions Web site at www.ucpressjournals.com/<br />

reprintinfo.asp. doi:10.1525/bio.2012.62.4.8<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 367


Articles<br />

and to define a range <strong>of</strong> possibilities and uncertainties in<br />

quantitative and qualitative terms. The value <strong>of</strong> scenario<br />

studies lies in the process <strong>of</strong> embedding alternative states <strong>of</strong><br />

the future into a transparent problem-solving framework<br />

(Swart et al. 2004). In this way, scenario studies may help<br />

to anticipate change in systems characterized by high levels<br />

<strong>of</strong> irreducible uncertainty and low levels <strong>of</strong> controllability<br />

(Bennett et al. 2003, Peterson et al. 2003a) and evoke new<br />

integrative perspectives and novel concepts <strong>of</strong> ecological<br />

change (Carpenter and Folke 2006, Carpenter et al. 2006).<br />

Although environmental scenarios are developed for a wide<br />

range <strong>of</strong> specific purposes, scenario studies can serve three<br />

widely accepted functions: education and public information,<br />

scientific exploration, and decision support and strategic<br />

planning (Alcamo and Henricks 2008, Henrichs et al.<br />

2010).<br />

In many large science- and environmental-assessment<br />

programs, scenarios have been used to describe and underpin<br />

analyses <strong>of</strong> alternative futures. The Intergovernmental<br />

Panel on Climate Change’s (IPCC) emission scenarios<br />

(Naki enovi and Swart 2000) and the Millennium Ecosystem<br />

Assessment’s scenarios (MA 2005) are perhaps the most well<br />

known, but there are many others (e.g., Sala et al. 2000,<br />

Raskin et al. 2005). Several <strong>LTER</strong> sites, for example, are<br />

deeply involved in scenario studies and even more plan to<br />

be. A recent <strong>LTER</strong> Network–sponsored workshop brought<br />

together 32 social and ecological researchers representing<br />

16 <strong>LTER</strong> sites from around the United States that were<br />

actively engaged in some aspect <strong>of</strong> scenario studies for the<br />

region surrounding their sites (Thompson and Foster 2009).<br />

The participants reaffirmed what may seem self-evident:<br />

<strong>LTER</strong>-based science has several characteristics amenable<br />

to developing regional socioecological scenarios. Credible<br />

socioecological scenarios require a deep understanding <strong>of</strong><br />

long-term environmental dynamics—a signature strength<br />

<strong>of</strong> <strong>LTER</strong> science—and a tight coupling <strong>of</strong> ecological and<br />

social research—an emerging strength <strong>of</strong> and direction for<br />

the <strong>LTER</strong> Network (Collins et al. 2011, Robertson et al. 2012<br />

[this issue]).<br />

Looking forward to the next 30 years <strong>of</strong> <strong>LTER</strong>—and,<br />

more generally, to other research programs concerned<br />

with understanding socioecological systems—we argue<br />

that the application <strong>of</strong> scenario studies can advance prescient<br />

thinking in an era <strong>of</strong> unprecedented rates <strong>of</strong> global<br />

change. For example, an overarching goal set forth in the<br />

<strong>LTER</strong> Network’s Strategic and Implementation Plan is to<br />

use its deep understanding <strong>of</strong> complex socioecological systems<br />

to help anticipate ecological, evolutionary, and social<br />

responses to future environmental change and to inform<br />

societal strategies to adapt to this change (<strong>LTER</strong> Network<br />

2011). This aligns seamlessly with the primary functions <strong>of</strong><br />

scenario studies (i.e., underpinning decision processes, collaborative<br />

learning, and scientific exploration). More specifically,<br />

scenarios can enable site-based research programs to<br />

explore possible as well as desirable and sustainable future<br />

states <strong>of</strong> their respective regions. And through sustained<br />

partnerships with land managers, policymakers, and other<br />

stakeholders, participatory scenarios can increase the societal<br />

relevance <strong>of</strong> their research. Moreover, in scenario studies<br />

in which the socioecological future <strong>of</strong> their regions is<br />

formally considered, new research needs can be identified<br />

while research in temporal and spatial dimensions is scaled<br />

up. Finally, developing and analyzing future scenarios<br />

would provide a platform for working with many scientific<br />

networks, including the National Ecological Observatory<br />

Network (NEON) and Urban Long Term Research Areas<br />

(ULTRA).<br />

Driscoll and colleagues (2012) explore the novel ways in<br />

which scientists have delivered their research findings to<br />

policymakers and managers where they can inform decisions.<br />

They describe several important communications<br />

approaches that allow new science to span boundaries and<br />

to address new challenges. In the present article, we examine<br />

scenario studies as one such boundary-spanning approach,<br />

using several examples <strong>of</strong> scenario studies from <strong>LTER</strong> sites<br />

in order to identify approaches that may be more broadly<br />

applicable. More specifically, we evaluate one emerging and<br />

four mature scenario studies in terms <strong>of</strong> three questions that<br />

address the value <strong>of</strong> scenarios for advancing socioecological<br />

science, programs, and outreach:<br />

A science question: How do the scenario studies relate the past to<br />

the future? Related to this question are those <strong>of</strong> what attributes<br />

<strong>of</strong> the socioecological system will change a lot, what<br />

attributes will change a little, and why. To articulate a range<br />

<strong>of</strong> alternative futures in a plausible and credible manner, it is<br />

necessary to understand the relevant history and trajectory<br />

<strong>of</strong> environmental and social change <strong>of</strong> the system <strong>of</strong> interest<br />

and its component parts. Evaluating future scenarios in<br />

light <strong>of</strong> recent changes, then, leads to an understanding <strong>of</strong><br />

system attributes that are more or less resilient or vulnerable<br />

to future change. This feature <strong>of</strong> scenario studies is especially<br />

valuable as socioecological change approaches “tipping<br />

points” and the need to anticipate and mitigate future<br />

change becomes acute (Scheffer et al. 2001, Rockström et al.<br />

2009).<br />

A programmatic question: How do the scenario studies relate to the<br />

more traditional long-term science occurring within <strong>LTER</strong>? Related<br />

to this question is how they can advance science synthesis.<br />

Scenarios take many forms and, as is shown below, they may<br />

have narrow or expansive thematic scope. But in all the case<br />

studies, scenarios are either informed by or are used as a<br />

platform for applying the core long-term research coming<br />

from the individual <strong>LTER</strong> sites. Consequently, scenarios can<br />

be a compelling approach to science synthesis and for crosssite<br />

comparative analyses.<br />

An outreach question: How do the scenario studies affect the<br />

region and regional society through real-world changes or capacity<br />

building? Future scenarios draw together stakeholders<br />

affected by the hypothesized future changes. By engaging<br />

368 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


in scenario studies, scientists can place themselves in a new<br />

relationship with society. It can be a means for linking<br />

science and decisionmaking in order to support future<br />

and hopefully sustainable trajectories for socioecological<br />

systems, whether they are wild landscapes, natural resource<br />

lands, or urban areas.<br />

Case studies<br />

The case studies showcase diverse ways in which scenario<br />

studies are being employed as a vehicle to understand<br />

the drivers <strong>of</strong> socioecological change, to engage with<br />

regional stakeholders, and to consider shared socioecological<br />

futures. Each is distinct in its motivation, methodology,<br />

and outcomes. To provide a consistent organization for<br />

evaluation, we describe each using a seven-part conceptual<br />

framework (figure 1, table 1) that includes (1) a trigger<br />

or context (i.e., what precipitated the study and in what<br />

environment [e.g., academic, pr<strong>of</strong>essional] it occured), (2) a<br />

goal (i.e., what the leaders and participants in the scenario<br />

study hoped to achieve), (3) construction or methodology<br />

(i.e., what scenario construction method was used), (4) collaboration<br />

(i.e., who was involved in scenario development<br />

[e.g., regional or national stakeholders, experts]), (5) scenarios<br />

(i.e., what scenarios [i.e., topics] were generated and<br />

what the mode or representation [e.g., narratives, visuals,<br />

maps] was), (6) use (i.e., in what ways the scenarios<br />

were used after they were completed [e.g., research, planning<br />

and decisionmaking, education and training]), and<br />

(7) impacts (i.e., what social impacts the application <strong>of</strong> the<br />

scenarios yielded [e.g., it improved the network, influenced<br />

policy decisions, changed pr<strong>of</strong>essional practice, or increased<br />

capacity]).<br />

Figure 1. Analytical framework for evaluating diverse forms <strong>of</strong> scenario studies<br />

conducted throughout the US Long Term Ecological Research Network.<br />

Articles<br />

Northern Highlands Lake District scenarios: The North Temperate<br />

Lakes <strong>LTER</strong> site. The Northern Highlands Lake District<br />

(NHLD) is located in northern Wisconsin. During the past<br />

several decades, the population <strong>of</strong> the region has expanded<br />

significantly, which has resulted in the construction <strong>of</strong> new<br />

housing and infrastructure and a marked increase in tourism<br />

and recreation, which are important components <strong>of</strong><br />

the region’s economy (Peterson et al. 2003b). These developments<br />

have placed additional pressure on the lakes and<br />

surrounding forests, which has caused concern about the<br />

long-term sustainability <strong>of</strong> these resources and the economic<br />

activity they support. Partially on the basis <strong>of</strong> the social-<br />

and natural-science infrastructure developed by scientists<br />

at the North Temperate Lakes (NTL) <strong>LTER</strong> site, the NHLD<br />

was selected as one <strong>of</strong> a few pilot study regions for the<br />

Millennium Ecosystem Assessment Scenarios Working<br />

Group (Carpenter 2008). Participants in the working group<br />

were drawn from local businesses, nongovernmental organizations<br />

(NGOs), tribal organizations, and government, as<br />

well as technical experts from the NTL, to engage in wideranging<br />

conversations about the vulnerabilities and potential<br />

strengths <strong>of</strong> the region, as well as the fears and hopes for<br />

the landscape’s future. Through an iterative process, facilitators<br />

distilled diverse storylines into four composite scenarios<br />

that portrayed alternative drivers <strong>of</strong> change over the<br />

coming 25 years. They commissioned an artist to develop<br />

illustrations for the storylines and allowed the stakeholders<br />

to refine the scenarios. The four scenarios represented very<br />

different pathways that the NHLD might follow with regard<br />

to future population trends, zoning requirements, and recreation<br />

policies. They were designed to be internally coherent,<br />

logically consistent, and plausible. In this way, it was possible<br />

to explore each scenario individually,<br />

as well as the similarities<br />

and differences among them.<br />

The scenarios were presented as<br />

illustrated stories in a short book<br />

and on a Web site. After the scenarios<br />

were settled, simulation<br />

modeling was begun in order<br />

to couple the qualitative stories<br />

to the anticipated quantitative<br />

changes in ecological systems.<br />

Together, the scenarios and outputs<br />

from simulations were used<br />

to spark conversations about the<br />

future in meetings with individual<br />

decisionmakers, in large<br />

public meetings, and in online<br />

forums. The facilitators also ran<br />

an online survey to gather public<br />

reactions to the scenarios. In<br />

subsequent years, the scenarios<br />

have been used in several university<br />

classes through role-playing<br />

adaptive-management games.<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 369


Articles<br />

Table 1. Features <strong>of</strong> the US Long Term Ecological Research (<strong>LTER</strong>) Network sites and their associated projects discussed in this article.<br />

Scenarios: topic (1)<br />

and format (2) Uses Impacts<br />

Method (1) and<br />

steps (2) Collaboration<br />

Triggers and<br />

context Goals<br />

Project Title, <strong>LTER</strong> site, space,<br />

and record length<br />

1. Increased network<br />

2. Influenced policy<br />

choices<br />

3. Increased<br />

outreach capacity<br />

<strong>of</strong> NTL<br />

1. Role-playing<br />

adaptive-management<br />

games<br />

2. Simulation models<br />

1. Landscape change<br />

2. Narratives (stories);<br />

artistic renditions;<br />

book; Web page<br />

Input from<br />

regional<br />

stakeholders<br />

1. Workshops<br />

2. Iteratively<br />

review scenarios<br />

and combine<br />

scenario elements<br />

1. Create contrasting<br />

scenarios<br />

2. Stakeholder<br />

outreach and<br />

collaborative learning<br />

3. Provide input for<br />

<strong>LTER</strong><br />

1. Academic<br />

2. Pilot scenario<br />

study for Millennium<br />

Ecosystem<br />

Assessment<br />

Northern Highlands Lake District<br />

<strong>LTER</strong> site: North Temperate<br />

Lakes (NTL)<br />

Space: 5300 square kilometers<br />

(km2 )<br />

Time: 25 years forward<br />

1. Influenced forest<br />

practices in North<br />

America<br />

1. Development <strong>of</strong><br />

forest management<br />

plan<br />

2. Begin plan<br />

implementation<br />

3. Public<br />

demonstration project<br />

1. Forest landscape<br />

structure and function<br />

under alternative<br />

disturbance regimes<br />

2. Narratives<br />

(technical); maps;<br />

report<br />

Research-<br />

Forest<br />

Service<br />

managers<br />

collaboration<br />

1. Assess fire<br />

history<br />

2. Write<br />

alternative<br />

scenarios<br />

3. Formally<br />

analyze<br />

4. Elicit feedback<br />

1. Create contrasting<br />

scenarios<br />

2. Test history as<br />

management guide<br />

3. Provide input for<br />

management and policy<br />

1. Pr<strong>of</strong>essional<br />

(forest planning)<br />

2. Alternative forest<br />

management<br />

Blue River Landscape Plan and<br />

Study<br />

<strong>LTER</strong> site: Andrews Forest (AND)<br />

Space: 23 km2 Time: 500 years back<br />

200 years forward<br />

1. Incorporated into<br />

general plan<br />

2. Influenced<br />

policies<br />

3. Built planners and<br />

citizens’ capacity<br />

1. Revision <strong>of</strong><br />

general plan<br />

2. Pr<strong>of</strong>essional<br />

training, fundraising<br />

3. Primary and<br />

university classes<br />

1. Urban development<br />

(economy,<br />

environment, society,<br />

infrastructure)<br />

2. Narratives<br />

(technical, stories);<br />

illustrations; graphs;<br />

report<br />

Input from<br />

regional<br />

stakeholders<br />

1. Review<br />

scenario literature<br />

2. Elicit vision<br />

statements<br />

3. Formally<br />

analyze<br />

4. Elicit feedback<br />

1. Create sustainability<br />

vision and contrasting<br />

scenarios<br />

2. Stakeholder<br />

outreach and<br />

collaborative learning<br />

3. Provide input for<br />

General Plan<br />

1. Pr<strong>of</strong>essional<br />

(urban planning)<br />

2. Revision <strong>of</strong> the<br />

city’s general plan<br />

2. Future Vision and Scenarios<br />

for Phoenix<br />

<strong>LTER</strong> site: Central Arizona<br />

Phoenix (CAP)<br />

Space: 1344 km2 Time: 40 years forward<br />

1. Incorporated into<br />

report to governor’s<br />

panel on climate<br />

change<br />

2. Informed<br />

decisions<br />

(communities)<br />

1. Modeling <strong>of</strong><br />

ecological impacts;<br />

planning<br />

2. Energy projects<br />

3. <strong>University</strong> classes<br />

Expert-driven 1. Climate and<br />

biophysical<br />

parameters<br />

2. Reports; interactive<br />

Web mapping<br />

1. Downscale<br />

climate scenarios<br />

2. Model<br />

biophysical<br />

parameters<br />

1. Create contrasting<br />

scenarios<br />

2. Stakeholder<br />

outreach and<br />

collaborative learning<br />

3. Create research and<br />

communication tool on<br />

climate change<br />

1. Academic<br />

2. Societal need for<br />

regional information<br />

regarding climate<br />

change<br />

Scenarios Network for Alaska/<br />

Arctic Planning (SNAP)<br />

<strong>LTER</strong> site: Bonanza Creek (BNZ)<br />

Space: AK / 1.7 million km2 Time: 20 years back<br />

100 years forward<br />

1. Forest conversion,<br />

energy production,<br />

climate mitigation,<br />

conservation, etc.<br />

2. Narratives<br />

(technical), maps,<br />

graphs, visualizations<br />

Input from<br />

national<br />

and regional<br />

stakeholders<br />

1. Elicit narratives<br />

2. Parameterize<br />

interactive<br />

simulation models<br />

1. Create contrasting<br />

scenarios<br />

2. Provide input<br />

for cross-site <strong>LTER</strong><br />

comparative research<br />

1. Academic<br />

2. Identify regional<br />

characteristics that<br />

alter continentalscale<br />

drivers <strong>of</strong><br />

global change<br />

North American Forest Scenarios<br />

<strong>LTER</strong> sites: Harvard Forest,<br />

<strong>Coweeta</strong>, NTL, AND, BNZ<br />

Space: 15 landscapes /<br />

5600–27,000 km2 Time: 50 years forward<br />

370 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org<br />

This is an emerging project whose use and<br />

impacts have yet to be realized.


Overall, the process <strong>of</strong> scenario planning increased the<br />

connectivity <strong>of</strong> the NTL group to other groups actively<br />

engaged in the region. The process also had the immediate<br />

effect <strong>of</strong> increasing the extent <strong>of</strong> networking among diverse<br />

stakeholders who had long histories in the NHLD but little<br />

prior interaction. This connectivity improved the outreach<br />

capacity <strong>of</strong> the NTL and introduced the science to more<br />

potential users.<br />

Blue River Landscape Plan and Study scenarios: The Andrews<br />

Forest <strong>LTER</strong> site. In the early 1990s, forest policy in the<br />

Pacific Northwest abruptly shifted from several decades <strong>of</strong><br />

management in which commodity (timber) production was<br />

emphasized to one that was focused on species conservation<br />

(Duncan and Thompson 2006). The Northwest Forest Plan<br />

(NWFP) was the policy instrument that codified this change.<br />

The NWFP was itself the product <strong>of</strong> a scenario approach<br />

in which scientists, including several from the Andrews<br />

Forest (AND) <strong>LTER</strong> site, prepared 10 management scenarios<br />

(alternatives) for the Clinton administration (FEMAT 1993),<br />

with one selected as the basis <strong>of</strong> the NWFP. Within the<br />

plan, 10 large parcels (40,000–200,000 hectares) <strong>of</strong> federal<br />

forestlands were delineated as adaptive management areas<br />

(AMAs), in which scientists and land managers were charged<br />

with examining alternative strategies to meeting conservation<br />

and timber-production goals. At the Central Cascades<br />

AMA, which contains the AND site, <strong>LTER</strong> scientists teamed<br />

with land managers <strong>of</strong> the Willamette National Forest<br />

to construct a landscape-scale forest-management plan.<br />

They used science-based visioning to describe an ecologically<br />

desirable but unconventional approach to long-term<br />

management that was informed by historic fire regimes<br />

and landscape dynamics rather than by standard reservedesign<br />

criteria. The resulting scenario, called the Blue River<br />

Landscape Plan and Study (BRLP; Ecoshare 2011), patterned<br />

timber harvests on the historical wildfire regime (Cissel<br />

et al. 1999). The BRLP’s resulting “disturbance-based” or<br />

“historical range <strong>of</strong> variability” approach began with a<br />

dendrochronology-based fire-history study that spanned<br />

the previous 500 years (Weisberg 1998), which was used to<br />

semiquantitatively generate a map <strong>of</strong> three fire-regime types<br />

distinguished by fire frequency and severity and constrained<br />

in part by topography. A forest planner used this fire-regime<br />

map and a map <strong>of</strong> the then-present distribution <strong>of</strong> forest<br />

age classes to project harvests 200 years into the future. The<br />

plan included novel approaches to individual-tree and patch<br />

retention and to cutting-rotation length to emulate the<br />

historical wildfire severity and frequency. This disturbancebased<br />

approach to landscape management was contrasted<br />

with expected future management under the NWFP, which<br />

is based on the management <strong>of</strong> unharvested reserves and<br />

matrix land (i.e., actively harvested areas between reserves),<br />

for its conservation value for selected species (Cissel et al.<br />

1999). Public reaction to the disturbance-based plan was<br />

assessed through surveys and field-tour discussions. The<br />

public had a significant level <strong>of</strong> acceptance, although the<br />

Articles<br />

concept and vocabulary were unfamiliar to many (Shindler<br />

and Mallon 2009). The BRLP has since served as an actual<br />

management plan intended for implementation, as well as<br />

a demonstration project used for critical discourse within<br />

science, land-management, and public circles. The BRLP<br />

helped reveal to the various stakeholders that an understanding<br />

<strong>of</strong> natural-disturbance regimes can lead to a viable<br />

approach to harvesting and the conservation <strong>of</strong> biodiversity<br />

and ecosystem functions. Lessons being learned from the<br />

BRLP have been widely communicated and may be used for<br />

new applications as society charts the management <strong>of</strong> public<br />

lands through a dynamic future socioecological system<br />

(Spies and Duncan 2009).<br />

Future vision and scenarios for Phoenix: Central Arizona–Phoenix<br />

<strong>LTER</strong> site. In autumn 2009, the city <strong>of</strong> Phoenix, the Central<br />

Arizona–Phoenix (CAP) <strong>LTER</strong> site, and the School <strong>of</strong><br />

Sustainability at Arizona State <strong>University</strong> initiated a research<br />

project entitled “The Future <strong>of</strong> Phoenix: Crafting Sustainable<br />

Development Strategies.” (Wiek et al. 2012). The purpose <strong>of</strong><br />

the combined visioning and scenario study was to create—<br />

through a collaboration <strong>of</strong> expert facilitators, scientists,<br />

and stakeholders—a vision and contrasting scenarios that<br />

captured a spectrum <strong>of</strong> possible future developments <strong>of</strong><br />

Phoenix. The vision described Phoenix as a desirable and<br />

sustainable socioecological system, as was determined by<br />

stakeholder agreement and sustainability principles (Gibson<br />

2006). The contrasting scenarios described alternative—less<br />

desirable and less sustainable—future states and represented<br />

what might happen if the vision is not achieved (Withycombe<br />

and Wiek 2011). In the vision that resulted from this participatory<br />

research process, Phoenix was described as being<br />

comprised <strong>of</strong> vibrant communities, where healthy food,<br />

clean water, fresh air, excellent educational opportunities,<br />

satisfying jobs, and public transit options are available to<br />

all citizens; where strong local businesses take advantage <strong>of</strong><br />

local assets to build a diverse and community-oriented urban<br />

economy; where governance is open and transparent and<br />

reflects the values <strong>of</strong> all people, regardless <strong>of</strong> their power or<br />

influence; and where the urban ecological system is preserved<br />

and cared for, so that it can be sustained for generations to<br />

come. In the contrasting scenarios, two alternative future<br />

states <strong>of</strong> Phoenix are described: Phoenix behind the times, in<br />

which the city acknowledges critical challenges from climate<br />

change to environmental degradation and social tensions<br />

but cannot seem to keep pace with other regions in creating<br />

a healthy socioecological system, and Phoenix overwhelmed,<br />

in which the city ignores long-term challenges, upholds the<br />

growth paradigm, and continues to overextend its capacities<br />

and to jeopardize sustainable future development pathways.<br />

The vision was presented in three formats: stories with photographs,<br />

a narrative, and a map <strong>of</strong> priorities. The scenarios<br />

were presented as newspaper cover pages with illustrations<br />

and as narratives. Visioning and scenario construction combined<br />

several methods, such as consistency analysis, diversity<br />

analysis, sustainability appraisal, and trade-<strong>of</strong>f analysis<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 371


Articles<br />

(Wiek et al. 2006, Withycombe and Wiek 2011). In the study,<br />

the knowledge, preferences, and values <strong>of</strong> a broad range <strong>of</strong><br />

experts and regional stakeholders were elicited, deliberated,<br />

and integrated. The core stakeholder-engagement activities<br />

were 15 meetings in all parts <strong>of</strong> the city, designed to elicit<br />

vision statements, and a visioning workshop, in which the<br />

elicited vision statements were revisited, systemically linked,<br />

and reprioritized on the basis <strong>of</strong> potential conflicts. More<br />

than 100 citizens, city planners, business representatives,<br />

nonpr<strong>of</strong>it organization members, and researchers participated<br />

in this workshop. The vision and the scenarios are now<br />

being used by the city administration in planning, training,<br />

and fund raising, as well as for teaching purposes in public<br />

schools and at Arizona State <strong>University</strong>. Overall, the project<br />

spurred a rich public discussion about the future directions<br />

in which Phoenix may be headed and about the degree to<br />

which these directions align with a desirable and sustainable<br />

future vision. The substance <strong>of</strong> the vision and scenarios has<br />

been incorporated into the current public-hearing draft <strong>of</strong><br />

the city’s general plan (City <strong>of</strong> Phoenix 2010), which is the<br />

city’s most important guide for long-term planning and<br />

development. It provides direction for planning and allows<br />

researchers, policy analysts, and stakeholders to evaluate the<br />

effectiveness <strong>of</strong> policies and actions.<br />

The Scenarios Network for Alaska Planning: The Bonanza Creek<br />

<strong>LTER</strong> site. Arctic and boreal forests are warming about twice<br />

as fast as the global average, creating widespread concern and<br />

interest in the patterns and consequences <strong>of</strong> climate change,<br />

especially among northern residents. At the Bonanza Creek<br />

(BNZ) <strong>LTER</strong> site, scenarios have been used as both a research<br />

and a communications tool to explore the consequences <strong>of</strong><br />

recent and projected climate change in Alaska. Rapid change<br />

experienced in Alaska has focused public attention on these<br />

scenarios, which in turn has led to the establishment <strong>of</strong><br />

a research partnership: the Scenarios Network for Alaska<br />

Planning (SNAP; www.snap.uaf.edu), which comprises the<br />

BNZ, the <strong>University</strong> <strong>of</strong> Alaska, and several state and federal<br />

agencies, local communities, and nonpr<strong>of</strong>its. SNAP has<br />

developed scenarios <strong>of</strong> future climate as high-resolution<br />

maps <strong>of</strong> mean monthly historical and projected temperature<br />

and precipitation (between 1900 and 2100) that account for<br />

the known effects <strong>of</strong> topography and the movement <strong>of</strong> air<br />

masses. By downscaling the five global circulation models<br />

that were shown to perform best in the far north and by providing<br />

outputs for three emission scenarios defined by the<br />

IPCC, SNAP <strong>of</strong>fers stakeholders a range <strong>of</strong> possible climate<br />

futures, rather than a single prediction. Moreover, discussion<br />

and interpretation <strong>of</strong> uncertainty play a large role in SNAP<br />

projects. The actual technical work <strong>of</strong> downscaling the<br />

climate scenarios is an expert-driven process. However, all<br />

SNAP projects are collaborations between SNAP researchers<br />

and land managers or other stakeholders. These stakeholders<br />

help determine how the data will be linked to landscape<br />

models, existing data sets, and local knowledge and<br />

how the resulting scenarios will be used and interpreted.<br />

For example, in collaboration with US Fish and Wildlife<br />

Service, The Nature Conservancy, and other partners, SNAP<br />

researchers have used climate-change scenarios to model<br />

potential biome shifts and changes in the ranges <strong>of</strong> endemic<br />

and invasive species. In the resulting report (Murphy et al.<br />

2010), barren-ground caribou, Alaska marmots, trumpeter<br />

swans, and reed canary grass were used as indicator species<br />

to assess multiple possible futures, given the possible range<br />

<strong>of</strong> climate-change impacts. In addition, the climate scenarios<br />

have been used directly by Alaska communities in order to<br />

inform decisionmaking about the future sustainability <strong>of</strong><br />

hydroelectric generation and other energy project plans<br />

(Cherry et al. 2010).<br />

The American Forest Futures Projects: The Harvard Forest,<br />

Andrews Forest, Bonanza Creek, <strong>Coweeta</strong>, and North Temperate<br />

Lakes <strong>LTER</strong> sites. Scenario planning is the focus <strong>of</strong> American<br />

Forest Futures Projects, an emerging group <strong>of</strong> cross-site<br />

<strong>LTER</strong> projects that are being advanced across the major<br />

forested regions <strong>of</strong> the United States in collaboration with<br />

major NSF-sponsored programs (including, e.g., five <strong>LTER</strong><br />

Network sites, NEON, ULTRA), the US Forest Service, the<br />

US Geological Survey, the Smithsonian Institution, and<br />

several NGOs (e.g., The Nature Conservancy, the Ecological<br />

Society <strong>of</strong> America, the Heinz Center). The projects were<br />

designed to address pressing scientific questions related to<br />

how and why forested regions vary in their socioecological<br />

responses to global change (www.wildlandsandwoodlands.<br />

org/node/133). In a parallel thrust, the researchers designed<br />

the projects as a means to advance several burgeoning<br />

network-level <strong>LTER</strong> Network priorities, such as national-<br />

and regional-scale stakeholder engagement, development<br />

<strong>of</strong> models and visualization tools, and communication with<br />

policymakers. The projects rely on tiered scenarios developed<br />

at national and regional scales to address at least four<br />

dominant themes: economic development, energy exploitation<br />

(e.g., bioenergy), climate mitigation and adaptation,<br />

and landscape-scale conservation.<br />

The researchers have convened a national advisory board<br />

<strong>of</strong> experts, consisting <strong>of</strong> federal agency and national NGO<br />

representatives, to develop national-scale scenarios describing<br />

the anticipated drivers and the changes associated with<br />

each <strong>of</strong> the themes. After they are finalized, the suite <strong>of</strong><br />

national-scale scenarios will be reinterpreted at regional<br />

scales around the country by engaging regional stakeholders<br />

(e.g., natural-resource managers, scientists, agency <strong>of</strong>ficials,<br />

conservation pr<strong>of</strong>essionals) in the development <strong>of</strong> narratives<br />

in which the local manifestation <strong>of</strong> each <strong>of</strong> the national<br />

scenarios is described. This process <strong>of</strong> downscaling global<br />

scenarios to local scales can be an effective approach for<br />

participatory capacity building and for motivating behavioral<br />

changes in local stakeholders and decisionmakers (Shaw<br />

et al. 2009). The regional scenarios will, in turn, be used to<br />

define land-use assumptions and to parameterize spatially<br />

interactive landscape-simulation models (e.g., Thompson<br />

et al. 2011) within study a series <strong>of</strong> study landscapes<br />

372 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


(5600–27,000 sqaure kilometers) dispersed throughout five<br />

forest regions: the Northeast, the Lake States, the Southeast,<br />

the Pacific Northwest, and Alaska. Using the scenarios and<br />

simulation outputs, the group plans to conduct a series <strong>of</strong><br />

within- and across-region comparisons to evaluate how ecosystem<br />

attributes and services change in response to similar<br />

national scenarios <strong>of</strong> global change along multiple social<br />

and ecological gradients, including productivity and topography,<br />

land tenure and demographics, land-use history and<br />

policies, disturbance regimes, and climate. This approach for<br />

coupling qualitative and quantitative scenarios has informed<br />

prescient planning and policy and has generated a rich set <strong>of</strong><br />

fundamental research questions (see, e.g., Spies et al. 2007,<br />

Schmitt Olabisi et al. 2010). Although it is still in its early<br />

stages, the American Forest Futures Projects have already<br />

spurred dialogue among dozens <strong>of</strong> agencies and stakeholder<br />

groups in an effort to describe an envelope <strong>of</strong> plausible<br />

futures for forested landscapes across the country.<br />

Discussion <strong>of</strong> the case studies<br />

From these case studies, we <strong>of</strong>fer responses to our initial<br />

three questions:<br />

How do the scenario studies relate the past to the future? Unlike<br />

predictive modeling, in scenario studies, no level <strong>of</strong> confidence<br />

is asserted that any particular changes will occur.<br />

Instead, several possible changes are integrated into a set<br />

<strong>of</strong> potential future pathways in which the major drivers are<br />

logical and consistent across scenarios. In doing so, scenario<br />

studies provide useful contexts for addressing questions<br />

about how past change may or may not help understand<br />

future change, given a range <strong>of</strong> possible societal dynamics<br />

(e.g., population growth, shifting demographics, land-use<br />

change) and possible environmental dynamics (e.g., climate,<br />

invasive species, major disturbance events). The BRLP, for<br />

example, was based on the concept that important elements<br />

<strong>of</strong> the future will represent an extension <strong>of</strong> historic trajectories<br />

<strong>of</strong> change and also incorporated a detailed understanding<br />

<strong>of</strong> historic landscape change over many centuries in<br />

response to disturbance by wildfire, flood, and landslides.<br />

The future landscape is expected to include continued<br />

responses <strong>of</strong> forest and stream systems to those past disturbance<br />

regimes, as well as future regimes resulting from direct<br />

and indirect human influences and natural processes. The<br />

vision <strong>of</strong> management based on historical ecosystem dynamics<br />

acted as a medium for discussing these complexities<br />

among scientists, land managers, and the public. By examining<br />

a range <strong>of</strong> plausible futures, scientists and stakeholders<br />

could consider the range <strong>of</strong> social and ecological uncertainty<br />

and learn about the attributes that drive change, even if<br />

they do not know exactly what level <strong>of</strong> change will occur.<br />

The value <strong>of</strong> considering how historical trends may diverge<br />

along multiple alternative pathways was also evident in the<br />

NHLD, where big shifts in the status quo were foreseen from<br />

both internal and external forces. Internally, concern was<br />

focused on political gridlock and a lack <strong>of</strong> capacity to make<br />

