What Are Ecosystem Services, and How Do They Help Our Planet?

Protecting diverse ecosystems and the natural benefits that they provide is essential to the future of life on our planet and the well-being of humanity.

Those services, which are often called ecosystem services, include providing resources such as food and water, maintaining habitats that support biodiversity, offering opportunities for recreation, and helping to regulate human-caused impacts like climate change.

Healthy, diverse ecosystems are responsible for the air we breathe, the food we eat, and the natural places that we visit to nurture our bodies and minds. They support the species that, in turn, sustain human life. “Every morsel of food, every sip of water, the air we breathe is the result of work done by another species,” says Enric Sala , a National Geographic Explorer-in-Residence and the leader of National Geographic’s Pristine Seas Initiative. “Without them, there is no us.”

When it comes to mitigating the impacts of climate change, ecosystem services really shine. Studies have shown that the natural world can provide one-third of all climate mitigation efforts. Land and marine ecosystems currently absorb about half of the human-generated carbon dioxide emissions, with forests alone removing 2.6 billion tons of carbon from the atmosphere each year.

Ecosystems must stay intact and healthy to receive the myriad of benefits that they provide. Campaign for Nature is a global effort to raise awareness of the threats facing our natural world and inspire world leaders to take action to protect 30 percent of the planet by 2030.

The protection of 30 percent of the planet by 2030 is a critical milestone toward protecting 50 percent of the planet by 2050, a benchmark that scientists say will ensure the health and diversity of ecosystems across the globe. The goal is challenging yet attainable—and our future depends on it.

The National Geographic Society is a global nonprofit organization that uses the power of science, exploration, education and storytelling to illuminate and protect the wonder of our world. Since 1888, National Geographic has pushed the boundaries of exploration, investing in bold people and transformative ideas, providing more than 15,000 grants for work across all seven continents, reaching 3 million students each year through education offerings, and engaging audiences around the globe through signature experiences, stories and content.

To learn more, visit www.nationalgeographic.org or follow us on Instagram , LinkedIn, and Facebook .

Ecosystem Services and Their Main Types Essay

  • To find inspiration for your paper and overcome writer’s block
  • As a source of information (ensure proper referencing)
  • As a template for you assignment

Ecosystem services (ES) refer to all the benefits people receive from ecosystems. They differ from most other goods and services because they are much more complex in function, interaction, and impact. Ecosystem services enable people to obtain resources from the land itself through proper care and use of resources. Four types of ecosystem services are currently distinguished: providing, regulating, cultural, and supporting (Pearce, 2020). Each class allows humanity to sustain itself with vital resources such as food, firewood, water, etc.

The development of ecosystem services not only satisfies human life needs but also works for a sense of aesthetics. For example, cultural services have enabled people to breed new species of ornamental plants, which they use to decorate their homes and grounds. Many communities place a high value on preserving historically significant landscapes (“cultural landscapes”) or culturally notable species (animals or plants). Thus, the ecosystem influences human cultural development by identifying people with certain landscapes, forests, mountains, or fields.

Generally, the term “indigenous peoples” refers to peoples who have faced colonization or conquest and therefore find themselves a minority (or “non-dominant” part of the population) within a state formed by an incoming dominant group (Amnesty International, 2019). Indigenous peoples have inherent characteristics, such as Historical continuity with pre-colonial societies (societies before conquest) and strong ties to territories and adjacent natural resources. They are also characterized by different social, economic, or political systems and a rare language, culture, and beliefs.

The notion of unpopular indigenous peoples in society is considered a rudimentary phenomenon. Despite this, recently, young representatives of indigenous peoples, who are almost forgotten, have begun to remind themselves of their rights by standing up for them. For example, the story of Alba Veronica Yacabalcchia, who belongs to the Guatemalan indigenous Maya Kiche (United Nations, 2022). She has long defended the rights of her people through the recognition of her people’s language, the translation of necessary documentation for her fellow citizens, and the fight against discrimination based on the nationality of more “recognized” people. This is just one of the thousands of examples of young people drawing public attention to their people through news and government coverage.

Amnesty International. (2019). Indigenous peoples . Amnesty.org. Web.

Pearce, R. (2020). What are ecosystem services? Earth.org – Past | Present | Future. Web.

United Nations. (2022). Fighting for indigenous rights in guatemala . OHCHR.org. Web.

  • Cashion Water Quality: Spatial Distribution of Water Pollution Incidents
  • Tropical Rain Forest
  • “To Build Fire” in Nulato Village, Alaska
  • Ornamental Fancy Style in China and Australia Market
  • Rolland’s Model on the Phases of Illness
  • The Tundra Biome and Its Keystone Species
  • Balanced Ecosystem in Connection to Art and Science
  • Beach Zonation and Tarpon Springs’ Man-Made Beaches
  • Fisheries and Sharks in Trinidad and Tobago
  • The Biome and Ecosystem Concepts and Importance
  • Chicago (A-D)
  • Chicago (N-B)

IvyPanda. (2023, December 18). Ecosystem Services and Their Main Types. https://ivypanda.com/essays/ecosystem-services-and-their-main-types/

"Ecosystem Services and Their Main Types." IvyPanda , 18 Dec. 2023, ivypanda.com/essays/ecosystem-services-and-their-main-types/.

IvyPanda . (2023) 'Ecosystem Services and Their Main Types'. 18 December.

IvyPanda . 2023. "Ecosystem Services and Their Main Types." December 18, 2023. https://ivypanda.com/essays/ecosystem-services-and-their-main-types/.

1. IvyPanda . "Ecosystem Services and Their Main Types." December 18, 2023. https://ivypanda.com/essays/ecosystem-services-and-their-main-types/.

Bibliography

IvyPanda . "Ecosystem Services and Their Main Types." December 18, 2023. https://ivypanda.com/essays/ecosystem-services-and-their-main-types/.

IvyPanda uses cookies and similar technologies to enhance your experience, enabling functionalities such as:

  • Basic site functions
  • Ensuring secure, safe transactions
  • Secure account login
  • Remembering account, browser, and regional preferences
  • Remembering privacy and security settings
  • Analyzing site traffic and usage
  • Personalized search, content, and recommendations
  • Displaying relevant, targeted ads on and off IvyPanda

Please refer to IvyPanda's Cookies Policy and Privacy Policy for detailed information.

Certain technologies we use are essential for critical functions such as security and site integrity, account authentication, security and privacy preferences, internal site usage and maintenance data, and ensuring the site operates correctly for browsing and transactions.

Cookies and similar technologies are used to enhance your experience by:

  • Remembering general and regional preferences
  • Personalizing content, search, recommendations, and offers

Some functions, such as personalized recommendations, account preferences, or localization, may not work correctly without these technologies. For more details, please refer to IvyPanda's Cookies Policy .

To enable personalized advertising (such as interest-based ads), we may share your data with our marketing and advertising partners using cookies and other technologies. These partners may have their own information collected about you. Turning off the personalized advertising setting won't stop you from seeing IvyPanda ads, but it may make the ads you see less relevant or more repetitive.

Personalized advertising may be considered a "sale" or "sharing" of the information under California and other state privacy laws, and you may have the right to opt out. Turning off personalized advertising allows you to exercise your right to opt out. Learn more in IvyPanda's Cookies Policy and Privacy Policy .

Encyclopedia Britannica

  • History & Society
  • Science & Tech
  • Biographies
  • Animals & Nature
  • Geography & Travel
  • Arts & Culture
  • Games & Quizzes
  • On This Day
  • One Good Fact
  • New Articles
  • Lifestyles & Social Issues
  • Philosophy & Religion
  • Politics, Law & Government
  • World History
  • Health & Medicine
  • Browse Biographies
  • Birds, Reptiles & Other Vertebrates
  • Bugs, Mollusks & Other Invertebrates
  • Environment
  • Fossils & Geologic Time
  • Entertainment & Pop Culture
  • Sports & Recreation
  • Visual Arts
  • Demystified
  • Image Galleries
  • Infographics
  • Top Questions
  • Britannica Kids
  • Saving Earth
  • Space Next 50
  • Student Center
  • Introduction

History of concepts and methods

Identification, quantification, and valuation, role in policy and management.

apple pollination

  • What are the abiotic and biotic components of the biosphere?

A large school of mano (sharks) called Galapagos sharks at Maro Reef in the Papahanaumokuakea Marine National Monument. Galapagos shark (Carcharhinus galapagensis)a worldwide species of requiem shark, family Carcharhinidae.

ecosystem services

Our editors will review what you’ve submitted and determine whether to revise the article.

  • Food and Agriculture Organisation of the United Nations - AGP - Biodiversity and Ecosystem Services
  • The National Wildlife Federation - Ecosystem Services
  • Table Of Contents

ecosystem services , outputs, conditions, or processes of natural systems that directly or indirectly benefit humans or enhance social welfare. Ecosystem services can benefit people in many ways, either directly or as inputs into the production of other goods and services. For example, the pollination of crops provided by bees and other organisms contributes to food production and is thus considered an ecosystem service. Another example is the attenuation of flooding in residential areas provided by riparian buffers and wetlands .

Because ecosystem services are not usually bought and sold directly in markets , market activities do not fully reflect the benefits provided by those services. Unregulated markets thus promote excessive depletion of natural capital (e.g., the biotic and abiotic components of ecosystems) and ecosystem services. The United Nations Millennium Ecosystem Assessment (2005), which evaluated the consequences of ecosystem change, concluded that humans have degraded the ability of Earth’s ecosystems to support social welfare. In response, ecosystem services analyses promote policy decisions that recognize the full range of benefits and costs associated with actions that affect those services. Most formal evaluations of ecosystem services examine the consequences of changes to specific services in certain geographical areas for particular beneficiary groups; very few analyses approach ecosystem services from universal perspectives (e.g., all the services provided by wetlands across the planet).

The concept of “ecosystem services” emerged during the 1970s and gained increasing recognition in the following decades. However, the idea that natural systems support human welfare is much older. Relationships between deforestation and water supply were documented by Plato as early as 400 bce . Economists in the 18th and 19th centuries recognized the value provided by land and other natural resources as productive assets. Man and Nature (1864) by George Perkins Marsh is recognized as one of the founding works of the U.S. conservation movement and was among the first writings to formally characterize relationships between natural and social systems. Marsh argued that extensive damage to natural systems by human actions would diminish human welfare.

Methods to characterize the structure and function of natural systems are grounded in centuries of work by natural scientists. Of particular relevance to ecosystem services analysis are modern ecological concepts, models, and methods developed during and after the 20th century. Quantification of ecosystem service values has its foundation in formal economic methods for nonmarket valuation, which have been refined extensively since their initial development by environmental and resource economists in the 1940s. By the early 21st century, ecosystem services analyses paid greater attention to issues such as the complex relationships between ecological and socioeconomic systems, how changes in those relationships affect human welfare, to what extent the values of different types of services can and should be quantified in monetary terms, and the most-suitable approaches to quantify the different types of services.

Two criteria distinguish ecosystem services from other ecosystem conditions or processes. First, an ecosystem service must be linked to an identifiable set of human beneficiaries. The service can be an aspect or consequence of an ecological condition and can directly or indirectly benefit or profit the beneficiaries. Second, physical and institutional access constraints must not prevent people from realizing those benefits. For example, increases in fish abundance can enhance the welfare of those engaged in commercial or recreational fishing, but only if those increases occur in areas where fishing occurs. Conditions or processes of ecosystems that cannot be linked to the welfare of identifiable beneficiary groups are not ecosystem services. For example, changes in fish abundance in areas not used by humans and that have no direct or indirect effect on human benefits are not ecosystem services.