Articles<br />

collective decisions. Externally, concern was focused on the<br />

demographic and economic drivers <strong>of</strong> massive development<br />

in the Northern Highland. These internal and external social<br />

drivers were thought to affect the resilience <strong>of</strong> the region to<br />

climate change and other biophysical drivers. In the NHLD<br />

scenarios that were seen as optimistic by most respondents,<br />

the social and political system <strong>of</strong> the region changed in ways<br />

that facilitated resilient responses to large-scale biophysical<br />

changes.<br />

Evaluating multiple scenarios that are informed but not<br />

constrained by history can have great advantages over the<br />

predictive modeling <strong>of</strong> future conditions, which can pose<br />

serious challenges even in seemingly straightforward analyses<br />

and when long-term data exist. For example, we live in<br />

a period <strong>of</strong> high concern over climate change, yet in many<br />

cases, climate-change signals are difficult to interpret amid<br />

the temporal variability <strong>of</strong> climate at multiple temporal<br />

scales, even with 50-year historical records (Jones et al. 2012<br />

[in this issue]). In Alaska, where the impacts <strong>of</strong> climate<br />

change have been felt by communities for many years, gathering<br />

information about how land managers and residents<br />

are experiencing and reacting to change <strong>of</strong>fers crucial information<br />

about how to address future change. Those who are<br />

living with changes, such as increased fire risk, treeline shift,<br />

or severe coastal erosion, can <strong>of</strong>fer information about how<br />

these changes are already being experienced and managed<br />

and can make specific requests about what kind <strong>of</strong> climate<br />

information and what models would be most useful. The<br />

SNAP scenarios span a range <strong>of</strong> potential climate futures,<br />

informed by the best available science, but without the<br />

false precision <strong>of</strong> predictive models. The climate scenarios<br />

can, in turn, be integrated into socioecological models and<br />

narrative stories that define potential futures in a way that<br />

managers and residents can relate to and plan for. The use <strong>of</strong><br />

scenario studies is therefore a productive means for relating<br />

historical landscape change to forward-looking analyses <strong>of</strong><br />

unpredictable socioecological systems.<br />

How do the scenario studies relate to the more traditional<br />

long-term science occurring within the <strong>LTER</strong> Network? Scenario<br />

studies are <strong>of</strong>ten a form <strong>of</strong> synthesis <strong>of</strong> both biophysical and<br />

social science. Framing plausible alternative depictions <strong>of</strong><br />

the future and evaluating them promotes highly integrative<br />

thinking. The social, political, and land-management contexts<br />

in which <strong>LTER</strong> sites are embedded increasingly demand<br />

a level <strong>of</strong> integrated analysis that is uncommon in traditional<br />

scientific research but that is central to scenario studies. In<br />

each <strong>of</strong> the case studies, the strengths <strong>of</strong> traditional <strong>LTER</strong><br />

science—notably, monitoring and evaluating the impacts<br />

<strong>of</strong> long-term environmental change—were integrated into<br />

the scenario studies. For example, in the American Forest<br />

Futures Projects, the future response <strong>of</strong> forest ecosystems to<br />

climate and land-use change are evaluated over time using<br />

ecosystem-process models that are developed and constrained<br />

by <strong>LTER</strong>. National and regional stakeholders define<br />

the narratives that guide their assumptions about land use<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 373


Articles<br />

and other societal responses, thus linking the scenarios to<br />

the legacy <strong>of</strong> “hard science” traditionally upheld by <strong>LTER</strong>.<br />

Likewise, in Alaska, where climate change is a topic that has<br />

reached a relatively high level <strong>of</strong> public awareness, SNAP<br />

and the BNZ are already engaged in the type <strong>of</strong> synthesis<br />

that forges connections between <strong>LTER</strong> and exogenous social<br />

and economic pressures. In working with groups such as the<br />

National Park Service, the Fairbanks North Star Borough<br />

Climate Change Task Force (www. investfairbanks.com/<br />

Taskforces/climate.php), and Alaska’s governor’s Subcabinet<br />

on Climate Change (www.snap.uaf.edu/projects/governorssubcabinet-climate-change-0),<br />

SNAP has received input from<br />

partners who are concerned about the interplay among<br />

multiple ecological factors in the context <strong>of</strong> budgets, short<br />

planning cycles, and public criticism. Scenario planning has<br />

proven to be a useful tool in this context, because it allows<br />

scientists to <strong>of</strong>fer the best available information in a way that<br />

incorporates uncertainty and that allows planners to assess<br />

risk on their own terms.<br />

Importantly, in all <strong>of</strong> the case studies discussed here,<br />

knowledge transfer and syntheses flowed in both directions,<br />

whereby the science informed the scenarios and, in turn,<br />

the scenarios informed the science. For example, within<br />

the NTL program, the scenarios contributed to a shift<br />

toward explicit long-term thinking about socioecological<br />

change that has influenced research and the planning <strong>of</strong> site<br />

activities. Recent projects have been focused on thresholds<br />

for abrupt change in fisheries, food webs, and invasivespecies<br />

dynamics, for example. Climate change has been a<br />

long-standing interest at the NTL that is also serving as an<br />

organizing focus for hypothesis-driven long-term research.<br />

Similarly, the Future Vision and Scenarios for Phoenix<br />

study serves as a pilot project and will be continued with<br />

even more emphasis on the integration <strong>of</strong> previous CAP<br />

work. To this end, a formal synthesis workshop will be conducted<br />

in order to make the links to the other CAP working<br />

groups explicit and to plan future contributions. The study<br />

has also been proven to stimulate collaboration between<br />

CAP researchers and other research groups at Arizona State<br />

<strong>University</strong> that are already engaged or interested in scenario<br />

studies in the CAP region (across different spatial levels and<br />

topic areas).<br />

How do the scenario studies impact the region and the<br />

regional communities through real-world changes or capacity<br />

building? Scenario studies are a distinctive form <strong>of</strong> science<br />

engagement with society and one that has only recently<br />

been employed in <strong>LTER</strong>. They are much more participatory<br />

than traditional outreach that consists <strong>of</strong> a delivery <strong>of</strong> scientific<br />

findings to policymakers, managers, and the public.<br />

For example, the Future Vision and Scenarios for Phoenix<br />

study has served as a powerful medium to enhance the<br />

engagement between CAP researchers and regional stakeholders.<br />

Although CAP research has included stakeholder<br />

engagement since its inception in 1997, the dominant mode<br />

<strong>of</strong> operation has been one-way elicitation. The scenario<br />

study, in contrast, has demonstrated how more interactive<br />

engagement can enhance the interest and ownership for<br />

the challenges as well as potential solutions across different<br />

stakeholder groups. The ongoing scenario work further<br />

expands the participatory-research methodology through<br />

advanced collaborative visualization, walking audits, and<br />

exploration courses. Similarly, the BRLP scenario has proven<br />

to be very useful process for generating discussion among<br />

varied stakeholders—pro- and antilogging groups alike—<strong>of</strong><br />

the dynamics <strong>of</strong> forest ecosystems and the relevance <strong>of</strong><br />

natural variability for future management. Having a management<br />

scenario based on historic disturbance regimes<br />

has fostered public understanding <strong>of</strong> landscape dynamism<br />

at a time when some elements <strong>of</strong> the public seek to freeze<br />

components <strong>of</strong> the landscape as though it were a museum<br />

diorama.<br />

<strong>LTER</strong> scenario studies have also led to novel uses <strong>of</strong><br />

technology to engage the public. Taking just a few examples<br />

from the case studies, the American Forest Futures Projects<br />

are developing a Web-based course modeled after the highly<br />

successful <strong>LTER</strong> Network–sponsored “From Yardstick to<br />

Gyroscope” class (http://news.lternet.edu/article170.html),<br />

in which the science <strong>of</strong> scenario studies will be taught to<br />

university students and landscape planners. As part <strong>of</strong> the<br />

BNZ SNAP project, a Web-based community charts tool<br />

has been developed (www.snap.uaf.edu/community-charts)<br />

that allows Alaska residents to compare a range <strong>of</strong> possible<br />

future climate scenarios for their own village (from among<br />

more than 350 in the state) and a Google Earth interface<br />

(www.snap.uaf.edu/google-earth-maps) that lets users zoom<br />

in on their own region and define which models they would<br />

like to explore. At CAP, as part <strong>of</strong> the Future Vision and<br />

Scenarios for Phoenix study, the Decision Theater at Arizona<br />

State <strong>University</strong> is to be used to evaluate the effectiveness<br />

<strong>of</strong> those participatory processes for capacity building and<br />

decisionmaking.<br />

Overall, the scenario studies presented here demonstrate<br />

how considering our shared socioecological future<br />

can motivate sustained engagement between science and<br />

society. We believe that the long-term dimension <strong>of</strong> the US<br />

<strong>LTER</strong> Network makes it a highly suitable venue for scenario<br />

studies. Indeed, the 30-year history <strong>of</strong> the <strong>LTER</strong> program<br />

provides time for development <strong>of</strong> social networks with key<br />

parts <strong>of</strong> society (e.g., decisionmakers, NGOs, government<br />

agencies with lands responsibilities, interested members <strong>of</strong><br />

the public). Finally, members <strong>of</strong> the <strong>LTER</strong> Network’s science<br />

community live and work in their studied landscapes and<br />

are themselves stakeholders with a deep personal stake in<br />

its future.<br />

Conclusions<br />

As the case studies show, scenarios can be an effective<br />

approach for synthesizing science in a major research<br />

program and can lead to an improved understanding<br />

<strong>of</strong> socioecological change. Scenario studies <strong>of</strong>fer a flexible<br />

framework for integrating the best available ecological<br />

374 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


science with the myriad <strong>of</strong> uncertainties that are inherent<br />

in global change. Using scenarios, large ecological research<br />

programs, such as the <strong>LTER</strong> Network, NEON, and ULTRA,<br />

can lead societal discussion regarding the future <strong>of</strong> their<br />

landscapes. The core strengths <strong>of</strong> the <strong>LTER</strong> Network in<br />

particular—its history <strong>of</strong> long-term, place-based studies; its<br />

community <strong>of</strong> scholars committed to integrative research<br />

across disciplines and service to society; and its diversity <strong>of</strong><br />

landscapes, stakeholders, and disturbance regimes—make<br />

it ideally suited to leading scenario studies in each <strong>of</strong> the<br />

landscapes in which <strong>LTER</strong> sites are present. As such, we suggest<br />

that scenario studies be advanced in collaboration with<br />

many other research groups and agencies as a network-wide<br />

activity to promote research in socioecological systems and<br />

cross-site comparative analyses across the network.<br />

Acknowledgments<br />

Support for this work was provided by US National Science<br />

Foundation Long Term Ecological Research Network awards<br />

to Harvard <strong>University</strong>, Arizona State <strong>University</strong>, Oregon<br />

State <strong>University</strong>, the <strong>University</strong> <strong>of</strong> Alaska, and the <strong>University</strong><br />

<strong>of</strong> Wisconsin and by the US Forest Service.<br />

References cited<br />

Alcamo J, Henrichs T. 2008. Towards guidelines for environmental scenario<br />

analysis. Pages 13–35 in Alcamo J, ed. Environmental Futures: The<br />

Practice <strong>of</strong> Environmental Scenario Analysis. Elsevier.<br />

Bennett EM, Carpenter SR, Peterson GD, Cumming GS, Zurek M, Pingali P.<br />

2003. Why global scenarios need ecology. Frontiers in Ecology and the<br />

Environment 1: 322–329.<br />

Bestelmeyer S, Dailey S, Elser M, Hembree P, Landis C, O’Connell K,<br />

Simmons B, Sommer S, Steiner S. 2005. Handbook for <strong>LTER</strong> Education.<br />

US Long Term Ecological Research Network.<br />

Bolte JP, Hulse DW, Gregory SV, Smith C. 2007. Modeling biocomplexity:<br />

Actors, landscapes and alternative futures. Environmental Modeling<br />

and S<strong>of</strong>tware 22: 570–579.<br />

Carpenter S[R]. 2008. Seeking adaptive change in Wisconsin’s ecosystems.<br />

Pages 407–421 in Waller DM, Rooney TP, eds. The Vanishing<br />

Present: Wisconsin’s Changing Lands, Waters, and Wildlife. <strong>University</strong><br />

<strong>of</strong> Chicago Press.<br />

Carpenter SR, Folke C. 2006. Ecology for transformation. Trends in Ecology<br />

and Evolution 21: 309–315.<br />

Carpenter SR, Bennett EM, Peterson GD. 2006. Scenarios for ecosystem<br />

services: An overview. Ecology and Society 11: 29.<br />

Cherry JE, Walker S, Fresco N, Trainor S, Tidwell A. 2010. Impacts<br />

<strong>of</strong> Climate Change and Variability on Hydropower in Southeast<br />

Alaska: Planning for a Robust Energy Future. <strong>University</strong> <strong>of</strong> Alaska,<br />

Fairbanks. (31 January 2012; http://ine.uaf.edu/accap/documents/seak_<br />

report_final.pdf )<br />

Cissel JH, Swanson FJ, Weisberg PJ. 1999. Landscape management using<br />

historical fire regimes: Blue River, Oregon. Ecological Applications 9:<br />

1217–1231.<br />

City <strong>of</strong> Phoenix. 2010. Phoenix General Plan Update: Transitioning to<br />

a Sustainable Future. Public Hearing Draft. December 2010. City<br />

<strong>of</strong> Phoenix, Phoenix, Arizona. (8 September 2011; http://phoenix.gov/<br />

citygovernment/planres/cityplan/planphx/index.html)<br />

Collins SL, et al. 2011. An integrated conceptual framework for long-term<br />

social-ecological research. Frontiers in Ecology and the Environment<br />

9: 351–357.<br />

Driscoll CT, Lambert KF, Chapin FS III, Nowak DJ, Spies TA, Swanson FJ,<br />

Kittredge DB Jr, Hart CM. 2012. Science and society: The role <strong>of</strong> longterm<br />

studies in environmental stewardship. BioScience 62: 354–366.<br />

Articles<br />

Duncan SL, Thompson JR. 2006. Forest plans and ad hoc scientist groups in<br />

the 1990s: Coping with the Forest Service viability clause. Forest Policy<br />

and Economics 9: 32–41.<br />

[FEMAT] Forest Ecosystem Management Assessment Team. 1993.<br />

Forest Ecosystem Management: An Ecological, Economic, and Social<br />

Assessment. US Forest Service.<br />

Gibson RB. 2006. Sustainability assessment: Basic components <strong>of</strong> a<br />

practical approach. Impact Assessment and Project Appraisal 24:<br />

170–182.<br />

Henrichs T, Zurek M, Eickout B, Kok K, Raudsepp-Hearne C, Ribeiro T,<br />

van Vuuren D, Volkery A. 2010. Scenario development and analysis for<br />

forward-looking ecosystem assessments. Pages 151–220 in Ash N, et al.,<br />

eds. Ecosystems and Human Well-Being: A Manual for Assessment<br />

Practitioners. Island Press.<br />

[Ecoshare] Ecoshare: Interagency Clearinghouse <strong>of</strong> Ecological Information.<br />

2011. Blue River Landscape Strategy. Ecoshare. (31 January 2012;<br />

http://ecoshare.info/projects/central-cascade-adaptive-managementpartnership/landscape-studies/blue-river-landscape-strategy)<br />

Iverson Nassauer J, Corry RC. 2004. Using normative scenarios in landscape<br />

ecology. Landscape Ecology 19: 343–356.<br />

Jones JA, et al. 2012. Ecosystem processes and human influences regulate<br />

streamflow response to climate change at long-term ecological research<br />

sites. BioScience 62: 390–404.<br />

Kahn H. 1962. Thinking about the unthinkable. Horizon Press.<br />

[<strong>LTER</strong> Network] US Long Term Ecological Research Network. 2011. Strategic<br />

and Implementation Plan: Long-term Ecological Research Network<br />

(<strong>LTER</strong>): Research and Education. <strong>LTER</strong> Network. (31 January 2012;<br />

http://intranet2.lternet.edu/sites/intranet2.lternet.edu/files/documents/<br />

<strong>LTER</strong>%20History/Planning%20Documents/<strong>LTER</strong>_SIP_Dec_05_2010.<br />

pdf )<br />

Mahmoud M, et al. 2009. A formal framework for scenario development in<br />

support <strong>of</strong> environmental decision-making. Environmental Modeling<br />

and S<strong>of</strong>tware 24: 798–808.<br />

[MA] Millennium Ecosystem Assessment. 2005. Ecosystems and Human<br />

Well-being: Scenarios. Island Press.<br />

Murphy K, Huettmann F, Fresco N, Morton J. 2010. Connecting Alaska<br />

Landscapes into the Future. US Fish and Wildlife Service. (8 September<br />

2011; www.snap.uaf.edu/files/SNAP-connectivity-2010-complete.pdf )<br />

Naki enovi N, Swart R. 2000. Emissions Scenarios. Intergovernmental<br />

Panel on Climate Change.<br />

Peterson GD, Cumming GS, Carpenter S. 2003a. Scenario planning: A<br />

tool for conservation in an uncertain world. Conservation Biology 17:<br />

358–366.<br />

Peterson GD, Beard TD Jr, Beisner BE, Bennett EM, Carpenter SR,<br />

Cumming GS, Dent CL, Havlicek TD. 2003b. Assessing future ecosystem<br />

services: A case study <strong>of</strong> the Northern Highlands Lake District,<br />

Wisconsin. Conservation Ecology 7 (3, Art. 1).<br />

Raskin P, Monks F, Riberio T, van Vuuren D, Zurek M. 2005. Global scenarios<br />

in historical perspective. Pages 35–45 in Carpenter S, Pingali P,<br />

Bennett E, Zurek M, eds. Ecosystems and Human Well-being, Vol. 2:<br />

Scenarios. Island Press.<br />

Reid WV, Chen D, Goldfarb L, Hackmann H, Lee YT, Mokhele K, Ostrom<br />

E, Raivio K, Rockström J, Schellnhuber HJ, Whyte A. 2010. Earth system<br />

science for global sustainability: Grand challenges. Science 330:<br />

916–917.<br />

Robertson GP, et al. 2012. Long-term ecological research in a humandominated<br />

world. Bioscience 62: 342–353.<br />

Rockström J, et al. 2009. A safe operating space for humanity. Nature 461:<br />

472–475.<br />

Sala OE, et al. 2000. Global biodiversity scenarios for the year 2100. Science<br />

287: 1770–1774.<br />

Scheffer M, Carpenter S, Foley JA, Folke C, Walker B. 2001. Catastrophic<br />

shifts in ecosystems. Nature 413: 591–596.<br />

Schmitt Olabisi LK, Kapuscinski AR, Johnson KA, Reich PB, Stenquist B,<br />

Draeger KJ. 2010. Scenario visioning and participatory system dynamics<br />

modeling to investigate the future: Lessons from Minnesota 2050.<br />

Sustainability 2: 2686–2706.<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 375


Articles<br />

Shaw A, Sheppard S, Burch S, Flanders D, Wiek A, Carmichael J, Robinson J,<br />

Cohen S. 2009. Making local futures tangible—Synthesizing, downscaling,<br />

and visualizing climate change scenarios for participatory capacity<br />

building. Global Environmental Change 19: 447–463.<br />

Shindler B, Mallon A. 2009. Public acceptance <strong>of</strong> disturbance-based<br />

forest management: A study <strong>of</strong> the Blue River Landscape Strategy in<br />

the Central Cascades adaptive management area. US Department <strong>of</strong><br />

Agriculture, US Forest Service, Pacific Northwest Research Station.<br />

Research Report no. PNW-RP-581.<br />

Spies TA, Duncan SL. 2009. Old Growth in a New World: A Pacific<br />

Northwest Icon Reexamined. Island Press.<br />

Spies TA, Johnson KN, Burnett KM, Ohmann JL, McComb BC, Reeves GC,<br />

Bettinger P, Kline JD, Garber-Yonts B. 2007. Cumulative ecological and<br />

socioeconomic effects <strong>of</strong> forest policies in coastal Oregon. Ecological<br />

Applications 17: 5–17.<br />

Swart RJ, Raskin P, Robinson J. 2004. The problem <strong>of</strong> the future:<br />

Sustainability science and scenario analysis. Global Environmental<br />

Change 14: 137–146.<br />

Thompson J[R], Foster D[R]. 2009. Report to the <strong>LTER</strong> Network Office<br />

on the Scenarios <strong>of</strong> Future Landscape Change Working Group<br />

Meeting April 1–2, 2009. (8 September 2011; http://lno.lternet.edu/files/<br />

lno/Workship%20Report%20-%20Landscape%20Change_Thompson_<br />

Foster.pdf )<br />

Thompson JR, Foster DR, Scheller R, Kittredge D[B Jr]. 2011. The influence<br />

<strong>of</strong> land use and climate change on forest biomass and composition in<br />

Massachusetts, USA. Ecological Applications 21: 2425–2444.<br />

Wack P. 1985. Scenarios: Uncharted waters ahead. Harvard Business Review<br />

63: 73–89.<br />

Weaver PM, Rotmans J. 2006. Integrated sustainability assessment: What<br />

is it, why do it and how? International Journal <strong>of</strong> Innovation and<br />

Sustainable Development 1: 284–303.<br />

Weisberg PJ. 1998. Fire history, fire regimes, and development <strong>of</strong> forest<br />

structure in the central western Oregon Cascades. PhD dissertation.<br />

Oregon State <strong>University</strong>, Corvallis.<br />

Wiek A, Binder CR, Scholz RW. 2006. Functions <strong>of</strong> scenarios in transition<br />

processes. Futures 38: 740–766.<br />

Wiek A, Ness B, Brand FS, Schweizer-Ries P, Farioli F. 2012. From complex<br />

systems analysis to transformational change: A comparative appraisal<br />

<strong>of</strong> sustainability science projects. Sustainability Science. doi:10.1007/<br />

s11625-011-0148-y<br />

Withycombe L, Wiek A. 2011. What if the future does not play out as imagined<br />

in the vision? Advances in sustainability planning and research<br />

through contrasting, non-intervention scenarios. Project Report. School<br />

<strong>of</strong> Sustainability, Arizona State <strong>University</strong>. (http://start.lab.asu.edu/)<br />

Xiang W-N, Clarke KC. 2003. The use <strong>of</strong> scenarios in land-use planning.<br />

Environment and Planning 30: 885–909.<br />

Jonathan R. Thompson (thompsonjr@si.edu) is affiliated with the Smithsonian<br />

Institution’s Conservation Biology Institute, in Front Royal, Virginia. Arnim<br />

Wiek is affiliated with the School <strong>of</strong> Sustainability at Arizona State <strong>University</strong>,<br />

in Tempe. Frederick J. Swanson and Thomas A. Spies are affiliated with<br />

the US Forest Service’s Pacific Northwest Research Station, in Corvallis,<br />

Oregon. Stephen R. Carpenter is affiliated with the Center for Limnology<br />

at the <strong>University</strong> <strong>of</strong> Wisconsin, Madison. Nancy Fresco is affiliated with the<br />

School <strong>of</strong> Natural Resources and Agricultural Science at the <strong>University</strong> <strong>of</strong><br />

Alaska, Fairbanks. Teresa Hollingsworth is affiliated with the Boreal Ecology<br />

Cooperative Research Unit at the <strong>University</strong> <strong>of</strong> Alaska, Fairbanks, and with<br />

the US Forest Service’s Pacific Northwest Research Station, also in Fairbanks.<br />

David R. Foster is affiliated with Harvard <strong>University</strong>’s Harvard Forest, in<br />

Petersham, Massachusetts.<br />

376 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


Past, Present, and Future Roles<br />

<strong>of</strong> Long-Term Experiments in the<br />

<strong>LTER</strong> Network<br />

AlAn K. KnApp, MelindA d. SMith, SArAh e. hobbie, Scott l. collinS, tiMothy J. FAhey,<br />

Gretchen J. A. hAnSen, douGlAS A. lAndiS, KiMberly J. lA pierre, Jerry M. Melillo,<br />

tiMothy r. SeAStedt, GAiuS r. ShAver, And JAcKSon r. WebSter<br />

Articles<br />

The US National Science Foundation–funded Long Term Ecological Research (<strong>LTER</strong>) Network supports a large (around 240) and diverse portfolio<br />

<strong>of</strong> long-term ecological experiments. Collectively, these long-term experiments have (a) provided unique insights into ecological patterns and<br />

processes, although such insight <strong>of</strong>ten became apparent only after many years <strong>of</strong> study; (b) influenced management and policy decisions; and<br />

(c) evolved into research platforms supporting studies and involving investigators who were not part <strong>of</strong> the original design. Furthermore, this<br />

suite <strong>of</strong> long-term experiments addresses, at the site level, all <strong>of</strong> the US National Research Council’s Grand Challenges in Environmental Sciences.<br />

Despite these contributions, we argue that the scale and scope <strong>of</strong> global environmental change requires a more-coordinated multisite approach to<br />

long-term experiments. Ideally, such an approach would include a network <strong>of</strong> spatially extensive multifactor experiments, designed in collaboration<br />

with ecological modelers that would build on and extend the unique context provided by the <strong>LTER</strong> Network.<br />

Keywords: climate change, global change, long-term research, <strong>LTER</strong> Network, multifactor experiments<br />

More than 30 years ago, Odum (1977) described ecology<br />

as a uniquely integrative and synthetic scientific<br />

endeavor, and the diversity <strong>of</strong> research approaches employed<br />

today reflects this perspective (e.g., Rees et al. 2001).<br />

Collectively, ecologists conduct research that includes virtually<br />

every ecosystem type on Earth, with studies that span<br />

a broad range <strong>of</strong> spatial and temporal scales. Moreover,<br />

the most successful research programs generally employ a<br />

mixture <strong>of</strong> complementary approaches, including retrospective<br />

studies, observations, experiments (natural and<br />

manipulative), gradient studies, synthetic analyses, and<br />

modeling. The research portfolios <strong>of</strong> the 26 sites within the<br />

US National Science Foundation (NSF)–funded Long Term<br />

Ecological Research (<strong>LTER</strong>) Network (http://lternet.edu)<br />

exemplify this spatially and temporally extensive and multifaceted<br />

approach to ecology (Hobbie et al. 2003), although<br />

the degree to which the sites allocate time and resources to<br />

particular research approaches varies on the basis <strong>of</strong> the<br />

questions being addressed, site-specific constraints, and the<br />

culture <strong>of</strong> the discipline (e.g., oceanographers study systems<br />

differently than do forest ecologists).<br />

Site-based manipulative experiments have long been a<br />

cornerstone <strong>of</strong> ecological research. Although long-term<br />

experiments (box 1) have a foundational history in ecology<br />

(e.g., the Rothamstad fertilization study begun in 1856),<br />

short-term experiments are much more common for a<br />

variety <strong>of</strong> reasons (Tilman 1989). Short-term experiments<br />

provide information on how a system is regulated at the<br />

time and place <strong>of</strong> the manipulation (i.e., the initial limiting<br />

factors and their interactions). However, in complex systems<br />

with multiple components operating on different time scales<br />

<strong>of</strong> response, the initial trajectories <strong>of</strong> response <strong>of</strong> either the<br />

whole system or <strong>of</strong> individual components (e.g., single species)<br />

will not necessarily indicate either the direction or the<br />

magnitude <strong>of</strong> long-term change. Long-term experiments,<br />

when coupled with measurements <strong>of</strong> key processes, will be<br />

more likely to enable interpretation <strong>of</strong> the entire trajectory<br />

<strong>of</strong> system response, as well as the trajectories <strong>of</strong> change<br />

in system components (i.e., species or pools <strong>of</strong> organic<br />

matter and elements). A key difference between long-term<br />

and short-term experiments is that long-term experiments<br />

provide insight into the causes <strong>of</strong> the changes in the slope<br />

<strong>of</strong> responses, the causes <strong>of</strong> the inflection points, and the<br />

magnitude <strong>of</strong> the long-term change, whereas short-term<br />

experiments are focused only on the initial trajectories.<br />

Thus, long-term experiments can address mechanisms and<br />

temporal dynamics in a complementary fashion, particularly<br />

when the experiments are combined with other research<br />

approaches (e.g., observational and gradient studies). In<br />

doing so, they can help elucidate how historical influences<br />

BioScience 62: 377–389. ISSN 0006-3568, electronic ISSN 1525-3244. © 2012 by American Institute <strong>of</strong> Biological Sciences. All rights reserved. Request<br />

permission to photocopy or reproduce article content at the <strong>University</strong> <strong>of</strong> California Press’s Rights and Permissions Web site at www.ucpressjournals.com/<br />

reprintinfo.asp. doi:10.1525/bio.2012.62.4.9<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 377


Articles<br />

Box 1. Definition <strong>of</strong> long-term ecological experiments.<br />

We define experiments simply as planned or active manipulations<br />

<strong>of</strong> a single factor or <strong>of</strong> several factors (drivers or components<br />

<strong>of</strong> ecological systems) in order to test hypotheses.<br />

Such manipulations may occur at spatial scales that vary from<br />

small plots to whole systems (watersheds or entire lakes). This<br />

definition excludes consideration <strong>of</strong> natural field experiments<br />

such as those involving floods, droughts, hurricanes, wildfires,<br />

and other uncontrolled natural events, unless they can serendipitously<br />

be incorporated into a planned experimental design<br />

(see the text).<br />

More challenging is the definition <strong>of</strong> long term. Ideally,<br />

ecological or life-history criteria (e.g., process and turnover<br />

rates, life span <strong>of</strong> organisms) or the time scales <strong>of</strong> ecological<br />

phenomena would be used to define an experiment as long<br />

term. Operationally, however, the duration <strong>of</strong> experiments,<br />

which typically require financial resources to be initiated and<br />

continued, is to a large extent constrained by funding cycles.<br />

Given the Long Term Ecological Research (<strong>LTER</strong>) Network<br />

focus <strong>of</strong> this review, we view a long-term experiment as one<br />

that is planned to exceed or already exceeds six years in length,<br />

which would span two traditional <strong>LTER</strong> Network funding<br />

cycles and is also a sufficient length to allow both mechanisms<br />

and temporal dynamics to be identified. This duration is also<br />

consistent with the requisite period <strong>of</strong> continuous measurements<br />

for support by the US National Science Foundation’s<br />

Long-Term Research in Environmental Biology program.<br />

that shape the present can be merged with a mechanistic<br />

understanding <strong>of</strong> contemporary processes to enable forecasting<br />

the future <strong>of</strong> ecological systems (Carpenter 2002). In<br />

contributing to this understanding, long-term experiments<br />

are critical for informing and validating mechanistic models<br />

and for linking short-term process studies with long-term<br />

observations to improve models for forecasting future<br />

scenarios.<br />

Long-term experiments have been an integral part <strong>of</strong><br />

the NSF-funded <strong>LTER</strong> program since its inception in 1980<br />

(Callahan 1984). Indeed, the <strong>LTER</strong> program’s early mission<br />

statement specifically called for the creation <strong>of</strong> a legacy <strong>of</strong><br />

“well-designed and well-documented long-term observations,<br />

experiments, and archives <strong>of</strong> samples and specimens”<br />

(Hobbie et al. 2003, p. 22). Although there are certainly<br />

many valuable long-term experiments worldwide (Rees et al.<br />

2001), a large and diverse collection has been developed<br />

within the US <strong>LTER</strong> Network. Now, after more than three<br />

decades <strong>of</strong> research, an assessment <strong>of</strong> the types <strong>of</strong> long-term<br />

experiments within the <strong>LTER</strong> Network and their contributions<br />

is warranted. Below, we review the motivation for and<br />

the role <strong>of</strong> long-term experiments in the <strong>LTER</strong> Network,<br />

we summarize the collective lessons learned from these<br />

long-term experiments, and we envision the future role<br />

for long-term experiments in advancing ecological science<br />

and addressing critical questions facing society as the <strong>LTER</strong><br />

Network continues into its fourth decade.<br />

Early long-term experiments <strong>of</strong> the <strong>LTER</strong> Network<br />

Long-term experiments served as the basis for the establishment<br />

<strong>of</strong> the research programs at several <strong>of</strong> the initial <strong>LTER</strong><br />

sites funded in the 1980s. For example, the AND, CWT, and<br />

HBR <strong>LTER</strong> sites (see table 1 for site abbreviations) began as<br />

US Department <strong>of</strong> Agriculture (USDA) Forest Service experimental<br />

forests. At these sites, the responses <strong>of</strong> hydrology,<br />

energy flow, and nutrient cycling to changes in forest structure<br />

caused by disturbance had been quantified experimentally<br />

using paired watersheds (e.g., Hewlett and Helvey 1970) and<br />

the small-watershed approach (Likens 1985) for a number<br />

<strong>of</strong> years prior to their being included in the <strong>LTER</strong> program<br />

(figure 1). Similarly, the USDA Agricultural Research Service<br />

established long-term livestock grazing experiments at several<br />

<strong>of</strong> their experimental ranges in the early 1900s, two <strong>of</strong> which<br />

became <strong>LTER</strong> program sites (JRN and SGS; see figure 1). The<br />

response <strong>of</strong> arctic tundra vegetation to long-term manipulations<br />

<strong>of</strong> nutrient availability and temperature (Chapin et al.<br />

1995) also provided the foundation for continuing studies <strong>of</strong><br />

global-change effects at the ARC site (figure 1).<br />

Research programs at several other founding <strong>LTER</strong> sites<br />

were built around newly established long-term experiments<br />

designed to test a diverse set <strong>of</strong> hypotheses, with the expectation<br />

<strong>of</strong> a long-term funding commitment. For example, the<br />

role and mechanisms whereby soil resources regulate community<br />

structure have been examined in plot-level manipulations<br />

at the CDR site since its inception (figure 1). The<br />

long-term consequences <strong>of</strong> nitrogen (N) deposition, climate<br />

warming, and hurricanes for nutrient cycling, carbon (C)<br />

dynamics, and forest productivity have been assessed at HFR<br />

(Foster et al. 1997). Moreover, experiments evaluating the<br />

interactions <strong>of</strong> fire, ungulate grazing, and climate variability<br />

in driving patterns and processes in tallgrass prairie formed<br />

the basis for long-term experiments at KNZ (figure 1).<br />

Finally, the core experiment in row-crop agriculture at KBS<br />

involved the long-term imposition <strong>of</strong> a gradient <strong>of</strong> varying<br />

management inputs into midwestern US cropping systems.<br />

In all cases, the expectation that key ecological responses to<br />

these manipulations might not become evident for many<br />

years justified the long-term nature <strong>of</strong> the experimental<br />

design for these <strong>LTER</strong> programs.<br />

Long-term experiments today in the <strong>LTER</strong> Network<br />

In order to provide a more comprehensive overview <strong>of</strong><br />

long-term experiments in the <strong>LTER</strong> Network, we surveyed<br />

each <strong>of</strong> the 26 <strong>LTER</strong> sites (table 1) to quantify the number,<br />

types, and durations <strong>of</strong> long-term experiments supported<br />

by the <strong>LTER</strong> Network. In addition, we convened a two-day<br />

working-group meeting with participants from <strong>of</strong> a subset<br />

<strong>of</strong> <strong>LTER</strong> sites in February 2011. On the basis <strong>of</strong> responses<br />

(from 100% <strong>of</strong> the sites) to the survey questions, a total <strong>of</strong><br />

239 long-term experiments (completed and ongoing) were<br />

identified within the <strong>LTER</strong> Network (table 1). Of these, less<br />

than 10% exceed 30 years in duration, which indicates that<br />

most long-term experiments were initiated during the time<br />

frame <strong>of</strong> the <strong>LTER</strong> Network (figure 2). Diversity among the<br />

378 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


Articles<br />

Table 1. Selected results from a questionnaire survey <strong>of</strong> the 26 sites in the Long-Term Ecological Research (<strong>LTER</strong>) Network,<br />

site acronyms, and site age.<br />

Site name Site acronym<br />

experiments was manifest in the wide array <strong>of</strong> resources<br />

and environmental factors manipulated (e.g., soil nutrients,<br />

water, and temperature were the most commonly manipulated,<br />

carbon dioxide [CO 2 ] more rarely), as well as alterations<br />

in biotic attributes (e.g., species removal and additions,<br />

herbivory) and disturbance regimes (e.g., fire, logging, windstorms,<br />

pathogen outbreaks; figure 2). However, in only 28%<br />

(66 <strong>of</strong> 239) <strong>of</strong> the long-term experiments was more than<br />

one <strong>of</strong> these factors manipulated concurrently (table 1).<br />

As befits the chronic nature <strong>of</strong> many global-change drivers<br />

(Smith et al. 2009), the number <strong>of</strong> sites that employed press<br />

(i.e., continuous alteration) treatments in long-term experiments<br />

was 80%, with 75% also supporting long-term pulse<br />

experiments. Collectively, all <strong>of</strong> the Grand Challenges for the<br />

environment identified by the National Research Council<br />

(NRC 2001) are being addressed across the <strong>LTER</strong> Network,<br />

Year established<br />

in <strong>LTER</strong> Network<br />

Number <strong>of</strong><br />

long-term<br />

experiments<br />

Number <strong>of</strong><br />

multifactor<br />

experiments<br />

Multisite<br />

experiments<br />

(yes [y] or no [n])<br />

Andrews Experimental Forest AND 1980 23 1 y<br />

Arctic ARC 1987 10 7 y<br />

Baltimore Ecosystem Study BES 1997 1 0 n<br />

Bonanza Creek Experimental Forest BNZ 1987 18 3 y<br />

California Current Ecosystem CCE 2004 2 0 n<br />

Cedar Creek Ecosystem Science Reserve CDR 1981 36 9 y<br />

Central Arizona-Phoenix CAP 1997 2 0 n<br />

<strong>Coweeta</strong> Hydrological Laboratory CWT 1980 5 1 y<br />

Florida Coastal Everglades FCE 2000 5 1 n<br />

<strong>Georgia</strong> Coastal GCE 2000 1 0 n<br />

Harvard Forest HFR 1988 18 2 y<br />

Hubbard Brook Experimental Forest HBR 1987 5 1 n<br />

Jornada Basin JRN 1981 15 2 y<br />

Kellogg Biological Station KBS 1987 12 8 y<br />

Konza Prairie Biological Station KNZ 1980 18 8 y<br />

Luquillo Experimental Forest LUQ 1988 4 1 y<br />

McMurdo Dry Valleys MCM 1993 12 3 n<br />

Moorea Coral Reef MCR 2004 1 1 n<br />

Niwot Ridge NWT 1980 8 6 y<br />

North Temperate Lakes NTL 1980 6 2 n<br />

Palmer Station PAL 1991 0 0 n<br />

Plum Island Ecosystems PIE 1998 3 2 y<br />

Santa Barbara Coastal SBC 2000 1 0 y<br />

Sevilleta SEV 1988 12 3 y<br />

Shortgrass Steppe SGS 1981 17 4 y<br />

Virginia Coast Reserve VCR 1987 4 1 y<br />

Note: The principal investigators (PIs) were asked to provide information about long-term ecological experiments (as defined in box 1) at their site.<br />

Among other questions, the PIs were asked to report the number <strong>of</strong> long-term experiments, the number <strong>of</strong> experiments that included more than one<br />

manipulated factor, and if any <strong>of</strong> these long-term experiments were conducted at multiple sites. Additional results from this survey are provided in<br />

figure 2 and in the text.<br />

although long-term experiments related to infectious disease<br />

are notably underrepresented (figure 2). Finally, a majority<br />

<strong>of</strong> the sites (16 <strong>of</strong> 26) participated in multisite or networklevel<br />

experiments (i.e., experiments that spanned multiple<br />

sites), which may or may not have included other <strong>LTER</strong><br />

sites. Notable examples include the Nutrient Network, the<br />

Long-Term Intersite Decomposition Team (LIDET), and the<br />

International Tundra Experiment.<br />

Our assessment <strong>of</strong> long-term experiments during the<br />

working-group meeting revealed that many <strong>of</strong> the earliest<br />

long-term experiments were designed as single-factor<br />

manipulations with the goal <strong>of</strong> gaining an improved understanding<br />

<strong>of</strong> the long-term nature <strong>of</strong> change (primarily at<br />

the process level) in response to either short-term (pulse) or<br />

long-term (press) manipulations. In the early years <strong>of</strong> <strong>LTER</strong>,<br />

assessing recovery from disturbance using pulse experiments<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 379


Articles<br />

Figure 1. Examples <strong>of</strong> the diversity <strong>of</strong> long-term experiments established or already ongoing in the earliest cohorts <strong>of</strong> sites<br />

(more than 20 years old; see table 1) in the Long Term Ecological Research (<strong>LTER</strong>) Network. (a) Small watersheds with<br />

different forest harvest treatments at the HBR site (photograph: Hubbard Brook Research Foundation); (b) snowfence<br />

experiment designed to alter soil moisture and growing-season length at the NWT site (photograph: Mark Williams, Niwot<br />

Ridge <strong>LTER</strong> Program, <strong>University</strong> <strong>of</strong> Colorado); (c) a long-term ungrazed exclosure (left) adjacent to a grazed grassland at the<br />

SGS site (photograph: Alan K. Knapp); (d) warming, nutrient addition, and shade experiments at the ARC site (photograph:<br />

Sarah E. Hobbie); (e) biodiversity manipulations at the CDR site (photograph: David Tilman); (f) watershed-scale fire<br />

experiments at the KNZ site (photograph: Alan K. Knapp). Site abbreviations are defined in table 1.<br />

(i.e., a single perturbation or manipulation event) was the<br />

most common type <strong>of</strong> experiment (Callahan 1984), such as<br />

deforestation studies at HBR and CWT or wind-disturbance<br />

studies at HRF. These studies were motivated by the need<br />

to develop or test theory in order to enable a deeper understanding<br />

<strong>of</strong> the ways in which key drivers altered ecosystem<br />

structure and function. The goal was to complement historical<br />

observations with knowledge and insight that were<br />

unattainable from traditional short-term experiments and<br />

space-for-time studies. Other experiments were initiated<br />

to understand the impacts <strong>of</strong> anthropogenic disturbances<br />

such as logging, grazing, and acid deposition. Although the<br />

initial focus was on disturbances, the more-recent trend is<br />

for long-term experiments to have stronger relevance to<br />

global-change drivers or to be linked explicitly to policy and<br />

major issues in natural-resource management (figure 3).<br />

The working group concluded that, given the number and<br />

diversity <strong>of</strong> long-term experiments in the <strong>LTER</strong> Network,<br />

Callahan’s (1984) vision <strong>of</strong> the <strong>LTER</strong> Network as one that<br />

could address the “ serious contradiction between the time<br />

scales <strong>of</strong> many ecological phenomena and the support to<br />

finance their study” (p. 363) has been fulfilled.<br />

Synergies between the <strong>LTER</strong> Network and long-term<br />

experiments<br />

Although long-term experiments have been supported by<br />

many other programs and have been conducted successfully<br />

at a wide range <strong>of</strong> sites around the globe, long-term<br />

experiments conducted at <strong>LTER</strong> sites have many important<br />

advantages. First, long-term measurements at <strong>LTER</strong> sites<br />

provide essential context for interpreting experimental<br />

responses. Long-term observations can also provide pretreatment<br />

and reference-system data, as well as the basis<br />

for identifying ecological surprises, such as extreme climatic<br />

events (Lindenmayer et al. 2010, Smith 2011). Such<br />

data are critical for interpreting experimental results over<br />

380 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


Figure 2. (a) The number <strong>of</strong> long-term experiments across<br />

the 26 Long-Term Ecological Research (<strong>LTER</strong>) Network<br />

sites according to their duration. Experiments more than<br />

30 years in duration were initiated prior to the existence<br />

<strong>of</strong> the <strong>LTER</strong> Network, whereas those less than 6 years<br />

in duration have been recently initiated with plans for<br />

continuation for more than 6 years. (b) Proportion <strong>of</strong> sites<br />

in the <strong>LTER</strong> Network with long-term experiments in which<br />

different types <strong>of</strong> factors were manipulated. Resource<br />

manipulations are in the left side <strong>of</strong> the panel, and other<br />

types <strong>of</strong> manipulations are shown in the right side (Biotic =<br />

species removals/additions, herbivory, etc.; Disturbance =<br />

fire, hurricane simulation, etc.; Land use = agriculture,<br />

forest clear-cutting, etc.; Other = salinity, soil pH, etc.).<br />

(c) Proportion <strong>of</strong> <strong>LTER</strong> Network sites with long-term<br />

experiments in which the National Research Council’s<br />

Grand Challenges in Environmental Sciences (NRC 2001)<br />

were addressed. These data are from a survey <strong>of</strong> all 26<br />

<strong>LTER</strong> Network–site principal investigators (see table 1).<br />

Articles<br />

a background <strong>of</strong> natural variability—for separating signal<br />

from noise; however, these data sets are usually impossible<br />

to maintain through typical funding programs. At the<br />

HBR site, for example, researchers were able to quantify<br />

the magnitude <strong>of</strong> soil calcium (Ca) depletion caused by<br />

twentieth-century acid deposition only through 40 years<br />

<strong>of</strong> continuous biogeochemical measurements (Likens et al.<br />

1996). These observations, in turn, inspired a long-term<br />

experiment in which soil Ca availability was returned to<br />

preindustrial levels (figure 3), with the long-term trajectory<br />

<strong>of</strong> forest biomass and demography providing the context<br />

for interpreting responses to the experimental treatment.<br />

The dramatic recovery <strong>of</strong> growth, health, and regeneration<br />

<strong>of</strong> sugar maple (Acer saccharum) in response to a moderate<br />

increase in soil Ca provided further evidence that humaninduced<br />

soil Ca depletion has caused widespread decline<br />

in this highly valued forest tree and in forest productivity<br />

overall (Juice et al. 2006). In this case, the deployment <strong>of</strong> a<br />

highly sensitive tracer in the Ca treatment provided additional<br />

insights into the pathways <strong>of</strong> Ca flux through forested<br />

watersheds (Dasch et al. 2006).<br />

In addition to long-term observations revealing patterns<br />

and inspiring experimentation to elucidate mechanisms,<br />

post-experiment monitoring at <strong>LTER</strong> sites can improve<br />

understanding and prompt the design <strong>of</strong> new studies<br />

(Janzen 2009). For example, a decade following the cessation<br />

<strong>of</strong> 10 years <strong>of</strong> experimental N additions at the<br />

CDR site, species richness had recovered, whereas other<br />

aspects <strong>of</strong> community structure had not. This led to a<br />

short-term experiment in which the manipulation <strong>of</strong> soil<br />

N availability, plant litter, and seed dispersal revealed the<br />

mechanisms related to recovery from chronic N fertilization<br />

(Clark and Tilman 2010). Similarly, ecosystem recovery was<br />

monitored for 10 years following 6 years <strong>of</strong> experimental<br />

acidification <strong>of</strong> Little Rock Lake (part <strong>of</strong> the NTL site) in<br />

northern Wisconsin (figure 3). This monitoring revealed<br />

that total zooplankton biomass recovered in one year, but<br />

community composition exhibited sustained differences.<br />

Approximately 40% <strong>of</strong> zooplankton taxa exhibited a lag in<br />

recovery after pH returned to the levels at which a biological<br />

response was first observed (Frost et al. 2006). Continued<br />

monitoring <strong>of</strong> the recovery <strong>of</strong> this system revealed complex<br />

trophic interactions and hysteresis effects—responses that<br />

would not have appeared with only a year or two <strong>of</strong> posttreatment<br />

measurements. Such opportunities arise because<br />

<strong>LTER</strong> Network sites provide consistent access to instrumentation<br />

and infrastructure, stable funding levels, and<br />

well-trained personnel who are available to continue the<br />

measurements. Furthermore, the information-management<br />

capabilities <strong>of</strong> <strong>LTER</strong> Network sites ensure that data will be<br />

collected using consistent protocols and will be archived<br />

and readily available for use by other researchers (Ingersoll<br />

1997). Easily accessible data with well-documented metadata<br />

facilitate cross-site comparisons that are integral for<br />

expanding the results <strong>of</strong> experiments to broader spatial<br />

scales.<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 381


Articles<br />

Figure 3. Examples <strong>of</strong> long-term experiments related to global change in the Long-Term Ecological Research (<strong>LTER</strong>)<br />

Network. (a) Application <strong>of</strong> calcium silicate to an experimental watershed at the HRB site to mechanistically examine<br />

acid-rain impacts (photograph: US Forest Service); (b) the multifactor BioCON experiment at the CDR site with<br />

manipulations <strong>of</strong> carbon dioxide, nitrogen, and biodiversity (photograph: David Tilman); (c) Lake acidification curtain<br />

in a lake at the NTL site to assess whole-system impacts <strong>of</strong> changes in pH (photograph: Carl J. Watras); (d) soil-warming<br />

experiment at the HFR site (plot delineated by the lack <strong>of</strong> snow in winter; photograph: Jerry M. Melillo); (e) automated<br />

trace-gas chambers for assessing the impacts <strong>of</strong> different management practices in field crops at the KBS site (photograph:<br />

Julie E. Doll, Kellogg Biological Station <strong>LTER</strong> Program); (f) litter-exclusion experiment at the CWT site designed to allow<br />

the examination <strong>of</strong> the importance <strong>of</strong> terrestrial energy inputs into streams, which also has relevance for evaluating the<br />

impacts <strong>of</strong> mountaintop mining practices (photograph: Sue L. Eggert). Site abbreviations are defined in table 1.<br />

The value <strong>of</strong> long-term ecological experiments<br />

On the basis <strong>of</strong> the working-group meeting and the broader<br />

site survey, we highlight below the diverse array <strong>of</strong> insights<br />

and contributions provided by long-term ecological experiments<br />

within the <strong>LTER</strong> Network. Our goal is to extend our<br />

review beyond the core scientific benefits <strong>of</strong> this suite <strong>of</strong><br />

studies to include unique opportunities and broader impacts<br />

afforded by these long-term experiments.<br />

Insights into long-term responses. Obviously, but not trivially,<br />

one <strong>of</strong> the greatest scientific benefits <strong>of</strong> long-term site-based<br />

experiments is that they elucidate how ecological systems<br />

respond to experimental treatments over the long term.<br />

Indeed, such experiments have repeatedly demonstrated<br />

that long-term responses to treatments can differ markedly<br />

from short-term responses (Tilman 1989, Debinski and Holt<br />

2000). For example, in the BioCON experiment at the CDR<br />

site (biodiversity, CO 2 , and N are manipulated; see figure 3),<br />

Reich and colleagues (2006) tested the hypothesis that the<br />

availability <strong>of</strong> N would constrain the response <strong>of</strong> productivity<br />

in an N-poor grassland community to elevated CO 2 .<br />

Although this hypothesized interaction between N and CO 2<br />

eventually became apparent, it was not until the fourth year<br />

<strong>of</strong> treatment. If the experiment had been discontinued before<br />

the fourth year (i.e., within the time frame <strong>of</strong> typical funding<br />

cycles), researchers might have concluded that N availability<br />

had no effect on the response <strong>of</strong> these communities to elevated<br />

CO 2 . Similarly, in a manipulation <strong>of</strong> soil temperature<br />

at the HFR site (figure 3), the conclusion regarding the influence<br />

<strong>of</strong> warming on soil respiration would have been very<br />

different had the experiment ended after its first few years<br />

(figure 4). In that experiment, warming strongly increased<br />

382 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


Figure 4. Long-term dynamics <strong>of</strong> soil carbon-dioxide<br />

release from the soil-warming experiment at the Harvard<br />

Forest Long-Term Ecological Research site. The data shown<br />

are the proportional change in carbon lost between soils<br />

in the heated plots (H, heated by soil-warming cables)<br />

and those not warmed (but disturbed similarly to the<br />

heated plots, DC). Note that from 1991 to 1997, there was<br />

a strong effect <strong>of</strong> soil warming, but this effect diminished<br />

substantially as the experiment continued. See Melillo and<br />

colleagues (2002) for further details.<br />

soil respiration (by about 25%) in the first five years <strong>of</strong> the<br />

study (Melillo et al. 2002). However, by the tenth year <strong>of</strong><br />

treatment, the warming stimulation had declined to less than<br />

5% above ambient controls. Conclusions based on the initial<br />

results <strong>of</strong> this experiment would have led to an overestimate<br />

<strong>of</strong> the positive feedback to climate warming resulting from<br />

enhanced soil organic matter decomposition. Elucidation<br />

<strong>of</strong> the mechanisms underlying these responses was made<br />

possible by coupling these single-factor studies with those in<br />

which soil and N were simultaneously manipulated, whereas<br />

additional microbial studies enhanced ecosystem-scale<br />

understanding (Contosta et al. 2011).<br />

Occasionally, experimentally induced shifts in ecosystems<br />

can take years to appear, which can lead to ecological<br />

surprises (Lindenmayer et al. 2010), as exemplified by the<br />

long-term phosphorus (P)–addition experiment conducted<br />

at the Upper Kuparuk River, Alaska (part <strong>of</strong> the ARC site).<br />

In this case, although P fertilization stimulated epilithic<br />

algal production immediately, a major increase in bryophyte<br />

production (with subsequent effects on nutrient cycling and<br />

higher trophic levels) became apparent only after a decade<br />

<strong>of</strong> treatment—a response that was completely unexpected<br />

(figure 5; Slavik et al. 2004). In other instances, the lack <strong>of</strong><br />

response even after a decade or more <strong>of</strong> treatments may<br />

lead to an important new understanding <strong>of</strong> patterns and<br />

processes. For example, inspired by the dramatic response<br />

<strong>of</strong> desert grasslands to an exclusion <strong>of</strong> small mammals in<br />

Portal, Arizona (e.g., Brown and Heske 1990), replicate<br />

Articles<br />

Figure 5. Long-term changes in epilitihic chlorophyll<br />

(top panel, in micrograms [µg] per square centimeter<br />

[cm 2 ]) and bryophyte cover (bottom panel, percentage)<br />

in reference and phosphorus-fertilized reaches <strong>of</strong> the<br />

Kuparuk River at the Arctic Long Term Ecological Research<br />

site. In this long-term experiment, bryophyte cover<br />

was negligible in both reference and fertilized reaches<br />

prior to 1992 but increased dramatically in response to<br />

chronic phosphorus additions afterward, with concurrent<br />

reductions in epilithic chlorophyll. The chlorophyll<br />

data are not shown in 1988 because the samples were<br />

contaminated by green algal filaments. Means are shown,<br />

and the error bars represent the positive standard error.<br />

The data are updated from Slavik and colleagues (2004).<br />

small mammal exclosures were established in a number<br />

<strong>of</strong> grassland and shrubland sites, including the SEV site.<br />

Although no significant differences in vegetation composition<br />

and dynamics have been observed to date at the SEV<br />

site (cf. Báez et al. 2006), major changes have occurred in<br />

other arid-land ecosystems (Meserve et al. 2003). The differences<br />

in responses among sites reflect the degree to which<br />

top-down control <strong>of</strong> community structure can vary among<br />

arid-land ecosystems and can prompt new experiments to<br />

better understand context dependence in the function <strong>of</strong><br />

communities and ecosystems.<br />

Besides elucidating how ecological responses to experimental<br />

treatments can change over time, long-term experiments<br />

can provide unique opportunities to uncover the<br />

mechanisms underlying such dynamics. In some instances,<br />

initial system responses may be dominated by the disturbance<br />

associated with initiating an experiment or because <strong>of</strong><br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 383


Articles<br />

legacies <strong>of</strong> pretreatment conditions. Therefore, it may take<br />

a while before system responses are indicative <strong>of</strong> ecological<br />

processes. For example, patterns <strong>of</strong> nitrate loss from agricultural<br />

systems that vary in management intensity are notably<br />

variable, in part because many studies are initiated before<br />

treatments have fully equilibrated or are conducted over relatively<br />

short periods <strong>of</strong> time. This has led to widely conflicting<br />

reports and little agreement about the best management<br />

practices to decrease nitrate loss. In contrast, the long-term<br />

cropping systems experiment at the KBS site (figure 3) has<br />

enabled comparisons <strong>of</strong> nitrate losses from conventional, notill,<br />

low-input, and organic cropping systems. Each <strong>of</strong> these<br />

treatments had six years to equilibrate before measurement,<br />

and they have been assessed for 11 years, which accounts for<br />

interannual variability (Syswerda et al. 2008). This study has<br />

revealed consistent and marked differences in nitrate losses,<br />

with the low-input and organic systems having about half the<br />

nitrate losses <strong>of</strong> the conventionally managed systems.<br />

Long-term experiments can reveal the importance <strong>of</strong> indirect<br />

effects in ecosystem processes not apparent in the short<br />

term, as well as responses that may change over the long term<br />

(Tilman 1989). At the NTL site, the experimental acidification<br />

<strong>of</strong> a small seepage lake (figure 3) produced numerous<br />

changes driven by indirect interactions related to changes in<br />

the food web rather than by direct consequences <strong>of</strong> lower<br />

pH. For example, the rotifer Keratella taurocephela has displayed<br />

low acid tolerance in laboratory studies but increased<br />

in density throughout the acidification experiment as a result<br />

<strong>of</strong> decreased invertebrate-predator abundance (Gonzalez and<br />

Frost 1994). Keratella taurocephela also underwent morphological<br />

changes in the acidified basin as a result <strong>of</strong> reduced<br />

predation pressure. Responses driven by food-web interactions<br />

were <strong>of</strong>ten the opposite <strong>of</strong> expectations based on laboratory<br />

studies <strong>of</strong> pH tolerance and generally exhibited a time lag;<br />

these unexpected results would not have been observed on time<br />

scales shorter than the response time <strong>of</strong> all trophic levels.<br />

Other opportunities from long-term experiments. Unavoidably,<br />

long-term experiments occur against a backdrop <strong>of</strong> longterm<br />

trends; stochastic ambient conditions, including<br />

climate variability and extremes and infrequent disturbances;<br />

and changes in the abundances <strong>of</strong> predators and<br />

pathogens at scales greater than the experimental units.<br />

Although the inability to control ambient conditions can be<br />

challenging, variability in background conditions can sometimes<br />

also prove fortuitous (Tilman 1989). For example,<br />

long-term experiments can be particularly valuable if they<br />

coincide with climate or weather extremes that provide new<br />

ecological understanding. Of course, the longer an experiment<br />

is conducted, the greater the chances that ambient<br />

conditions will vary in ways that produce insights and even<br />

inspire new experiments. During the Ca-addition studies<br />

at the HBR site (figure 3), an intense ice storm damaged<br />

sugar maple trees, which allowed researchers to document<br />

improved wound repair as one <strong>of</strong> the major responses to<br />

the alleviation <strong>of</strong> Ca deficiency and, presumably, a principal<br />

mechanism underlying increased growth rates (Huggett<br />

et al. 2007). At the CWT site, Yeakley and colleagues<br />

(2003) designed an experiment to investigate the importance<br />

<strong>of</strong> riparian Rhododendron species to nutrient export<br />

to streams. They instrumented treatment and reference<br />

hillslopes, made two years <strong>of</strong> pretreatment measurements,<br />

and then removed all Rhododendron stems from a 10-meter<br />

(m) strip along 30 m <strong>of</strong> stream. Less than two months<br />

later, Hurricane Opal downed most <strong>of</strong> the large trees on<br />

the reference hillslope. Over the course <strong>of</strong> the study, they<br />

found that the hurricane impact on canopy trees had far<br />

greater effects on nutrient export than did the experimental<br />

Rhododendron removal. Finally, a wildfire at the KNZ site<br />

in 1996 reset most <strong>of</strong> the long-term fire treatments in the<br />

60 watersheds at the site (figure 1) by burning them all on<br />

the same date. This afforded the opportunity to sample<br />

a large number <strong>of</strong> sites affected by the same fire but with<br />

a wide array <strong>of</strong> longer-term fire histories. This sampling<br />

revealed the importance <strong>of</strong> the amount <strong>of</strong> time since a prior<br />

fire in determining aboveground net primary productivity<br />

responses to fire (figure 6) and helped researchers interpret<br />

results from other experiments regarding the role <strong>of</strong> soil N<br />

in postfire responses (Blair 1997, Knapp et al. 1998).<br />

Scientists conducting long-term experiments can also<br />

take advantage <strong>of</strong> dramatic changes in community structure,<br />

such as those resulting from extirpation or biological invasion.<br />

For example, at the KBS site, long-term monitoring<br />

<strong>of</strong> predaceous lady beetles (Coccinellidae) in experimental<br />

agricultural treatments has revealed three exotic species<br />

additions since 1988, including the multicolored Asian lady<br />

beetle (Harmonia axyridis). The arrival <strong>of</strong> the soybean aphid<br />

(Aphis glycines) in 2000 reunited these two Asian species in<br />

a new context and resulted in surprising dynamics. Prior to<br />

2000, H. axyridis was a common species; however, after 2000,<br />

it became dominant (figure 7). Moreover, it demonstrated<br />

classic predator–prey cycling with high abundances following<br />

years <strong>of</strong> aphid outbreak (Heimpel et al. 2010); process-level<br />

studies demonstrated strong top-down control <strong>of</strong> A. glycines<br />

by coccinellids (Costamagna and Landis 2006). A further<br />

surprise was that biological control <strong>of</strong> the soybean aphid<br />

was regulated by the structure <strong>of</strong> the surrounding landscape:<br />

Suppression was positively correlated with landscape<br />

diversity at the 1.5-kilometer scale (Gardiner et al. 2009).<br />

Finally, long-term experiments can provide opportunities<br />

to address novel questions that were not part <strong>of</strong><br />

the original motivation for the experiment. Long-term<br />

N-enrichment studies at the ARC site were initially established<br />

as part <strong>of</strong> a broader suite <strong>of</strong> treatments designed<br />

to assess resource limitation and the response <strong>of</strong> tundra<br />

ecosystems to global-change factors (figure 1; Chapin et al.<br />

1995). After 20 years <strong>of</strong> adding N, the researchers shifted<br />

their focus toward asking how much <strong>of</strong> the cumulative N<br />

added still remained in the plots and how this affected C<br />

pools (Mack et al. 2004). Somewhat surprisingly, these plots<br />

had lost significant C, despite greater C inputs (net primary<br />

production) and aboveground C stocks with N addition. Net<br />

384 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


Figure 6. Response in aboveground net primary production<br />

(ANPP, in grams [g] per square meter [m 2 ]) to a wildfire at<br />

the Konza Prairie (KNZ) Long-Term Ecological Research<br />

site. The KNZ site includes a number <strong>of</strong> watersheds that are<br />

experimentally burned at different frequencies (from annual<br />

fire to no fire). A wildfire burned most <strong>of</strong> the site in 1991<br />

(the no-fire data are from areas that escaped the wildfire),<br />

which allowed the researchers to gain insight into how<br />

long-term fire history (including the time since the last fire)<br />

influenced ANPP responses. The mean for each time interval<br />

is displayed, and the error bars represent the standard error.<br />

Source: Adapted from Knapp and colleagues (1988).<br />

ecosystem C losses resulted from soil C losses in deeper<br />

horizons, presumably from enhanced decomposition caused<br />

by N addition.<br />

Broader contributions <strong>of</strong> long-term experiments. Beyond the<br />

basic ecological understanding that results from individual<br />

experiments, long-term experiments can be valuable by contributing<br />

to synthetic activities, testing and inspiring ecological<br />

theory and models across systems, and informing policy<br />

and management. With respect to synthesis, combining the<br />

results <strong>of</strong> multiple studies in meta-analyses or data syntheses<br />

can elucidate how factors such as climate, species composition,<br />

and edaphic conditions interact with treatments to<br />

influence ecological responses, which adds value to the original,<br />

individual experiments. For example, although numerous<br />

studies have demonstrated declines in plant species richness<br />

in response to N addition, synthesis <strong>of</strong> such results across a<br />

number <strong>of</strong> <strong>LTER</strong> sites revealed that the relative magnitude<br />

<strong>of</strong> species loss varied greatly across sites; warmer sites and<br />

those with high cation-exchange-capacity (CEC) soils exhibited<br />

lower declines in richness than sites that were colder<br />

or had low CEC (Clark et al. 2007). Sites with high abundance<br />

<strong>of</strong> C 4 grasses had a greater productivity response to N<br />

addition than did other sites, which was in turn associated<br />

with greater proportional declines in species richness.<br />

Long-term experiments conducted at multiple sites,<br />

although less common, are particularly powerful ways to<br />

Articles<br />

achieve general, synthetic understanding <strong>of</strong> ecological processes.<br />

Such studies can identify important influences on<br />

ecological processes at large spatial scales. They are more<br />

powerful than ad hoc meta-analytical syntheses, because they<br />

can eliminate variation in experimental methodologies across<br />

sites. For example, the LIDET studied long-term decomposition<br />

(10 years) by deploying common substrates across 27<br />

sites, including 16 <strong>LTER</strong> Network sites (Harmon et al. 2009).<br />

The LIDET experiment yielded new insights into decomposition<br />

processes including a new understanding <strong>of</strong> the rates<br />

and controls <strong>of</strong> decomposition <strong>of</strong> more slowly decomposing<br />

litter fractions. Such insights would have been difficult to<br />

derive from the meta-analysis <strong>of</strong> published decomposition<br />

studies because most <strong>of</strong> those studies last one year at most<br />

and because substrate variability confounds site-to-site variability<br />

in factors such as climate (Adair et al. 2010).<br />

Long-term experiments also take on added value when<br />

they inform and are informed by ecological theory and<br />

models. Tilman’s (1982) study <strong>of</strong> resource competition<br />

that developed the resource-ratio hypothesis <strong>of</strong> competitive<br />

interactions was enriched by the close interplay between<br />

long-term experiments and the development <strong>of</strong> theory at the<br />

CDR site. Similarly, empirical research on resource limitation<br />

at the ARC site has occurred in close connection with<br />

the development <strong>of</strong> theory on multiple-resource limitation<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 385<br />

Mean abundance <strong>of</strong> Harmonia axyridis<br />

(number per trap per week)<br />

Figure 7. Abundance <strong>of</strong> adult multicolored Asian lady<br />

beetles (Harmonia axyridis) at the Kellogg Biological<br />

Station Long Term Ecological Research site from 1989 to<br />

2010. This invasive species was first detected in 1994 and,<br />

until 1999, its mean abundance was approximately 0.2<br />

adults per trap per week (a). Following the arrival <strong>of</strong> the<br />

soybean aphid (Aphis glycines), major aphid outbreaks<br />

(arrows) occurred every other year from 2001 to 2005,<br />

prompting strong numerical responses by Harmonia in the<br />

subsequent years and overall greater mean abundance (b).<br />

Since 2007, no aphid outbreaks have occurred, and a<br />

new pattern <strong>of</strong> intermediate Harmonia abundance may<br />

be forming (c). The data are updated from Heimpel and<br />

colleagues (2010).