The first step in most ecosystem services assessments is to identify the services to be evaluated. This requires an understanding of how beneficiaries are affected directly or indirectly by changes in ecological conditions and processes and can be assisted by the development of a conceptual causal chain. Such a chain clarifies the links between human actions and ecological effects and the subsequent changes to ecosystem services and the associated human benefits. For example, a causal chain related to mechanical forest thinning would first identify the primary effects of thinning on ecological conditions such as forest structure. Those changes are then linked to impacts on ecosystem outputs, conditions, or processes that affect the welfare of identifiable beneficiary groups; those are the ecosystem services. Forest thinning, for instance, could change forest structure in such a way that there is a reduction in fire risk to populated areas. Causal chains often include multiple steps between the original human action and the effects on final ecosystem services. One of the challenges of ecosystem services analysis is identifying the many ways that ecosystem services affect different beneficiary groups.

essay on ecosystem service

A common second step in an ecosystem service assessment is to quantify one or more of the services identified in the causal chain. Quantification requires an understanding of the measures of quantity, quality, or other biophysical features that are most directly relevant to human welfare, followed by the use of appropriate methods to obtain those measures. Some ecosystem services are easily quantified, such as the quantity of timber produced in a given forest area. Others are more difficult to measure, either because of limitations in biophysical models or data (e.g., fish produced by tidal wetlands) or because the service is inherently difficult to quantify (e.g., aesthetic features of natural landscapes).

The third step in an ecosystem service assessment is to determine the consequences for social welfare. This is often conducted by using formal economic valuation methods grounded in neoclassical economic theory, although noneconomic techniques may also be applied. Economic valuation quantifies the value of an ecosystem service in commensurable (typically monetary) units; that value is well-defined only for a specific quantity of the ecosystem service, measured from a particular baseline. Values calculated for services at a biome or planetary scale (for which changes and baselines are often poorly defined) are generally considered invalid—or at least highly inaccurate—by economists. The accuracy of valuation also depends on the ability of an analysis to account for factors that influence the contribution of ecosystem services to welfare. For example, all else equal, per-unit values often increase as a service becomes more scarce (reflecting diminishing marginal utility ) and are frequently affected by spatial factors such as the distance of beneficiaries from the service.

wetland

Valuation also requires analysts to distinguish between intermediate and final services. Final services affect welfare directly, whereas intermediate services underpin final services but are not valued directly. For example, biochemical processes in wetlands can reduce the delivery of excess nutrients such as nitrogen to nearby waters (an intermediate service), thereby reducing algae growth and improving water clarity (a final service to swimmers and others who prefer clear water). The economic value of a final service, if correctly measured, should incorporate the value of all intermediate inputs used in its production. Some ecosystem services serve as both final and intermediate services.

Additionally, ecosystem services can provide different types of values, some related to the use of the services by humans and others unrelated to human use; these are known to economists as use values and nonuse values, respectively. Use values are related to observable behaviours through which people directly or indirectly use or enjoy ecosystem services. Nonuse values are values unrelated to observed behaviours; these include values generated by the knowledge that something in nature exists, that it can be passed on to future generations, or that it is available to benefit other people (i.e., existence, bequest , and altruistic values).

There are increasing worldwide efforts to incorporate information on ecosystem services into public and private decisions. In the United States , for example, the President’s Committee of Advisors on Science and Technology (PCAST) 1998 report, Teaming with Life: Investing in Science to Understand and Use America’s Living Capital , highlighted the value and decline of the country’s living capital and called for the government to develop sustainable strategies for conservation and management of biodiversity and ecosystems. A 2011 PCAST report, Sustaining Environmental Capital: Protecting Society and the Economy , recommended actions to integrate ecosystem services information into federal decisions, including increased use of ecosystem services valuation. Those initiatives accompany multiple agency-specific mandates and initiatives to consider ecosystem services.

Payments for ecosystem services (PES) initiatives, which incentivize the provision of ecosystem services by private suppliers, are emerging worldwide, including national programs in countries such as China and Costa Rica . Many other countries have integrated ecosystem services information into governmental planning processes, including multiple examples in the European Union and Latin America . International initiatives include the World Bank’s Wealth Accounting and Valuation of Ecosystem Services (WAVES) program and The Economics of Ecosystems and Biodiversity (TEEB) project. Despite progress in such efforts, many areas of ecosystem services analysis and policy integration remain at a proof-of-concept stage; systematic and formal use of information on ecosystem services to guide public and private decisions is not yet commonplace.

What Are Ecosystem Services?

How are ecosystem services defined, what are the unique services ecosystems provide.

Ecosystem services are all the processes and outputs that nature provides us with. These include provisioning services (food, water), regulating services (waste water treatment, pollution control), supporting services (shelter), and cultural services (recreation and tourism).

Our planet is blessed with many types of ecosystems, including terrestrial, marine, freshwater, forest and grassland. An ecosystem is a dynamic community that comprises living organisms, such as microorganisms, plants and animals, as well as non-living environments, each interacting with one another. Each ecosystem and its components (water, living organisms, soil) play a key role in maintaining our wellbeing and health.

Our planet earth and it's diverse ecosystems

Human societies have benefited both directly and indirectly from nature for centuries. From clean air to drinking water; weather and climate control to natural crop pollination, nature provides us with critical services for our wellbeing. As you can see, we cannot survive without these goods, outputs, and processes that natural and managed ecosystems provide for us.

Recommended Video for you:

Over the years, we have defined ecosystem services in many ways. For instance, Costanza defined ecosystem services in 1997 as the processes and outputs provided to us during the transformation of natural resources. Similarly, in 2005, the Millennium Ecosystem Assessment (MEA) report defined ecosystem services as ‘the benefits people get from ecosystems.’ More recently , we have defined them as the contributions of ecosystem structure and function, along with a combination of other inputs, to human well-being.

Also Read: What Causes Biodiversity Loss?

Overall, there are four major types of ecosystem services: provisioning, regulating, supporting and cultural.

Provisioning Services

Our environment provides us with raw materials, food, shelter, energy and other resources that are essential for our physical wellbeing and various economic activities. Together, these materialistic resources are known as provisioning services. Some examples of provisioning services are water, food, raw materials and medicinal resources.

Water: Freshwater ecosystems play a huge role in regulating and purifying water. Similarly, forests and vegetation help improve the quality of water by acting as filters.

Food: Ecosystems provide us with the right conditions and environments to grow vegetables, fruit, pulses, rice and other important food types. On the other hand, humans also rely on freshwater and marine resources, as well as wild animals living in forested ecosystems for meat.

The many provisioning resources and services ecosystems provide.

Raw materials: We all need raw materials for construction, whether it is for our homes or offices. Ecosystems provide us with valuable resources, such as timber (from forests) and biofuels from plants, which are the primary materials used in construction activities.

Medicinal resources: Nature provides many resources, including plants and animals that we humans use as traditional medicines. For instance, if you have an upset stomach, you might immediately reach out for ginger from the fridge. When we feel the flu coming on, we immediately start making chicken soup! Such resources are all sourced from plants and animals are not only used in common households, but are also sourced by pharmaceutical companies as raw material for medicine.

Regulating Services

These are the many services that nature provides while acting as regulators. For instance, many natural resources, e.g., air, water, soil, flood and disease, require some regulation to ensure their quality, all of which are enabled by ecosystems.

Treatment of wastewater: Wetlands are classic examples of ecosystems that provide vital filtering services of animal and human waste, which purifies water.

Air quality and carbon sequestration: Trees and forests play a critical role in our lives. They provide shade, promote rainfall, influence the availability of water and help regulate air pollution. Forested ecosystems play a huge role in regulating climate and weather by storing carbon and other greenhouse gases. As trees grow older, they remove more carbon from the atmosphere, which helps keep our planet cool.

Key examples of regulatory services provided by nature.

Prevent or moderate extreme events: Many ecosystems can regulate natural disasters, such as floods and storms, by acting as barriers. The most classic example is that of mangrove ecosystems, which help reduce the impact of tsunamis.

Pollination: Pollination is perhaps one of the most critical services that biodiversity provides. Two of nature’s key pollinators are insects (particularly bees) and wind; without these two pollinators, growing agricultural crops would be immensely challenging for us as humans. In fact, researchers have found that pollinators improve crop yields by approximately 75% worldwide!

Supporting Services

These are services that ecosystems provide besides those outlined above. For instance, ecosystems provide shelter and habitats for countless plant and animal species while also maintaining their diversity. Nature also provides them with food and water, which are vital for their survival. Ecosystems help maintain genetic diversity on our planet, which is why we have such a broad and spectacular variety of life forms on Earth.

Trees provide homes to birds and the big blue ocean to countless marines species such as dolphins and whales.

Cultural Services

Our planet’s ecosystems provide us with many non-materialistic goods and services. For instance, nature provides us with green spaces that we use for walking and picnics, as well as land and seascapes that allow for tourism-related activities. Several communities across the world even consider forests sacred and, in many countries, they worship certain tree species.

As you can see, ecosystems provide humans and other life forms with several services that are necessary for survival. However, these critical services are under threat, as human populations have been proliferating over the past few centuries. The mindless abuse and overconsumption of nature’s resources have contributed to the rapid extinction of thousands of species, in addition to causing widespread deforestation, leading to climate change and air, water and soil pollution.

Unfortunately, it is hard to quantify or place a price tag on these services. However, in the past, researchers have been able to quantify how many humans depend on these provisioning services. For instance, a study revealed that forest provisioning ecosystem services accounted for 44% of the average household income of families in Zambia. So, as you can see, these services are immensely valuable, which is why we must ensure that we protect them at all costs. After all, humanity relies on these services to survive and thrive!

Also Read: How Vital Is The Ocean Ecosystem?

  • Costanza, R., d'Arge, R., de Groot, R., Farber, S., Grasso, M., Hannon, B., … van den Belt, M. (1997, May). The value of the world's ecosystem services and natural capital. Nature. Springer Science and Business Media LLC.
  • Ecosystems AND HUMAN WELL-BEING. millenniumassessment.org
  • Burkhard, B., & Maes, J. (Eds.). (2017). Mapping Ecosystem Services. (B. Burkhard & J. Maes, Eds.), []. Pensoft Publishers.
  • Klein, A.-M., Vaissière, B. E., Cane, J. H., Steffan-Dewenter, I., Cunningham, S. A., Kremen, C., & Tscharntke, T. (2006, October 27). Importance of pollinators in changing landscapes for world crops. Proceedings of the Royal Society B: Biological Sciences. The Royal Society.
  • Kalaba, F. K., Quinn, C. H., & Dougill, A. J. (2013, August 4). Contribution of forest provisioning ecosystem services to rural livelihoods in the Miombo woodlands of Zambia. Population and Environment. Springer Science and Business Media LLC.

Tamanna holds a Master’s degree in Ecology and Environmental Sciences and has been working in the field of wildlife conservation for over six years now. She studies wild Asian elephants (their behavior and genetics, interactions with humans) for a living, and thinks it’s the coolest job in the world. She spends most of her free time soaking her feet in the cold waters of the Bay of Bengal.

A male elephant without tusk is walking through the Udawalawe National Park in Sri Lanka(Bene_A)S

What Role Do Elephants Play In Ecosystems?

Different Species Interact With One Another

Species Interactions: How Do Different Species Interact With One Another?

She-bear and Cubs in the summer forest(Sergey Uryadnikov)S

What Are Habitats And Niches?

essay on ecosystem service

Do We Need Deserts On The Planet?

water cycle

What Are Biogeochemical Cycles?

Natural resources

What Natural Resources Are On The Verge Of Exhaustion?

essay on ecosystem service

Are Mushrooms More Similar to Humans than Plants?

essay on ecosystem service

Photosynthesis: The Biochemistry Behind How Plants Make Their Food

essay on ecosystem service

Where Do Fish Come From In New Lakes And Ponds?

essay on ecosystem service

Why Is Volcanic Soil So Fertile?

essay on ecosystem service

Biomes of the World: Oceans, Mountains, Forests, Grasslands, Deserts and More!

essay on ecosystem service

How Can Animals Drink Water From Dirty Ponds And Not Get Sick?

U.S. flag

An official website of the United States government

Here's how you know

Official websites use .gov A .gov website belongs to an official government organization in the United States.

Secure .gov websites use HTTPS A lock ( ) or https:// means you’ve safely connected to the .gov website. Share sensitive information only on official, secure websites.

Home

  •   Facebook
  •   Twitter
  •   Linkedin
  •   Digg
  •   Reddit
  •   Pinterest
  •   Email

Latest Earthquakes |    Chat Share Social Media  

Ecosystems, Ecosystem Services, and Biodiversity

Biodiversity —the variety of life on Earth—provides vital services that support and improve human health and well-being. Ecosystems, which are composed of living things that interact with the physical environment, provide numerous essential benefits to people. These benefits, termed ecosystem services, encompass four primary functions: provisioning materials, such as food and fiber; regulating critical parts of the environment, such as water quality and erosion control; providing cultural services, such as recreational opportunities and aesthetic value; and providing supporting services, such as nutrient cycling.   Climate change  poses many threats and potential disruptions to ecosystems and biodiversity, as well as to the ecosystem services on which people depend.

Building on the findings of the Third National Climate Assessment (NCA3),  this chapter provides additional evidence that climate change is significantly impacting ecosystems and biodiversity in the United States. Mounting evidence also demonstrates that climate change is increasingly compromising the ecosystem services that sustain human communities, economies, and well-being. Both human and natural systems respond to change, but their ability to respond and thrive under new conditions is determined by their adaptive capacity, which may be inadequate to keep pace with rapid change. Our understanding of climate change impacts and the responses of biodiversity and ecosystems has improved since NCA3. The expected consequences of climate change will vary by region, species, and ecosystem type. Management responses are evolving as new tools and approaches are developed and implemented; however, they may not be able to overcome the negative impacts of climate change. Although efforts have been made since NCA3 to incorporate climate adaptation strategies into natural resource management, significant work remains to comprehensively implement climate-informed planning. This chapter presents additional evidence for climate change impacts to biodiversity, ecosystems, and ecosystem services, reflecting increased confidence in the findings reported in NCA3. The chapter also illustrates the complex and interrelated nature of climate change impacts to biodiversity, ecosystems, and the services they provide.

Citation Information

Publication Year 2018
Title Ecosystems, Ecosystem Services, and Biodiversity
DOI
Authors Douglas Lipton, Madeleine A. Rubenstein, Sarah R. Weiskopf, Shawn L. Carter, Jay Peterson, Lisa Crozier, Michael Fogarty, Sarah Gaichas, Kimberly J. W. Hyde, Toni Lyn Morelli, Jeffrey Morisette, Hassan Moustahfid, Roldan Munoz, Rajendra Poudel, Michelle D. Staudinger, Charles Stock, Laura Thompson, Robin S. Waples, Jake Weltzin
Publication Type Report
Publication Subtype Other Government Series
Index ID
Record Source
USGS Organization Office of Science Quality and Integrity

Related Content

Madeleine rubenstein, national science lead, national casc, sarah weiskopf, ph.d., biologist, national casc, shawn l carter, ph.d., senior scientist, national casc, toni lyn morelli, ph.d., research ecologist, northeast casc, michelle staudinger, ph.d. (former employee), science coordinator, northeast casc, laura thompson, ph.d., research ecologist, national casc, jake f. weltzin, senior science advisor.

REVIEW article

Ecosystem services—current challenges and opportunities for ecological research.

\r\nKlaus Birkhofer*&#x;

  • 1 Biodiversity Unit, Department of Biology, Lund University, Lund, Sweden
  • 2 Department of Animal Ecology, Justus Liebig University, Gießen, Germany
  • 3 Centre for Environmental and Climate Research, Lund University, Lund, Sweden

The concept of ecosystem services was originally developed to illustrate the benefits that natural ecosystems generate for society and to raise awareness for biodiversity and ecosystem conservation. In this article we identify major challenges and opportunities for ecologists involved in empirical or modeling ecosystem service research. The first challenge arises from the fact that the ecosystem service concept has not been generated in the context of managed systems. Ecologists need to identify the effect of anthropogenic interventions in order to propose practices to benefit service-providing organisms and associated services. The second challenge arises from the need to evaluate relationships between indicators of ecosystem services that are collected in ecological studies while accounting for uncertainties of ecological processes that underlie these services. We suggest basing the assessment of ecosystem services on the utilization of sets of indicators that cover aspects of service-providing units, ecosystem management and landscape modification. The third challenge arises from the limited understanding of the nature of relationships between services and a lack of a general statistical framework to address these links. To manage ecosystem service provisioning, ecologists need to establish whether services respond to a shared driver or if services are directly linked to each other. Finally, studies relating biodiversity to ecosystem services often focus on services at small spatial or short temporal scales, but research on the protection of services is often directed toward services providing benefits at large spatial scales. Ecological research needs to address a range of spatial and temporal scales to provide a multifaceted understanding of how nature promotes human well-being. Addressing these challenges in the future offers a unique opportunity for ecologists to act as promoters for the understanding about how to conserve benefits gained from nature.

Introduction

The concepts of nature's services ( Westman, 1977 ) or ecosystem services ( Ehrlich and Ehrlich, 1981 ) were originally developed to draw attention to the benefits that ecosystems generate for society and to raise awareness for biodiversity conservation. Since ecosystem services by definition depend on ecological functions, revealing their value should in theory entice managers and policy makers to safeguard those functions. In an early attempt, Costanza et al. (1997) estimated the monetary value of 17 ecosystem services to range from US$16–54 trillion per year, initiating a wave of research on how to value ecosystem services ( De Groot et al., 2002 ; Engel et al., 2008 ; TEEB, 2010 ). Although the valuation of ecosystem services is complex and controversial, the concept has had major consequences for the development of environmental research and policies in the last decades. The Millennium Ecosystem Assessment, in a global assessment of the status and drivers of past and expected future changes in the delivery of ecosystem services, demonstrated the urgent need for research in this field ( Millennium Ecosystem Assessment, 2005 ).

There is a range of definitions for ecosystem services based on diverging views on how they are generated and linked to human well-being (see Vihervaara et al., 2010 and Seppelt et al., 2011 for reviews) leading to alternative classification schemes ( De Groot, 2006 ; Boyd and Banzhaf, 2007 ; Zhang et al., 2007 ; Fisher et al., 2009 ; De Groot et al., 2010 ; Haines-Young and Potschin, 2010 ). The term “ecosystem services” was originally intended to highlight both direct and indirect benefits humans obtained from nature ( Daily, 1997 ). The risk of double counting in economic valuation later motivated some researchers to advocate that the term should be restricted to the final benefits obtained by humans ( Boyd and Banzhaf, 2007 ). De Groot et al. (2002) , for example, integrated information from ecology and economics to propose a comprehensive concept that described, classified, and valued ecosystem functions and the resulting final goods and services provided by natural and semi-natural systems. However, the Millennium Ecosystem Assessment (2005) explicitly considered supporting ecosystem services as ecosystem functions underlying other ecosystem services, i.e., provisioning services (products obtained from ecosystems, e.g., food, fiber, and water), regulating services (benefits obtained from regulation of ecosystem processes, e.g., climate regulation, flood regulation) and cultural services (non-material benefits people obtain from ecosystems, e.g., recreational, aesthetic and spiritual benefit). In contrast, the global initiative “The Economics of Ecosystems and Biodiversity” ( TEEB, 2010 ) to value biodiversity, considered supporting services as ecological processes, but instead added habitat services as an additional concept.

Ecologists have an important role in ecosystem service research, because services irrespective of the definition and classification are related to organisms and their interactions with the environment ( Feld et al., 2009 ). Hence, the focus of an ecologist is particularly at the role of biodiversity and ecosystem functions underpinning the services and goods directly appreciated by humans, i.e., the intermediate ecosystem services in the terminology of Fisher et al. (2009) . It is these functions which remain invisible and risk being underprovided if research does not reveal their contribution to the final services. For example, several ecosystem services are linked to distinct groups of organisms (“service-providing units”; Luck et al., 2003 ). Examples include biological control of pests (performed by natural enemies) and pollination (performed by pollinating insects) which both contribute to agricultural yields, carbon sequestration (performed by soil organisms) that contributes to climate regulation, reduction of water flows (performed by vegetation) that contributes to flood control and the intrinsic value of biodiversity ( Mace et al., 2012 ). Changes in population size or community composition of these service-providing units in response to anthropogenic activities often affect intermediate and therefore also final ecosystem services ( Raffaelli and White, 2013 ). In fact, human impact has been identified as the main driver of changes in ecosystems and associated services ( Millennium Ecosystem Assessment, 2005 ). Consequently, information about effects of land-use change on service-providing units and associated ecosystem services is increasingly demanded by managers and policy makers in order to promote the sustainable use and continuous provision of services (e.g., by the conference of the parties to the convention on biological diversity; CBD, 2010 ).

Understanding interactions between ecosystem properties and processes is a basic domain of ecology and is crucial to map and manage final ecosystem services. However, there are major challenges facing ecologists engaged in this field. First, ecosystem services are by definition determined by the interaction between ecological and social systems, because only ecosystem processes that contribute to the fulfillment of human needs are ecosystem services. This requires ecologists to work with scientists from other disciplines when trying to understand how ecosystems contribute to human welfare. Second, attempts to use the concept to quantify management consequences on ecosystem functions and resulting changes in the economic value of goods and services may oversimplify complex interactions in social-ecological systems ( Norgaard, 2010 ). For example, monetization of nature's services may result in better management of some services, but still underestimates the value of preserving ecosystem functions for long-term sustainability ( Sterner and Persson, 2008 ). Some services may also fail to become incorporated into an optimization framework, such as conservation of biodiversity per se , because they are not transactable ( Mace et al., 2012 ). A fundamental understanding of the ecosystem processes responsible for ecosystem services, including the contribution of organisms to these processes, is a necessary part of ecosystem service research and involves both challenges and opportunities to ecologists (e.g., Hails and Ormerod, 2013 ).

Challenges and Opportunities

By understanding the links between natural and social systems, ecosystem service research aims at developing more sustainably managed ecosystems ( Daily et al., 2009 ). Although this framework may appear oversimplified ( Braat and De Groot, 2012 ), and ecological-economic modeling may better represent social-ecological systems ( Reyers et al., 2013 ), it shows the inherently cross-disciplinary character of ecosystem service research. Here we focus on some selected conceptual, methodological, and statistical challenges arising in empirical ecological studies and associated modeling approaches to ecosystem service research based on our experiences as ecologists and landscape planners ( Garibaldi et al., 2013 ; Lundin et al., 2013 ; Setälä et al., 2014 ; Ekroos et al., 2014 ; Früh-Müller et al., 2014 ). We further provide recommendations about how to deal with these challenges by highlighting opportunities for ecologists to contribute to ecosystem service research in the future. In the following sections, we discuss challenges for ecologists in ecosystem service research when dealing with anthropogenic modifications of ecosystems (challenge 1), assessment of services (challenge 2) including statistical pitfalls and issues of causality when analyzing relationships between multiple ecosystem services (challenge 3) and spatial and temporal scales at which services are provided and/or managed (challenge 4; see Table 1 for an overview of the challenges addressed).

www.frontiersin.org

Table 1. Selected challenges and sub-challenges discussed in this article, with opportunities for ecologists to contribute to improved recommendations regarding the management of ecosystem services .

Challenge 1: Understanding Anthropogenically Modified Systems

Initial accounts focused on ecosystem services provided by natural systems ( Westman, 1977 ; Daily, 1997 ), while ecosystem services associated to managed ecosystems have only received attention later (e.g., Tylianakis et al., 2007 ). Ecologists need to communicate that the concept of ecosystem services is useful to understand how management of human-modified landscapes affect both the production of goods and environmental externalities. For ecologists working in such anthropogenically modified systems (e.g., agricultural landscapes, production forest or urban areas) challenges arise in (i) the identification of human impact on service-providing units and associated ecosystem services and (ii) considering effects of landscapes surrounding land units that provide ecosystem services.