Articles<br />

(Herbert et al. 1999). Multiple-resource-limitation theory, in<br />

turn, has influenced the development <strong>of</strong> Earth-system models<br />

that explore the implications <strong>of</strong> N constraints on C cycling at<br />

a global scale (e.g., Gerber et al. 2010). Given the number and<br />

diversity <strong>of</strong> long-term experiments in the <strong>LTER</strong> Network, the<br />

opportunity certainly exists for additional theory and model<br />

development capable <strong>of</strong> linking a broad array <strong>of</strong> patterns and<br />

processes across temporal and spatial scales.<br />

There are broader practical outcomes <strong>of</strong> many long-term<br />

experiments when they influence the design <strong>of</strong> policy and<br />

management strategies. For example, the litter-exclusion<br />

experiment at the CWT site (figure 3) clearly demonstrated<br />

the importance <strong>of</strong> tree-leaf litter to the organisms living in<br />

headwater streams (Wallace et al. 1997). The importance<br />

<strong>of</strong> maintaining the integrity <strong>of</strong> headwater streams has been<br />

used as an argument against mountaintop mining practices<br />

in the central Appalachian region (Meyer and Wallace 2001)<br />

and cited in court cases (e.g., P. C. Chambers, US District<br />

Judge, Memorandum Opinion and Order on Civil Action<br />

No. 3:05-0784 in the Huntington Division <strong>of</strong> the US District<br />

Court <strong>of</strong> the Southern District <strong>of</strong> West Virginia, 23 March<br />

2007). Experimental demonstration <strong>of</strong> the dramatic impacts<br />

<strong>of</strong> deforestation on soil and surface-water chemistry at the<br />

HBR site (Bormann et al. 1974) raised awareness <strong>of</strong> the<br />

consequences <strong>of</strong> intensive forest harvest for environmental<br />

quality and influenced the development <strong>of</strong> best management<br />

practices for these forests. Simulations <strong>of</strong> hurricanes and<br />

mortality resulting from pests and pathogens demonstrated<br />

that the ecosystem consequences <strong>of</strong> salvage and restoration<br />

management <strong>of</strong> forests—both post-windstorm and in<br />

advance <strong>of</strong> insect infestation or disease—are <strong>of</strong>ten much<br />

greater than the impacts <strong>of</strong> the disturbances themselves and<br />

led to the argument that leaving forests alone is <strong>of</strong>ten a viable<br />

management alternative from an ecological perspective<br />

(Foster and Orwig 2006). Finally, the results from long-term<br />

experiments, coupled with gradient studies, field observations,<br />

and data from monitoring networks, have been used<br />

to estimate the critical loads <strong>of</strong> N for freshwater and terrestrial<br />

ecosystems <strong>of</strong> the United States (Pardo et al. 2011).<br />

Long-term experiments can provide a unique opportunity<br />

to directly evaluate the consequences <strong>of</strong> policy changes<br />

using an adaptive-management approach. In adaptivemanagement<br />

experiments, the researcher attempts to learn<br />

about managed systems by experimentally changing policy<br />

and assessing the outcomes on the appropriate time scale.<br />

At the NTL site, several policies were enacted as part <strong>of</strong> a<br />

long-term experiment designed to reduce P inputs and to<br />

control algal blooms in Lake Mendota, Wisconsin. Initial<br />

changes to land-use practices revealed little effect <strong>of</strong> management<br />

on the P inputs to the lake, possibly because <strong>of</strong> resuspension<br />

<strong>of</strong> P stored in sediments. Subsequent manipulation<br />

<strong>of</strong> the Lake Mendota food web to create a trophic cascade<br />

was more effective in reducing algal blooms, especially when<br />

they were accompanied by fishing-regulation changes and<br />

additional land-use practices implemented in the 1990s and<br />

2000s (Carpenter et al. 2006).<br />

Future policy can also be informed by results from<br />

long-term experiments. Work at the KBS site suggests<br />

that agricultural nitrous oxide (N 2 O) emissions (figure 3)<br />

increase exponentially with increasing rates <strong>of</strong> N fertilizer<br />

use, after the rates exceed the N-uptake capacity <strong>of</strong> the crop.<br />

In intensive agricultural systems, application <strong>of</strong> N in excess<br />

<strong>of</strong> plant needs is rather common, because fertilizer is inexpensive<br />

relative to commodity prices, and producers tend to<br />

hedge against the risk <strong>of</strong> insufficient N in order to achieve<br />

maximum yields. Using the concept <strong>of</strong> maximum return<br />

to N, Millar and colleagues (2010) developed a transparent<br />

N 2 O-reduction protocol suitable for incentivizing N 2 O<br />

reductions without affecting crop yields. They estimated<br />

that if the protocol were widely adopted as a part <strong>of</strong> future C<br />

cap-and-trade markets, the protocol could reduce N 2 O from<br />

fertilized row-crop agriculture by more than 50%.<br />

Long-term experiments as platforms for new and unplanned<br />

research. Because <strong>of</strong> their multiyear nature, long-term<br />

experiments can serve as platforms for research that goes<br />

well beyond the goals originally envisioned. As a result, they<br />

provide opportunities for novel studies that take advantage<br />

<strong>of</strong> imposed treatments, as well as for more detailed<br />

process-level studies to uncover mechanisms behind longterm<br />

patterns. For example, the soil-warming studies at the<br />

HFR site have become a platform for new studies <strong>of</strong> soil<br />

N; the consequences <strong>of</strong> garlic mustard invasion, an exotic<br />

species that inhibits mycorrhizae and the growth <strong>of</strong> some<br />

tree seedlings; and for microbial studies <strong>of</strong> the mechanisms<br />

responsible for the pattern <strong>of</strong> increased heterotrophic soil<br />

respiration followed by a diminishing response to warming<br />

(Bradford et al. 2008). This latter research showed that the<br />

apparent acclimation <strong>of</strong> soil respiration at the ecosystem<br />

scale results from the combined effects <strong>of</strong> reductions in soil<br />

C pools and microbial biomass and changes in the thermal<br />

response <strong>of</strong> microbial respiration. Mass-specific respiration<br />

rates were lower when seasonal temperatures were higher,<br />

which suggests that rate reductions under experimental<br />

warming probably occurred through temperature-induced<br />

changes in the microbial community.<br />

The long-term nutrient-addition studies that were conducted<br />

at multiple <strong>LTER</strong> sites to understand the role <strong>of</strong> N<br />

limitation, as well as to test ecosystem response to enhanced N<br />

deposition, have similarly provided a rich template for more<br />

short-term mechanistic studies. At the NWT site, for example,<br />

these studies have identified the importance <strong>of</strong> plant species<br />

traits in affecting community change (Suding et al. 2006)<br />

and demonstrated how plant–microbial feedbacks influence<br />

species coexistence (Ashton et al. 2008). Experiments initially<br />

designed to study vegetation change have been used to provide<br />

important insights as to how increased N availability can<br />

affect soil C storage (Neff et al. 2002).<br />

Long-term experiments are increasingly serving as platforms<br />

for ecological metagenomics studies. At the SEV site,<br />

for example, microbial ecologists are examining the metagenomic<br />

responses <strong>of</strong> rhizosphere microbes in a fully crossed<br />

386 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


multifactor experiment that includes increased winter rainfall,<br />

N amendment, and nighttime warming (Collins et al.<br />

2010). At the CDR site, metagenomic studies have shown<br />

a marked divergence <strong>of</strong> microbial communities in grassland<br />

communities developed under ambient as opposed to<br />

elevated CO 2 in terms <strong>of</strong> both composition and function,<br />

with communities under elevated CO 2 exhibiting increased<br />

abundance <strong>of</strong> the genes involved in labile C decomposition<br />

and C and N fixation (He et al. 2010).<br />

A call for a new generation <strong>of</strong> long-term experiments<br />

We have argued that well-designed long-term experiments<br />

focused on key questions and important processes can have<br />

tremendous value for individual sites. These long-term<br />

experiments can provide the understanding necessary for<br />

forecasting, coping with, and mitigating the consequences <strong>of</strong><br />

human-driven global changes to the environment (Collins<br />

et al. 2011). However, the scope and pace <strong>of</strong> change occurring<br />

in ecological systems today—and forecast for the future—<br />

are, by all accounts, unprecedented in human history (Palmer<br />

et al. 2004, Solomon et al. 2007, Smith et al. 2009). Because<br />

<strong>of</strong> the global scale <strong>of</strong> altered biogeochemical cycles from<br />

increased atmospheric CO 2 concentrations, nutrient enrichment<br />

and depletion, climatic change (means and extremes;<br />

Smith 2011), exotic species introductions, and land-use<br />

change, all ecosystems are, and will continue to be, affected<br />

by these alterations (Solomon et al. 2007). Independently<br />

conducted site-based studies in which global-change factors<br />

(i.e., N or temperature) were manipulated and modeling<br />

studies both show that responses to these global-change factors<br />

may vary dramatically among ecosystems—from little or<br />

no response to a substantial change in function (e.g., Weltzin<br />

et al. 2003). But experimental ecologists need to think beyond<br />

the site level and provide a more complete understanding <strong>of</strong><br />

how and why ecosystems differ in their susceptibility and<br />

sensitivity to these changes. Such understanding is critical for<br />

forecasting the ecological consequences <strong>of</strong> global change at<br />

regional to continental spatial scales and yet it is a challenge<br />

that ecologists have not fully met (Smith et al. 2009).<br />

This lack <strong>of</strong> comprehensive understanding occurs, in part,<br />

because ecologists have historically conducted disparate,<br />

largely independent experiments that tend to differ markedly<br />

with regard to (a) what is manipulated, (b) how much<br />

and for how long manipulations occur, and (c) what (and<br />

how) response variables are measured. This is certainly true<br />

<strong>of</strong> long-term experiments conducted in the <strong>LTER</strong> Network.<br />

As a result, syntheses across these studies can be difficult,<br />

since there is no way <strong>of</strong> knowing how much these different<br />

approaches contribute to the range <strong>of</strong> responses observed<br />

among ecosystems (Knapp et al. 2004). As Callahan (1984)<br />

pointed out more than 25 years ago, “even similar projects<br />

are not <strong>of</strong>ten comparable unless effort and resources have<br />

been devoted to making them so. This inherent tendency<br />

away from comparability becomes more prominent among<br />

projects conducted at locations that are geographically and<br />

biologically disjunct” (p. 363). Recently, renewed calls have<br />

Articles<br />

been made for establishing unified sampling protocols<br />

and conducting simultaneous multisite, multifactor experiments<br />

to address the most pressing global-change issues<br />

(Janzen 2009, Luo et al. 2011; e.g., The Nutrient Network,<br />

http://nutnet.science.oregonstate.edu). We echo and extend<br />

these calls by proposing that these multisite experiments<br />

should be planned as long-term experiments capable <strong>of</strong><br />

elucidating both ecological dynamics and ecological mechanisms.<br />

As is appropriate, these should take advantage <strong>of</strong> the<br />

long-term observations and contextual understanding extant<br />

within the <strong>LTER</strong> Network as well as from nonnetwork sites<br />

and other existing and emerging observatories (Carpenter<br />

2008, Robertson et al. 2012 [in this issue]).<br />

Designing multisite experiments will not be without challenges.<br />

For example, how one scales treatment levels across<br />

disparate ecosystems can influence how these different ecosystems<br />

respond to “common treatments.” For experiments<br />

that alter CO 2 , treatments can be designed to increase CO 2<br />

by either a constant amount or a constant proportion with<br />

an expectation <strong>of</strong> comparable results because essentially all<br />

terrestrial ecosystems have very similar atmospheric CO 2<br />

concentrations. But for resources such as water and N, initial<br />

stocks and turnover rates can vary by orders <strong>of</strong> magnitude<br />

among mesic or xeric and fertile or infertile ecosystems.<br />

Therefore, experiments designed to increase or decrease<br />

resources by a constant proportion <strong>of</strong> availability and those<br />

designed to change them by a constant amount will likely<br />

lead to very different responses among sites. As was noted<br />

above, long-term data from observations and monitoring<br />

at <strong>LTER</strong> sites and the rich array <strong>of</strong> shorter-term processlevel<br />

studies that each site has conducted can be invaluable<br />

for providing the appropriate context for planning such<br />

experiments, as well as for interpreting their results. Longterm<br />

data can also provide the breadth <strong>of</strong> understanding<br />

necessary for devising more-detailed studies that are <strong>of</strong>ten<br />

necessary for elucidating key mechanisms. Other issues that<br />

must be addressed, and that <strong>LTER</strong> scientists grapple with<br />

today, include determining the end point for long-term<br />

experiments (e.g., when do the costs outweigh the new<br />

knowledge gained by continuing an experiment?) and how<br />

to optimally balance the trade-<strong>of</strong>f between the spatial scale<br />

<strong>of</strong> manipulations and the number <strong>of</strong> replicates. Small-plotbased<br />

long-term experiments, such as those at the CDR site,<br />

can be designed to be statistically robust with many replicates.<br />

This is not feasible with large-scale whole-watershed<br />

manipulations, such as those at the HBR site (figure 1). The<br />

scale <strong>of</strong> manipulations also determines the degree to which<br />

long-term experiments can serve as platforms for additional<br />

research. These are important trade<strong>of</strong>fs in design that need<br />

to be considered for future long-term experiments.<br />

We further advocate that such multisite long-term experiments<br />

should be designed in close collaboration with ecological<br />

modelers in order to alleviate the more vexing<br />

uncertainties in today’s models (Luo et al. 2011). Scenario<br />

planning (Peterson et al. 2003) can also be a valuable tool<br />

for designing experiments capable <strong>of</strong> providing information<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 387


Articles<br />

relevant to stakeholders in addition to ecologists. These longterm<br />

experiments should be as large in scale as is feasible in<br />

order to permit the evaluation <strong>of</strong> ecological processes not possible<br />

in small-plot experiments (Smith et al. 2009). Designing<br />

large-scale experiments will present many unique challenges,<br />

since they are not simply big versions <strong>of</strong> the small-scale<br />

manipulations that ecologists have conducted in the past.<br />

Innovation in experimental design, statistical analysis, and<br />

engineering will be necessary. But as was noted above, with<br />

space set aside to accommodate unanticipated use, these networks<br />

<strong>of</strong> large-scale experiments would also serve as valuable<br />

research platforms for the broader ecological community. A<br />

highly coordinated and spatially extensive network <strong>of</strong> multifactor<br />

long-term experiments that adopts such a comparative<br />

approach would provide understanding and quantitative<br />

response data on ecological change at a temporal and spatial<br />

scale heret<strong>of</strong>ore unavailable to ecologists. Such a network<br />

<strong>of</strong> experiments would complement the monitoring-based<br />

approach that the emerging National Ecological Observatory<br />

Network (NEON) has adopted by providing experimentally<br />

defined units <strong>of</strong> accelerated or amplified ecological change.<br />

Data from multisite, long-term experiments can be fused<br />

with models (Luo et al. 2011) to make more-robust forecasts<br />

for a broad range <strong>of</strong> ecosystems. Such forecasts can then be<br />

tested against the real-time tracking <strong>of</strong> ecological change<br />

provided by NEON and other observatory networks—and,<br />

<strong>of</strong> course, by the <strong>LTER</strong> Network.<br />

For long-term experimental networks as they are envisioned<br />

above to be realized, it is clear that clever designs,<br />

collaborations between ecologists and sensor and infrastructure<br />

engineers (Collins et al. 2006), and efficient<br />

deployment will be necessary from both scientific and economic<br />

perspectives. But as society demands the knowledge<br />

needed to cope with and mitigate global changes, ecologists<br />

would be remiss in forgoing the opportunity to build on the<br />

legacy <strong>of</strong> long-term experiments reviewed above to meet<br />

these challenges.<br />

Acknowledgments<br />

We thank the principal investigators <strong>of</strong> the 26 Long Term<br />

Ecological Research Network sites for providing information<br />

regarding long-term experiments at their sites, Steve<br />

Carpenter and Tim Kratz for insight regarding North<br />

Temperate Lake–site experiments, and David Foster for providing<br />

editorial input and oversight. Robert Holt, Charles<br />

Driscoll, and an anonymous reviewer provided helpful comments<br />

on an earlier version <strong>of</strong> the manuscript.<br />

References cited<br />

Adair EC, Hobbie SE, Hobbie RK. 2010. Single-pool exponential decomposition<br />

models: Potential pitfalls in their use in ecological studies.<br />

Ecology 91: 1225–1236.<br />

Ashton IW, Miller AE, Bowman WD, Suding KN. 2006. Nitrogen preferences<br />

and plant-soil feedbacks as influenced by neighbors in the alpine<br />

tundra. Oecologia 156: 625–636.<br />

Báez S, Collins SL, Lightfoot D, Koontz TL. 2006. Bottom-up regulation<br />

<strong>of</strong> plant community structure in an aridland ecosystem. Ecology 87:<br />

2746–2754.<br />

Blair JM. 1997. Fire, N availability and plant response in grasslands: A test <strong>of</strong><br />

the transient maxima hypothesis. Ecology 78: 2359–2368.<br />

Bormann FH, Likens GE, Siccama TG, Pierce RS, Eaton JS. 1974. The export<br />

<strong>of</strong> nutrients and recovery <strong>of</strong> stable conditions following deforestation<br />

at Hubbard Brook. Ecological Monographs 44: 255–277.<br />

Bradford MA, Davies CA, Frey SD, Maddox TR, Melillo JM, Mohan JE,<br />

Reynolds JF, Treseder KK, Wallenstein MD. 2008. Thermal adaptation<br />

<strong>of</strong> soil microbial respiration to elevated temperature. Ecology Letters<br />

11: 1316–1327.<br />

Brown JH, Heske EJ. 1990. Control <strong>of</strong> a desert–grassland transition by a<br />

keystone rodent guild. Science 250: 1705–1707.<br />

Callahan JT. 1984. Long-term ecological research. BioScience 34:<br />

363–367.<br />

Carpenter SR. 2002. Ecological futures: Building an ecology <strong>of</strong> the long now.<br />

Ecology 83: 2069–2083.<br />

———. 2008. Emergence <strong>of</strong> ecological networks. Frontiers in Ecology 6: 228.<br />

Carpenter SR, Lathrop RC, Nowak P, Bennett EM, Reed T, Soranno PA. 2006.<br />

The ongoing experiment: Restoration <strong>of</strong> Lake Mendota and its watershed.<br />

Pages 236–256 in Magnuson JJ, Kratz TK, Benson BJ, eds. Long-<br />

Term Dynamics <strong>of</strong> Lakes in the Landscape. Oxford <strong>University</strong> Press.<br />

Chapin FS III, Shaver GR, Giblin AE, Nadelh<strong>of</strong>fer KJ, Laundre JA. 1995.<br />

Responses <strong>of</strong> arctic tundra to experimental and observed changes in<br />

climate. Ecology 76: 694–711.<br />

Clark CM, Cleland EE, Collins SL, Fargione JE, Gough L, Gross KL,<br />

Pennings SC, Suding KN, Grace JB. 2007. Environmental and plant<br />

community determinants <strong>of</strong> species loss following nitrogen enrichment.<br />

Ecology Letters 10: 596–607.<br />

Collins SL, et al. 2006. New opportunities in ecological sensing using wireless<br />

sensor networks. Frontiers in Ecology and the Environment 4:<br />

402–407.<br />

Collins SL, Fargione JE, Crenshaw CL, Nonaka E, Elliott JR, Xia Y, Pockman<br />

WT. 2010. Rapid plant community responses during the summer<br />

monsoon to nighttime warming in a northern Chihauhuan Desert<br />

grassland. Journal <strong>of</strong> Arid Environments 74: 611–617.<br />

Collins SL, et al. 2011. An integrated conceptual framework for long-term<br />

social–ecological research. Frontiers in Ecology and the Environment<br />

9: 351–357.<br />

Contosta AR, Frey SD, Cooper AB. 2011. Seasonal dynamics <strong>of</strong> soil respiration<br />

and N mineralization in chronically warmed and fertilized soils.<br />

Ecosphere 2 (3, Art. 36).<br />

Costamagna AC, Landis DA. 2006. Predators exert top-down control <strong>of</strong><br />

soybean aphid across a gradient <strong>of</strong> agricultural management systems.<br />

Ecological Applications 16: 1619–1628.<br />

Dasch AA, Blum JD, Eagar C, Fahey TJ, Driscoll CT, Siccama TG. 2006. The<br />

relative uptake <strong>of</strong> Ca and Sr into tree foliage using a whole-watershed<br />

calcium addition. Biogeochemistry 80: 21–41.<br />

Debinski DM, Holt RD. 2000. A survey and overview <strong>of</strong> habitat fragmentation<br />

experiments. Conservation Biology 14: 342–355.<br />

Franklin JF, Bledsoe CS, Callahan JT. 1990. Contributions <strong>of</strong> the long-term<br />

ecological research program. BioScience 40: 509–523.<br />

Frost TM, Fischer JM, Klug JL, Arnott SE, Montz PK. 2006. Trajectories <strong>of</strong><br />

zooplankton recovery in the Little Rock Lake whole-lake acidification<br />

experiment. Ecological Applications 16: 353–367.<br />

Foster DR, Orwig DA. 2006. Preemptive and salvage harvesting <strong>of</strong> New<br />

England forests: When doing nothing is a viable alternative. Conservation<br />

Biology 20: 959–970.<br />

Foster DR, Aber JD, Melillo JM, Bowden RD, Bazzaz FA. 1997. Forest<br />

response to disturbance and anthropogenic stress: Rethinking the 1938<br />

Hurricane and the impact <strong>of</strong> physical disturbance vs. chemical and<br />

climate stress on forest ecosystems. BioScience 47: 437–445.<br />

Gardiner MM, Landis DA, Gratton C, DiFonzo CD, O’Neal M, Chacon<br />

JM, Wayo MT, Schmidt NP, Mueller EE, Heimpel GE. 2009. Landscape<br />

diversity enhances the biological control <strong>of</strong> an introduced crop pest in<br />

the north-central U.S. Ecological Applications 19: 143–154.<br />

Gerber S, Hedin LO, Oppenheimer M, Pacala SW, Shevliakova E. 2010.<br />

Nitrogen cycling and feedbacks in a global dynamic land model. Global<br />

Biogeochemical Cycles 24 (Art. GB1001). doi:10.1029/2008GB003336<br />

388 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


Gonzalez MJ, Frost TM. 1994. Comparisons <strong>of</strong> laboratory bioassays and a<br />

whole-lake experiment: Rotifer responses to experimental acidification.<br />

Ecological Applications 4: 69–80.<br />

Harmon ME, Silver WL, Fasth B, Chen H, Burke IC, Parton WJ, Hart SC,<br />

Currie WS, LIDET. 2009. Long-term patterns <strong>of</strong> mass loss during the<br />

decomposition <strong>of</strong> leaf and fine root litter: An intersite comparison.<br />

Global Change Biology 15: 1320–1338.<br />

He Z, Xu M, Deng Y, Kang S, Kellogg L, Wu L, Van Nostrand JD, Hobbie<br />

SE, Reich PB, Zhou J. 2010. Metagenomic analysis reveals a marked<br />

divergence in the structure <strong>of</strong> belowground microbial communities at<br />

elevated CO 2 . Ecology Letters 13: 564–575.<br />

Heimpel GE, Frelich LE, Landis DA, Hopper KR, Hoelmer K, Sezen Z,<br />

Asplen MK, Wu K. 2010. European buckthorn and Asian soybean<br />

aphid as part <strong>of</strong> an extensive invasional meltdown in North America.<br />

Biological Invasions 12: 2913–2931.<br />

Herbert DA, Rastetter EB, Shaver GR, Ågren GI. 1999. Effects <strong>of</strong> plant<br />

growth characteristics on biogeochemistry and community composition<br />

in a changing climate. Ecosystems 2: 367–382.<br />

Hobbie JE, Carpenter SR, Grimm NB, Gosz JR, Seastedt TR. 2003. The US<br />

long-term ecological research program. BioScience 53: 21–32.<br />

Huggett BA, Schaberg PG, Hawley GJ, Eagar C. 2007. Long-term calcium<br />

addition increases growth release, wound closure, and health <strong>of</strong> sugar<br />

maple (Acer saccharum) trees at the Hubbard Brook Experimental<br />

Forest. Canadian Journal <strong>of</strong> Forest Research 37: 1692–1700.<br />

Janzen HH. 2009. Long-term ecological sites: Musings on the future, as seen<br />

(dimly) from the past. Global Change Biology 15: 2770–2778.<br />

Juice SM, Fahey TJ, Siccama TG, Driscoll CT, Denny EG, Eagar C, Cleavitt<br />

NL, Minocha R, Richardson AD. 2006. Response <strong>of</strong> sugar maple to calcium<br />

addition to northern hardwood forest. Ecology 87: 1267–1280.<br />

Knapp AK, Briggs JM, Hartnett DC, Collins SL. 1998. Grassland Dynamics:<br />

Long-Term Ecological Research in Tallgrass Prairie. Oxford <strong>University</strong><br />

Press.<br />

Knapp AK, et al. 2004. Searching for generality in ecology: Post-hoc<br />

synthesis <strong>of</strong> long-term data from North American grasslands and<br />

South African savannas. Frontiers in Ecology and the Environment 2:<br />

483–491.<br />

Likens GE, Driscoll CT, Buso DC. 1996. Long-term effects <strong>of</strong> acid rain:<br />

Response and recovery <strong>of</strong> a forest ecosystem. Science 272: 244–246.<br />

Lindenmayer DB, Likens GE, Krebs CJ, Hobbs RJ. 2010. Improved probability<br />

<strong>of</strong> detection <strong>of</strong> ecological “surprises.” Proceedings <strong>of</strong> the National<br />

Academy <strong>of</strong> Sciences 107: 21957–21962.<br />

Luo Y, et al. 2011. Coordinated approaches to quantify long-term ecosystem<br />

dynamics in response to global change. Global Change Biology 17:<br />

843–854.<br />

Mack MC, Schuur EAG, Bret-Harte MS, Shaver GR, Chapin FS III. 2004.<br />

Ecosystem carbon storage in arctic tundra reduced by long-term<br />

nu trient fertilization. Nature 431: 440–443.<br />

Melillo JM, Steudler PA, Aber JD, Newkirk K, Lux H, Bowles FP, Cartricala<br />

C, Magill AH, Ahrens T, Morrisseau S. 2002. Soil warming and carboncycle<br />

feedbacks to the climate system. Science 298: 2173–2176.<br />

Meserve PL, Kelt DA, Milstead WB, Gutiérrez JR. 2003. Thirteen years <strong>of</strong><br />

shifting top-down and bottom-up control. BioScience 53: 633–646.<br />

Meyer JL, Wallace JB. 2001. Lost linkages and lotic ecology: Rediscovering<br />

small streams. Pages 295–317 in Press MC, Huntly NJ, Levin S, eds.<br />

Ecology: Achievement and Challenge. Blackwell Science.<br />

Millar N, Robertson GP, Grace PR, Gehl RJ, Hoben JP. 2010. Nitrogen fertilizer<br />

management for nitrous oxide (N 2 O) mitigation in intensive corn<br />

(Maize) production: An emissions reduction protocol for US Midwest<br />

agriculture. Mitigation and Adaptation Strategies for Global Change<br />

15: 185–204.<br />

Neff JC, Townsend AR, Gleixner G, Lehman SJ, Turnbull J, Bowman WD.<br />

2002. Variable effects <strong>of</strong> nitrogen additions on the stability and turnover<br />

<strong>of</strong> soil carbon. Nature 419: 915–917.<br />

[NRC] National Research Council. 2001. Grand Challenges in Environmental<br />

Sciences. National Academies Press.<br />

Odum EP. 1977. The emergence <strong>of</strong> ecology as a new integrative discipline.<br />

Science 195: 1289–1293.<br />

Articles<br />

Pardo LH, et al. 2011. Effects <strong>of</strong> nitrogen deposition and empirical nitrogen<br />

critical loads for ecoregions <strong>of</strong> the United States. Ecological Applications<br />

21: 3049–3082.<br />

Peterson GD, Cumming GS, Carpenter SR. 2003. Scenario planning: A<br />

tool for conservation in an uncertain future. Conservation Biology 17:<br />

358–366.<br />

Rees M, Condit R, Crawley M, Pacala S, Tilman D. 2001. Long-term studies<br />

<strong>of</strong> vegetation dynamics. Science 293: 650–655.<br />

Reich PB, Hobbie SE, Lee T, Ellsworth DS, West JB, Tilman D, Knops JMH,<br />

Naeem S, Trost J. 2006. Nitrogen limitation constrains sustainability <strong>of</strong><br />

ecosystem response to CO 2 . Nature 440: 922–925.<br />

Robertson GP, et al. 2012. Long-term ecological research in a humandominated<br />

world. BioScience 62: 342–353.<br />

Slavik K, Peterson BJ, Deegan LA, Bowden WB, Hershey AE, Hobbie JE.<br />

2004. Long-term responses <strong>of</strong> the Kuparuk River ecosystem to phosphorus<br />

fertilization. Ecology 85: 939–954.<br />

Smith MD. 2011. An ecological perspective on extreme climatic events: A<br />

synthetic definition and framework to guide future research. Journal <strong>of</strong><br />

Ecology 99: 656–663.<br />

Smith MD, Knapp AK, Collins SL. 2009. A framework for assessing ecosystem<br />

dynamics in response to chronic resource alterations induced by<br />

global change. Ecology 30: 3279–3289.<br />

Solomon SD, Qin M, Manning Z, Chen M, Marquis KB, Avery T, Tignor M,<br />

Miller HL, eds. 2007. Climate Change 2007: The Physical Science Basis.<br />

Cambridge <strong>University</strong> Press.<br />

Suding KN, Miller A, Bechtold H, Bowman W. 2006. The consequences<br />

<strong>of</strong> species loss on ecosystem nitrogen cycling depends on community<br />

composition dynamics. Oecologia 149: 141–149.<br />

Syswerda SP, Basso B, Hamilton SK, Tausig JB, Robertson GP. 2012.<br />

Long-term nitrate loss along an agricultural intensity gradient in the<br />

upper Midwest USA. Agriculture, Ecosystems and Environment 149:<br />

10–19.<br />

Tilman D. 1982. Resource Competition and Community Structure.<br />

Princeton <strong>University</strong> Press.<br />

———. 1989. Ecological experimentation: Strengths and conceptual problems.<br />

Pages 136–157 in Likens GE, ed. Long-Term Studies in Ecology. Springer.<br />

———. 1996. Biodiversity: Population versus ecosystem stability. Ecology<br />

77: 350–363.<br />

Wallace JB, Eggert SL, Meyer JL, Webster JR. 1997. Multiple trophic<br />

levels <strong>of</strong> a forest stream linked to terrestrial litter inputs. Science 277:<br />

102–104.<br />

Weltzin JF, et al. 2003. Assessing the response <strong>of</strong> terrestrial ecosystems to<br />

potential changes in precipitation. BioScience 53: 941–952.<br />

Yeakley JA, Coleman DC, Haines BL, Kloeppel BD, Meyer JL, Swank WT,<br />

Argo BW, Deal JM, Taylor SF. 2003. Hillslope nutrient dynamics following<br />

upland riparian vegetation disturbance. Ecosystems 6: 154–167.<br />

Alan K. Knapp (aknapp@colostate.edu, alan.knapp@colostate.edu) is affiliated<br />

with the Graduate Degree Program in Ecology and with the Department<br />

<strong>of</strong> Biology at Colorado State <strong>University</strong>, Fort Collins. Melinda D. Smith<br />

and Kimberly J. La Pierre are affiliated with the Department <strong>of</strong> Ecology<br />

and Evolutionary Biology at Yale <strong>University</strong>, in New Haven, Connecticut.<br />

Sarah E. Hobbie is affiliated with the Department <strong>of</strong> Ecology, Evolution<br />

and Behavior at the <strong>University</strong> <strong>of</strong> Minnesota, St. Paul. Scott L. Collins is<br />

affiliated with the Department <strong>of</strong> Biology at the <strong>University</strong> <strong>of</strong> New Mexico,<br />

in Albuquerque. Timothy J. Fahey is affiliated with the Department <strong>of</strong><br />

Natural Resources at Cornell <strong>University</strong>, in Ithaca, New York. Gretchen<br />

J. A. Hansen is affiliated with the Center for Limnology at the <strong>University</strong> <strong>of</strong><br />

Wisconsin–Madison. Douglas A. Landis is affiliated with the Department <strong>of</strong><br />

Entomology and with the Great Lakes Bioenergy Research Center at Michigan<br />

State <strong>University</strong>, in East Lansing. Jerry M. Melillo and Gaius R. Shaver are<br />

affiliated with The Ecosystems Center, Marine Biological Laboratory, in Woods<br />

Hole, Massachusetts. Timothy R. Seastedt is affiliated with the Department <strong>of</strong><br />

Ecology and Evolutionary Biology and with the Institute <strong>of</strong> Arctic and Alpine<br />

Research, at the <strong>University</strong> <strong>of</strong> Colorado, Boulder. Jackson R. Webster is affiliated<br />

with the Department <strong>of</strong> Biological Sciences at Virginia Tech, Blacksburg.<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 389


Articles<br />

Ecosystem Processes and Human<br />

Influences Regulate Streamflow<br />

Response to Climate Change at<br />

Long-Term Ecological Research Sites<br />

Julia a. Jones, irena F. Creed, Kendra l. HatCHer, robert J. Warren, Mary betH adaMs, Melinda H.<br />

benson, eMery boose, Warren a. broWn, JoHn l. CaMpbell, alan CoviCH, david W. CloW, CliFFord n.<br />

daHM, Kelly elder, CHelCy r. Ford, nanCy b. GriMM, donald l. HensHaW, Kelli l. larson, evan s.<br />

Miles, KatHleen M. Miles, stepHen d. sebestyen, adaM t. sparGo, asa b. stone, JaMes M. vose,<br />

and MarK W. WilliaMs<br />

Analyses <strong>of</strong> long-term records at 35 headwater basins in the United States and Canada indicate that climate change effects on streamflow are not<br />

as clear as might be expected, perhaps because <strong>of</strong> ecosystem processes and human influences. Evapotranspiration was higher than was predicted by<br />

temperature in water-surplus ecosystems and lower than was predicted in water-deficit ecosystems. Streamflow was correlated with climate variability<br />

indices (e.g., the El Niño–Southern Oscillation, the Pacific Decadal Oscillation, the North Atlantic Oscillation), especially in seasons when<br />

vegetation influences are limited. Air temperature increased significantly at 17 <strong>of</strong> the 19 sites with 20- to 60-year records, but streamflow trends were<br />

directly related to climate trends (through changes in ice and snow) at only 7 sites. Past and present human and natural disturbance, vegetation<br />

succession, and human water use can mimic, exacerbate, counteract, or mask the effects <strong>of</strong> climate change on streamflow, even in reference basins.<br />

Long-term ecological research sites are ideal places to disentangle these processes.<br />

Keywords: precipitation/run<strong>of</strong>f ratio, trend, succession, socioecological systems, Budyko curve<br />

Although many factors affect streamflow, recent concerns<br />

have been focused on the effects <strong>of</strong> climate change on<br />

streamflow. Increasing temperature, more-severe storms,<br />

advanced snowmelt, and declining snow cover are associated<br />

with increased drought and flooding (Groisman<br />

et al. 2004, Stewart et al. 2005, Huntington 2006, Barnett<br />

et al. 2008, Karl et al. 2009, McDonald et al. 2011, USDOI<br />

2011). Nevertheless, many human actions, natural disturbance<br />

effects, and ecosystem processes complicate, mitigate,<br />

and potentially counteract the climate effects on streamflow<br />

(Meybeck 2003, Jones 2011). Relevant human actions<br />

include ongoing disturbance and legacies <strong>of</strong> past disturbance,<br />

as well as global climate change. Understanding how<br />

climate change, social systems, and ecosystem processes<br />

affect streamflow is critical for mitigating conflicts between<br />

economic development and environmental conservation.<br />

Long-term studies <strong>of</strong> headwater basins, the source areas<br />

for water supplies, provide an informative starting point<br />

for understanding the effects <strong>of</strong> climate, social factors, and<br />

ecosystem processes on streamflow (figure 1). The US Forest<br />

Service (USFS) Experimental Forests and Ranges (EFRs) and<br />

the US Department <strong>of</strong> Agriculture Agricultural Research<br />

Service (ARS) established long-term studies in small basins<br />

throughout the United States beginning in the early 1900s<br />

(see supplemental table S1, available online at http://dx.doi.<br />

org/10.1525/bio.2012.62.4.10). Four EFRs (AND, CWT, HBR,<br />

LUQ; for the site-name abbreviations, see table S1) and one<br />

ARS site (JRN) became member sites <strong>of</strong> the US Long Term<br />

Ecological Research (<strong>LTER</strong>) Network as early as 1980. Some<br />

<strong>LTER</strong> Network sites also utilize streamflow records from the<br />

US Geological Survey (USGS) National Water Information<br />

Service database. Some headwater basin studies participate<br />

in the USGS Hydrologic Benchmark Network; the USGS<br />

Water, Energy, and Biogeochemical Budgets (WEBB) program;<br />

or the Canadian HydroEcological Landscapes and<br />

Processes (HELP) program. Climate and hydrologic data<br />

from many sites have been publicly available since the 1990s<br />

(www.fsl.orst.edu/climhy).<br />

BioScience 62: 390–404. ISSN 0006-3568, electronic ISSN 1525-3244. © 2012 by American Institute <strong>of</strong> Biological Sciences. All rights reserved. Request<br />

permission to photocopy or reproduce article content at the <strong>University</strong> <strong>of</strong> California Press’s Rights and Permissions Web site at www.ucpressjournals.com/<br />

reprintinfo.asp. doi:10.1525/bio.2012.62.4.10<br />

390 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


Figure 1. Selected study basins from US Long Term Ecological Research (<strong>LTER</strong>) sites, US Forest Service Experimental<br />

Forests and Ranges, and US Geological Survey (USGS) Water, Energy, and Biogeochemical Budgets (WEBB) sites.<br />

Although they are less numerous than, for example, the<br />

USGS reference sites used in studies <strong>of</strong> climate change (e.g.,<br />

P<strong>of</strong>f et al. 2007), <strong>LTER</strong> study sites provide unique insights<br />

Articles<br />

into the interacting effects <strong>of</strong> social systems, ecosystems,<br />

and climate change on hydrology. In common with all ecosystems<br />

on Earth, the study basins have a long history <strong>of</strong><br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 391


Articles<br />

natural disturbances and human impacts. Most <strong>of</strong> the study<br />

basins experienced no management during the period <strong>of</strong><br />

record, but all <strong>of</strong> them are experiencing succession from past<br />

human disturbances; many continue to experience natural<br />

disturbances; and a few have agriculture, forestry, or residential<br />

development. Moreover, these study sites have matching<br />

records <strong>of</strong> climate drivers and hydrologic responses<br />

from (in most cases) relatively small areas. Therefore, these<br />

sites and analyses provide a unique opportunity to compare<br />

hydrologic responses to climate drivers over multiple<br />

decades and to interpret the responses on the basis <strong>of</strong> concurrent<br />

studies <strong>of</strong> social and ecosystem processes.<br />

A conceptual model <strong>of</strong> social systems, climate,<br />

ecosystems, and water<br />

The fundamental premise <strong>of</strong> the present article is that<br />

streamflow responds to ecosystem processes, which, in<br />

turn, respond both to climate drivers and to social drivers<br />

(figure 2). Social systems are a primary driver <strong>of</strong> streamflow<br />

through human water use and regulation and may indirectly<br />

affect streamflow through human-induced climate change.<br />

However, even in headwater ecosystems lacking human<br />

residents, social factors, including population dynamics,<br />

economic development, political conflicts, and resource<br />

policies, produce ecosystem disturbances, including forest<br />

harvest or clearance, grazing, agriculture, mining, and fire.<br />

In turn, these disturbances influence ecological succession,<br />

evapotranspiration, and streamflow.<br />

Climate drivers—especially precipitation, temperature,<br />

snow and ice, and extreme events—also create ecosystem<br />

disturbances (e.g., wildfires, floods, wind and ice storms).<br />

Ecosystems continuously respond to disturbances (both<br />

human and natural) through ecological succession, and disturbances<br />

and responses differ among biomes. Ecosystems,<br />

social systems, and climate also respond to streamflow.<br />

Headwaters provide water for downstream ecosystems and<br />

communities, and ecosystem processes drive climate through<br />

evapotranspiration and energy exchange (figure 2).<br />

Social systems<br />

Economic development<br />

Political stability<br />

Demographic trends<br />

Resource policies<br />

Institutions<br />

Ecosystems<br />

Disturbance<br />

Succession<br />

Biomes<br />

Water use<br />

Streamflow<br />

Climate systems<br />

Precipitation<br />

Temperature<br />

Snow/ice dynamics<br />

Extreme events<br />

Figure 2. Conceptual model <strong>of</strong> social, ecological, and climate<br />

influences on streamflow.<br />

Many <strong>of</strong> the study sites experienced social effects on<br />

ecosystem processes prior to becoming <strong>LTER</strong> sites. Most <strong>of</strong><br />

the temperate forest in the eastern half <strong>of</strong> the United States<br />

and Canada and tropical forests in Puerto Rico experienced<br />

forest harvest, land clearance for agriculture, and grazing,<br />

followed by land abandonment (Swank and Crossley 1988,<br />

Foster and Aber 2004). Sites in the southwestern United<br />

States experienced intensive grazing <strong>of</strong> domestic animals<br />

during the past four centuries (Peters et al. 2006). Boreal<br />

forest, temperate wet forest, and tundra sites experienced<br />

varying fire regimes, mostly driven by climate but, in some<br />

cases, affected by prehistoric peoples (e.g., Weisberg and<br />

Swanson 2003). Some sites contain agriculture, forestry, and<br />

urban development.<br />

In this study, we primarily examine the relationship between<br />

climate and streamflow on the basis <strong>of</strong> energy exchange. In<br />

cases in which streamflow behavior cannot be explained<br />

purely by climate, other hydrologic processes, as well as past<br />

and present human and natural disturbance effects on ecosystems,<br />

are considered as possible explanations.<br />

Study sites and questions<br />

In this study, we examined long-term records <strong>of</strong> air temperature<br />

(T), precipitation (P), and streamflow (Q) from<br />

35 basins at US <strong>LTER</strong> Network sites, USFS EFRs, USGS<br />

WEBB sites, and Canadian sites (table S1, figure 3a, 3b).<br />

The study sites were headwater basins that have mostly<br />

experienced no management since their records began.<br />

Nevertheless, the study basins have undergone succession in<br />

response to earlier natural or human disturbance, and some<br />

basins experienced natural disturbance during the periods<br />

<strong>of</strong> record. The records <strong>of</strong> T, P, and Q were obtained from<br />

the US <strong>LTER</strong> Network’s Climate and Hydrology Database<br />

Projects (ClimDB/HydroDB; http://climhy.lternet.edu), supplemented<br />

by USGS streamflow data (http://co.water.usgs.<br />

gov/lochvale, http://ny.cf.er.usgs.gov/hbn) and data from the<br />

Canadian HELP program (table S1).<br />

The study basins ranged from 0.1 to 10,000 square kilometers;<br />

2 were less than 10 hectares (ha), 5 were 10–100 ha;<br />

10 were 100–1000 ha; 5 were 1000–10,000 ha; 8 were<br />

10,000–100,000 ha; 2 were 100,000–1 million ha; and 2 were<br />

undefined (FCE, MCM; see table S1). The two largest basins<br />

were near ARC and GCE; eight additional large basins are<br />

located in the vicinity <strong>of</strong> CAP, HFR, JRN, KBS, NTL, OLY,<br />

PIE, and SEV. The smallest basins (smaller than 100 ha) are<br />

mostly associated with USFS or Canadian sites, at AND, BES,<br />

CWT, DOR, FER, HBR, MAR, and TLW.<br />

The study sites represent many biomes (potential vegetation)<br />

(figure 4). More than half <strong>of</strong> the sites were temperate<br />

forest (CAS, CWT, DOR, HBR, HFR, FER, KEJ, MAR,<br />

MRM, NTL, PIE, TLW), temperate wet forest (AND, CAR,<br />

MAY, OLY), and boreal forest (ELA, FRA, LVW, TEN,<br />

UPC). The remainder includes tundra or cold desert (ARC,<br />

MCM, NWT), warm desert (CAP, JRN, SEV), cool desert or<br />

woodland (BNZ), woodland or grassland (BES, FCE, GCE,<br />

KBS, KNZ, SBC), and tropical rainforest (LUQ). Actual<br />

392 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


a<br />

b<br />

ARC<br />

BNZ<br />

Legend<br />

Data analysis<br />

Budyko<br />

Trend<br />

Oscillations<br />

MAY<br />

CAR<br />

CAR<br />

CAR<br />

CAS<br />

ARC<br />

BNZ<br />

MAY<br />

CAR<br />

CAR<br />

CAR<br />

CAS<br />

UPC<br />

OLY<br />

AND<br />

SBC CAP<br />

OLY<br />

UPC<br />

AND<br />

SBC CAP<br />

Average annual P−PET (mm)<br />

< 49<br />

600–799<br />

50–99 800–999<br />

100–199 1000–1999<br />

200–399 2000–2999<br />

400–599 > 3.000<br />

MRM<br />

TEN<br />

LVW NWT<br />

FRA<br />

MRM<br />

SEV<br />

JRN<br />

TEN<br />

LVW NWT<br />

FRA<br />

SEV<br />

JRN<br />

ELA<br />

MAR<br />

KNZ<br />

KNZ<br />

NTL<br />

ELA<br />

MAR<br />

NTL<br />

KEJ<br />

TLW DOR<br />

HBR<br />

HFR<br />

KBS<br />

PIE<br />

BES<br />

FER<br />

BES<br />

FER<br />

Articles<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 393<br />

GCE<br />

TLW<br />

GCE<br />

CWT<br />

FCE<br />

DOR<br />

HFR<br />

KBS<br />

PIE<br />

FER<br />

BES<br />

CWT<br />

FCE<br />

KEJ<br />

HBR<br />

Figure 3. (a) Map <strong>of</strong> sites used in this analysis. The study-site characteristics and abbreviations are in supplemental table S1,<br />

available online at http://dx.doi.org/10.1525/bio.2012.62.4.10. The symbols indicate which <strong>of</strong> the three analyses (Budyko<br />

curve, trends, and correlation with climate indices [oscillations]) were conducted with data from that site. (b) Map <strong>of</strong> average<br />

annual precipitation (P) minus potential evapotranspiration (PET) in millimeters (mm) with study-site locations.<br />

LUQ<br />

LUQ


Articles<br />

Figure 4. Mean annual temperature, mean annual<br />

precipitation, and biomes <strong>of</strong> the study sites. The studysite<br />

abbreviations are in supplemental table S1, available<br />

online at http://dx.doi.org/10.1525/bio.2012.62.4.10. Data<br />

for the MCM site are not shown. Abbreviations: °C, degrees<br />

Celsius; mm, millimeters.<br />

and potential vegetation may differ because <strong>of</strong> disturbance<br />

(table S1).<br />

In our analyses, we examined three questions, using<br />

successively more-stringent requirements <strong>of</strong> data sets and<br />

interpreted these results in the light <strong>of</strong> ecological and social<br />

factors: (1) How is potential evapotranspiration (PET)<br />

related to actual evapotranspiration (AET) at each site, and<br />

how do these relationships compare with the theoretical<br />

Budyko curve (n = 30 sites)? (2) How is streamflow correlated<br />

with climate indices (e.g., the El Niño–Southern<br />

Oscillation [ENSO], the Pacific Decadal Oscillation [PDO],<br />

the North Atlantic Oscillation [NAO]; n = 21 sites)? (3) How<br />

have temperature, precipitation, and streamflow changed<br />

over time (n = 19 sites)?<br />

Energy- and water-balance relationships to observed<br />

water use<br />

The values <strong>of</strong> T, P, and Q from 30 sites with matched T, P, and<br />

Q (table S1, figure 3) were used to calculate PET and AET<br />

(i.e., P – Q). These values were plotted on the Budyko curve<br />

(Budyko 1974), which displays the relationship between<br />

PET and AET, each indexed by P (figure 5a). Thirty <strong>of</strong> the<br />

35 sites had data on T, P, Q, and basin area for a common<br />

10-year period (1993–2002), although a slightly adjusted<br />

period was used for 10 <strong>of</strong> the sites (figure 5b). PET was<br />

calculated from T (after Hamon 1963) on the basis <strong>of</strong> the<br />

number <strong>of</strong> daylight hours, mean monthly temperature, and<br />

the saturated vapor pressure. Annual PET was calculated as a<br />

sum <strong>of</strong> monthly values. The Budyko curve assumes that the<br />

water balance is Q = P – ET (evapotranspiration), with no<br />

significant losses to or gains from groundwater, and that the<br />

basins are at steady state, unaffected by vegetation dynamics<br />

(Donohue et al. 2007).<br />

The distribution <strong>of</strong> study basins on the Budyko curve<br />

reveals that observed water use in ecosystems in small basins<br />

deviated systematically from its expected dependence on<br />

energy and water balances. As was expected, observed ecosystem<br />

water use (AET ÷ P) was positively correlated to energy<br />

and water inputs to evapotranspiration (PET ÷ P) in sites<br />

with a water surplus (P ÷ PET) and insensitive to increases<br />

in energy at sites with a water deficit (P ÷ PET), following<br />

the theoretical Budyko curve (figure 5b). However, only 7 <strong>of</strong><br />

30 sites (ARC, DOR, FRA, HBR, KEJ, KNZ, and OLY) fell<br />

on the Budyko curve, where observed water use (AET ÷ P)<br />

was equal to predicted water use (PET ÷ P) (fi gure 5b). Of<br />

the 19 sites with a moisture surplus (P ÷ PET) that did not<br />

fall on the Budkyo curve, 14 were above it, with higher than<br />

expected evapotranspiration (AET ÷ P > PET ÷ P). Of the<br />

five sites with moisture deficits (P < PET), four fell below the<br />

Budyko curve, with lower than expected evapotranspiration<br />

(AET ÷ P < PET ÷ P) (figure 5b).<br />

This result may indicate that ecosystems evaporate, transpire,<br />

and store more water than would be expected on the<br />

basis <strong>of</strong> temperature and day length at wet sites and less than<br />

would be expected at dry sites. Ecosystem structure (e.g.,<br />

rooting depth, leaf area) and processes (e.g., adaptations to<br />

water deficits) may produce lower streamflow in wet sites and<br />

higher streamflow in dry sites than would be predicted from<br />

energy and water balances alone. However, other factors may<br />

also explain the departures <strong>of</strong> the sites from the theoretical<br />

Budyko curve. For instance, the PET value estimated from<br />

climate-station T records may not represent PET over entire<br />

basins, especially in mountain sites (e.g., AND, NWT, LVW).<br />

AET ÷ P is also considerably overestimated from P – Q<br />

in basins in which the groundwater recharge bypasses the<br />

stream gauge (Graham et al. 2010, Verry et al. 2011).<br />

When the annual values <strong>of</strong> T, P, and Q are plotted on the<br />

Budkyo curve, the interannual variation <strong>of</strong> AET relative<br />

to PET varies among biomes (figure 5c, 5d). Variation in<br />

AET ÷ P was less than in PET ÷ P at the desert sites (CAP,<br />

SEV) and at forested sites (AND, CAS, CWT, FER, HBR,<br />

MAR, NTL) (figure 5d). In contrast, at alpine sites (LVW,<br />

NWT), the interannual variation in AET ÷ P was large relative<br />

to the variation in PET ÷ P. This behavior <strong>of</strong> sites relative<br />

to the Budyko curve implies that ecosystems are capable <strong>of</strong><br />

adjusting AET to compensate for climate variability at desert,<br />

grassland, and forest sites, but less so at alpine sites.<br />

Ecosystems have more-similar rates <strong>of</strong> net primary productivity<br />

per unit precipitation in dry than in wet years<br />

(Huxman et al. 2004). Comparisons <strong>of</strong> long-term AET and<br />

PET from study basins to the theoretical Budyko curve<br />

(figure 5) suggest that AET varies in a narrower range than<br />

would be expected from energy and water balances alone,<br />

which underscores the importance <strong>of</strong> ecosystem process<br />

effects on streamflow.<br />

394 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


a<br />

Evaporative index (actual<br />

evapotranspiration ÷ precipitation (AET÷P))<br />

c<br />

Actual evapotranspiration ÷ precipitation<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

Water limit<br />

Energy limit<br />

Budyko curve<br />

0.2<br />

PET÷P1<br />

0.0<br />

energy limited water limited<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0<br />

Dryness index (potential evapotranspiration ÷ precipitation (PET÷P))<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

−0.2<br />

BA<br />

SBC<br />

BNZ<br />

A<br />

SEV<br />

CAP<br />

0.0<br />

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0 12.0<br />

Potential evapotranspiration ÷ precipitation<br />

Streamflow and regional climate oscillations<br />

We examined the relationship <strong>of</strong> three climate indices<br />

(ENSO, PDO, and NAO) to Q at 21 sites (tables S1 and S2).<br />

These indices measure multiyear or multidecadal oscillations<br />

<strong>of</strong> sea-surface temperatures and atmospheric-pressure<br />

differentials in the east–central tropical Pacific (ENSO), the<br />

northern Pacific (PDO), and the northern Atlantic (NAO).<br />

They are correlated with local and regional temperature,<br />

precipitation, and streamflow in the United States (Cayan<br />

et al. 1999, Barlow et al. 2001, Enfield et al. 2001). The sites<br />

included in this analysis had fewer than10 years <strong>of</strong> continuous<br />

(monthly) Q at one or more gauging stations, separated<br />

A<br />

Actual evapotranspiration ÷ precipitation<br />

Actual evapotranspiration ÷ precipitation<br />

ELA MARNTLBES<br />

GCE<br />

CAS<br />

TEN<br />

FER<br />

TLW<br />

CWT<br />

ARC<br />

AND<br />

UPC<br />

CAR<br />

KBS<br />

PIE<br />

MRM<br />

NWT<br />

DOR<br />

FRA<br />

OLY<br />

LVW<br />

KEJ<br />

HBR<br />

LUQ<br />

KNZ SBC BNZ<br />

Articles<br />

into a cool season (November–April) and a warm season<br />

(May–October). Correlated streamflow records (Pearson’s<br />

r > .80) were pooled at study sites with multiple stream<br />

gauges. Climate indices were obtained from online databases<br />

(NAO, www.cgd.ucar.edu/cas/jhurrell/indices.html; ENSO,<br />

www.cdc.noaa.gov/ClimateIndices/List; PDO, www.esrl.noaa.<br />

gov/psd/data/correlation/pdo.data). The streamflow–climate<br />

oscillation relationships were tested using generalized least<br />

squares models with autoregressive moving average functions;<br />

the models were evaluated with the Durbin–Watson test<br />

statistic and Akaike’s information criterion. The results are<br />

shown as the sign (+ or –) <strong>of</strong> the relationship <strong>of</strong> streamflow<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 395<br />

b<br />

d<br />

1.0<br />

SEV<br />

P−PET<br />

CAP<br />


Articles<br />

to climate oscillation variables that were significant at<br />

a < .10 in the models (table S2).<br />

The cool-season (November–April) and warm-season<br />

(May–October) streamflow values were significantly correlated<br />

with at least one climate index or interaction term<br />

at all sites except FRA, KNZ, NWT, and PIE (table S2).<br />

Significant correlations <strong>of</strong> streamflow were more frequent<br />

with ENSO (11, plus six interactions) and PDO (10, plus<br />

one interaction) than with NAO (4, plus six interactions)<br />

(table S2). Significant correlations were also slightly more<br />

frequent in winter (18) than in summer (14). These findings<br />

extend Greenland and colleagues’ (2003) analysis <strong>of</strong> climate<br />

indices, temperature, and precipitation at US <strong>LTER</strong> Network<br />

sites and corroborate the results <strong>of</strong> other studies. Molles<br />

and Dahm (1990) noted that streamflow in two rivers in<br />

New Mexico was significantly higher during El Niño (warm<br />

sea-surface temperatures in the eastern Pacific) than during<br />

La Niña conditions. Cayan and colleagues (1999) showed<br />

that days with high daily precipitation and streamflow<br />

were more frequent than average in the US Southwest and<br />

less frequent in the Northwest during El Niño by examining<br />

effects on snowpack accumulation and the subsequent<br />

melt. Enfield and colleagues (2001) found that sea-surface<br />

temperatures in the northern Atlantic are correlated with<br />

those in the northern Pacific and are associated with variations<br />

in streamflow in the Mississippi River and in Florida.<br />

Sea-surface temperature and pressure anomalies originating<br />

in the North Pacific affect cyclonic circulation over the East<br />

Coast and summer precipitation, streamflow, and drought<br />

(Barlow et al. 2001).<br />

These findings underscore the importance <strong>of</strong> separating<br />

the effects on streamflow <strong>of</strong> climate variability from longterm<br />

trends. Because the ENSO oscillation has a wavelength<br />

<strong>of</strong> 2–7 years, trends in climate and streamflow data sets over<br />

fewer than 20 years may simply reflect ENSO. Similarly, the<br />

PDO has a wavelength <strong>of</strong> 4–16 years (MacDonald et al. 2005),<br />

with mostly negative PDO in the 1950s to the mid-1970s<br />

and mostly positive PDO from 1976 to 1998. The NAO was<br />

predominantly negative between the 1950s and the early<br />

1970s and was mostly positive between the 1980s and the<br />

early 1990s. As a result, climate and streamflow trends from<br />

the 1950s to 2000 may be strongly affected by these climate<br />

oscillations. For example, at ARC, streamflow increased<br />

between 1988 and 2003, but declined between 1988 and 2008,<br />

so lengthening the record shifted the direction <strong>of</strong> apparent<br />

change. The lack <strong>of</strong> statistically significant increases in<br />

minimum temperature at the tundra ARC and boreal forest<br />

LVW sites (see the next section) may also be attributed to<br />

confounding effects <strong>of</strong> climate oscillations on these relatively<br />

short-term records.<br />

Climate oscillations influence ecosystem processes<br />

through streamflow and moisture. Streamflow was slightly<br />

more weakly correlated to climate indices in summer than<br />

in winter, perhaps because precipitation and streamflow are<br />

more closely related when ecosystems are dormant. ENSO<br />

is linked to aquatic-community structure in the Southwest<br />

(Sponseller et al. 2010), PDO is related to salmon returns<br />

in the Northwest (Mantua et al. 1997), and NAO is linked<br />

to stream salamander abundance in the Southeast (Warren<br />

and Bradford 2010). Headwater streamflow records are<br />

just beginning to be long enough to relate climate variability<br />

and trends to ecosystem processes and population<br />

dynamics.<br />

Climate and streamflow trends at long-term<br />

headwater basin study sites<br />

Nineteen sites had long-term records suitable for testing<br />

trends in T, P, and Q (table S1). Sites were included in the<br />

analysis if they had overlapping records <strong>of</strong> T, P, and Q<br />

that exceeded 20 years. The climate and streamflow record<br />

lengths used for trend estimation ranged from 20 to just<br />

over 60 years; five were 20–30 years; one was 30–40 years;<br />

four were 40–50 years; seven were 50–60 years; and two<br />

were more than 60 years (table S1). The records exceeding<br />

40 years are from USGS gauges and nearby climate stations<br />

(CAP, GCE, JRN, OLY, SBC, SEV), USFS EFRs that became<br />

US <strong>LTER</strong> Network sites (AND, CWT, HBR), other <strong>LTER</strong><br />

Network sites (HFR), and USFS EFRs (FER, MAR) that<br />

did not become <strong>LTER</strong> Network sites. The records less than<br />

40 years in length were USGS gauges and nearby climate<br />

stations at <strong>LTER</strong> Network sites, WEBB sites, and EFRs (ARC,<br />

FRA, LUQ, LVW, NTL, and NWT).<br />

Interannual trends in minimum and maximum daily T,<br />

P, Q, and run<strong>of</strong>f ratios (Q:P) were estimated using linear<br />

regression and the Mann–Kendall nonparametric trend test<br />

(Helsel and Hirsch 2002). In these analyses, we used the<br />

period <strong>of</strong> record or from 1950 onward. Linear regressions<br />

and Mann–Kendall tests produced almost identical results<br />

(Hatcher 2011). The water year was defined as 1 October to<br />

30 September. Tests were conducted using daily data. The<br />

daily P and Q values were log transformed before analysis.<br />

Data were tested for autocorrelation before analysis, and<br />

residuals from linear regression analyses were also tested<br />

for autocorrelation. Significant trends in annual T, P, and Q<br />

were defined as 10 or more days (out <strong>of</strong> 365) with significant<br />

trends (at a � .025) and no autocorrelation before regression<br />

or in the residuals, and an average slope <strong>of</strong> the trend in<br />

daily values exceeding its standard error.<br />

Annual minimum or maximum daily temperature<br />

increased significantly at 17 <strong>of</strong> the 19 sites (minimum temperatures<br />

increased at 13 sites and maximum temperatures<br />

increased at 7 sites), but only two sites experienced significant<br />

changes in precipitation over the period <strong>of</strong> available<br />

record (figure 6). Minimum daily temperatures increased by<br />

several degrees Celsius since 1980 at NWT and FRA, highelevation,<br />

snow-dominated sites in the Rocky Mountains,<br />

but not at the other high-elevation Rocky Mountain site<br />

(LVW). Minimum daily temperature also increased by several<br />

degrees Celsius since the 1950s at climate stations near<br />

JRN and SEV in New Mexico, since the 1950s at a southeastern<br />

temperate forest site (CWT), and since the 1960s<br />

at a northern hardwood site (MAR) but not at its neighbor<br />

396 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


Annual precipitation (mm)<br />

4000<br />

3500<br />

3000<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

a<br />

0<br />

−15.0 −10.0 −5.0 0.0 5.0 10.0 15.0 20.0 25.0<br />

Minimum daily temperature (°C)<br />

1950 est<br />

1960 est<br />

1970 est<br />

1980 est<br />

1990 est<br />

2010 est<br />

(NTL, for which the record began in 1990). Mean annual<br />

precipitation increased significantly at LUQ and NWT. The<br />

first day <strong>of</strong> spring (defined as the last day <strong>of</strong> freezing temperature)<br />

moved earlier by between 0.31 and 1.98 days per<br />

year—that is, by more than 15 days in 50 years—at eight<br />

sites (AND, ARC, CWT, FER, FRA, HBR, LUQ, MAR, NWT)<br />

(Hatcher 2011).<br />

Run<strong>of</strong>f ratios (Q:P) changed at 8 <strong>of</strong> 19 sites (figure 6b).<br />

Tundra and boreal forest sites with ice and permafrost (LVW,<br />

NWT) experienced increases in run<strong>of</strong>f ratios, and so did<br />

temperate deciduous forest sites in the northeastern United<br />

States (HBR, HFR, PIE), which have a seasonal snowpack. An<br />

increase in run<strong>of</strong>f ratio means either that AET has decreased,<br />

or that there is a net addition <strong>of</strong> water to the system, such as<br />

from melting ice or interbasin water transfers. The observed<br />

increases in run<strong>of</strong>f ratios at LVW and NWT may be associated<br />

with the melt <strong>of</strong> ice, snow, and permafrost in response<br />

to warming temperatures during seasons in which these<br />

ecosystems are dormant (not taking up water). However,<br />

warming did not result in increased run<strong>of</strong>f ratios at other<br />

sites with permafrost (ARC, which has a short record) or<br />

seasonal snowpacks (e.g., AND, FRA, MAR, NTL). Run<strong>of</strong>f<br />

Articles<br />

1950 est<br />

1960 est<br />

1970 est<br />

1980 est<br />

1990 est<br />

2010 est<br />

.1<br />

0<br />

A CAP<br />

−10.0 −5.0 0.0 5.0 10.0 15.0 20.0 25.0 30.0 35.0<br />

ratios did not change at most other sites, which mostly lack<br />

significant snow and ice (Hatcher 2011).<br />

Streamflow changes vary according to the season and differ<br />

among various biomes (figure 7). At undisturbed desert<br />

sites in Arizona (CAP) and New Mexico (SEV), streamflow<br />

did not change at any time <strong>of</strong> year (figure 7a, 7b). However,<br />

in a desert mountain basin northeast <strong>of</strong> JRN (New Mexico)<br />

and in a semiarid mountain basin near SBC (southern<br />

California) containing residential and urban development,<br />

streamflow increased during low-flow periods (figure 7c,<br />

7d). In a large basin in coastal <strong>Georgia</strong> containing agriculture<br />

and forest plantations (GCE), streamflow declined in<br />

early and late summer (figure 7e).<br />

At tundra sites on the North Slope <strong>of</strong> Alaska and in the<br />

Rocky Mountains (ARC, NWT; figure 7f, 7g), streamflow<br />

increased in early spring and late fall, during time periods<br />

adjacent to freezing periods. Streamflow increased in spring<br />

at boreal forest sites in the Rocky Mountains (FRA, LVW;<br />

figure 7h, 7i). At a temperate forest site in western North<br />

Carolina (CWT), where seasonal snowpacks do not form,<br />

streamflow did not change at any time <strong>of</strong> year (figure 7j), but<br />

at a temperate forest site in West Virginia (FER), streamflow<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 397<br />

Run<strong>of</strong>f ratio (discharge:precipitation)<br />

1<br />

.9<br />

.8<br />

.7<br />

.6<br />

.5<br />

.4<br />

.3<br />

.2<br />

b<br />

Maximum daily temperature (°C)<br />

Figure 6. (a) Multidecade trends in minimum daily temperature (in degrees Celsius [°C]) and precipitation (in millimeters<br />

[mm]) at long-term watershed study sites. Each site is designated by the minimum daily temperature and precipitation<br />

at the beginning and end <strong>of</strong> the decades spanning the period <strong>of</strong> record, based on the statistically significant trend<br />

in that variable over the period <strong>of</strong> record. The initial observation is connected to or contained within the 2010<br />

estimated (est) observation for each site. The radius <strong>of</strong> the 2010 estimated symbol is 0.5°C and 125 mm. Statistically<br />

significant increases in minimum daily temperature occurred at all sites except ARC, CAP, JRN, PIE, and SEV<br />

(the study-site characteristics and abbreviations are in supplemental table S1, available online at http://dx.doi.<br />

org/10.1525/bio.2012.62.4.10). Statistically significant changes in annual precipitation occurred only at LUQ and<br />

NWT. (b) Multidecade trends in maximum daily temperature and run<strong>of</strong>f ratios at long-term watershed study sites.<br />

Each site is designated by the maximum daily temperature and run<strong>of</strong>f ratio at the beginning and end <strong>of</strong> the decades<br />

spanning the period <strong>of</strong> record, based on the trend in that variable over the period <strong>of</strong> record. The initial observation is<br />

connected to or contained within the 2010 estimated observation for each site. The radius <strong>of</strong> the 2010 estimated symbol<br />

is 0.5°C and .0125. Statistically significant increases in maximum daily temperature occurred at CAP, FRA, JRN, LUQ,<br />

NWT, and SEV. Statistically significant changes in annual streamflow occurred at FER, GCE, HFR, HBR, JRN, LVW,<br />

MAR, NWT, NTL, PIE, and SEV.