Ecosystems that are managed to produce food, fuel or fiber or local public infrastructure comprise large proportions of the world's terrestrial surface, e.g., almost half is used for agricultural areas, and almost half of the human population inhabits urban ecosystems ( FAOSTAT, 2014 ). The consequences of human impact for biodiversity and ecosystem service delivery vary both qualitatively and quantitatively depending on system properties and land-use intensities. Consequently the management options to sustainably supply ecosystem services vary as much, requiring ecologists to widen the kind of ecological systems studied beyond the traditional domain of ecology. Intensive agricultural management, for example, may lead to high crop yields (final services), but intensively managed fields often have simplified communities of service-providing units and hence low levels of intermediate services such as biological control by natural enemies or pollination ( Médiène et al., 2011 ). To increase final service delivery, but also to compensate for the loss of intermediate services, anthropogenic management is often intensified (e.g., pesticide application; Médiène et al., 2011 ). Given the negative environmental externalities of some intensive management strategies (e.g., groundwater pollution and resource depletion), alternative management strategies that integrate intermediate services by promoting service-providing units are an opportunity to sustainably ensure crop production and to reduce reliance on anthropogenic interventions ( Bommarco et al., 2013 ). Only a comprehensive perspective, that considers the response of all components of agricultural systems (biodiversity, intermediate, and final ecosystem services) to management will help to communicate the overarching importance of ecosystem service management. Urbanization, as a second example, may lead to increases in plant diversity as a consequence of increased habitat heterogeneity, but due to habitat fragmentation negatively affect species that rely on large habitats ( Kowarik, 2011 ). Urban planning that considers the installation of green infrastructure in cities such as street trees and parks may benefit biodiversity and numerous ecosystem services (e.g., air filtration, water regulation, and noise reduction; Bolund and Hunhammar, 1999 ). Ecologists can directly contribute to ecosystem service research and support policy decisions, not only by evaluating human impact, but also by proposing anthropogenic interventions to benefit service-providing units and ecosystem services.

From a landscape perspective, the expansion of sites under human land use (e.g., agricultural fields, pastures and urban areas) at the cost of losing (semi-)natural land may lead to landscape simplification and fragmentation ( Tscharntke et al., 2005 ). In agricultural landscapes, for instance, arable fields provide the final service of crop production, but constitute disturbed and ephemeral habitats, while many species associated with intermediate services (e.g., pollinators or biocontrol agents) depend on less disturbed habitats in the surrounding landscape (e.g., hedges or uncultivated field borders; see also challenge 4 and Table 2 ). Wild bees are one such example ( Garibaldi et al., 2013 ), as these service providers maintain higher levels of crop pollination in the vicinity of semi-natural habitats ( Garibaldi et al., 2011 ). Such context dependency has also been shown for biological control, which is predicted to be higher in more complex landscapes ( Bianchi et al., 2006 ). Hence, to account for complex interactions with complementary habitat types or non-linear relationships to habitat area ( Jauker et al., 2009 , see also Hauck et al., 2013 ), a simple mapping from the extent of different habitat types may not suffice, but instead a spatially explicit landscape perspective on ecosystem services is needed.

www.frontiersin.org

Table 2. Examples for biological control measures from the literature and proposed categorization in service-providing units, ecosystem modification (e.g., an agricultural field) and landscape modification (i.e., an agricultural landscape; see also challenge 2) .

Challenge 2: Assessing Ecosystem Services

Instruments for assessing ecosystem services, including quantification, mapping and modeling, are a matter of debate in ecosystem service research (e.g., Carpenter et al., 2009 ; Feld et al., 2009 ; Hou et al., 2013 ). From the perspective of an ecologist challenges in assessing ecosystem services arise from the need (i) to evaluate relationships between services and the kind of measures usually collected in ecological studies (e.g., species richness) and (ii) to account for the characteristics of ecological processes (e.g., dynamics, feedbacks, and uncertainties) in statistical models focusing on service provision.

Final ecosystem services are often directly assessed, but such assessment does not provide information about contributing ecological processes or how management could be adapted to increase service provision. A mechanistic understanding of relationships between management and ecosystem services is required to transfer management recommendation outside the context where data were collected. This includes the assessment of the contribution of intermediate ecosystem services and how they are affected by management. The assessment of intermediate services is often more costly and time-consuming than for final services. This partly stems from the lack of proxies for ecological functions and the fact that links between ecosystem functions and final services may be context dependent, e.g., depend on spatial association ( Tixier et al., 2013 ) or ecosystem type ( Feld et al., 2009 ). Measures used to assess intermediate services include direct measures of intermediate services (e.g., pollination success; Kremen et al., 2002 ), indicators of service provision (e.g., dung removal by dung beetles; Gollan et al., 2013 ) and proxies that are indirectly linked to ecosystem services (e.g., proportion of semi-natural habitats in the surrounding of a focal field; Rusch et al., 2012 ). During the past decade, there has been considerable effort in developing instruments to perform assessments of ecosystem services, but measuring ecosystem services based on scientific standards is still not trivial ( Carpenter et al., 2009 ). For instance, predator species richness has been used to indicate levels of biological control (e.g., Duelli and Obrist, 2003 ), even though the effect of predator richness on prey is still debated ( Bruno and Cardinale, 2008 ). The direct quantification of intermediate services is challenging, as complex biotic interactions and environmental conditions may alter service provision. Biological control of arable weeds can, for example, be estimated by quantifying seed removal from seed cards (e.g., Jonason et al., 2013 ). However, such estimates are difficult to scale up to a whole field or farm. Pollination of potted plants, so called phytometers, is a promising technique to estimate pollination potential ( Woodcock et al., 2014 ), but uncertainty remains about how the pollination success of a small number of potted plants reflects pollination of crops (cf. Sih and Baltus, 1987 ). Ecologists, in collaboration with agricultural and forest scientists, thus need to identify scientifically sound ecological measures that are reliable indicators of ecosystem service provision.

As a first and simple step to account for the characteristics of processes underlying service provision, it is suggested here to choose among a small set of measures that form joint, reliable indicators of an individual service. The following example illustrates why the selection of a set of indicators may be superior to the use of a single indicator using the ecosystem service of biological control (see also Kandziora et al., 2013 ). Processes underlying the service of biological control are related to service-providing units (predators and parasitoids), units that provide a disservice (pests; Letourneau et al., 2009 ) and both groups of organisms are altered by anthropogenic interventions at the spatial scale of fields ( Médiène et al., 2011 ) and landscapes ( Bianchi et al., 2006 ). The assessment of biological control may therefore be improved if a small set of selected measures is included that covers aspects of service and disservice-providing units (e.g., pest consumption rates), ecosystem management (e.g., insecticide applications) and landscape modification (e.g., proportion of semi-natural habitats in the surrounding landscape; Table 2 ). Consideration of abiotic variables such as climate ( Diehl et al., 2013 ) or soil characteristics ( Birkhofer et al., 2008 ) will add to the explanatory power of this set of measures.

Ecosystem service research is particularly focused on predicting the consequences of future management options. Statistical models can be used to identify driving forces of changes in service provision and to predict system shifts and fluctuations in service provision as a consequence of environmental change and anthropogenic intervention ( Evans et al., 2012 ). Simple statistical models (e.g., regression) are based on interpolations along existing gradients and cannot provide predictions about levels of ecosystem services under future conditions outside of these gradients. In contrast, process-based models are based on the assumption that essential features of ecological processes can be extrapolated to conditions not currently observed. These models rely on knowledge about the dynamics of ecological processes, i.e., intermediate ecosystem services, including interactions, feedbacks, and uncertainties ( Nicholson et al., 2009 ). For example, models based on the food and nestling requirement of bees can be used to predict pollinator abundance across landscapes because fundamental assumptions about bee behavior hold under novel conditions ( Kennedy et al., 2013 ). In this context, climatic conditions deserve particular attention, since climate change will have a strong impact on service-providing units, intermediate and final ecosystem services ( Montoya and Raffaelli, 2010 ; Birkhofer and Wolters, 2012 ; Diehl et al., 2013 ). Predictions of future changes will only be possible if studies address this aspect by using mechanistic models (e.g., Schröter et al., 2005 ; Jönsson et al., 2014a ). For example, recommendations about forest management under a changing climate can be based on a dynamic vegetation model that uses basic characteristics of tree growth to predict consequences of alternative silvicultural regimes ( Jönsson et al., 2014b ). However, mechanistic models are never better than the theories and empirical data underpinning them and the development of models with predictive power is a challenge for ecologists.

The quantification of uncertainty in predictive modeling requires critical evaluation ( Cheaib et al., 2012 ) and ecologists need to account for uncertainty particularly if (i) multiple sources contribute to uncertainty (e.g., model and parameter uncertainty; Barry and Elith, 2006 ), (ii) uncertainties result from combinations of different sources (e.g., statistical relationships and expert knowledge; Krueger et al., 2012 ) and (iii) new information necessitates an update of the models (e.g., in Bayesian frameworks; Ricketts et al., 2008 ). Mapping of ecosystem services is fraught with multiple uncertainties stemming from uncertainty in the ability to capture relevant processes as well as translating and scaling mapped information ( Hou et al., 2013 ). The evaluation of uncertainty, the integration of knowledge about evolutionary aspects and human impacts into the development of process-based models and their coupling with socio-economic models are important fields of future research to which ecologists need to contribute (e.g., Polce et al., 2013 ; Van der Biest et al., 2014 ).

Challenge 3: Analyzing Relationships between Ecosystem Services

Ecosystem services may demonstrate joint variation, either synergistic or antagonistic, in space and time. The interpretation of such patterns between multiple ecosystem services (more than two) has become an intensively debated subject ( Cimon-Morin et al., 2013 ) and multi-ecosystem service models that link service provision and trade-offs are rapidly emerging (for a review see Nelson and Daily, 2010 ). Such joint variation may also concern relationships between beneficial ecosystem services and so called ecosystem disservices, for example environmental externalities such as water pollution ( Zhang et al., 2007 ). Ecologists can contribute to the analyses of joint variation of services and disservices by identifying the underlying mechanisms that explain relationships between services and their response patterns to environmental change. For instance, the marginal contribution of enhancing pollination on crop yield may partly depend on the level of other ecosystem services, with highest yield under a simultaneous increase of pollination and biological control ( Bos et al., 2007 ; Lundin et al., 2013 ). Rodríguez et al. (2006) and Bennett et al. (2009) argued that it will only be possible to make informed decisions and avoid unexpected outcomes if relationships between services are better understood. Alterations of a single ecosystem service by agricultural management can, for example, have unintended effects on other services and a better understanding of such unexpected relationships will safeguard human societies against the consequences of sudden regime-shifts in ecosystems (e.g., Gordon et al., 2008 ).

Improving the understanding of the relationships between ecosystem services poses two major challenges to ecological research: (i) drawing conclusions about relationships between ecosystem services by understanding if relationships are indirect through shared environmental drivers or direct because one ecosystem services causally affects another and (ii) solving issues of visualization and statistical testing when analyzing relationships between multiple (more than two) ecosystem services.

To be able to predict the consequences of environmental change as drivers of changes in ecosystem services, it is important to distinguish between indirect and direct relationships ( Bennett et al., 2009 ; Lautenbach et al., 2010 ). Both direct relationships (if services are related to each other) and indirect relationships (if services are related through a driver) can lead to synergies and trade-offs between the services ( Bennett et al., 2009 ). Ecosystem services may be directly and causally linked, because one ecosystem service directly interacts with another ecosystem services (Figure 1A , direct relationships). For example, fertility of agricultural soils (service 1) is directly and positively linked to crop yields (service 2; Lal, 2005 ). Given this direct relationship and assuming the absence of other driving forces, a manipulation of one service (e.g., increase soil fertility by adding manure) would directly increase or decrease the second service (e.g., increase crop yield). However, ecosystem services may be statistically associated, negatively or positively, because their underlying drivers are related (Figure 1B ; indirect relationship). Water retention (service 1) and landscape beauty (service 2), for example, may be statistically associated, because the proportion of urban area that reduces water retention (driver 1) may be negatively related to the proportion of semi-natural land (driver 2; Raudsepp-Hearne et al., 2010 ) that increases landscape beauty. In this case, manipulation of one service (e.g., increasing water retention by leaving out joints between paving stones in urban areas), will not affect the second service (e.g., landscape beauty). In contrast, manipulation of one driver (e.g., reduce the area of semi-natural land), may affect the second driver (e.g., increase the proportion of urban areas) in the absence of other drivers (e.g., other land-use types) and thereby affect both services (e.g., reduce water retention and landscape beauty). Finally, services may also be correlated because of independent responses to a common driver (Figure 1C , indirect relationship). Pollination of crop plants (service 1) and pest control in crop fields (service 2), for example, are both increased by the proportion of semi-natural habitats surrounding crop fields (driver 1; Bianchi et al., 2006 ; Garibaldi et al., 2011 ). In addition, pollination is affected by the proportion of nesting habitats (driver 2; Ricketts et al., 2008 ). Given this indirect relationship, increasing one service (e.g., pest control by augmentation of natural enemies) would not affect the other service (e.g., pollination). Manipulating the shared driver (e.g., increasing the availability of semi-natural habitats around a focal field by sowing flowering strips), will increase both services (e.g., pollination and pest control), while affecting the non-shared driver (e.g., availability of nesting habitats) will only affect one service (e.g., pollination).

www.frontiersin.org

Figure 1. Potential relationships between two ecosystem services (service 1 and service 2). (A) Direct relationships between ecosystem services and indirect relationships via (B) two associated drivers or (C) a shared driver.

In the literature, both types of relationships are frequently labeled “interactions” independent of their correlative or causal nature ( Seppelt et al., 2011 ). Services that show comparable or contrasting responses are then characterized in terms of synergies or trade-offs and grouped as “bundles” ( Raudsepp-Hearne et al., 2010 ). It is without doubt important to describe relationships between multiple services independent of what causes statistical associations ( Tallis et al., 2008 ; Power, 2010 ; Maskell et al., 2013 ). However, the ability to manage situations in which multiple drivers act on multiple services would benefit from an improved understanding of the relationships between individual services (indirect or direct), their relationships to drivers and the processes that affect both relationships ( Lautenbach et al., 2010 ). To manage ecosystem service provisioning, planners and decision-makers need to know if ecosystem services respond to a shared driver or if services are directly linked to each other. If services respond independently, but contrastingly to a single shared driver, better ecological understanding of the individual relationships between the driver and the services will help to identify management strategies that mitigate trade-offs between services. If services are directly linked to each other, improving management becomes more complicated as in addition to the relationship between services and the driver, interactions between services need to be considered. We therefore encourage ecologists to not only investigate the relationship between services and various drivers, but to also test for direct relationships between multiple ecosystem services. Conclusions about direct links between ecosystem services can be derived from studies using large, replicated datasets in approaches that implicitly model direct and indirect effects of anthropogenic interventions on service provision (e.g., structural equation models, Gamfeldt et al., 2013 ), but also from direct experimental tests of ecosystem service relationships (e.g., Lundin et al., 2013 ). Together, these approaches, coupled with the development of mechanistic models (e.g., InVEST model, http://www.naturalcapitalproject.org ), will contribute to an improved management of ecosystems for the provision of multiple services in the future ( Tixier et al., 2013 ).

To study relationships between two or three ecosystem services techniques such as correlation analysis (Figure 2A ; e.g., Raudsepp-Hearne et al., 2010 ) or linear mixed effect models (Figure 2B ; e.g., De Vries et al., 2013 ) can be used. Efficiency frontier analyses ( Nelson et al., 2008 ) or landscape optimization approaches ( Lautenbach et al., 2010 ) are then often used to identify solutions for the simultaneous provision of services. It may be important to consider multiple services in the same analytical framework, as it is likely that most services observed in a study are related to each other. Simple spider web or flower diagrams can be used to illustrate relationships between several services (Figure 2C ; e.g., Foley et al., 2005 ). For the purpose of relating multiple services to drivers in a single analytical framework, the frequent use of principal component analysis is notable (Figure 2D ; e.g., Raudsepp-Hearne et al., 2010 ; Maes et al., 2011 ; Maskell et al., 2013 ; Martín-López et al., 2014 ). However, since relationships between ecosystem services in response to a driver can be non-linear, asymptotic, unimodal or characterized by tipping points (e.g., Maskell et al., 2013 ), it should be noted that the quality of principal component analysis entirely depends on if relationships between variables are linear ( McCune et al., 2002 ). The use of this method should therefore be constraint to datasets in which relationships between multiple services are approximately linear (see also Quinn and Keough, 2002 ). Among the alternative methods to visualize trade-offs between multiple services in one analytical framework, principal coordinates analysis holds potential as it allows for the analysis of non-linear relationships ( Legendre and Legendre, 2012 ). Partial least squares regression is another promising technique to analyze relationships between intermediate ecosystem services based on empirical data (e.g., Haenlein and Kaplan, 2004 ).

www.frontiersin.org

Figure 2. Common approaches to study relationships between ecosystem services with (A) correlation analysis between two services (ES1, ES2), (B) a three dimensional surface plot that shows the modeled relationship between three services (ES1, ES2, ES3), (C) a spider web plot showing the relationship between six services in two land-use types (different colors; ES1, ES2, ES3, ES4, ES5, ES6) and (D) a principal component analysis showing the relationship between six services (vectors; ES1, ES2, ES3, ES4, ES5, ES6) in the ordination graph for sites of two different land-use types (different colors) .

Challenge 4: Considering Appropriate Spatial and Temporal Scales

Scale is a contentious issue in ecosystem service research, because ecological processes are fundamentally scale dependent ( Levin, 1992 ) and a large number of diverging approaches to study spatial scales in ecological research adds to this complexity (e.g., Blackburn and Gaston, 2002 ). This potentially impedes the integration of different research fields (e.g., Lima and Zollner, 1996 ) particularly in a multidisciplinary context such as ecosystem service research ( Cumming et al., 2013 ). Compared to spatial scales, temporal aspects have received remarkably little attention in ecosystem service research ( Kremen, 2005 ). Most of the existing ecological knowledge on ecosystem processes is based on investigations covering short periods of time (e.g., Cardinale et al., 2009 ). A better understanding of the (i) spatial and (ii) temporal scales at which the provision of ecosystem services is affected by environmental change or anthropogenic interventions is needed to satisfy the growing public and political demand for sustainable land use ( Tilman et al., 2002 ).

It is a major challenge for ecologists to scale up from experimental plots to scales that are relevant for the management of most ecosystem services ( Cardinale et al., 2012 ; Figure 3 ). These plot-level studies can often not account for the heterogeneity of complex landscapes and therefore may not provide adequate empirical data about ecosystem service provision from major land-use types in a landscape. Studies relating biodiversity to ecosystem services often focus on ecological processes and intermediate ecosystem services at small spatial scales ( Cardinale et al., 2012 ). In contrast, research on spatial prioritizations for the protection of ecosystem services ( Luck et al., 2012 ) is biased toward services providing benefits at large spatial scales. This suggests that there is a mismatch between species-oriented ecological research dealing with mechanisms underlying the provision of services, and conservation-oriented research identifying hotspots in space for particular services (Figure 3 ). Coupling these two research approaches is critically important to improve the understanding of ecosystem service provision across real landscapes.

www.frontiersin.org

Figure 3. Number of published cases analyzing (A) regulating and supporting ecosystem services with respect to biodiversity-ecosystem service relationships, and (B) spatial prioritization of ecosystem services . Number of cases based on number of syntheses presented in Cardinale et al. (2012) (A) and number of assessments in Luck et al. (2012) (B) respectively. Ecosystem services are classified according to the spatial scale at which ecological processes are linked to human benefits (local, regional or national/global; Kremen, 2005 ; Hein et al., 2006 ). The dark part of bars represents the most dominant ecosystem service among all cases/studies at a particular scale.

The scale of ecological processes is relevant to ecosystem service research, because of the need to account for the spatial relationship between generation and consumption of ecosystem services ( Fisher et al., 2009 ). One approach to account for this goal is based on spatially explicit modeling of ecological production functions (e.g., Kremen et al., 2007 ; Nelson et al., 2009 ; Jonsson et al., 2014 ). In addition, ecological functions underlying final ecosystem services may depend on the spatial scale at which management is applied (cf. Leibold et al., 2004 ). This generates context dependent responses of management interventions ( Tscharntke et al., 2012 ). For example, populations of service-providing units may only maintain viable sizes given that enough habitats are preserved across multiple land-owners ( Drechsler et al., 2010 ). Hence, to optimize ecosystem service provision at larger spatial scales, the identification of conditions under which land-owners benefit from co-operation will be an important future topic in ecosystem services research (e.g., Stallman, 2011 ; Sutherland et al., 2012 ; Cong et al., 2014 ). In addition, ecological research needs to cover the relevant spatial scales at which multiple ecosystem services are efficiently managed (see also Mastrangelo et al., 2014 ). Scaling up models for individual ecosystem services in space is certainly one of the major challenges ( Stuart and Gillon, 2013 ), but it is also crucial to account for relationships between services that are caused by interactions between services or anthropogenic interventions at different spatial scales (e.g., management by farmers at local scales and policy makers at broader scales, Tixier et al., 2013 ; see also challenge 3.3).

It is essential to understand the temporal dynamics of service provision for the development of sustainable management and conservation strategies. For example, the quality of provision of an ecosystem service may not only depend on its average provision over time, but also on its variation over time ( Mori et al., 2013 ). It is therefore important to assess the stability of ecosystem service provision in simplified ecosystems, where losses of ecosystem resilience to disturbances can be expected to be strongest ( Bengtsson et al., 2003 ; Tscharntke et al., 2012 ). In addition, lag-effects of management decisions may make ecosystem service losses only apparent a long time after the anthropogenic intervention ( Millennium Ecosystem Assessment, 2005 ). Such lag-effects may be further accentuated by climate change, where loss of biodiversity may reduce resilience of critical functions (cf. Elmqvist et al., 2003 ). We therefore need long-term estimates of ecosystem service provision to better understand how inter-annual variation in environmental conditions, such as climate change, affects the magnitude and stability of service provision. However, the time-span of ecological research is often constrained to a few years due to generally short funding periods. Such short research periods will fail to provide reliable estimates of altered behavior of service-providing units in response to climate change (e.g., Mooney et al., 2009 ). The few long-term studies, such as the Cedar Creek experiment in the US ( Siemann et al., 1998 ) or the Biodiversity Exploratories in Germany ( Fischer et al., 2010 ), deliver fundamental insights into biodiversity and ecosystem functioning over longer temporal scales. We call for more such approaches to get a better understanding of both long-term changes and temporal variability of ecosystem service provision.

Concluding Remarks

Although the ecosystem service concept is based on an ecological understanding of ecosystems, ecologists are confronted with a range of challenges when researching ecosystem services. This is partly explained by the wide variety of terms and definitions from different scientific disciplines as well as a lack of generally accepted assessment methods, difficulties with analytical and modeling methods and mismatches of spatial and temporal scales between service provision and anthropogenic interventions. Ecologists need to adapt their perspective and methods to a larger societal context for the improvement of ecosystem service research. Particular emphasis needs to be directed toward supporting decision makers with relevant information about service-providing units and mechanisms underlying the provision of services at appropriate temporal and spatial scales. To conclude, ecosystem service research is challenging for ecologists, but developing a multifaceted understanding of how nature promotes human well-being is crucial for the sustainable use of the earth's resources. Ecosystem service research offers ecologists the unique opportunity to act as promoters for the understanding of how to conserve and sustain benefits gained from nature.