Articles<br />

Proportion change per year<br />

relative to the mean daily flow<br />

Proportion change per year<br />

relative to the mean daily flow<br />

Proportion change per year<br />

relative to the mean daily flow<br />

Proportion change per year<br />

relative to the mean daily flow<br />

Proportion change per year<br />

relative to the mean daily flow<br />

a b<br />

CAP – Sycamore Creek – 1961–2010<br />

SEV – Jemez River – 1954–2010<br />

.10<br />

.015<br />

.05<br />

.00<br />

.01<br />

.005<br />

.00<br />

−.05<br />

−.005<br />

−.01<br />

O N D J F M A M J J A S O<br />

−.01<br />

O N D J F M A M J J A S O<br />

Month<br />

Month<br />

c JRN – Rio Ruidoso – 1950–2010<br />

d SBC – San Jose Creek – 1950–2010<br />

.015<br />

.08<br />

.01<br />

.06<br />

.005<br />

.04<br />

.00<br />

.02<br />

−.005<br />

.00<br />

−.01<br />

−.02<br />

−.015<br />

O N D J F M A M J J A S O<br />

−.04<br />

O N D J F M A M J J A S O<br />

Month<br />

Month<br />

e GCE – Ohoopee River – 1950–2009<br />

f ARC – Kuparuk River – 1971–2010<br />

.02<br />

.20<br />

.01<br />

.15<br />

.10<br />

.00<br />

.05<br />

−.01<br />

.00<br />

−.05<br />

−.02<br />

−.10<br />

−.03<br />

O N D J F M A M J J A S O<br />

−.15<br />

O N D J F M A M J J A S O<br />

Month<br />

Month<br />

g<br />

NWT – Green Lake 4 – 1981–2008<br />

FRA – East Saint Louis – 1976–2005<br />

.10<br />

.04<br />

.05<br />

.03<br />

.02<br />

.00<br />

.01<br />

−.05<br />

.00<br />

−.01<br />

−.10<br />

−.02<br />

−.15<br />

O N D J F M A M J J A S O<br />

−.03<br />

O N D J F M A M J J A S O<br />

Month<br />

Month<br />

i LVW – Loch Vale out let – 1984–2010<br />

j<br />

CWT – WS18 – 1950–2009<br />

.10<br />

.015<br />

.001<br />

.05<br />

.005<br />

.00<br />

.000<br />

−.005<br />

−.05<br />

−.010<br />

−.10<br />

O N D J F M A M J J A S O<br />

−.015<br />

O N D J F M A M J J A S O<br />

Month<br />

Month<br />

Figure 7. Daily changes in streamflow at 19 US Long Term Ecological Research sites, US Forest Service Experimental<br />

Forests and Ranges, and US Geological Survey Water, Energy, and Biogeochemical Budgets sites arranged by biome (from<br />

figure 4) as a function <strong>of</strong> the day <strong>of</strong> the water year (1 October to 30 September). (a–c) Desert sites (CAP, SEV, JRN);<br />

(d), (e) savanna sites (SBC, GCE); (f), (g) tundra sites (ARC, NWT); (h), (i) boreal forest sites (FRA, LVW); (j–p) temperate<br />

forest sites (CWT, FER, HFR, HBR, MAR, NTL, PIE); (q), (r) wet temperate forest sites (AND, OLY); (s) wet tropical<br />

forest site (LUQ). The vertical axis and the green line are the slope <strong>of</strong> regression <strong>of</strong> log-transformed streamflow for each<br />

day <strong>of</strong> the water year over the period <strong>of</strong> record (see supplemental table S1, available online at http://dx.doi.org/10.1525/<br />

bio.2012.62.4.10). The vertical axis units are the proportion change per year relative to the mean daily flow. The<br />

percentage change can be calculated as (1 + p) n , where p is the proportion change and n is the number <strong>of</strong> years. Note the<br />

different vertical axis scales. The horizontal black line represents no change (a proportion change <strong>of</strong> 0); the wiggly black<br />

lines are the upper and lower bounds on the 97.5% confidence interval. The red dots represent significant increases, and<br />

the blue dots represent significant decreases in daily streamflow, where a � .025.<br />

398 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org<br />

Proportion change per year<br />

relative to the mean daily flow<br />

Proportion change per year<br />

relative to the mean daily flow<br />

Proportion change per year<br />

relative to the mean daily flow<br />

Proportion change per year<br />

relative to the mean daily flow<br />

Proportion change per year<br />

relative to the mean daily flow<br />

h


Proportion change per year<br />

relative to the mean daily flow<br />

Proportion change per year<br />

relative to the mean daily flow<br />

Proportion change per year<br />

relative to the mean daily flow<br />

Proportion change per year<br />

relative to the mean daily flow<br />

k<br />

increased in the summer (figure 7k). Winter streamflow<br />

increased at three temperate forest sites in New England (HFR,<br />

HBR, PIE) and declined at one (NTL) (figure 7l, 7m, 7o, 7p).<br />

In addition, streamflow increased in March and decreased<br />

in April at HBR (figure 7m), and it increased in March and<br />

declined in summer at MAR (figure 7n). At wet temperate forest<br />

sites in Oregon (AND, OLY), streamflow declined in spring<br />

(figure 7q, 7r). Streamflow did not change at any time <strong>of</strong> year<br />

at a wet tropical forest site in Puerto Rico (LUQ) (figure 7s).<br />

PIE – Ipswich River – 1938–2010<br />

Articles<br />

.08<br />

FER – WS4 – 1952–2007<br />

.06<br />

HFR – Swift River – 1964–2010<br />

.06<br />

.04<br />

.04<br />

.02<br />

.02<br />

.00<br />

.00<br />

−.02<br />

−.02<br />

−.04<br />

−.04<br />

O N D J F M A M J J A S O<br />

−.06<br />

O N D J F M A M J J A S O<br />

Month<br />

Month<br />

m<br />

−.02<br />

O N D J F M A<br />

Month<br />

M J J A S O<br />

Social, ecological, and climate factors influencing<br />

streamflow trends<br />

Multiple social and ecological factors may explain the<br />

streamflow trends at long-term headwater basin sites, even<br />

though humans do not directly affect most <strong>of</strong> these sites<br />

( figure 2). Economic development, population growth, and<br />

the use <strong>of</strong> fossil-fuel resources have increased atmospheric<br />

carbon dioxide, warmed the Earth, contributed to more-<br />

intense precipitation events, and increased evapotranspiration<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 399<br />

Proportion change per year<br />

relative to the mean daily flow<br />

Proportion change per year<br />

relative to the mean daily flow<br />

.02<br />

AND – WS2 – 1958–2009<br />

.03<br />

OLY – Hoko River near Sekiu – 1962–2009<br />

.01<br />

.02<br />

.00<br />

.01<br />

.00<br />

−.01<br />

−.01<br />

−.02<br />

−.02<br />

−.03<br />

O N D J F M A M J J A S O<br />

−.03<br />

O N D J F M A M J J A S O<br />

Month<br />

Month<br />

Proportion change per year<br />

relative to the mean daily flow<br />

.04<br />

.02<br />

.00<br />

−.02<br />

−.04<br />

Proportion change per year<br />

relative to the mean daily flow<br />

Proportion change per year<br />

relative to the mean daily flow<br />

.04<br />

HBR – WS3 – 1958–2007<br />

.10<br />

MAR – S2 – 1961–2009<br />

.03<br />

.05<br />

.02<br />

.01<br />

.00<br />

.00<br />

−.05<br />

−.01<br />

−.10<br />

−.02<br />

O N D J F M A M J J A S O<br />

−.15<br />

O N D J F M A M J J A S O<br />

Month<br />

Month<br />

.01<br />

.00<br />

−.01<br />

−.02<br />

−.03<br />

−.04<br />

−.05<br />

−.06<br />

o<br />

O N D J F M A M J J A S O<br />

Month<br />

Figure 7. (Continued)<br />

NTL – Trout River near Trout Lake – 1990–2010<br />

.03<br />

.02<br />

.01<br />

.00<br />

−.01<br />

q r<br />

s<br />

LUQ – Espiritu Santo – 1975–2009<br />

−.06<br />

O N D J F M A<br />

Month<br />

M J J A S O<br />

l<br />

n<br />

p


Articles<br />

(Min et al. 2011, Pall et al. 2011), which in turn have been<br />

linked to increased flooding and drought (Barnett et al.<br />

2008, Karl et al. 2009). Yet direct climate-trend effects on<br />

streamflow in headwater basins may be mitigated by ecological<br />

processes, including disturbance, succession, and<br />

vegetation adaptations to water scarcity. In many cases, the<br />

vegetation—and, therefore, evapotranspiration in headwater<br />

basins—is affected by human activities, such as past<br />

logging, grazing, agriculture, and fire suppression. Moreover,<br />

in some headwater basins, land-use changes, including agriculture<br />

and exurban expansion, may mitigate or overwhelm<br />

climate-trend effects on streamflow.<br />

Some observed trends in streamflow appear to be direct<br />

effects <strong>of</strong> climate trends. For example, increased streamflow<br />

in fall and spring at ARC, FRA, and NWT is probably the<br />

result <strong>of</strong> the expanding period <strong>of</strong> thaw at these tundra and<br />

boreal forest sites (figure 7f, 7g, 7i). Increases in streamflow<br />

at these sites may be driven by permafrost melt; changes<br />

in the chemical composition <strong>of</strong> streamflow support this<br />

hypothesis (Bowden et al. 2008, Caine 2010). In addition,<br />

increased spring streamflow at temperate forest sites with<br />

seasonal snowpacks in New England (HBR) and in the<br />

upper Midwest (MAR) is probably the result <strong>of</strong> earlier<br />

snowmelt, whereas increased winter streamflow at temperate<br />

forest sites in New England (HFR) may be the result <strong>of</strong><br />

a shift from snow to rain. These responses are consistent<br />

with the results <strong>of</strong> published studies (Hodgkins et al. 2003,<br />

Stewart et al. 2005, Clow 2010, Campbell et al. 2011).<br />

Some observed trends in streamflow may be the result<br />

<strong>of</strong> biological responses to climate change. For example,<br />

declining streamflow at a woodland site (summer at GCE), a<br />

boreal forest site (LVW), temperate forest sites (MAR, NTL),<br />

and wet temperate forest sites (spring at AND, OLY) may<br />

be the result <strong>of</strong> increased evapotranspiration in response to<br />

warmer temperatures. Conifer forests, which occur at MAR,<br />

NTL, LVW, AND, and OLY, are adapted to photosynthesize<br />

and respire when conditions are favorable; warmer temperatures<br />

may lead to an earlier onset <strong>of</strong> transpiration and<br />

to declining streamflow (e.g., Moore KM 2010). Streamflow<br />

trends at wet temperate and boreal forest sites (LVW, AND,<br />

OLY; figure 7) are restricted to the immediate period <strong>of</strong><br />

snowmelt, and declines may reflect increased evapotranspiration<br />

(Moore KM 2010, Oishi et al. 2010, Campbell et al.<br />

2011). In contrast, streamflow trends are largest during the<br />

nonsnowmelt periods at temperate forest sites in the upper<br />

Midwest (NTL, MAR), where wetlands (bogs and lakes)<br />

occupy a large proportion <strong>of</strong> basin area (Verry et al. 2011).<br />

Increased evapotranspiration associated with declining ice<br />

cover (Magnuson et al. 2000) or increased radiation associated<br />

with decreased precipitation may account for declining<br />

flows at these sites.<br />

Some sites experienced no trends in streamflow, despite<br />

increases in temperature. For example, desert sites (CAP and<br />

SEV) and the wet tropical site (LUQ) experienced significant<br />

increases in minimum and maximum daily temperatures,<br />

no change in precipitation (figure 6), and almost no changes<br />

in streamflow (figure 7a, 7b, 7s). Vegetation adaptations to<br />

drought might explain the lack <strong>of</strong> a streamflow response to<br />

warming at the desert sites. At the wet tropical forest site,<br />

the effects <strong>of</strong> the 1996 Hurricane Hugo on leaf area, evapotranspiration,<br />

and streamflow (Scatena et al. 1996) may have<br />

overwhelmed climate-trend effects.<br />

Some sites experienced trends in streamflow that appear<br />

to be biological responses to past disturbances. For example,<br />

forest succession and declining evapotranspiration may<br />

explain increased summer streamflow at two temperate<br />

forest sites (FER, HBR), which were logged in the early<br />

nineteenth century (figure 7k, 7m). At a third temperate<br />

forest site (CWT), streamflow and run<strong>of</strong>f ratios have not<br />

changed, despite increases in air temperature (figure 6,<br />

figure 7j). Forest succession following disturbances in the<br />

early 1900s at all three sites (table S1; Swank et al. 2001,<br />

Adams et al. 2006) and associated changes in species composition<br />

or leaf area may have influenced streamflow trends.<br />

Analyses <strong>of</strong> long-term paired-basin experiment data (e.g.,<br />

Jones and Post 2004) indicate that streamflow continues to<br />

change over decades or centuries <strong>of</strong> forest succession after<br />

disturbance.<br />

Responses to land use and disturbance, such as advanced<br />

snowmelt or declining summer streamflow, may be misconstrued<br />

as responses to climate change. In paired-basin<br />

experiments (see the discussion <strong>of</strong> long-term experiments<br />

in Knapp and colleagues 2012 [in this issue]), forest harvest<br />

advanced the timing <strong>of</strong> peak snowmelt and associated<br />

streamflow by up to three weeks in temperate forest sites<br />

with a seasonal snowpack (AND, HBR); the effect lasted for<br />

more than 10 years (Jones and Post 2004). By 25 to 35 years<br />

after forest harvest in temperate forest basins (AND, CWT,<br />

HBR), summer streamflow declined by up to 30%–50%<br />

relative to the reference basins (Hornbeck et al. 1997, Swank<br />

et al. 2001, Jones and Post 2004). Regenerating species in<br />

early forest succession may transpire more water per unit <strong>of</strong><br />

leaf area and, in some cases, have greater total leaf area than<br />

the species that were removed, which would reduce summer<br />

streamflow (Swank et al. 2001, Moore GM et al. 2004).<br />

Historic legacies from past disturbance in these long-term<br />

studies demonstrate that streamflow and timing responses<br />

to forest disturbance are at least as large as responses associated<br />

with climate trends over the past 20–60 years at the<br />

study sites. Daily streamflow during the late summer and<br />

early fall increased by up to 300% in the 1–5-year period<br />

after experimental forest harvest (AND, CWT, HBR), but<br />

most daily changes were on the order <strong>of</strong> 50% or less (Jones<br />

and Post 2004). By comparison, trends in daily streamflow<br />

associated with climate trends at the 19 study sites were on<br />

the order <strong>of</strong> 0.005–0.05 <strong>of</strong> log(Q) per year. The lower value<br />

is equivalent to changes <strong>of</strong> 10%–25%, and the higher value<br />

is equivalent to more than a 100% change over 20–60 years,<br />

but changes <strong>of</strong> this magnitude are restricted to a few days<br />

per year (figure 7).<br />

Finally, some observed trends in streamflow may be<br />

direct human effects on the hydrologic cycle. For example,<br />

400 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


increased irrigation using groundwater or water imported<br />

from other basins may explain increasing streamflow during<br />

dry seasons at a desert site in New Mexico (JRN) and a<br />

savanna site in southern California (SBC), which have some<br />

agriculture and residential development (figure 7c, 7d).<br />

Increasing winter streamflow trends at a temperate forest<br />

site (PIE) may reflect urban expansion (Claessens et al.<br />

2006). Therefore, human effects on streamflow may mimic,<br />

exacerbate, counteract, or mask climate effects on streamflow,<br />

making it challenging to determine the vulnerability<br />

<strong>of</strong> human communities (sensu Polsky et al. 2007) to variations<br />

in water supply.<br />

Headwater basins in this study drain into major river<br />

systems that supply water to major agricultural areas and<br />

medium and large cities. Climate change is expected to<br />

increase the variability <strong>of</strong> future streamflow and to stress<br />

municipal water supplies (Milly et al. 2008, Covich 2010,<br />

McDonald et al. 2011, USDOI 2011). Long-term studies<br />

<strong>of</strong> headwater basins can help distinguish biophysical from<br />

social causes <strong>of</strong> variability in water supply and, hence,<br />

the relationships between ecological and social resilience<br />

(Adger 2000). Water scarcity may be perceived even in<br />

areas with abundant rainfall, where politics rather than<br />

true scarcity may govern water restrictions (Hill and Polsky<br />

2006). Meanwhile, residents, pr<strong>of</strong>essional policymakers,<br />

and academics in Phoenix (CAP) implicated population<br />

growth, climate change, and drought as the most important<br />

causes <strong>of</strong> water scarcity, rather than their own wateruse<br />

habits (Larson et al. 2009). Adding to this research,<br />

in this study, we suggest that rather complex interactions<br />

among historical social factors, ecosystem processes, and climate<br />

influence the long-term water supply from headwater<br />

basins.<br />

The role <strong>of</strong> information management<br />

Long-term ecological data are critical to answering societal<br />

questions <strong>of</strong> national concern and significance. Long-term<br />

data are the only way to distinguish trends from short-term<br />

variability in key environmental indicators, such as climate<br />

and streamflow. However, many long-term data remain<br />

inaccessible or difficult to access. Many valuable data sets are<br />

stored in inconvenient file formats with limited metadata.<br />

Variations in methods, variables, units, measurement scales,<br />

and quality-control annotation complicate data integration<br />

and prevent automated approaches to data synthesis.<br />

Until the 1990s, the difficulty <strong>of</strong> identifying, accessing,<br />

and integrating climate and hydrologic data from<br />

the <strong>LTER</strong> Network, EFRs, and related networks precluded<br />

cross-site studies. ClimDB/HydroDB (http://climhy.lternet.<br />

edu), a collaborative effort between <strong>LTER</strong> Network and<br />

USFS information managers, was initiated in 1997 to<br />

overcome these limitations. ClimDB/HydroDB is a Web<br />

harvester and data warehouse that provides uniform access<br />

and visualization <strong>of</strong> daily streamflow and meteorological<br />

data through a single portal. Participating sites manage<br />

original data within their local information systems but<br />

Articles<br />

periodically contribute data to the warehouse. Although<br />

the ClimDB/HydroDB approach is not a complete solution<br />

to data-access and -integration issues, it has served as an<br />

effective bridge technology between older, more rigid data-<br />

distribution models and modern service-oriented architectures<br />

(Henshaw et al. 2006).<br />

The <strong>LTER</strong> Network has made great strides in collecting,<br />

archiving, and integrating long-term data sets online,<br />

enabling synthesis activities such as this one, and providing<br />

an example for other environmental observatories.<br />

Information managers at <strong>LTER</strong> Network sites have led the<br />

development <strong>of</strong> metadata standards, data dictionaries, and<br />

s<strong>of</strong>tware for data integration. The <strong>LTER</strong> Network datamanagement<br />

system serves as a model for emerging national<br />

observatories and existing programs.<br />

Conclusions<br />

This study provides an example <strong>of</strong> the special kinds <strong>of</strong> science<br />

that are possible from networks <strong>of</strong> long-term study<br />

sites. Climate and streamflow records are sufficiently long<br />

that averages, variability, and trends can be meaningfully<br />

analyzed. The climate and streamflow properties are<br />

simple and comparable because they are consistently measured<br />

across sites. Climate and streamflow data are broadly<br />

relevant to ecosystem processes and ecosystem services.<br />

Above all, multisite synthesis is fostered by and contributes<br />

to an open, inclusive culture <strong>of</strong> science collaboration.<br />

This study showed that actual evapotranspiration was<br />

predicted by PET at only 7 <strong>of</strong> 30 sites with 10-year-long<br />

records (the Budyko curve). Taken individually, these departures<br />

might simply reflect the inability <strong>of</strong> a climate station<br />

to represent the conditions <strong>of</strong> a whole basin or the fact<br />

that streamflow depends on groundwater and other forms<br />

<strong>of</strong> storage, as well as precipitation and temperature. But<br />

taken collectively, the departures <strong>of</strong> these sites from the<br />

Budyko prediction suggest the intriguing hypothesis that<br />

water-scarce ecosystems evapotranspire less and waterabundant<br />

ecosystems evapotranspire more than would<br />

be predicted from their climates. Moreover, streamflow<br />

at many <strong>of</strong> these sites was significantly related to one or<br />

more climate index (ENSO, NAO, or PDO), which is not<br />

surprising, but the slightly more frequent significant correlations<br />

<strong>of</strong> streamflow with climate indices in winter than<br />

in summer imply that ecosystem processes mediate climate–<br />

streamflow coupling. Finally, 17 <strong>of</strong> 19 sites had significant<br />

increases in their minimum or maximum daily temperatures<br />

or both, but streamflow trends were directly related<br />

to climate trends at only 7 <strong>of</strong> the sites, all <strong>of</strong> which have<br />

permanent or seasonal ice and snow. In contrast, at other<br />

sites, and during certain seasons at these seven sites, streamflow<br />

trends were contrary to those expected from climate<br />

drivers.<br />

A key finding from this study is that the past and present<br />

human uses <strong>of</strong> ecosystems and human water-use practices<br />

can mimic, exacerbate, counteract, or mask the effects <strong>of</strong><br />

climate change on streamflow. Social factors, including<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 401


Articles<br />

past land management and disturbance and the resulting<br />

ecological succession, all influence trends in streamflow,<br />

even at sites that are considered to be “reference” basins. In<br />

other words, exogenous (climate) factors are not the only<br />

drivers <strong>of</strong> nonstationarity (e.g., Milly et al. 2008) in streamflow<br />

from headwaters: Ecosystem processes and their social<br />

drivers are also important controls. In order to understand<br />

reference conditions for natural flow regimes (e.g., P<strong>of</strong>f<br />

et al. 2007), we need to better understand how ecological<br />

processes and social drivers mediate the expression <strong>of</strong><br />

climate on streamflow. Long-term study sites, where all<br />

these processes are being studied, are ideal places for this<br />

ongoing work.<br />

Acknowledgments<br />

Funding for this work was provided by the US Long Term<br />

Ecological Research (<strong>LTER</strong>) Network and National Science<br />

Foundation grants to participating sites and by a Natural<br />

Sciences and Engineering Research Council <strong>of</strong> Canada<br />

Discovery Grant to IFC. We thank <strong>LTER</strong> Network information<br />

managers for the creation and maintenance <strong>of</strong><br />

ClimDB/HydroDB; the US Forest Service for the initial<br />

establishment and continued support <strong>of</strong> climate and basin<br />

measurements at many <strong>of</strong> the study sites; the US Geological<br />

Survey (USGS) Water, Energy, and Biogeochemical Budgets<br />

program, the USGS Hydrologic Benchmark Network, and<br />

the USGS National Water Information Service for provision<br />

<strong>of</strong> data; and the Networks <strong>of</strong> Centres <strong>of</strong> Excellence–<br />

Sustainable Forest Management Network–funded project on<br />

HydroEcological Landscapes and Processes (HELP) and the<br />

participating Canadian experimental basins from which data<br />

were contributed to the HELP project (Peter Tschaplinski for<br />

CAR, Tom Clair for KEJ, Fred Beall for TLW and MRM,<br />

Ray Hesslein for ELA, Peter Dillon for DOR, and Rita<br />

Winkler for UPC). We thank Merryl Albers, David R. Foster,<br />

Eveleyn Gaiser, Ann Giblin, Stephen P. Loheide II, Randy K.<br />

Kolka, Richard V. Pouyat, Sylvia Schaefer, Emily H. Stanley,<br />

Frederick J. Swanson, Will Wollheim, and three anonymous<br />

reviewers for comments on the manuscript.<br />

References cited<br />

Adams MB, DeWalle DR, Hom JL, eds. 2006. The Fernow Watershed<br />

Acidification Study. Springer.<br />

Adger WN. 2000. Social and ecological resilience: Are they related? Progress<br />

in Human Geography 24: 347–364.<br />

Barnett TP, et al. 2008. Human-induced changes in the hydrology <strong>of</strong> the<br />

western United States. Science 319: 1080–1083.<br />

Barlow M, Nigam S, Berbury EH. 2001. ENSO, Pacific decadal variability,<br />

and U.S. summertime precipitation, drought, and stream flow. Journal<br />

<strong>of</strong> Climate 14: 2105–2128.<br />

Bowden WB, Gooseff MN, Balser A, Green A, Peterson BJ, Bradford J.<br />

2008. Sediment and nutrient delivery from thermokarst features in the<br />

foothills <strong>of</strong> the North Slope, Alaska: Potential impacts on headwater<br />

stream ecosystems. Journal <strong>of</strong> Geophysical Research 113 (Art. G02026).<br />

doi:10.1029/2007JG000470<br />

Budyko MI. 1974. Climate and Life. Academic Press.<br />

Caine N. 2010. Recent hydrologic change in a Colorado alpine basin: An<br />

indicator <strong>of</strong> permafrost thaw? Annals <strong>of</strong> Glaciology 51: 130–134.<br />

Campbell JL, Driscoll CT, Pourmokhtarian A, Hayhoe K. 2011.<br />

Streamflow responses to past and projected future changes in climate<br />

at the Hubbard Brook Experimental Forest, New Hampshire,<br />

United States. Water Resources Research 47 (Art. W02514).<br />

doi:10.1029/2010WR009438<br />

Cayan DR, Redmond KT, Riddle LG. 1999. ENSO and hydrologic extremes<br />

in the western United States. Journal <strong>of</strong> Climate 12: 2881–2893.<br />

Claessens L, Hopkinson C, Rastetter E, Vallino J. 2006. Effect <strong>of</strong> historical<br />

changes in land use and climate on the water budget <strong>of</strong> an urbanizing<br />

watershed. Water Resources Research 42 (Art. W03426).<br />

doi:10.1029/2005WR004131<br />

Clow DW. 2010. Changes in the timing <strong>of</strong> snowmelt and streamflow<br />

in Colorado: A response to recent warming. Journal <strong>of</strong> Climate 23:<br />

2293–2306. doi: 10.1175/2009JCLI2951.1<br />

Covich AP. 2010. Adaptation to sustain high-quality freshwater supplies<br />

in response to climatic change. Resources For the Future. Issue Brief<br />

no. 10-05:1-22.<br />

Donohue RJ, Roderick ML, McVicar TR. 2007. On the importance<br />

<strong>of</strong> including vegetation dynamics in Budyko’s hydrological model.<br />

Hydrology and Earth System Sciences 11: 983–995.<br />

Enfield DB, Mestas-Nuñez AM, Trimble PJ. 2001. The Atlantic multidecadal<br />

oscillation and its relation to rainfall and river flows in the continental<br />

U.S. Geophysical Research Letters 28: 2077–2080.<br />

Foster DR, Aber JD, eds. 2004. Forests in Time: The Environmental<br />

Consequences <strong>of</strong> 1,000 Years <strong>of</strong> Change in New England. Yale <strong>University</strong><br />

Press.<br />

Graham CB, van Verseveld W, Barnard HR, McDonnell JJ. 2010. Estimating<br />

the deep seepage component <strong>of</strong> the hillslope and catchment water<br />

balance within a measurement uncertainty framework. Hydrological<br />

Processes 24: 3631–3647.<br />

Greenland D, Goodin DG, Smith RC. 2003. Climate Variability and<br />

Ecosystem Response at Long-Term Ecological Research Sites. Oxford<br />

<strong>University</strong> Press.<br />

Groisman PY, Knight RW, Karl TR, Easterling DR, Sun B, Lawrimore JH.<br />

2004. Contemporary changes <strong>of</strong> the hydrological cycle over the contiguous<br />

United States: Trends derived from in situ observations. Journal <strong>of</strong><br />

Hydrometeorology 5: 64–85.<br />

Hamon WR. 1963. Computation <strong>of</strong> direct run<strong>of</strong>f amounts from storm rainfall.<br />

International Association <strong>of</strong> Scientific Hydrological Publications<br />

63: 52–62.<br />

Hatcher KL. 2011. Interacting effects <strong>of</strong> climate, forest dynamics, landforms,<br />

and river regulation on streamflow trends since 1950: Examples from<br />

the Willamette basin and forested headwater sites in the US. Master’s<br />

thesis. Oregon State <strong>University</strong>, Corvallis.<br />

Helsel DR, Hirsch RM. 2002. Statistical methods in water resources.<br />

Techniques <strong>of</strong> Water Resources Investigations, Book 4, Chapter A3.<br />

US Geological Survey. (30 January 2012; http://pubs.usgs.gov/twri/<br />

twri4a3)<br />

Henshaw DL, Sheldon WM, Remillard SM, Kotwica K. 2006. ClimDB/<br />

HydroDB: A Web harvester and data warehouse approach to building<br />

a cross-site climate and hydrology database. In: Proceedings <strong>of</strong><br />

the 7th International Conference on Hydroscience and Engineering,<br />

Philadelphia, PA, September 2006. (30 January 2012; http://hdl.handle.<br />

net/1860/1434).<br />

Hill TD, Polsky C. 2006. Adaptation to drought in the context <strong>of</strong> suburban<br />

sprawl and abundant rainfall. Geographical Bulletin 47: 85–100.<br />

Hodgkins GA, Dudley RW, Huntington TG. 2003. Changes in the timing <strong>of</strong><br />

high flows in New England over the 20th century. Journal <strong>of</strong> Hydrology<br />

278: 244–252.<br />

Hornbeck JW, Martin CW, Eagar C. 1997. Summary <strong>of</strong> water yield<br />

experiments at Hubbard Brook Experimental Forest, New Hampshire.<br />

Canadian Journal <strong>of</strong> Forest Research 27: 2043–2052.<br />

Huntington TG. 2006. Evidence for intensification <strong>of</strong> the global water cycle:<br />

Review and synthesis. Journal <strong>of</strong> Hydrology 319: 83–95. doi:10.1016/<br />

j.jhydrol.2005.07.003<br />

Huxman TE, et al. 2004. Convergence across biomes to a common rain-use<br />

efficiency. Nature 429: 651–654.<br />

402 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


Jones JA. 2011. Hydrologic responses to climate change: Considering geographic<br />

context and alternative hypotheses. Hydrological Processes 25:<br />

1996–2000. doi:10.1002/hyp.8004<br />

Jones JA, Post DA. 2004. Seasonal and successional streamflow response<br />

to forest cutting and regrowth in the northwest and eastern United<br />

States. Water Resources Research 40 (Art. W05203). doi:10.1029/<br />

2003WR002952<br />

Karl TR, Melillo JM, Peterson TC, eds. 2009. Global Climate Change<br />

Impacts in the United States: A State <strong>of</strong> Knowledge Report from the U.S.<br />

Global Change Research Program. Cambridge <strong>University</strong> Press.<br />

Knapp AK, et al. 2012. Past, present, and future roles <strong>of</strong> long-term experiments<br />

in the <strong>LTER</strong> network. BioScience 62: 377–389.<br />

Larson KL, White D, Gober P, Harlan S, Wutich A. 2009. Divergent perspectives<br />

on water resource sustainability in a public-policy–science context.<br />

Environmental Science and Policy 12: 1012–1023.<br />

MacDonald GM, Case RA. 2005. Variations in the Pacific Decadal<br />

Oscillation over the past millennium. Geophysical Research Letters 32<br />

(Art. L08703). doi:10.1029/2005GL022478<br />

Magnuson JJ, et al. 2000. Historical trends in lake and river ice cover in the<br />

northern hemisphere. Science 289: 1743–1746.<br />

Mantua NJ, Hare SR, Zhang Y, Wallace JM, Francis RC. 1997. A Pacific interdecadal<br />

climate oscillation with impacts on salmon production. Bulletin<br />

<strong>of</strong> the American Meteorological Society 78: 1069–1079.<br />

McDonald RI, Green P, Balk D, Fekete BM, Revenga C, Todd M,<br />

Montgomery M. 2011. Urban growth, climate change, and freshwater<br />

availability. Proceedings <strong>of</strong> the National Academy <strong>of</strong> Sciences 108:<br />

6312–6317.<br />

Meybeck M. 2003. Global analysis <strong>of</strong> river systems: From Earth system<br />

controls to Anthropocene syndromes. Philosophical Transactions <strong>of</strong> the<br />

Royal Society B 358: 1935–1955. doi:10.1098/rstb.2003.1379<br />

Milly PCD, Betancourt J, Falkenmark M, Hirsch RM, Kundzewicz ZW,<br />

Lettenmaier DP, Stouffer RJ. 2008. Stationarity is dead: Whither water<br />

management? Science 319: 573–574.<br />

Min SK, Zhang X, Zwiers FW, Hegerl GC. 2011. Human contribution to<br />

more-intense precipitation extremes. Nature 470: 378–381.<br />

Molles MC, Dahm CN. 1990. A perspective on El Niño and La Niña:<br />

Global implications for stream ecology. Journal <strong>of</strong> the North American<br />

Benthological Society 9: 68–76.<br />

Moore GM, Bond BJ, Jones JA, Phillips N, Meinzer FC. 2004. Structural<br />

and compositional controls on transpiration in 40- and 450-year-<br />

old riparian forests in western Oregon, USA. Tree Physiology 24:<br />

481–491.<br />

Moore KM. 2010. Trends in streamflow from old growth forested watersheds<br />

in the western Cascades. Student research paper. Oregon State<br />

<strong>University</strong>, Corvallis. (30 January 2012; http://ir.library.oregonstate.edu/<br />

xmlui/handle/1957/15741)<br />

Oishi AC, Oren R, Novick KA, Palmroth S, Ka GG. 2010. Interannual invariability<br />

<strong>of</strong> forest evapotranspiration and its consequence to water flow<br />

downstream. Ecosystems 13: 421–436.<br />

Pall P, Aina T, Stone DA, Stott PA, Nozawa T, Hilberts AGJ, Lohmann D,<br />

Allen MR. 2011. Anthropogenic greenhouse gas contribution to flood<br />

risk in England and Wales in autumn 2000. Nature 470: 382–385.<br />

Peters DPC, Bestelmeyer BT, Herrick JE, Fredrickson EL, Monger HC,<br />

Havstad KM. 2006. Disentangling complex landscapes: New insights<br />

into arid and semiarid system dynamics. BioScience 56: 491–501.<br />

P<strong>of</strong>f NL, Olden LD, Merritt DM, Pepin DM. 2007. Homogenization <strong>of</strong><br />

regional river dynamics by dams and global biodiversity implications.<br />

Proceedings <strong>of</strong> the National Academy <strong>of</strong> Sciences 104: 5732–5737.<br />

Polsky C, Neff R, Yarnal B. 2007. Building comparable global change<br />

vulnerability assessments: The vulnerability scoping diagram. Global<br />

Environmental Change 17: 472–485.<br />

Scatena FN, Moya S, Estrada C, Chinea JD. 1996. The first five years in<br />

the reorganiation <strong>of</strong> aboveground biomass and nutrient use following<br />

Hurricane Hugo in the Bisley experimental watersheds, Luquillo experimental<br />

forest, Puerto Rico. Biotropica 28: 424-440.<br />

Sponseller RA, Grimm NB, Boulton AJ, Sabo JL. 2010. Responses <strong>of</strong> macroinvertebrate<br />

communites to long-term flow variability Sonoran Desert<br />

Articles<br />

stream. Global Change Biology 16: 2891–2900. doi:10.1111/j.1365-<br />

2486.2010.02200.x<br />

Stewart IT, Cayan DR, Dettinger MD. 2005. Changes toward earlier streamflow<br />

timing across western North America. Journal <strong>of</strong> Climate 18:<br />

1136–1155.<br />

Swank WT, Crossley DA Jr, eds. 1988. Forest Hydrology and Ecology at<br />

<strong>Coweeta</strong>. Vol. 66, Ecological Studies. Springer.<br />

Swank WT, Vose JM, Elliott KJ. 2001. Long-term hydrologic and water<br />

quality responses following commercial clearcutting <strong>of</strong> mixed hardwoods<br />

on a southern Appalachian catchment. Forest Ecology and<br />

Management 143: 163–178.<br />

[USDOI] US Department <strong>of</strong> the Interior. 2011. Reclamation: Managing<br />

Water in the West. USDOI.<br />

Verry ES, Brooks KN, Nichols DS, Ferris DR, Sebestyen SD. 2011. Watershed<br />

hydrology. Pages 193–212 in Kolka RK, Sebestyen SD, Verry ES, Brooks<br />

KN, eds. Peatland Biogeochemistry and Watershed Hydrology at the<br />

Marcell Experimental Forest. CRC Press.<br />

Warren RJ, Bradford MA. 2010. Seasonal climate trends, the North Atlantic<br />

Oscillation, and salamander abundance in the Southern Appalachian<br />

Mountain Region. Journal <strong>of</strong> Applied Meteorology and Climatology<br />

49: 1597–1603.<br />

Weisberg PJ, Swanson FJ. 2003. Regional synchroneity in fire regimes <strong>of</strong><br />

western Oregon and Washington, USA. Forest Ecology and Management<br />

172: 17–28.<br />

Julia A. Jones (jonesj@geo.oregonstate.edu) is a pr<strong>of</strong>essor <strong>of</strong> geosciences at<br />

Oregon State <strong>University</strong>, Corvallis, and co-principal investigator (PI) <strong>of</strong> the<br />

Andrews <strong>LTER</strong>L Network site; she studies hydrology and landscape ecology.<br />

Irena F. Creed is a pr<strong>of</strong>essor <strong>of</strong> <strong>biology</strong> and geography at the <strong>University</strong> <strong>of</strong><br />

Western Ontario, London, Ontario, Canada; she conducts biogeochemical<br />

modeling and analysis. Kendra L. Hatcher completed an MS in geography at<br />

Oregon State <strong>University</strong>, Corvallis; she studies physical geography and geographic<br />

information science. Robert J. Warren is a postdoctoral associate at the<br />

Yale School <strong>of</strong> Forestry and Environmental Studies, in New Haven, Connecticut,<br />

and studies organisms and climate variability at the <strong>Coweeta</strong> <strong>LTER</strong> Network<br />

site. Mary Beth Adams is a soil scientist at the US Forest Service (USFS)<br />

Northern Research Station, in Parsons, West Virginia, studying biogeochemistry<br />

at the Fernow Experimental Forest. Melinda H. Benson is an assistant<br />

pr<strong>of</strong>essor <strong>of</strong> geography at the <strong>University</strong> <strong>of</strong> New Mexico, in Albuquerque, studying<br />

ecosystem services and resilience. Emery Boose is an information manager<br />

for Harvard <strong>University</strong>’s Harvard Forest, in Petersham, Massachusetts; he<br />

studies climate and hydrology at the Harvard Forest <strong>LTER</strong> Network site.<br />

Warren A. Brown is a senior research associate at the Cornell Institute for<br />

Social and Economic Research at Cornell <strong>University</strong>, in Ithaca, New York; he<br />

works on demographic change. John L. Campbell is a research ecologist at the<br />

USFS Northeastern Research Station, in Durham, New Hampshire; he studies<br />

climate, hydrology, and biogeochemistry at the Hubbard Brook Experimental<br />

Forest, an <strong>LTER</strong> Network site. Alan Covich is a pr<strong>of</strong>essor at the Odum School<br />

<strong>of</strong> Ecology, at the <strong>University</strong> <strong>of</strong> <strong>Georgia</strong>, in Athens, and co-PI <strong>of</strong> the Luquillo<br />

Experimental Forest, an <strong>LTER</strong> Network site; he studies stream ecology. David<br />

W. Clow is a research hydrologist at the US Geological Survey’s Colorado Water<br />

Science Center, in Lakewood; he studies alpine hydrology and climate at the<br />

Loch Vale <strong>LTER</strong> Network site. Clifford N. Dahm is a pr<strong>of</strong>essor <strong>of</strong> <strong>biology</strong> at the<br />

<strong>University</strong> <strong>of</strong> New Mexico, in Albuquerque; lead scientist <strong>of</strong> the CALFED Bay-<br />

Delta Program; and co-PI <strong>of</strong> the Sevilleta <strong>LTER</strong> Network site; he studies stream<br />

ecology, hydrology, and biogeochemistry. Kelly Elder is a research hydrologist<br />

at the USFS Rocky Mountain Research Station, in Fort Collins, Colorado; he<br />

studies snow hydrology and climate at the Fraser Experimental Forest, part<br />

<strong>of</strong> the USFS Experimental Forests and Ranges program, in Fraser, Colorado.<br />

Chelcy R. Ford is a research scientist with the USFS Southern Research Station,<br />

<strong>Coweeta</strong> Hydrological Laboratory, in Otto, North Carolina; she studies forest<br />

hydrology at the <strong>Coweeta</strong> <strong>LTER</strong> Network site. Nancy B. Grimm is a pr<strong>of</strong>essor<br />

in the School <strong>of</strong> Life Sciences at Arizona State <strong>University</strong>, in Tempe, and co-PI<br />

<strong>of</strong> the Central Arizona–Phoenix <strong>LTER</strong> Network site; she studies stream ecology<br />

and urban systems. Donald L. Henshaw is an information technology specialist<br />

with the USFS Pacific Northwest Research Station in Portland, Oregon; he<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 403


Articles<br />

designed and manages the <strong>LTER</strong> Network’s Climate and Hydrology Database<br />

Projects. Kelli L. Larson is an assistant pr<strong>of</strong>essor in the School <strong>of</strong> Geographic<br />

Science and Urban Planning at Arizona State <strong>University</strong>, in Tempe; she studies<br />

water resource geography and governance. Evan S. Miles completed an MS<br />

in water resources at Oregon State <strong>University</strong>, in Corvallis; he studies hydrogeology<br />

and glaciers. Kathleen M. Miles is a PhD student in geography at<br />

Oregon State <strong>University</strong>, in Corvallis; she studies hydrology and climate change<br />

at the Andrews Experimental Forest, part <strong>of</strong> the <strong>LTER</strong> Network. Stephen D.<br />

Sebestyen is a research hydrologist with the USFS Northern Research Station,<br />

in Grand Rapids, Minnesota; he studies hydrology and biogeochemistry at<br />

Marcell Experimental Forest, in Grand Rapids. Adam T. Spargo is a research<br />

technician in the Department <strong>of</strong> Biology at the <strong>University</strong> <strong>of</strong> Western Ontario,<br />

in London, Ontario, Canada; his role includes working on developing a longterm<br />

experimental catchment database for Canada. Asa B. Stone is an assistant<br />

pr<strong>of</strong>essor at Central New Mexico Community College, in Albuquerque;<br />

she studies ecological psychology. James M. Vose is the project leader at the<br />

USFS Southern Research Station, <strong>Coweeta</strong> Hydrological Laboratory, in Otto,<br />

North Carolina; he studies watershed hydrology, biogeochemistry, and ecology<br />

at the <strong>Coweeta</strong> <strong>LTER</strong> Network site. Mark W. Williams is a pr<strong>of</strong>essor <strong>of</strong> geography<br />

at the <strong>University</strong> <strong>of</strong> Colorado, in Denver; a fellow <strong>of</strong> the Institute <strong>of</strong> Arctic<br />

and Alpine Research; and PI <strong>of</strong> the Niwot Ridge <strong>LTER</strong> Network site; he studies<br />

mountain ecology and alpine hydrology.<br />

404 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


Articles<br />

The Disappearing Cryosphere:<br />

Impacts and Ecosystem Responses<br />

to Rapid Cryosphere Loss<br />

Andrew G. FountAin, John L. CAmpbeLL, edwArd A. G. SChuur, ShAron e. StAmmerJohn,<br />

mArk w. wiLLiAmS, And huGh w. duCkLow<br />

The cryosphere—the portion <strong>of</strong> the Earth’s surface where water is in solid form for at least one month <strong>of</strong> the year—has been shrinking in response<br />

to climate warming. The extents <strong>of</strong> sea ice, snow, and glaciers, for example, have been decreasing. In response, the ecosystems within the cryosphere<br />

and those that depend on the cryosphere have been changing. We identify two principal aspects <strong>of</strong> ecosystem-level responses to cryosphere loss:<br />

(1) trophodynamic alterations resulting from the loss <strong>of</strong> habitat and species loss or replacement and (2) changes in the rates and mechanisms <strong>of</strong><br />

biogeochemical storage and cycling <strong>of</strong> carbon and nutrients, caused by changes in physical forcings or ecological community functioning. These<br />

changes affect biota in positive or negative ways, depending on how they interact with the cryosphere. The important outcome, however, is the<br />

change and the response the human social system (infrastructure, food, water, recreation) will have to that change.<br />

Keywords: cryosphere, ecosystem response, environmental observatories<br />

Global average air temperature has warmed by 1 degree<br />

Celsius (°C) over the past century, and in response, the<br />

cryosphere—the part <strong>of</strong> the Earth’s surface most influenced<br />

by ice and snow—is changing. Specifically, alpine glaciers<br />

are retreating, the expanse <strong>of</strong> Arctic sea ice has been shrinking,<br />

the thickness and duration <strong>of</strong> winter snowpacks are<br />

diminishing, permafrost has been melting, and the ice cover<br />

on lakes and rivers has been appearing later in the year and<br />

melting out earlier. Although these changes are relatively<br />

well documented, the ecological responses and long-term<br />

consequences that they initiate are not. Detailed studies<br />

have identified specific responses to individual components<br />

cryospheric changes (e.g., polar bear habitat and sea ice<br />

loss), but a more integrated view across many landscapes<br />

and types <strong>of</strong> changes has been lacking. In the present article,<br />

we draw largely—but not exclusively—from sites <strong>of</strong> the US<br />

Long Term Ecological Research (<strong>LTER</strong>) Network (the special<br />

section in this issue; see especially Robertson et al. 2012) to<br />

synthesize our current knowledge <strong>of</strong> ecosystem responses<br />

to the changing cryosphere in an attempt to infer broad<br />

responses and to anticipate the further range <strong>of</strong> changes that<br />

we might expect. We contend that place-based, long-term,<br />

interdisciplinary efforts, such as <strong>LTER</strong>-type projects, are<br />

the best suited for tracking such changes and for detecting<br />

and understanding their cascading effects throughout the<br />

ecosystem.<br />

The cryosphere<br />

For the purposes <strong>of</strong> this synthesis, the spatial extent <strong>of</strong> the<br />

cryosphere for the Northern Hemisphere includes the mean<br />

February extent <strong>of</strong> snow cover (measured between 1987<br />

and 2003) and the mean March extent <strong>of</strong> sea ice (measured<br />

between 1979 and 2003). For the Southern Hemisphere, we<br />

include the mean August and September extents <strong>of</strong> snow<br />

and sea ice, respectively. Broad statistics for the cryosphere<br />

and its changes are provided in table 1 and are depicted in<br />

figure 1.<br />

Permafrost (figure 2a) is widespread in the Arctic and<br />

boreal regions <strong>of</strong> the Northern Hemisphere, with the permafrost<br />

zone occupying about 24% <strong>of</strong> the exposed land area.<br />

Most <strong>of</strong> this (78%) occurs in lowlands below 1000 meters<br />

(m) <strong>of</strong> elevation, whereas deposits <strong>of</strong> alpine permafrost<br />

are widely distributed. Changes in permafrost are typically<br />

documented by two metrics: temperature and the depth to<br />

the permafrost, which is defined as the active layer, which<br />

in turn is the surface layer that thaws seasonally. Since the<br />

1980s, permafrost temperatures have generally increased<br />

between 0.5° and 2°C when measured at about 10 m, a<br />

depth at which seasonal variations cancel each other out and<br />

thus yield a seasonally constant value (Romanovsky et al.<br />

2007). At some Russian sites, where many data are available,<br />

the active layer increased by 1.7–5.5 centimeters (cm)<br />

per year over the 10-year period between 1997 and 2007<br />

BioScience 62: 405–415. ISSN 0006-3568, electronic ISSN 1525-3244. © 2012 by American Institute <strong>of</strong> Biological Sciences. All rights reserved. Request<br />

permission to photocopy or reproduce article content at the <strong>University</strong> <strong>of</strong> California Press’s Rights and Permissions Web site at www.ucpressjournals.com/<br />

reprintinfo.asp. doi:10.1525/bio.2012.62.4.11<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 405


Articles<br />

(Mazhitova et al. 2008), whereas other sites have shown little<br />

change (Zamolodchikov et al. 2008). However, recent data<br />

have shown that active-layer depth measurements alone may<br />

obscure the degradation <strong>of</strong> permafrost, because the ground<br />

Table 1. Components <strong>of</strong> the global cryosphere and their areal extent.<br />

Component Definition and remarks<br />

Snow Perennial or seasonal cover <strong>of</strong><br />

the land surface: 98% in Northern<br />

hemisphere<br />

Glaciers Perennial snow or ice that moves<br />

Alpine glaciers and ice caps<br />

surface subsides as permafrost thaws and internal ice melts.<br />

This subsidence process (called thermokarst) can radically<br />

restructure surface hydrology by altering the dynamics<br />

<strong>of</strong> water bodies, initiating or expanding surface channel<br />

incision, and drying surface soil<br />

Extent<br />

(in 10 6 km 2 ) <strong>LTER</strong> site(s) a<br />

1.9 (summer)<br />

45 (winter)<br />

ARC, AND, BNZ, CDR,<br />

KBS, KNZ, HBR, HFR,<br />

MCM, NTL, NWT,<br />

SGS, PIE,<br />

0.53 MCM, NWT, PAL<br />

Ice sheets 14 MCM, PAL<br />

Permafrost Subsurface Earth material remaining<br />

below 0 degrees Celsius for at least<br />

2 years<br />

23 ARC, BNZ, MCM,<br />

NWT<br />

Lake and river ice Seasonal cover <strong>of</strong> lakes and rivers ? ARC, AND, BNZ, CDR,<br />

KBS, KNZ, HBR, HFR,<br />

MCM, NTL, NWT,<br />

SGS, PIE<br />

Sea ice Perennial or seasonal cover <strong>of</strong> the<br />

ocean<br />

19–27 PAL<br />

Note: All extents are from table 4.1 <strong>of</strong> Solomon and colleagues (2007). The value for permafrost is the<br />

region in which permafrost occurs and includes both frozen and unfrozen soils.<br />

km 2 , square kilometers; <strong>LTER</strong>, long-term ecological research.<br />

a See table 2 for the site abbreviations.<br />

a b<br />

layers. Observations near Toolik<br />

Lake, Alaska, have shown rapid<br />

mass wasting <strong>of</strong> surface soils<br />

undergoing thaw, which resulted<br />

in an increased loading <strong>of</strong> suspended<br />

sediment in streams,<br />

with direct and indirect effects<br />

on stream biota (Bowden et al.<br />

2008). In the McMurdo Dry<br />

Valleys <strong>of</strong> Antarctica, enhanced<br />

incision <strong>of</strong> stream water into<br />

massive subsurface ice has caused<br />

one river to flow underground<br />

for some distance. At Niwot Ridge<br />

in Colorado, increasing water<br />

flow and solute concentrations in<br />

early autumn have been attributed<br />

to the melting <strong>of</strong> alpine<br />

permafrost (Caine 2010).<br />

One iconic and highly conspicuous<br />

feature <strong>of</strong> global warming<br />

is glacier recession (figure 2b).<br />

Figure 1. Approximate geographic limits <strong>of</strong> the cryosphere. (a) January climatology <strong>of</strong> Northern Hemisphere sea ice<br />

(measured between 1979 and 2005) and snow extent (measured between 1967 and 2005) with the North Pole referenced<br />

(the red dot). (b) September climatology <strong>of</strong> Southern Hemisphere sea ice (measured between 1979 and 2003) and snow<br />

extent (measured between 1987 and 2002) with the South Pole referenced (the red dot). Source: Reprinted from John<br />

Maurer, Atlas <strong>of</strong> the Cryosphere. National Snow and Ice Data Center (2007; http://nsidc.org/data/atlas).<br />

406 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


Articles<br />

Figure 2. Examples <strong>of</strong> the cryosphere. (a) Winter eastern forest, Mount Washington, New Hampshire (Photograph: Jerry<br />

and Marcy Monkman, www.ecophotography.com); (b) Melting sea ice and an iceberg, Charcot Island, Antarctic Peninsula<br />

(Photograph: Grace K. Saba, Rutgers <strong>University</strong>); (c) Massive ice exposed by degrading permafrost, Noatak National Preserve,<br />

Arkansas (Photograph: Edward A. G. Schuur); (d) Dana Glacier, Sierra Nevada, California. The top panel is the glacier in<br />

1883 (Photograph: I. C. Russell, US Geological Survey); the bottom panel is from 2004 (Photograph: Hassan Basagic).<br />

Glaciers have been receding worldwide since the end <strong>of</strong><br />

the Little Ice Age in the late 1800s, although regional and<br />

temporal variations in recession have occurred (Dyurgerov<br />

and Meier 2000). In recent decades, glacier-mass loss has<br />

accelerated, with the increased rate ascribed to increased<br />

temperatures. Glacier change in the United States reflects<br />

these global trends through area losses over the past century<br />

<strong>of</strong> 34%–56% in the Sierra Nevada and the Cascades <strong>of</strong><br />

Oregon and Washington and about 40% at Niwot Ridge, in<br />

the Colorado Front Range. In contrast to these observations<br />

and to those elsewhere in the alpine Southern Hemisphere,<br />

glaciers in the McMurdo Dry Valleys <strong>of</strong> Antarctica appear<br />

to be in equilibrium, since their positions have not changed<br />

since their observation began in 1993 (Fountain et al. 2006).<br />

The removal <strong>of</strong> water from long-term storage in glacial ice<br />

increases summer streamflow and global sea level. However,<br />

as the mass <strong>of</strong> glaciers decline, their ability to support summer<br />

streamflow declines, and they decrease in their ability to<br />

buffer the watersheds against drought.<br />

Sea ice (figure 2c) occurs in the Arctic; in the Southern<br />

Ocean surrounding Antarctica and the Baltic Sea; and in<br />

part <strong>of</strong> the northwest Pacific, from the Siberian coast down<br />

to the Japanese island <strong>of</strong> Hokkaido. Most sea ice forms<br />

and melts annually, with perennial multiyear ice restricted<br />

to the high latitudes <strong>of</strong> the Arctic and Antarctica. Since<br />

the advent <strong>of</strong> continuous satellite monitoring in the late<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 407


Articles<br />

1970s, widespread decreases in sea ice have been recorded<br />

throughout most <strong>of</strong> the Arctic at an average rate <strong>of</strong> 3% loss<br />

per decade (Comiso and Nishio 2008). In contrast, decreases<br />

in Antarctic sea ice have been regionally confined and juxtaposed<br />

against regions <strong>of</strong> increasing sea ice, such that the<br />

average rate <strong>of</strong> change overall is a slight increase <strong>of</strong> 1% per<br />

decade. Changes in sea ice alter the extent and distribution<br />

<strong>of</strong> foraging platforms for larger mammals and refuge habitat<br />

for smaller species. At the Palmer Peninsula, seasonal seaice<br />

cover has been shrinking at astonishing rates because <strong>of</strong><br />

increases in onshore winds driven by hemispheric changes<br />

in atmospheric circulation. The duration <strong>of</strong> sea-ice cover has<br />

declined by 85 days since 1978 (Stammerjohn et al. 2008).<br />

Seasonal lake and river ice occur in all temperate regions,<br />

with durations <strong>of</strong> days to months, whereas perennial<br />

ice cover is only found at extremely high latitudes and<br />

elevations. The date <strong>of</strong> lake-ice formation and breakup is<br />

commonly recorded for commercial purposes related to<br />

shipping, trapping, fishing, ice harvesting, and transportation<br />

and yields an extensive long-term record (Magnuson<br />

et al. 2000). Since 1846, lake-ice duration in the Northern<br />

Hemisphere has decreased by 12 days per century, which<br />

is equivalent to a warming <strong>of</strong> 1.2°C per century. A 20-year<br />

record <strong>of</strong> ice thickness in late March on an alpine lake in<br />

the Niwot Ridge <strong>LTER</strong> site shows a consistent thinning<br />

<strong>of</strong> the ice cover at 2.0 cm per year (Caine 2002). Ice cover<br />

exhibits strong control over exchanges in gases and material,<br />

solar radiation, and heat between aquatic habitats and<br />

the atmosphere. The duration <strong>of</strong> ice exerts a pr<strong>of</strong>ound<br />

influence on the patterns <strong>of</strong> water circulation and thermal<br />

stratification, which are closely linked to the life cycles <strong>of</strong><br />

aquatic organisms and to the biogeochemical cycling <strong>of</strong> the<br />

ecosystem.<br />

Snow (figure 2d) is the largest component <strong>of</strong> the cryosphere<br />

in areal extent. About 98% <strong>of</strong> the snow-covered land<br />

on Earth is in the Northern Hemisphere, which contains<br />

nearly half <strong>of</strong> the planet’s land surface. In the Southern<br />

Hemisphere, over 99% <strong>of</strong> the snow cover is confined to<br />

Antarctica and largely consists <strong>of</strong> perennial snow. In the<br />

Northern Hemisphere, strong negative trends in the extent<br />

<strong>of</strong> snow cover have been observed over recent decades (Déry<br />

and Brown 2007). Increased snowfall and snow depth have<br />

been reported at the highest-elevation sites <strong>of</strong> the western<br />

United States (Williams et al. 1996); however, most locations<br />

in the Mountain West have experienced snowpack<br />

declines, and concern has risen about streamflow, water<br />

yields, and water supply. In the Pacific Northwest, extensive<br />

snow- covered regions are now deemed at risk in terms <strong>of</strong><br />

their capacity to provide reliable water yields because <strong>of</strong><br />

atmospheric warming, altitudinal shifts in the distribution<br />

<strong>of</strong> snow and rain, and declining winter snowpacks (Nolin<br />

and Daly 2006). Winter snow depths have also been decreasing<br />

throughout the northeastern United States. For example,<br />

at the Hubbard Brook <strong>LTER</strong> site in New Hampshire,<br />

the maximum snow depth has declined by 25 cm (7 cm<br />

water equivalent), and snow cover duration has decreased<br />

by 21 days over the past 53 years (Campbell et al. 2010),<br />

which has led to major changes in terrestrial and aquatic<br />

ecosystems.<br />

One simple metric in the attempt to capture potential<br />

ecosystem vulnerability to changes in the cryosphere across<br />

ecosystems is the duration <strong>of</strong> frost and freezing temperatures<br />

(table 2, figure 3). Frost days are those with long-term mean<br />

daily minimum temperatures below 0°C; freeze days are<br />

those with long-term mean daily maximum temperatures<br />

below 0°C. Vulnerability can be thought <strong>of</strong> as susceptible<br />

to increased or decreased frost or freezing periods. For<br />

example, ecosystems that do not experience frost, such<br />

as those in the tropics, are highly vulnerable to cold temperatures.<br />

Significant ecosystem changes could be expected<br />

if the climate were to cool, making frost commonplace.<br />

Alternatively, ecosystems accustomed to long frozen periods,<br />

such as polar and high alpine ecosystems, are highly vulnerable<br />

to warm temperatures. Those ecosystems exposed to<br />

moderate periods <strong>of</strong> frost or freezing would be expected to<br />

be less vulnerable to changes in temperature. We focus on<br />

the warming climate, and as such, the tropical ecosystems<br />

will not be directly exposed to cryospheric losses, whereas<br />

polar and high alpine ecosystems may be the most vulnerable<br />

to such change. In table 2 and figure 3, we can see the<br />

vulnerability <strong>of</strong> the major ecosystem research sites under<br />

study by US scientists.<br />

Ecosystem responses to the loss <strong>of</strong> snow and ice<br />

The various components <strong>of</strong> the cryosphere provide physical<br />

habitat for diverse organisms. Iconic examples include<br />

polar bears and penguins in sea ice and pikas in rock glaciers<br />

(rock debris with ice filling the void spaces between<br />

the rocks; the mass flows downhill), but many other species<br />

ranging in size from microbes to whales inhabit permafrost,<br />

glaciers, sea-ice, and snow-covered landscapes. As these<br />

habitats shrink and disappear, resident species are forced to<br />

migrate, <strong>of</strong>ten tracking the distribution <strong>of</strong> receding frozen<br />

habitats across the landscape. Since different organisms<br />

respond and move at different rates (e.g., trees versus penguins),<br />

cryosphere recession can have many consequences:<br />

the fragmentation <strong>of</strong> animal and plant communities and<br />

the development <strong>of</strong> new assemblages, disruption <strong>of</strong> seasonally<br />

synchronized phenological connections among species,<br />

and losses in biodiversity and the associated changes<br />

in ecosystem function (Parmesan 2006). Although these<br />

processes are occurring at unprecedented rates in response<br />

to rapid climate warming, it has required decades <strong>of</strong> coordinated<br />

observations to document significant change and<br />

to uncover the mechanisms linking climate forcing to ecosystem<br />

responses.<br />

Prolonged, systematic studies <strong>of</strong> this type are a key contribution<br />

<strong>of</strong> <strong>LTER</strong>. The <strong>LTER</strong> Network <strong>of</strong> sites facilitates longterm<br />

observations, experiments, and comparative studies<br />

that enable us to identify common processes and mechanisms<br />

across diverse ecosystems. The highly interdisciplinary<br />

nature <strong>of</strong> <strong>LTER</strong> helps to quickly reveal interpretations <strong>of</strong><br />

408 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


Table 2. Cryosphere processes at US Long Term Ecological Research (<strong>LTER</strong>) Network sites and related <strong>LTER</strong> sites.<br />

Site Abbreviation Record<br />

Continental<br />

glacier<br />

causes and consequences and needed adjustments <strong>of</strong> monitoring<br />

approaches to catch signals <strong>of</strong> previously unmonitored<br />

or unanticipated system behaviors. Here, we present<br />

some notable examples <strong>of</strong> ecological and biogeochemical<br />

changes in response to cryosphere loss.<br />

Changes in populations and trophodynamic implications. Decadalscale<br />

declines or distributional shifts in snow- and icedependent<br />

species are now extensive and well documented<br />

(Chapin et al. 2005). When ice-dependent species suffer<br />

habitat loss, the changes in frozen habitats (glaciers, sea<br />

ice, snowpacks, and permafrost) impose both bottom-up<br />

and top-down forcings on terrestrial and aquatic ecosystems.<br />

Ice loss affects ecosystems directly through the loss<br />

<strong>of</strong> physical habitat and through alterations in thermal<br />

conditions and indirectly by altering light and nutrient<br />

supply to primary producers. Both Arctic and Antarctic<br />

sea ice harbor a resident microbial community <strong>of</strong> diatoms,<br />

other phytoplankton, bacteria, and protozoan grazers that<br />

contributes to the total primary production <strong>of</strong> polar seas.<br />

Like sea ice, ice and rock glaciers are habitats for specially<br />

adapted species that may disappear as glaciers retreat and<br />

their cold, glacier-fed streams disappear. The American pika<br />

Alpine<br />

glacier Sea ice Lake ice<br />

Continuous<br />

permafrost<br />

Discontinuous<br />

permafrost<br />

Seasonal<br />

snow<br />

McMurdo MCM 1 1988–2009 X X X X<br />

Arctic Tundra ARC 1 1988–2008 X X X<br />

Niwot Ridge NWT 1 1952–2006 X X X X<br />

Bonanza Creek BNZ 1 1988–2009 X X<br />

Palmer PAL 1 1989–2010 X X X X<br />

Loch Vale LVW 2 1993–2008 X X X X<br />

Marcell MAR 1961–2010 X<br />

North Temperate<br />

Lakes<br />

NTL 1 1978–2010 X X<br />

Hubbard Brook HBR 1,3 1964–2007 X<br />

Harvard Forest HFR 1 1964–2002 X<br />

Kellogg KBS 1 1988–2010 X<br />

Shortgrass Steppe SGS 1 1969–2010 X<br />

Sevilleta SEV 1,4 1991–2010 X<br />

Articles<br />

Fernow FER 3 1951–2007 X<br />

Konza KNZ 1 1982–2011 X<br />

Plum Island PIE 1 1950–2004 X X<br />

<strong>Coweeta</strong> CWT 1,3 1950–2009 X<br />

Baltimore BES 1 2000–2009 X<br />

Andrews Forest AND 1,3 1957–2007 X X<br />

Jornada JRN 1,3 1983–2009 X<br />

Olympic OLY 5 1962–2009 X<br />

Transient<br />

snow<br />

Note: The superscript number next to the abbreviation refers to the sponsoring agency for that site: 1, US <strong>LTER</strong> Network site; 2, US Geological Survey<br />

Water, Energy, and Biochemical Budgets site; 3, US Department <strong>of</strong> Agriculture Experimental Forest and Range site; 4, National Wildlife Refuge; 5, State<br />

Experimental Forest. Record refers to the length <strong>of</strong> the air-temperature record used to estimate frost and freezing duration at each location.<br />

(Ochotona princeps)—although it is not as well known or<br />

charismatic as penguins or polar bears—is attaining new<br />

status as a poster child for glacier loss and climate change.<br />

Pikas do not hibernate and use subsurface microclimates<br />

in rocky debris to persist where surface conditions would<br />

preclude their survival. Despite this adaptation, some local<br />

pika extinctions in the Northwest have been linked to cold<br />

exposure caused by a loss <strong>of</strong> insulating snow cover (Ray et al.<br />

2012). Permafrost thaw also results in habitat disappearance<br />

for its resident species. Because permafrost occurs in so<br />

many different habitats in different stages <strong>of</strong> development,<br />

its loss may trigger primary or secondary successions.<br />

Widespread past and projected future reductions in<br />

snow-cover extent, duration, depth, and water equivalent<br />

can also have extensive ecological repercussions. Many plant<br />

and animal species are adapted to snow-cover conditions<br />

and will perish if they are unable to migrate or tolerate<br />

less snow cover. Even so, not all animals that live in cold<br />

environments respond negatively to reductions in snow<br />

cover. For example, ungulates such as white-tailed deer,<br />

mule deer, elk, and caribou expend less energy and are less<br />

susceptible to predation when snowpacks are shallower.<br />

Some <strong>of</strong> the species most susceptible to snow-cover loss<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 409


Articles<br />

Figure 3. Duration <strong>of</strong> frost and freezing periods at <strong>LTER</strong> and related long-term research<br />

sites, from climate records. See table 2 for the site abbreviations. Frost days are those<br />

with long-term mean daily minimum temperatures below 0 degrees Celsius (°C); freeze<br />

days are those with long-term mean daily maximum temperatures below 0°C.<br />

are those that overwinter below ground, since snow insulates<br />

the subsurface and moderates its temperature. The<br />

shortgrass steppe in the western United States receives little<br />

snowfall and therefore presents an endpoint in the spectrum<br />

<strong>of</strong> snow cover. At this semiarid, high-elevation site,<br />

snowfall from November to late February has little effect<br />

on ecological processes; however, large snowfalls in March<br />

and April (after the ground has thawed) strongly influence<br />

the subsequent productivity by controlling the availability<br />

<strong>of</strong> water and nutrients (Cayan et al. 2001). Some plants<br />

are photosynthetically active in shallow spring snowpacks,<br />

giving them a competitive advantage in regions with short<br />

growing seasons (Starr and Oberbauer 2003). As the climate<br />

warms, the disappearance <strong>of</strong> snow cover and the increased<br />

length <strong>of</strong> the growing season may benefit some plants,<br />

provided that other requirements for growth are not limiting.<br />

Many plants in alpine and tundra regions are reliant<br />

on snow for water and nutrients and therefore are found<br />

in the greatest abundance where the range <strong>of</strong> snow depths<br />

is optimal (Walker et al. 1993). Although the snow-free<br />

period will lengthen in a warmer climate, the lack <strong>of</strong> snow<br />

cover during colder months will increase soil temperature<br />

variation, making roots susceptible to winter injury. Soil<br />

freezing can directly and adversely affect roots by causing<br />

cellular damage and can also sever fine roots through frost<br />

heaving. Reduced nutrient uptake as a result <strong>of</strong> root injury<br />

has been shown to lower nutrient retention and to increase<br />

hydrologic fluxes from soils during the growing season<br />

(Fitzhugh et al. 2001).<br />

Changes in habitat and productivity<br />

regimes can ripple up<br />

the trophic ladder, as is demonstrated<br />

extensively in marine<br />

food webs. Changing snow<br />

and ice conditions alter habitat<br />

suitability for many bird species<br />

(e.g., petrels, Adelie penguins<br />

[Pygoscelis adeliae]) and<br />

limit the physical space available<br />

for habitation (Micol and<br />

Jouventin 2001, Weimerskirch<br />

et al. 2003). The huge populations<br />

<strong>of</strong> krill in Antarctic marginal<br />

sea-ice zones serve in turn<br />

as a major food resource for a<br />

suite <strong>of</strong> large predators, including<br />

seabirds, seals, and whales.<br />

Sea-ice microbial communities<br />

serve as a principal food source<br />

for juvenile krill, which also<br />

hide from predators in underice<br />

cryptic spaces. Therefore, the<br />

regional decline in the duration<br />

and extent <strong>of</strong> sea-ice cover in the<br />

Bellingshausen and Amundsen<br />

Seas has resulted in declining<br />

abundance and ranges <strong>of</strong> Antarctic krill (Euphausia superba),<br />

possibly the most numerous metazoan species on Earth.<br />

Atkinson and colleagues (2004) documented large-scale,<br />

order-<strong>of</strong>- magnitude declines in krill populations over 50 years<br />

in the South Atlantic sector <strong>of</strong> the Antarctic seas. Meanwhile,<br />

the number <strong>of</strong> salps—pelagic, gelatinous, ice-avoiding tunicates<br />

with few predators—has increased; they have, in effect,<br />

replaced krill as an intermediate species in Antarctic marine<br />

food chains. One <strong>of</strong> the best-studied examples <strong>of</strong> the response<br />

<strong>of</strong> predator populations to sea-ice loss is the Adelie penguin,<br />

a true Antarctic penguin with strong fidelity to sea ice as a<br />

platform for foraging activity (Ducklow et al. 2007). Since<br />

1975, Adelie penguins nesting near Palmer Station have<br />

declined by about 80% in response to a host <strong>of</strong> environmental<br />

changes, including habitat loss and altered food availability<br />

(figure 4). Fraser and H<strong>of</strong>mann (2003) demonstrated that<br />

penguin chicks weighing less than 300 grams at fledging<br />

had a reduced probability <strong>of</strong> surviving past the first year.<br />

They suggested that changes in the sea-ice season shifted the<br />

period <strong>of</strong> maximum krill stocks away from the penguins’<br />

peak foraging season. Over the same period, two subpolar<br />

species—chinstrap (Pygoscelis antarcticus) and gentoo penguins<br />

(Pygoscelis papua)—have successfully immigrated to the<br />

region and now constitute half <strong>of</strong> the total penguin population<br />

in the region. The mechanisms behind these shifts and<br />

their long-term outcome are unclear (Trivelpiece et al. 2011).<br />

The recent loss <strong>of</strong> sea ice could boost primary productivity in<br />

the Arctic Ocean by a factor <strong>of</strong> two or three. In the northern<br />

Bering Sea, primary-productivity changes caused by warming<br />

410 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


Year<br />

Figure 4. Populations <strong>of</strong> ice-dependent (Adelie) and icetolerant<br />

(chinstrap and gentoo) penguins near Palmer<br />

Station, Antarctica, measured between 1976 and 2009.<br />

Source: Adapted from Ducklow and colleagues (2007).<br />

and sea-ice loss have resulted in a dramatic reorganization <strong>of</strong><br />

the ecosystem (Grebmeier et al. 2006). This shallow marine<br />

ecosystem was formerly characterized by high primary productivity<br />

and efficient export to the bottom, which supports a<br />

high stock <strong>of</strong> benthic prey for diving ducks and walruses. With<br />

the loss <strong>of</strong> sea ice and the warming <strong>of</strong> the water column, the<br />

export <strong>of</strong> surface productivity into the benthos has declined,<br />

which has caused a switch from a system with top predators<br />

sustained by benthic prey to one dominated by pelagic fish.<br />

Changes in biogeochemical cycles. Changes in the extent, seasonality,<br />

and duration <strong>of</strong> cryosphere components affect the<br />

cycling <strong>of</strong> nutrients in land and ocean ecosystems. Glacier<br />

retreat and rock glacier shrinkage expose new landscapes<br />

that are typically carbon poor yet nutrient rich because<br />

<strong>of</strong> rock weathering. Microbial life—particularly nitrogen<br />

fixers—occupy these new landscapes (Nemergut et al. 2007),<br />

which increases the nitrate levels <strong>of</strong> streams and lakes down<br />

valley. These conditions are transient and slowly change<br />

as higher plants occupy the landscape over time scale <strong>of</strong><br />

decades to centuries. High alpine waters are typically oligotrophic<br />

and are therefore susceptible to ecological changes<br />

that result from increases in nitrogen export from the land<br />

(Baron et al. 2009). Williams and colleagues (2007) characterized<br />

the nutrient content in the outflow <strong>of</strong> the Green<br />

Lake 5 rock glacier, located in the Green Lakes Valley <strong>of</strong> the<br />