Author Contributions

All authors contributed to the manuscript by reviewing literature, discussing and developing ideas during two workshops, writing text sections and revising sections written by other authors. Eva Diehl and Klaus Birkhofer had the initial idea for the manuscript and wrote general parts of the manuscript (abstract, introduction and conclusion) together with Henrik G. Smith. Henrik G. Smith and Volkmar Wolters contributed to all sections in the manuscript. Andrea Früh-Müller contributed to challenge 4. Eva Diehl contributed to challenges 1-3. Franziska Machnikowski contributed to the introduction and challenge 4. Jesper Andersson contributed to the introduction. Johan Ekroos contributed to challenge 4. Klaus Birkhofer contributed to challenges 2-3. Keiko Sasaki contributed to the introduction and challenges 2-3. Lovisa Nilsson contributed to challenges 1 and 4. Maj Rundlöf contributed to the introduction and challenge 2. Viktoria L. Mader contributed to challenges 1-2.

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

We thank Johan G. Zaller, Quang Bao Le and Visa Nuutinen for very helpful comments that helped improving earlier versions of the manuscript. The strategic research environment BECC and the strong research environment SAPES funded by the Swedish research council Formas are greatly acknowledged for financial support. This article was written within the scope of the “ÖkoService” project and the “JAGUAR” project funded by the German Federal Ministry of Education and Research (BMBF).

Barry, S., and Elith, J. (2006). Error and uncertainty in habitat models. J. Appl. Ecol . 43, 413–423. doi: 10.1111/j.1365-2664.2006.01136.x

CrossRef Full Text | Google Scholar

Bastian, O., Grunewald, K., and Sybre, R. U. (2013). “Klassifikation von ÖSD,” in Ökosystemdienstleistungen: Konzept, Methoden und Fallbeispiele , eds K. Grunewald and O. Bastian (Berlin: Springer Spektrum), 48–56.

Google Scholar

Bengtsson, J., Angelstam, P., Elmqvist, T., Emanuelsson, U., Folke, C., Ihse, M., et al. (2003). Reserves, resilience and dynamic landscapes. Ambio 32, 389–396. doi: 10.1579/0044-7447-32.6.389

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

Bennett, E. M., Peterson, G. D., and Gordon, L. J. (2009). Understanding relationships among multiple ecosystem services. Ecol. Lett . 12, 1394–1404. doi: 10.1111/j.1461-0248.2009.01387.x

Bianchi, F., Booij, C., and Tscharntke, T. (2006). Sustainable pest regulation in agricultural landscapes: a review on landscape composition, biodiversity and natural pest control. Proc. Biol. Sci . 273, 1715–1727. doi: 10.1098/rspb.2006.3530

Birkhofer, K., Bezemer, T. M., Bloem, J., Bonkowski, M., Christensen, S., Dubois, D., et al. (2008). Long-term organic farming fosters below and aboveground biota: implications for soil quality, biological control and productivity. Soil Biol. Biochem . 40, 2297–2308. doi: 10.1016/j.soilbio.2008.05.007

Birkhofer, K., and Wolters, V. (2012). The global relationship between climate, net primary production and the diet of spiders. Glob. Ecol. Biogeogr . 21, 100–108. doi: 10.1111/j.1466-8238.2011.00654.x

Blackburn, T. M., and Gaston, K. J. (2002). Scale in macroecology. Glob. Ecol. Biogeogr . 11, 185–189. doi: 10.1046/j.1466-822X.2002.00290.x

Bolund, P., and Hunhammar, S. (1999). Ecosystem services in urban areas. Ecol. Econ . 29, 293–301. doi: 10.1016/S0921-8009(99)00013-0

Bommarco, R., Kleijn, D., and Potts, S. G. (2013). Ecological intensification: harnessing ecosystem services for food security. Trends Ecol. Evol . 28, 230–238. doi: 10.1016/j.tree.2012.10.012

Bos, M. B., Veddeler, D., Bogdanski, A. K., Klein, A. M., Tscharntke, T., Steffan-Dewenter, I., et al. (2007). Caveats to quantifying ecosystem services: fruit abortion blurs benefits from crop pollination. Ecol. Appl . 17, 1841–1849. doi: 10.1890/06-1763.1

Boyd, J., and Banzhaf, S. (2007). What are ecosystem services? The need for standardized environmental accounting units. Ecol. Econ . 63, 616–626. doi: 10.1016/j.ecolecon.2007.01.002

Braat, L. C., and De Groot, R. (2012). The ecosystem services agenda: bridging the worlds of natural science and economics, conservation and development, and public and private policy. Ecosys. Serv . 1, 4–15. doi: 10.1016/j.ecoser.2012.07.011

Bruno, J. F., and Cardinale, B. J. (2008). Cascading effects of predator richness. Front. Ecol. Environ . 6, 539–546. doi: 10.1890/070136

Cardinale, B., Duffy, J. E., Srivastava, D., Loreau, M., Thomas, M., and Emmerson, M. (2009). “Towards a food web perspective on biodiversity and ecosystem functioning,” in Biodiversity, Ecosystem Functioning, and Human Wellbeing. An Ecological and Economic Perspective , ed S. Naeem (Oxford, New York: Oxford University Press), 105–120. doi: 10.1093/acprof:oso/9780199547951.003.0008

CrossRef Full Text

Cardinale, B. J., Duffy, J. E., Gonzalez, A., Hooper, D. U., Perrings, C., Venail, P., et al. (2012). Biodiversity loss and its impact on humanity. Nature 486, 59–67. doi: 10.1038/nature11148

Carpenter, S. R., Mooney, H. A., Agard, J., Capistrano, D., DeFries, R. S., Díaz, S., et al. (2009). Science for managing ecosystem services: beyond the millennium ecosystem assessment. Proc. Natl. Acad. Sci. U.S.A . 106, 1305–1312. doi: 10.1073/pnas.0808772106

CBD. (2010). COP 10 Decision X/2: The Strategic Plan for Biodiversity 2011-2020 and the Aichi Biodiversity Targets . Available online at: http://www.cbd.int/decision/cop/?id=12268

Chaplin-Kramer, R., O'Rourke, M. E., Blitzer, E. J., and Kremen, C. (2011). A meta-analysis of crop pest and natural enemy response to landscape complexity. Ecol. Lett . 14, 922–932. doi: 10.1111/j.1461-0248.2011.01642.x

Cheaib, A., Badeau, V., Boe, J., Chuine, I., Delire, C., Dufrêne, E., et al. (2012). Climate change impacts on tree ranges: model intercomparison facilitates understanding and quantification of uncertainty. Ecol. Lett . 15, 533–544. doi: 10.1111/j.1461-0248.2012.01764.x

Cimon-Morin, J., Darveau, M., and Poulin, M. (2013). Fostering synergies between ecosystem services and biodiversity in conservation planning: a review. Biol. Conserv . 166, 144–154. doi: 10.1016/j.biocon.2013.06.023

Cong, R. G., Smith, H. G., Olsson, O., and Brady, M. (2014). Managing ecosystem services for agriculture: will landscape-scale management pay? Ecol. Econ . 99, 53–62. doi: 10.1016/j.ecolecon.2014.01.007

Cortesero, A. M., Stapel, J. O., and Lewis, W. J. (2000). Understanding and manipulating plant attributes to enhance biological control. Biol. Control 17, 35–49. doi: 10.1006/bcon.1999.0777

Costanza, R., d'Arge, R., De Groot, R., Farberk, S., Grasso, M., Limburg, K., et al. (1997). The value of the world's ecosystem services and natural capital. Nature 387, 253–260.

Cumming, G. S., Olsson, P., Chapin, F. S. III, and Holling, C. S. (2013). Resilience, experimentation, and scale mismatches in social-ecological landscapes. Landscape Ecol . 28, 1139–1150. doi: 10.1007/s10980-012-9725-4

Daily, G. C. (1997). Nature's Services. Societal Dependence on Natural Ecosystems . Washington, DC: Island Press.

Daily, G. C., Polasky, S., Goldstein, J., Kareiva, P. M., Mooney, H. A., Pejchar, L., et al. (2009). Ecosystem services in decision making: time to deliver. Front. Ecol. Environ . 7, 21–28. doi: 10.1890/080025

De Groot, R. (2006). Function-analysis and valuation as a tool to assess land use conflicts in planning for sustainable, multi-functional landscapes. Landsc. Urban Plan . 75, 175–186. doi: 10.1016/j.landurbplan.2005.02.016

De Groot, R., Alkemade, R., Braat, L., Hein, L., and Willemen, L. (2010). Challenges in integrating the concept of ecosystem services and values in landscape planning, management and decision making. Ecol. Complex . 7, 260–272. doi: 10.1016/j.ecocom.2009.10.006

De Groot, R., Wilson, M. A., and Boumans, R. M. (2002). A typology for the classification, description and valuation of ecosystem functions, goods and services. Ecol. Econ . 41, 393–408. doi: 10.1016/S0921-8009(02)00089-7

De Vries, F. T., Thebault, E., Liiri, M., Birkhofer, K., Tsiafouli, M. A., Bjornlund, L., et al. (2013). Soil food web properties explain ecosystem services across European land use systems. Proc. Natl. Acad. Sci. U.S.A . 110, 14296–14301. doi: 10.1073/pnas.1305198110

Diehl, E., Sereda, E., Wolters, V., and Birkhofer, K. (2013). Effects of predator specialization, host plant and climate on biological control of aphids by natural enemies: a meta-analysis. J. Appl. Ecol . 50, 262–270. doi: 10.1111/1365-2664.12032

Drechsler, M., Wätzold, F., Johst, K., and Shogren, J. F. (2010). An agglomeration payment for cost-effective biodiversity conservation in spatially structured landscapes. Resour. Energy Econ . 32, 261–275. doi: 10.1016/j.reseneeco.2009.11.015

Duelli, P., and Obrist, M. K. (2003). Biodiversity indicators: the choice of values and measures. Agric. Ecosyst. Environ . 98, 87–98. doi: 10.1016/S0167-8809(03)00072-0

Millennium Ecosystem Assessment. (2005). Ecosystems and Human Well-being. General Synthesis: a Report of the Millennium Ecosystem Assessment . Washington, DC: Island Press.

Ehrlich, P. R., and Ehrlich, A. H. (1981). Extinction. The Causes and Consequences of the Disappearance of Species . New York, NY: Random House.

Ekroos, J., Olsson, O., Rundlöf, M., Wätzold, F., and Smith, H. G. (2014). Optimizing agri-environment schemes for biodiversity, ecosystem services or both? Biol.Conserv . 172, 65–71. doi: 10.1016/j.biocon.2014.02.013

Elmqvist, T., Folke, C., Nystrom, M., Peterson, G., Bengtsson, J., Walker, B., et al. (2003). Response diversity, ecosystem change, and resilience. Front. Ecol. Environ . 1, 488–494. doi: 10.1890/1540-9295(2003)001[0488:RDECAR]2.0.CO;2

Engel, S., Pagiola, S., and Wunder, S. (2008). Designing payments for environmental services in theory and practice: an overview of the issues. Ecol. Econ . 65, 663–674. doi: 10.1016/j.ecolecon.2008.03.011

European Commission. (2014). Mapping and Assessment of Ecosystems and their Services – Indicators for Ecosystem Assessments under Action 5 of the EU Biodiversity Strategy to 2020 .