Colorado Front Range. The nitrate concentrations from<br />

the rock glacier are among the highest reported for highelevation<br />

surface waters. These extreme nitrate concentrations<br />

appear to be characteristic <strong>of</strong> rock glacier outflows in<br />

the Rocky Mountains (Williams et al. 2007). Fluorescence<br />

index values and dissolved organic matter (DOM) measurements<br />

are consistent with a switch from terrestrial DOM in<br />

the summer to an increasingly aquatic-like microbial source<br />

during the autumn months. Glacier melting has also been<br />

implicated in the regulation <strong>of</strong> phytoplankton species composition<br />

in Antarctic coastal regions where diatoms—the<br />

Articles<br />

preferred food for Antarctic krill (see above)—are being<br />

replaced by less-palatable cryptophytes. Glacial inputs<br />

change light availability by stabilizing the surface-water<br />

column and possibly stimulate growth selectively by adding<br />

limiting micronutrients (Dierssen et al. 2002). Melting<br />

glaciers and sea ice also transfer airborne pollutants stored<br />

in the snow and ice to the marine environment.<br />

Perhaps the most important result from the reduction in<br />

duration <strong>of</strong> lake ice in a warming climate is less-frequent<br />

oxygen-depletion events and the associated adverse biological<br />

consequences (Prowse et al. 2006). For river ice, large<br />

fluxes <strong>of</strong> allocthonous detrital material and nutrients are<br />

flushed into the river water column because <strong>of</strong> channel scour<br />

during ice breakup and flooding. Geomorphically, at the<br />

Pine Island <strong>LTER</strong> site in coastal Massachusetts, the formation<br />

and transport <strong>of</strong> river ice are important factors in determining<br />

salt marsh platform elevation and have implications<br />

for responses to rising sea level. The delivery <strong>of</strong> sediment to<br />

the marsh through ice rafting (Wood et al. 1989), the compression<br />

<strong>of</strong> the marsh surface as a function <strong>of</strong> ice thickness,<br />

and the scour <strong>of</strong> vegetation are winter processes that will<br />

change as less river ice forms and its transport into fringing<br />

salt marshes declines in the coming decades.<br />

In cold regions, nutrient cycling is closely coupled with<br />

snowpack dynamics, with much <strong>of</strong> the annual export <strong>of</strong><br />

stream nutrients occurring in winter, when biological uptake<br />

is low. Changes in the snowpack alter hydrology, which<br />

affects the amount, timing, and magnitude <strong>of</strong> spring snowmelt.<br />

The resulting changes in streamflow not only affect<br />

nutrient transport but, when they are combined with<br />

changes in temperature, also affect aquatic habitats, causing<br />

potential shifts in species assemblages. Nutrients accumulate<br />

in the snowpack over winter and are released in an ionic<br />

pulse during the first portion <strong>of</strong> snowmelt (Johannessen and<br />

Henriksen 1978). Although snowmelt can be an important<br />

source <strong>of</strong> nutrients and water early in the growing season,<br />

it can also cause episodic acidification in areas with high<br />

atmospheric deposition and poorly buffered soils (e.g.,<br />

Schaefer et al. 1990). The soils beneath the snowpack are also<br />

an important source <strong>of</strong> nutrients during winter and early<br />

spring. The snowpack regulates soil temperatures, keeping<br />

them warm enough for many biologically mediated reactions.<br />

Snow fence experiments, which enhance winter snow<br />

accumulation, have shown that higher rates <strong>of</strong> soil microbial<br />

respiration and nitrogen mineralization occur under deeper<br />

snowpacks in subalpine forest and Arctic tundra environments<br />

because <strong>of</strong> warmer soil temperatures that result from<br />

the thermally insulating effects <strong>of</strong> the snow (Schimel et al.<br />

2004). In contrast, in boreal spruce and temperate hardwood<br />

forests, thin winter snowpacks increase the frequency and<br />

depth <strong>of</strong> soil freezing, which results in elevated summer<br />

nitrogen emissions that are probably a result <strong>of</strong> reduced<br />

nitrate uptake by damaged roots and by root decomposition<br />

(Fitzhugh et al. 2001, Maljanen et al. 2010). Fluxes <strong>of</strong> carbon<br />

dioxide (CO 2 ) mirror those <strong>of</strong> nitrogen, and the timing<br />

and magnitude <strong>of</strong> the nitrogen and CO 2 fluxes in all cases<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 411


Articles<br />

depends on plant species. Elevated nitrogen mineralization<br />

contributes significantly to fluxes <strong>of</strong> greenhouse gases and<br />

to the cycling <strong>of</strong> nitrogen and carbon. In the mixed-grass<br />

prairie <strong>of</strong> North America, greater snowpack increased soil<br />

moisture by midsummer, which resulted in increased soil<br />

respiration (Chimner and Welker 2005) and increased plant<br />

invasions (Blumenthal et al. 2008).<br />

But most <strong>of</strong> the attention to the biogeochemical consequences<br />

<strong>of</strong> cryosphere loss has been focused on the potential<br />

for changes in carbon storage and release. Arctic permafrost<br />

contains twice the CO 2 found in the atmosphere, which<br />

dramatically demonstrates the potential for altering the<br />

climate as further warming occurs. Site-specific information<br />

can provide some indication as to the future release<br />

rate <strong>of</strong> carbon from thawing permafrost. A recent group<br />

<strong>of</strong> studies was focused on an upland thermokarst site near<br />

Denali National Park in Alaska, where changes in plant and<br />

soil processes were studied as a function <strong>of</strong> time since the<br />

thermokarst disturbance was initiated. The studies showed<br />

that increased permafrost thaw and ground-surface subsidence<br />

increased net and gross primary productivity as<br />

plant growth was stimulated by a thaw (Vogel et al. 2010).<br />

Plant species composition changed along with changes in<br />

plant growth rates as graminoid-dominated moist acidic<br />

tundra shifted to shrub-dominated tundra with increased<br />

rates <strong>of</strong> thawing. The increased carbon uptake by plants<br />

initially <strong>of</strong>fset the greater ecosystem respiration, such that<br />

this thermokarst was a net sink <strong>of</strong> carbon 15 years after the<br />

initiation <strong>of</strong> the thaw, even though decomposition <strong>of</strong> older<br />

carbon deep in the soil was already taking place (Schuur<br />

et al. 2009). Over more decades <strong>of</strong> thaw, plant growth rates<br />

remained high, but increased old soil carbon losses eventually<br />

<strong>of</strong>fset the greater carbon uptake, and this thermokarst<br />

became a net source <strong>of</strong> carbon to the atmosphere.<br />

In a contrasting study <strong>of</strong> lowland thermokarst in three<br />

Canadian peatlands, the carbon accumulation in surface<br />

soil organic matter was higher in unfrozen bogs and in<br />

areas where permafrost had degraded than in areas where<br />

permafrost was intact (Turetsky et al. 2007). This growth<br />

in surface soil carbon accumulation was consistent with<br />

the Alaskan upland study, but the equivalent net ecosystem<br />

carbon exchange measurements were not available to<br />

determine whether the thawed permafrost peat ecosystems<br />

were overall net sources or sinks <strong>of</strong> carbon. Permafrost<br />

thaw in this lowland system promoted the release <strong>of</strong> methane<br />

(CH 4 ) because waterlogged conditions predominated<br />

in Sphagnum moss lawns that replaced the feather moss<br />

(Hylocomium splendens) and black spruce (Picea mariana)<br />

forest in locations where permafrost degraded. This CH 4<br />

release was hypothesized to potentially <strong>of</strong>fset the observed<br />

surface soil carbon accumulation for at least for 70 years,<br />

until plant and ecosystem succession in the moss lawn created<br />

conditions more like those in the unfrozen bogs, which<br />

stored surface soil carbon but released only small amounts<br />

<strong>of</strong> CH 4 . The release <strong>of</strong> CH 4 is a common pathway <strong>of</strong> carbon<br />

loss in lowland thermokarst, where drainage is restricted<br />

( Myers-Smith et al. 2007), and CH 4 has 25 times greater heattrapping<br />

capacity than CO 2 on a century timescale. However,<br />

decreased total carbon emissions in anaerobic systems can<br />

partially <strong>of</strong>fset the increased radiative forcing <strong>of</strong> CH 4 release,<br />

which possibly makes the net radiative forcing <strong>of</strong> increased<br />

carbon losses in lowland and upland thermokarst more<br />

similar than what the difference in heat-trapping capacity<br />

between CO 2 and CH 4 would initially suggest.<br />

The oceanic sink for anthropogenic CO 2 is large and<br />

lessens the potential greenhouse effect by limiting CO 2<br />

accumulation in the atmosphere. The current (2009) net<br />

annual carbon uptake by the ocean is 2.3 ± 0.4 petagrams<br />

(Pg) <strong>of</strong> carbon per year, compared to 2.4 Pg <strong>of</strong> carbon<br />

per year for land, but the land uptake is partially <strong>of</strong>fset by<br />

the 1.1 ± 0.7 Pg <strong>of</strong> carbon per year in releases caused by<br />

deforestation (Le Quéré et al. 2009). As the ocean warms and<br />

its inventory <strong>of</strong> CO 2 increases, the oceanic sink is expected<br />

to weaken. Oceanic CO 2 uptake is governed by gas exchange<br />

across the air–sea interface, so the regional allocation <strong>of</strong><br />

CO 2 uptake is primarily a function <strong>of</strong> the area <strong>of</strong> sea surface<br />

involved. This fraction has been predicted to increase as ice<br />

melts and productivity increases, which will expose new<br />

ocean areas to solar irradiance (Peck et al. 2010). The Arctic<br />

Ocean constitutes just 3% <strong>of</strong> the total ocean area and is<br />

mostly covered by sea ice, which blocks gas exchange, but it<br />

accounts for 5%–14% <strong>of</strong> the total ocean CO 2 uptake. New<br />

observations suggest, however, that the recent dramatic loss<br />

<strong>of</strong> sea ice has been accompanied by decreased rather than<br />

increased CO 2 uptake (Cai et al. 2010), which is counter to<br />

current understanding and predictions. The rapid, diverse,<br />

and complex changes wrought by cryosphere loss constitute<br />

a major scientific challenge that demands new large-scale<br />

observing systems on land and in the ocean to provide new<br />

observational infrastructure as a resource for coordinated<br />

experimental studies performed by the <strong>LTER</strong> Network and<br />

other scientists.<br />

Effects on humans from the loss <strong>of</strong> snow and ice<br />

Cryosphere loss will result in far-reaching social, economic,<br />

and geopolitical impacts, but a detailed treatment<br />

is beyond the scope <strong>of</strong> this synthesis. Most attention has<br />

been devoted to the impacts associated with a loss <strong>of</strong> snow<br />

cover, glacier melting, and sea-level rise, which are treated<br />

elsewhere (Kundzewicz et al. 2007). Thawing permafrost<br />

will also have important social consequences, because it can<br />

destabilize engineered structures and can cause destructive<br />

slides, flows, and slumps. Changes in snow cover can<br />

have important consequences for humans and may affect<br />

many diverse activities, including agriculture, recreation,<br />

tourism, engineering, commerce, and energy production.<br />

For example, the New Hampshire ski industry has abandoned<br />

low-elevation ski areas in the southern part <strong>of</strong> the<br />

state since the 1970s, in part because <strong>of</strong> climate warming,<br />

in favor <strong>of</strong> ski areas at higher elevations in the north<br />

(Hamilton et al. 2003). Skiing contributes about $1 billion<br />

annually to the economy <strong>of</strong> Utah, but recent climate change<br />

412 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


evaluations <strong>of</strong> the ski industry there suggest that it is at risk<br />

in the next several decades (Lazar and Williams 2008). A<br />

similar impact is anticipated for the Pacific Northwest. The<br />

most important effect is the influence on streamflow. In<br />

many semiarid regions <strong>of</strong> the world, such as the southwestern<br />

United States, snowmelt from mountain snowpacks is<br />

the dominant source <strong>of</strong> water for human consumption and<br />

irrigation. Therefore, changes in the amount and timing <strong>of</strong><br />

snowmelt in mountainous areas could affect stream ecosystem<br />

services, such as drinking- water supply, wastewater<br />

assimilation, and hydropower. Lesser amounts <strong>of</strong> snow could<br />

also have an impact on agriculture and the ability to produce<br />

food, both through an increased occurrence <strong>of</strong> drought<br />

and through an inadequate supply <strong>of</strong> water for irrigation.<br />

Some evidence suggests that changes in snowmelt may also<br />

increase the risk <strong>of</strong> forest fires. In the western United States,<br />

earlier snowmelt dates correspond to increased wildfire frequency,<br />

because soils and vegetation are becoming drier and<br />

the period <strong>of</strong> potential ignition is lengthening (Westerling<br />

et al. 2006). Estimates <strong>of</strong> sea-level rise to 2100 have been<br />

continually revised upward since the 2007 report <strong>of</strong> the<br />

Intergovernmental Panel on Climate Change (Solomon<br />

et al. 2007) as new data and modeling have been developed.<br />

At the time <strong>of</strong> this writing, the rise in sea level by the end <strong>of</strong><br />

the century is projected to be about 1 m (Pfeffer et al. 2008).<br />

The economic cost <strong>of</strong> a 1-m rise in sea level is estimated to<br />

exceed $1 trillion (Anth<strong>of</strong>f et al. 2010), with enormous social<br />

and political dislocations as residents <strong>of</strong> low-lying regions<br />

are forced to move to higher ground.<br />

The Arctic has emerged as a key laboratory for the<br />

study <strong>of</strong> climate change impacts on human communities,<br />

partly because it is host to the world’s largest indigenous<br />

population that maintains a subsistence lifestyle (K<strong>of</strong>inas<br />

et al. 2010) and partly because <strong>of</strong> the rapidly manifesting<br />

impacts on infrastructure, transportation, and international<br />

relations. The complex interplay among climate, biogeochemical,<br />

ecological, and sociopolitical factors responding to<br />

cryosphere loss in the Arctic and around the world demands<br />

new levels <strong>of</strong> interdisciplinary collaboration and new models<br />

for scientific study (Driscoll et al. 2012 [in this issue]).<br />

A system-level understanding <strong>of</strong> the global cryosphere is<br />

fundamental to predicting the future course <strong>of</strong> the Earth’s<br />

socioecological system and to laying out a course for human<br />

social, political, and economic adaptation to climate change.<br />

As was demonstrated in this article and others in this issue,<br />

socioecological ecosystem science as pioneered by the US<br />

<strong>LTER</strong> Network is a key component <strong>of</strong> our current and future<br />

understanding <strong>of</strong> cryospheric change.<br />

Conclusions<br />

Earth is distinguished in the solar system by the coexistence<br />

<strong>of</strong> water in its three phases: solid (frozen), liquid (melted),<br />

and gas (evaporated). The solid phase—the global cryosphere<br />

in all its components: glaciers; snow; permafrost; sea,<br />

lake, and river ice—is arguably the most rapidly changing<br />

element <strong>of</strong> the Earth system. Cryosphere loss can be viewed<br />

Articles<br />

as a planetary-scale redistribution <strong>of</strong> solid water into its<br />

liquid and gas phases. This large-scale reorganization will<br />

trigger changes in the balance <strong>of</strong> positive and negative<br />

feedbacks in the climate system (e.g., changes in planetary<br />

albedo), with far-reaching consequences for ecosystems<br />

and society, including changes in sea level, precipitation,<br />

and water availability. The current geophysical rates <strong>of</strong><br />

cryosphere loss are now well documented but our lack in<br />

understanding <strong>of</strong> the relevant mechanisms limits our ability<br />

to predict the future course <strong>of</strong> change, with potentially grave<br />

consequences for society. In particular, we lack long-term<br />

observations and system-level experiments in which the<br />

linkages between changes in physical habitat and climate<br />

on one hand and ecosystem structure and biogeochemical<br />

functions on the other are addressed. <strong>LTER</strong> Network sites<br />

have pioneered coordinated observations and experimental<br />

manipulations <strong>of</strong> ecosystems and elemental cycles (Knapp<br />

et al. 2012 [in this issue]). An expansion <strong>of</strong> our fundamental<br />

knowledge <strong>of</strong> the phenologies and processes governing ecosystem<br />

responses to climate change is a necessary first step<br />

in creating future scenarios <strong>of</strong> change and human responses<br />

to it (Thompson et al. 2012 [in this issue]). This new understanding<br />

will continue to come from <strong>LTER</strong> sites situated in<br />

all the major cryosphere systems (table 1).<br />

Acknowledgments<br />

The authors wish to thank the funding provided by the<br />

National Science Foundation’s (NSF) Long Term Ecological<br />

Research (<strong>LTER</strong>) Network for supporting our long-term<br />

studies, in which we track the ecosystem response to the disappearing<br />

cryosphere. The <strong>LTER</strong> Network <strong>of</strong>fice supported<br />

the workshop from which this article is derived. Table 2 and<br />

figure 3 were kindly contributed by Julia A. Jones, Kendra<br />

L. Hatcher, and Evan S. Miles. Valuable contributions were<br />

made by Mike Antolin, Anne Giblin, John Hobbie, John<br />

Magnusson, Anne Nolin, Bruce Peterson, Bill Sobczak, and<br />

Will Wolheim. We greatly appreciate all who participated.<br />

Fred Swanson <strong>of</strong>fered many helpful suggestions, and David<br />

Foster carefully edited the manuscript (and it is much<br />

improved). NSF <strong>LTER</strong> Site Grants OPP 0823101, OPP<br />

1115245, DEB 1114804, DEB-1026415, DEB-0620579, and<br />

DEB-1027341 supported the authors during the preparation<br />

<strong>of</strong> this article.<br />

References cited<br />

Anth<strong>of</strong>f D, Nicholls R, Tol RSJ. 2010. The economic impact <strong>of</strong> substantial<br />

sea-level rise. Mitigation and Adaptation Strategies for Global Change<br />

15: 321–335.<br />

Atkinson A, Siegel V, Pakhomov E, Rothery P. 2004. Long-term decline in<br />

krill stock and increase in salps within the Southern Ocean. Nature 432:<br />

100–103.<br />

Baron JS, Schmidt TM, Hartman MD. 2009. Climate-induced changes in<br />

high elevation stream nitrate dynamics. Global Change Biology 15:<br />

1777–1789.<br />

Blumenthal D, Chimner RA, Welker KM, Morgan JA. 2008. Increased snow<br />

facilitates plant invasion in mixedgrass prairie. New Phytologist 179:<br />

440–448.<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 413


Articles<br />

Bowden WB, Gooseff MN, Balser A, Green A, Peterson BJ, Bradford J.<br />

2008. Sediment and nutrient delivery from thermokarst features in the<br />

foothills <strong>of</strong> the North Slope, Alaska: Potential impacts on headwater<br />

stream ecosystems. Journal <strong>of</strong> Geophysical Research 113 (Art. G02026).<br />

doi:10.1029/2007JG000470.<br />

Cai W-J, et al. 2010. Decrease in the CO 2 uptake capacity in an ice-free<br />

Arctic Ocean basin. Science 329: 556–559.<br />

Caine N. 2002. Declining ice thickness on an alpine lake is generated by<br />

increased winter precipitation. Climatic Change 54: 463–470.<br />

———. 2010. Recent hydrologic change in a Colorado alpine basin: An<br />

indicator <strong>of</strong> permafrost thaw? Annals <strong>of</strong> Glaciology 51: 130–134.<br />

Campbell JL, Ollinger SV, Flerchinger GN, Wicklein H, Hayhoe K, Bailey<br />

AS. 2010. Past and projected future changes in snowpack and soil frost<br />

at the Hubbard Brook Experimental Forest, New Hampshire, USA.<br />

Hydrological Processes 24: 2465–2480.<br />

Cayan DR, Kammerdiener SA, Dettinger MD, Caprio JM, Peterson DH.<br />

2001. Changes in the onset <strong>of</strong> spring in the western United States.<br />

Bulletin <strong>of</strong> the American Meteorological Society 82: 399–415.<br />

Chapin FS III, et al. 2005. Polar systems. Pages 717–743 in Hassan R, Scholes<br />

R, Ash N, eds. Ecosystems and Human Well-being: Current State and<br />

Trends. Island Press.<br />

Chimner RA, Welker JM. 2005. Ecosystem respiration responses to experimental<br />

manipulations <strong>of</strong> winter and summer precipitation in a<br />

Mixedgrass Prairie, WY, USA. Biogeochemistry 73: 257–270.<br />

Comiso JC, Nishio F. 2008. Trends in the sea ice cover using enhanced and<br />

compatible AMSR-E, SSM/I, and SMMR data. Journal <strong>of</strong> Geophysical<br />

Research 113 (Art. C02S07). doi:10.1029/2007JC004257.<br />

Déry SJ, Brown RD. 2007. Recent Northern Hemisphere snow cover extent<br />

trends and implications for the snow-albedo feedback. Geophysical<br />

Research Letters 34 (Art. L22504). doi:22510.21029/22007GL031474.<br />

Dierssen HM, Smith RC, Vernet M. 2002. Glacial meltwater dynamics<br />

in coastal waters west <strong>of</strong> the Antarctic peninsula. Proceedings <strong>of</strong> the<br />

National Academy <strong>of</strong> Sciences 99: 1790–1795.<br />

Driscoll CT, Lambert KF, Chapin FS III, Nowak DJ, Spies TA, Swanson<br />

FJ, Kitteridge DB Jr, Hart CM. 2012. Science and society: The role<br />

<strong>of</strong> long-term studies in environmental stewardship. BioScience 62:<br />

354–366.<br />

Ducklow HW, Baker K, Martinson DG, Quetin LB, Ross RM, Smith RC,<br />

Stammerjohn SE, Vernet M, Fraser W. 2007. Marine pelagic ecosystems:<br />

The west Antarctic peninsula. Philosophical Transactions <strong>of</strong> the Royal<br />

Society <strong>of</strong> London B 362: 67–94.<br />

Dyurgerov MB, Meier MF. 2000. Twentieth century climate change:<br />

Evidence from small glaciers. Proceedings <strong>of</strong> the National Academy <strong>of</strong><br />

Sciences 97: 1406–1411.<br />

Fitzhugh RD, Driscoll CT [Jr], Gr<strong>of</strong>fman PM, Tierney GL, Fahey TJ, Hardy<br />

JP. 2001. Effects <strong>of</strong> soil freezing disturbance on soil solution nitrogen,<br />

phosphorus, and carbon chemistry in a northern hardwood ecosystem.<br />

Biogeochemistry 56: 215–238.<br />

Fountain AG, Nylen TH, MacClune KL, Dana GL. 2006. Glacier mass balances<br />

(1993–2001), Taylor Valley, McMurdo Dry Valleys, Antarctica.<br />

Journal <strong>of</strong> Glaciology 52: 451–462.<br />

Fraser WR, H<strong>of</strong>mann EE. 2003. A predator’s perspective on causal links<br />

between climate change, physical forcing and ecosystem response.<br />

Marine Ecology Progress Series 265: 1–15.<br />

Grebmeier JM, Overland JE, Moore SE, Farley EV, Carmack EC, Cooper LW,<br />

Frey KE, Helle JH, McLaughlin FA, McNutt SL. 2006. A major ecosystem<br />

shift in the northern Bering Sea. Science 311: 1461–1464.<br />

Hamilton LC, Rohall DE, Brown BC, Hayward GF, Keim BD. 2003. Warming<br />

winters and New Hampshire’s lost ski areas: An integrated case study.<br />

International Journal <strong>of</strong> Sociology and Social Policy 23: 52–73.<br />

Johannessen M, Henriksen A. 1978. Chemistry <strong>of</strong> snow meltwater: Changes<br />

in concentration during melting. Water Resources Research 14:<br />

615–619.<br />

Knapp AK, et al. 2012. Past, present, and future roles <strong>of</strong> long-term experiments<br />

in the <strong>LTER</strong> network. BioScience 62: 377–389.<br />

K<strong>of</strong>inas GP, Chapin FS III, BurnSilver S, Schmidt JI, Fresco NL, Kielland<br />

K, Martin S, Springsteen A, Rupp TS. 2010. Resilience <strong>of</strong> Athabascan<br />

subsistence systems to interior Alaska’s changing climate. Canadian<br />

Journal <strong>of</strong> Forest Research 40: 1347–1359.<br />

Kundzewicz ZW, Mata LJ, Arnell NW, Döll P, Kabat P, Jiménez B, Miller K,<br />

Oki T, Sen Z, Shiklomanov IA. 2007. Freshwater resources and their<br />

management. Pages 173–210 in Parry ML, Canziani OF, Palutik<strong>of</strong> JP,<br />

van der Linden PJ, Hanson CE, eds. Climate Change 2007: Impacts,<br />

Adaptation and Vulnerability. Contribution <strong>of</strong> Working Group II to the<br />

Fourth Assessment Report <strong>of</strong> the Intergovernmental Panel on Climate<br />

Change. Cambridge <strong>University</strong> Press.<br />

Lazar B, Williams M. 2008. Climate change in western ski areas: Potential<br />

changes in the timing <strong>of</strong> wet avalanches and snow quality for the<br />

Aspen ski area in the years 2030 and 2100. Cold Regions Science and<br />

Technology 51: 219–228.<br />

Le Quéré C, et al. 2009. Trends in the sources and sinks <strong>of</strong> carbon dioxide.<br />

Nature Geoscience 2: 831–836.<br />

Magnuson JJ, et al. 2000. Historical trends in lake and river cover in the<br />

Northern Hemisphere. Science 289: 1743–1746.<br />

Maljanen M, Alm J, Martikainen PJ, Repo T. 2010. Prolongation <strong>of</strong> soil frost<br />

resulting from reduced snow cover increases nitrous oxide emissions<br />

from boreal forest soil. Boreal Environment Research 15: 34–42.<br />

Maurer J. 2007. Atlas <strong>of</strong> the Cryosphere. National Snow and Ice Data Center.<br />

(26 January 2012; http://nsidc.org/data/atlas).<br />

Mazhitova G, Malkova G, Chestnykh O, Zamolodchikov D. 2008. Recent<br />

decade thaw depth dynamics in the European Russian Arctic based on<br />

the Circumpolar Active Layer Monitoring (CALM) data. Pages 433–438<br />

in Kane DL, Hinkel KM, eds. Proceedings <strong>of</strong> the Ninth International<br />

Conference on Permafrost. Institute <strong>of</strong> Northern Engineering, <strong>University</strong><br />

<strong>of</strong> Alaska, Fairbanks.<br />

Micol T, Jouventin P. 2001. Long-term population trends in seven Antarctic<br />

seabirds at Pointe Géologie (Terre Adélie): Human impact compared<br />

with environmental change. Polar Biology 24: 175–185.<br />

Myers-Smith IH, McGuire AD, Harden JW, Chapin FS III. 2007. Influence<br />

<strong>of</strong> disturbance on carbon exchange in a permafrost collapse and adjacent<br />

burned forest. Journal <strong>of</strong> Geophysical Research 112: 11.<br />

Nemergut DR, Anderson SP, Cleveland CC, Martin AP, Miller AE, Seimon<br />

A, Schmidt SK. 2007. Microbial community succession in unvegetated,<br />

recently-deglaciated soils. Microbial Ecology 53: 110–122.<br />

Nolin AW, Daly C. 2006. Mapping “at risk” snow in the Pacific Northwest.<br />

Journal <strong>of</strong> Hydrometeorology 7: 1164–1171.<br />

Parmesan C. 2006. Ecological and evolutionary responses to recent climate<br />

change. Annual Review <strong>of</strong> Ecology, Evolution, and Systematics 37:<br />

637–669.<br />

Peck LS, Barnes DKA, Cook AJ, Fleming AH, Clarke A. 2010. Negative feedback<br />

in the cold: Ice retreat produces new carbon sinks in Antarctica.<br />

Global Change Biology 16: 2614–2623.<br />

Pfeffer WT, Harper JT, O’Neel S. 2008. Kinematic constraints on glacier<br />

contributions to 21st-century sea-level rise. Science 321: 1340–1343.<br />

Prowse TD, Wrona FJ, Reist JD, Gibson JJ, Hobbie JE, Lévesque LMJ,<br />

Vincent WF. 2006. Climate change effects on hydroecology <strong>of</strong> Arctic<br />

freshwater ecosystems. Ambio: A Journal <strong>of</strong> the Human Environment<br />

35: 347–358.<br />

Ray C, Beever E, Loarie S. 2012. Retreat <strong>of</strong> the American pika: Up the mountain<br />

or into the void? In Brodie JF, Post E, Doak D, eds. Conserving<br />

Wildlife Populations in a Changing Climate. <strong>University</strong> <strong>of</strong> Chicago<br />

Press. Forthcoming.<br />

Robertson GP, et al. 2012. Long-term ecological research in a humandominated<br />

world. BioScience 62: 342–353.<br />

Romanovsky VE, Gruber S, Instanes A, Jin H, Marchenko SS, Smith SL,<br />

Trombotto D, Walter KM. 2007. Frozen ground. Pages 181–200 in<br />

United Nations Environment Programme (UNEP). Global Outlook for<br />

Ice and Snow. UNEP.<br />

Schaefer DA, Driscoll CT Jr, van Dreason R, Yatsko CP. 1990. The episodic<br />

acidification <strong>of</strong> Adirondack lakes during snowmelt. Water Resources<br />

Research 26: 1639–1647.<br />

Schimel JP, Bilbrough C, Welker JM. 2004. Increased snow depth affects<br />

microbial activity and nitrogen mineralization in two Arctic tundra<br />

communities. Soil Biology and Biogeochemistry 36: 217–227.<br />

414 BioScience • April 2012 / Vol. 62 No. 4 www.biosciencemag.org


Schuur EAG, Vogel JG, Crummer KG, Lee H, Sickman JO, Osterkamp TE.<br />

2009. The effect <strong>of</strong> permafrost thaw on old carbon release and net<br />

carbon exchange from tundra. Nature 459: 556–559.<br />

Solomon SD, Qin M, Manning Z, Chen M, Marquis KB, Avery T, Tignor M,<br />

Miller HL, eds. 2007. Climate Change 2007: The Physical Science Basis.<br />

Cambridge <strong>University</strong> Press.<br />

Stammerjohn SE, Martinson DG, Smith RC, Iannuzzi RA. 2008. Sea ice<br />

in the western Antarctic Peninsula region: Spatio-temporal variability<br />

from ecological and climate change perspectives. Deep Sea Research<br />

55: 2041–2058.<br />

Starr G, Oberbauer SF. 2003. Photosynthesis <strong>of</strong> Arctic evergreens under<br />

snow: Implications for tundra ecosystem carbon balance. Ecology 84:<br />

1415–1420.<br />

Thompson JR, Wiek A, Swanson FJ, Carpenter SR, Fresco N, Hollingsworth<br />

T, Spies TA, Foster DR. 2012. Scenario studies as a synthetic and integrative<br />

research activity for long-term ecological research. BioScience<br />

62: 367–376.<br />

Trivelpiece WZ, Hinke JT, Miller AK, Reiss CS, Trivelpiece SG, Watters GM.<br />

2011. Variability in krill biomass links harvesting and climate warming<br />

to penguin population changes in Antarctica. Proceedings <strong>of</strong> the<br />

National Academy <strong>of</strong> Sciences 108: 7625–7628.<br />

Turetsky MR, Wieder RK, Vitt DH, Evans RJ, Scott KD. 2007. The disappearance<br />

<strong>of</strong> relict permafrost in boreal north America: Effects on peatland<br />

carbon storage and fluxes. Global Change Biology 13: 1922–1934.<br />

Vogel JG, Schuur EAG, Trucco C, Lee H. 2010. The carbon cycling response <strong>of</strong><br />

arctic tundra to permafrost thaw and thermokarst development. Journal<br />

<strong>of</strong> Geophysical Research 114 (Art. G4). doi:10.1029/2008JG000901<br />

Walker DA, Halfpenny JC, Walker MD, Wessman CA. 1993. Long-term<br />

studies <strong>of</strong> snow–vegetation interactions. BioScience 43: 287–301.<br />

Weimerskirch H, Inchausti P, Guinet C, Barbraud C. 2003. Trends in bird<br />

and seal populations as indicators <strong>of</strong> a system shift in the Southern<br />

Ocean. Antarctic Science 15: 249–256.<br />

Articles<br />

Westerling AL, Hidalgo HG, Cayan DR, Swetnam TW. 2006. Warming and<br />

earlier spring increase western U.S. forest wildfire activity. Science 313:<br />

940–943.<br />

Williams MW, Losleben M, Caine N, Greenland D. 1996. Changes in<br />

climate and hydrochemical responses in a high-elevation catchment<br />

in the Rocky Mountains, USA. Limnology and Oceanography 41:<br />

939–946.<br />

Williams MW, Knauf M, Cory R, Caine N, Liu F. 2007. Nitrate content and<br />

potential microbial signature <strong>of</strong> rock glacier outflow. Earth Surface<br />

Processes and Landforms 32: 1032–1047.<br />

Wood ME, Kelley JT, Belknap DF. 1989. Patterns <strong>of</strong> sediment accumulation<br />

in the tidal marshes <strong>of</strong> Maine. Estuaries 12: 237–246.<br />

Zamolodchikov D, Kotov A, Karelin D, Razzhivin V. 2008. Recent climate<br />

and active layer changes in northeast Russia: Regional output<br />

<strong>of</strong> Circumpolar Active Layer Monitoring (CALM). Pages 2021–2026<br />

in Kane DL, Hinkel KM, eds. Proceedings <strong>of</strong> the Ninth International<br />

Conference on Permafrost. Institute <strong>of</strong> Northern Engineering, <strong>University</strong><br />

<strong>of</strong> Alaska, Fairbanks.<br />

Andrew G. Fountain (andrew@pdx.edu) is a pr<strong>of</strong>essor in the Department <strong>of</strong><br />

Geology at Portland State <strong>University</strong> in Portland, Oregon. John L. Campbell is<br />

an ecologist at the US Department <strong>of</strong> Agriculture Forest Service in Durham, New<br />

Hampshire. Edward A. G. Schuur is an associate pr<strong>of</strong>essor in the Department<br />

<strong>of</strong> Biology at the <strong>University</strong> <strong>of</strong> Florida, Gainesville. Sharon E. Stammerjohn is<br />

an assistant pr<strong>of</strong>essor in the Department <strong>of</strong> Ocean Sciences at the <strong>University</strong><br />

<strong>of</strong> California, Santa Cruz. Mark W. Williams is a pr<strong>of</strong>essor in the Department<br />

<strong>of</strong> Geography at the <strong>University</strong> <strong>of</strong> Colorado, Boulder. Hugh W. Ducklow is a<br />

senior scientist and director <strong>of</strong> The Ecosystems Center at the Marine Biological<br />

Laboratory in Woods Hole, Massachusetts.<br />

www.biosciencemag.org April 2012 / Vol. 62 No. 4 • BioScience 415

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

Saved successfully!

Ooh no, something went wrong!