Evans, M. R., Norris, K. J., and Benton, T. G. (2012). Predictive ecology: systems approaches. Philos. Trans. R. Soc. B 367, 163–169. doi: 10.1098/rstb.2011.0191

FAOSTAT. Food and Agriculture Organization of the United Nations 2014. Resources – Land (Land area 2011). The Statistics Division of the FAO . Available online at: http://faostat.fao.org/

Feld, C. K., Martins da Silva, P., Paulo Sousa, J., De Bello, F., Bugter, R., Grandin, U., et al. (2009). Indicators of biodiversity and ecosystem services: a synthesis across ecosystems and spatial scales. Oikos 118, 1862–1871. doi: 10.1111/j.1600-0706.2009.17860.x

Fischer, M., Bossdorf, O., Gockel, S., Hänsel, F., Hemp, A., Hessenmöller, D., et al. (2010). Implementing large-scale and long-term functional biodiversity research: the biodiversity exploratories. Basic Appl. Ecol . 11, 473–485. doi: 10.1016/j.baae.2010.07.009

Fisher, B., Turner, R. K., and Morling, P. (2009). Defining and classifying ecosystem services for decision making. Ecol. Econ . 68, 643–653. doi: 10.1016/j.ecolecon.2008.09.014

Foley, J. A., DeFries, R., Asner, G. P., Barford, C., Bonan, G., Carpenter, S. R., et al. (2005). Global consequences of land use. Science 309, 570–574. doi: 10.1126/science.1111772

Früh-Müller, A. J., Wegmann, M., and Koellner, T. (2014). Flood exposure and settlement expansion since pre-industrial times in 1850 until 2011 in north Bavaria, Germany. Reg. Environ. Change . doi: 10.1007/s10113-014-0657-1

Gamfeldt, L., Snäll, T., Bagchi, R., Jonsson, M., Gustafsson, L., Kjellander, P., et al. (2013). Higher levels of multiple ecosystem services are found in forests with more tree species. Nat. Commun . 4, 1340. doi: 10.1038/ncomms2328

Garibaldi, L. A., Steffan-Dewenter, I., Kremen, C., Morales, J. M., Bommarco, R., Cunningham, S. A., et al. (2011). Stability of pollination services decreases with isolation from natural areas despite honey bee visits. Ecol. Lett . 14, 1062–1072. doi: 10.1111/j.1461-0248.2011.01669.x

Garibaldi, L. A., Steffan-Dewenter, I., Winfree, R., Aizen, M. A., Bommarco, R., Cunningham, S. A., et al. (2013). Wild pollinators enhance fruit set of crops regardless of honey bee abundance. Science 339, 1608–1611. doi: 10.1126/science.1230200

Geiger, F., Bengtsson, J., Berendse, F., Weisser, W. W., Emmerson, M., Morales, M. B., et al. (2010). Persistent negative effects of pesticides on biodiversity and biological control potential on European farmland. Basic Appl. Ecol . 11, 97–105. doi: 10.1016/j.baae.2009.12.001

Gollan, J. R., De Bruyn, L. L., Reid, N., and Wilkie, L. (2013). Monitoring the ecosystem service provided by dung beetles offers benefits over commonly used biodiversity metrics and a traditional trapping method. J. Nat. Conserv . 21, 183–188. doi: 10.1016/j.jnc.2012.12.004

Gordon, L. J., Peterson, G. D., and Bennett, E. M. (2008). Agricultural modifications of hydrological flows create ecological surprises. Trends Ecol. Evol . 23, 211–219. doi: 10.1016/j.tree.2007.11.011

Haenlein, M., and Kaplan, A. M. (2004). A beginner's guide to partial least squares analysis. Understand. Stat . 3, 283–297. doi: 10.1207/s15328031us0304_4

Hails, R. S., and Ormerod, S. J. (2013). Editorial: ecological science for ecosystem services and the stewardship of natural capital. J. Appl. Ecol . 50, 807–810. doi: 10.1111/1365-2664.12127

Haines-Young, R., and Potschin, M. (2010). “The links between biodiversity, ecosystem services and human well-being,” in Ecosystem Ecology. A New Synthesis , eds D. G. Raffaelli and C. Frid (Cambridge, New York: Cambridge University Press), 110–139.

Hauck, J., Görg, C., Varjopuro, R., Ratamäki, O., and Jax, K. (2013). Benefits and limitations of the ecosystem services concept in environmental policy and decision making: some stakeholder perspectives. Environ. Sci. Policy 25, 13–21. doi: 10.1016/j.envsci.2012.08.001

Hein, L., van Koppen, K., De Groot, R., and van Ierland, E. C. (2006). Spatial scales, stakeholders and the valuation of ecosystem services. Ecol. Econ . 57, 209–228. doi: 10.1016/j.ecolecon.2005.04.005

Hou, Y., Burkhard, B., and Müller, F. (2013). Uncertainties in landscape analysis and ecosystem service assessment. J. Environ. Manag . 127, S117. doi: 10.1016/j.jenvman.2012.12.002

Ingegno, B. L., Ferracini, C., Gallinotti, D., Alma, A., and Tavella, L. (2013). Evaluation of the effectiveness of Dicyphus errans (Wolff) as predator of Tuta absoluta (Meyrick). Biol. Control 67, 246–252. doi: 10.1016/j.biocontrol.2013.08.002

Jauker, F., Diekötter, T., Schwarzbach, F., and Wolters, V. (2009). Pollinator dispersal in an agricultural matrix: opposing responses of wild bees and hoverflies to landscape structure and distance from main habitat. Landscape Ecol . 24, 547–555. doi: 10.1007/s10980-009-9331-2

Jonason, D., Smith, H. G., Bengtsson, J., and Birkhofer, K. (2013). Landscape simplification promotes weed seed predation by carabid beetles ( Coleoptera; Carabidae ). Landscape Ecol . 28, 478–494. doi: 10.1007/s10980-013-9848-2

Jönsson, A. M., Anderbrant, O., Holmér, J., Johansson, J., Schurgers, G., Svensson, G. P., et al. (2014a). Enhanced science-stakeholder communication to improve ecosystem model performance for climate change impact assessments. Ambio . doi: 10.1007/s13280-014-0553-4. [Epub ahead of print].

Jönsson, A. M., Lagergren, F., and Smith, B. (2014b). Forest management facing climate change – an ecosystem model analysis of adaptation strategies. Mitig. Adapt. Strateg. Glob. Change . doi: 10.1007/s11027-013-9487-6

Jonsson, M., Bommarco, R., Ekbom, B., Smith, H. G., Bengtsson, J., Caballero-Lopez, B., et al. (2014). Ecological production functions for biological control services in agricultural landscapes. Method. Ecol. Evol . 5, 243–252. doi: 10.1111/2041-210X.12149

Kandziora, M., Burkhard, B., and Müller, F. (2013). Interactions of ecosystem properties, ecosystem integrity and ecosystem service indicators – A theoretical matrix exercise. Ecol. Indic . 28, 54–78. doi: 10.1016/j.ecolind.2012.09.006

Kennedy, C. M., Lonsdorf, E., Neel, M. C., Williams, N. M., Ricketts, T. H., Winfree, R., et al. (2013). A global quantitative synthesis of local and landscape effects on wild bee pollinators in agroecosystems. Ecol. Lett . 16, 584–599. doi: 10.1111/ele.12082

Kowarik, I. (2011). Novel urban ecosystems, biodiversity, and conservation. Environ. Pollut . 159, 1974–1983. doi: 10.1016/j.envpol.2011.02.022

Kremen, C. (2005). Managing ecosystem services: what do we need to know about their ecology? Ecol. Lett . 8, 468–479. doi: 10.1111/j.1461-0248.2005.00751.x

Kremen, C., Williams, N. M., Aizen, M. A., Gemmill-Herren, B., LeBuhn, G., Minckley, R., et al. (2007). Pollination and other ecosystem services produced by mobile organisms: a conceptual framework for the effects of land-use change. Ecol. Lett . 10, 299–314. doi: 10.1111/j.1461-0248.2007.01018.x

Kremen, C., Williams, N. M., and Thorp, R. W. (2002). Crop pollination from native bees at risk from agricultural intensification. Proc. Natl. Acad. Sci. U.S.A . 99, 16812–16816. doi: 10.1073/pnas.262413599

Krueger, T., Page, T., Hubacek, K., Smith, L., and Hiscock, K. (2012). The role of expert opinion in environmental modelling. Environ. Model. Softw . 36, 4–18. doi: 10.1016/j.envsoft.2012.01.011

Lal, R. (2005). Enhancing crop yields in the developing countries through restoration of the soil organic carbon pool in agricultural lands. Land Degrad. Dev . 17, 197–209. doi: 10.1002/ldr.696

Langellotto, G. A., and Denno, R. F. (2004). Responses of invertebrate natural enemies to complex-structured habitats: a meta-analytical synthesis. Oecologia 139, 1–10. doi: 10.1007/s00442-004-1497-3

Lautenbach, S., Volk, M., Gruber, B., Dormann, C. F., Strauch, M., and Seppelt, R. (2010). “Quantifying ecosystem service trade-offs,” in International Congress on Environmental Modelling and Software Modelling for Environment's Sake, Fifth Biennial Meeting , eds D. A. Swayne, W. Yang, A. A. Voinov, A. Rizzoli, and T. Filatova (Ottawa, ON: International Environmental Modelling and Software Society (iEMSs)).

Legendre, P., and Legendre, L. (2012). Numerical Ecology . Amsterdam, Boston: Elsevier.

Leibold, M. A., Holyoak, M., Mouquet, N., Amarasekare, P., Chase, J. M., Hoopes, M. F., et al. (2004). The metacommunity concept: a framework for multi-scale community ecology. Ecol. Lett . 7, 601–613. doi: 10.1111/j.1461-0248.2004.00608.x

Letourneau, D. K., Jedlicka, J. A., Bothwell, S. G., and Moreno, C. R. (2009). Effects of natural enemy biodiversity on the suppression of arthropod herbivores in terrestrial ecosystems. Annu. Rev. Ecol. Evol. Syst . 40, 573–592. doi: 10.1146/annurev.ecolsys.110308.120320

Levin, S. A. (1992). The problem of pattern and scale in ecology. Ecology 73, 1943–1967. doi: 10.2307/1941447

Lima, S. L., and Zollner, P. A. (1996). Towards a behavioral ecology of ecological landscapes. Trends Ecol. Evol . 11, 131–135. doi: 10.1016/0169-5347(96)81094-9

Luck, G. W., Chan, K. M., and Klien, C. J. (2012). Identifying spatial priorities for protecting ecosystem services. F1000 Res . 1:17 doi: 10.12688/f1000research.1-17.v1

Luck, G. W., Daily, G. C., and Ehrlich, P. R. (2003). Population diversity and ecosystem services. Trends Ecol. Evol . 18, 331–336. doi: 10.1016/S0169-5347(03)00100-9

Lundin, O., Smith, H. G., Rundlöf, M., and Bommarco, R. (2013). When ecosystem services interact: crop pollination benefits depend on the level of pest control. Proc. Biol. Sci . 280:20122243. doi: 10.1098/rspb.2012.2243

Mace, G. M., Norris, K., and Fitter, A. H. (2012). Biodiversity and ecosystem services: a multilayered relationship. Trends Ecol. Evol . 27, 19–26. doi: 10.1016/j.tree.2011.08.006

Maes, J., Paracchini, M. L., and Zulian, G. (2011). A European Assessment of the Provision of Ecosystem Services. Towards an Atlas of Ecosystem Services . Luxembourg: Publications Office.

Martín-López, B., Gómez-Baggethun, E., García-Llorente, M., and Montes, C. (2014). Trade-offs across value-domains in ecosystem services assessment. Ecol. Indic . 37, 220–228. doi: 10.1016/j.ecolind.2013.03.003

Maskell, L. C., Crowe, A., Dunbar, M. J., Emmett, B., Henrys, P., Keith, A. M., et al. (2013). Exploring the ecological constraints to multiple ecosystem service delivery and biodiversity. J. Appl. Ecol . 50, 561–571. doi: 10.1111/1365-2664.12085

Mastrangelo, M. E., Weyland, F., Villarino, S. H., Barral, M. P., Nahuelhual, L., and Lateraa, P. (2014). Concepts and methods for landscape multifunctionality and a unifying framework based on ecosystem services. Landscape Ecol . 29, 345–358. doi: 10.1007/s10980-013-9959-9

McCune, B., Grace, J. B., and Urban, D. L. (2002). Analysis of Ecological Communities . Gleneden Beach, OR: MjM Software Design.

Médiène, S., Valantin-Morison, M., Sarthou, J. P., Tourdonnet, S., Gosme, M., Bertrand, M., et al. (2011). Agroecosystem management and biotic interactions: a review. Agron. Sustain. Dev . 31, 491–514. doi: 10.1007/s13593-011-0009-1

Menalled, F. D., Lee, J. C., and Landis, D. A. (1999). Manipulating carabid beetle abundance alters prey removal rates in corn fields. Biocontrol 43, 441–456. doi: 10.1023/A:1009946004251

Mols, C. M., Visser, M. E., and Jones, P. (2007). Great tits ( Parus major ) reduce caterpillar damage in commercial apple orchards. PloS ONE 2:e202. doi: 10.1371/journal.pone.0000202

Montoya, J. M., and Raffaelli, D. (2010). Climate change, biotic interactions and ecosystem services. Philos. Trans. R. Soc. B 365, 2013–2018. doi: 10.1098/rstb.2010.0114

Mooney, H., Larigauderie, A., Cesario, M., Elmquist, T., Hoegh-Guldberg, O., Lavorel, S., et al. (2009). Biodiversity, climate change, and ecosystem services. Curr. Opin. Environ. Sustain . 1, 46–54. doi: 10.1016/j.cosust.2009.07.006

Mori, A. S., Furukawa, T., and Sasaki, T. (2013). Response diversity determines the resilience of ecosystems to environmental change. Biol. Rev . 88, 349–364. doi: 10.1111/brv.12004

Nelson, E., and Daily, G. C. (2010). Modelling ecosystem services in terrestrial systems. F1000 Biol. Rep . 2:53. doi: 10.3410/B2-53

Nelson, E., Mendoza, G., Regetz, J., Polasky, S., Tallis, H., Cameron, D., et al. (2009). Modeling multiple ecosystem services, biodiversity conservation, commodity production, and tradeoffs at landscape scales. Front. Ecol. Environ . 7, 4–11. doi: 10.1890/080023

Nelson, E., Polasky, S., Lewis, D. J., Plantinga, A. J., Lonsdorf, E., White, D., et al. (2008). Efficiency of incentives to jointly increase carbon sequestration and species conservation on a landscape. Proc. Natl. Acad. Sci. U.S.A . 105, 9471–9476. doi: 10.1073/pnas.0706178105

Nicholson, E., Mace, G. M., Armsworth, P. R., Atkinson, G., Buckle, S., Clements, T., et al. (2009). Priority research areas for ecosystem services in a changing world. J. Appl. Ecol . 46, 1139–1144. doi: 10.1111/j.1365-2664.2009.01716.x

Norgaard, R. B. (2010). Ecosystem services: from eye-opening metaphor to complexity blinder. Ecol. Econ . 69, 1219–1227. doi: 10.1016/j.ecolecon.2009.11.009

Östman, Ö., Ekbom, B., and Bengtsson, J. (2003). Yield increase attributable to aphid predation by ground-living polyphagous natural enemies in spring barley in Sweden. Ecol. Econ . 45, 149–158. doi: 10.1016/S0921-8009(03)00007-7

Polce, C., Termansen, M., Aguirre-Gutiérrez, J., Boatman, N. D., Budge, G. E., Crowe, A., et al. (2013). Species distribution models for crop pollination: a modelling framework applied to Great Britain. PloS ONE 8:e76308. doi: 10.1371/journal.pone.0076308

Power, A. G. (2010). Ecosystem services and agriculture: tradeoffs and synergies. Philos. Trans. Roy. Soc. B 365, 2959–2971. doi: 10.1098/rstb.2010.0143

Quinn, G. P., and Keough, M. J. (2002). Experimental Design and Data Analysis for Biologists . Cambridge, New York: Cambridge University Press. doi: 10.1017/CBO9780511806384

Raffaelli, D., and White, P. C. (2013). Ecosystems and their services in a changing world. Adv. Ecol. Res . 48, 1–70. doi: 10.1016/B978-0-12-417199-2.00001-X

Raudsepp-Hearne, C., Peterson, G. D., and Bennett, E. M. (2010). Ecosystem service bundles for analyzing tradeoffs in diverse landscapes. Proc. Natl. Acad. Sci. U.S.A . 107, 5242–5247. doi: 10.1073/pnas.0907284107

Reyers, B., Biggs, R., Cumming, G. S., Elmqvist, T., Hejnowicz, A. P., and Polasky, S. (2013). Getting the measure of ecosystem services: a social–ecological approach. Front. Ecol. Environ . 11, 268–273. doi: 10.1890/120144

Ricketts, T. H., Regetz, J., Steffan-Dewenter, I., Cunningham, S. A., Kremen, C., Bogdanski, A., et al. (2008). Landscape effects on crop pollination services: are there general patterns? Ecol. Lett . 11, 499–515. doi: 10.1111/j.1461-0248.2008.01157.x

Rodríguez, J. P., Beard, T. D. Jr. Bennett, E. M., Cumming, G. S., Cork, S. J., Agard, J., et al. (2006). Trade-offs across space, time, and ecosystem services. Ecol. Soc . 11, 28.

Rusch, A., Valantin-Morison, M., Roger-Estrade, J., and Sarthou, J. P. (2012). Using landscape indicators to predict high pest infestations and successful natural pest control at the regional scale. Landsc. Urban Plan . 105, 62–73. doi: 10.1016/j.landurbplan.2011.11.021

Schmidt, M. H., Lauer, A., Purtauf, T., Thies, C., Schaefer, M., and Tscharntke, T. (2003). Relative importance of predators and parasitoids for cereal aphid control. Proc. Biol. Sci . 270, 1905–1909. doi: 10.1098/rspb.2003.2469

Schröter, D., Cramer, W., Leemans, R., Prentice, I. C., Araújo, M. B., Arnell, N. W., et al. (2005). Ecosystem service supply and vulnerability to global change in Europe. Science 310, 1333–1337. doi: 10.1126/science.1115233

Seppelt, R., Dormann, C. F., Eppink, F. V., Lautenbach, S., and Schmidt, S. (2011). A quantitative review of ecosystem service studies: approaches, shortcomings and the road ahead. J. Appl. Ecol . 48, 630–636. doi: 10.1111/j.1365-2664.2010.01952.x

Setälä, H., Bardgett, R. D., Birkhofer, K., Brady, M., Byrne, L., De Ruiter, P. C., et al. (2014). Urban and agricultural soils: conflicts and trade-offs in the optimization of ecosystem services. Urban Ecosyst . 17, 239–253. doi: 10.1007/s11252-013-0311-6

Shrestha, R. B., and Parajulee, M. N. (2013). Potential cotton aphid, Aphis gossypii , population suppression by arthropod predators in upland cotton. Insect Sci . 20, 778–788. doi: 10.1111/j.1744-7917.2012.01583.x

Siemann, E., Tilman, D., Haarstad, J., and Ritchie, M. (1998). Experimental tests of the dependence of arthropod diversity on plant diversity. Am. Nat . 152, 738–750. doi: 10.1086/286204

Sih, A., and Baltus, M. S. (1987). Patch size, pollinator behavior, and pollinator limitation in catnip. Am. Nat . 68, 1679–1690.

Stallman, H. R. (2011). Ecosystem services in agriculture: determining suitability for provision by collective management. Ecol. Econ . 71, 131–139. doi: 10.1016/j.ecolecon.2011.08.016

Sterner, T., and Persson, U. M. (2008). An even sterner review - introducing relative prices into the discounting debate. Rev. Environ. Econ. Policy 2, 61–76. doi: 10.1093/reep/rem024

Stuart, D., and Gillon, S. (2013). Scaling up to address new challenges to conservation on US farmland. Land Use Policy 31, 223–236. doi: 10.1016/j.landusepol.2012.07.003

Sutherland, L. A., Gabriel, D., Hathaway-Jenkins, L., Pascual, U., Schmutz, U., Rigby, D., et al. (2012). The ‘Neighbourhood Effect’: a multidisciplinary assessment of the case for farmer co-ordination in agri-environmental programmes. Land Use Policy 29, 502–512. doi: 10.1016/j.landusepol.2011.09.003

Tallis, H., Kareiva, P., Marvier, M., and Chang, A. (2008). An ecosystem services framework to support both practical conservation and economic development. Proc. Natl. Acad. Sci. U.S.A . 105, 9457–9464. doi: 10.1073/pnas.0705797105

TEEB. (2010). The Economics of Ecosystems and Biodiversity: Mainstreaming the Economics of Nature: A Synthesis of the Approach, Conclusions and Recommendations of TEEB . Malta: Progress Press.

Tilman, D., Cassman, K. G., Matson, P. A., Naylor, R., and Polasky, S. (2002). Agricultural sustainability and intensive production practices. Nature 418, 671–677. doi: 10.1038/nature01014

Tixier, P., Peyrard, N., Aubertot, J. N., Gaba, S., Radoszycki, J., Caron-Lormier, G., et al. (2013). Modelling interaction networks for enhanced ecosystem services in agroecosystems. Adv. Ecol. Res . 49, 437–480. doi: 10.1016/B978-0-12-420002-9.00007-X

Tscharntke, T., Klein, A. M., Kruess, A., Steffan-Dewenter, I., and Thies, C. (2005). Landscape perspectives on agricultural intensification and biodiversity – ecosystem service management. Ecol. Lett . 8, 857–874. doi: 10.1111/j.1461-0248.2005.00782.x

Tscharntke, T., Tylianakis, J. M., Rand, T. A., Didham, R. K., Fahrig, L., Batáry, P., et al. (2012). Landscape moderation of biodiversity patterns and processes – eight hypotheses. Biol. Rev . 87, 661–685. doi: 10.1111/j.1469-185X.2011.00216.x

Tylianakis, J. M., Tscharntke, T., and Lewis, O. T. (2007). Habitat modification alters the structure of tropical host–parasitoid food webs. Nature 445, 202–205. doi: 10.1038/nature05429

Van der Biest, K., D'Hondt, R., Jacobs, S., Landuyt, D., Staes, J., Goethals, P., et al. (2014). EBI: an index for delivery of ecosystem service bundles. Ecol. Indic . 37, 252–265. doi: 10.1016/j.ecolind.2013.04.006

Veres, A., Petit, S., Conord, C., and Lavigne, C. (2013). Does landscape composition affect pest abundance and their control by natural enemies? a review. Agric. Ecosyst. Environ . 166, 110–117. doi: 10.1016/j.agee.2011.05.027

Vihervaara, P., Rönkä, M., and Walls, M. (2010). Trends in ecosystem service research: early steps and current drivers. Ambio 39, 314–324. doi: 10.1007/s13280-010-0048-x

Westman, W. E. (1977). How much are nature's services worth? Science 197, 960–964. doi: 10.1126/science.197.4307.960

Woodcock, T. S., Pekkola, L. J., Dawson, C., Gadallah, F. L., and Kevan, P. G. (2014). Development of a Pollination Service Measurement (PSM) method using potted plant phytometry. Environ. Monit. Assess . 186, 5041–5057. doi: 10.1007/s10661-014-3758-x

Zhang, W., Ricketts, T. H., Kremen, C., Carney, K., and Swinton, S. M. (2007). Ecosystem services and dis-services to agriculture. Ecol. Econ . 64, 253–260. doi: 10.1016/j.ecolecon.2007.02.024

Keywords: anthropogenic interventions, biodiversity conservation, ecosystem services, multiple services, service-providing units, spatial scale, temporal scale, trade-offs

Citation: Birkhofer K, Diehl E, Andersson J, Ekroos J, Früh-Müller A, Machnikowski F, Mader VL, Nilsson L, Sasaki K, Rundlöf M, Wolters V and Smith HG (2015) Ecosystem services—current challenges and opportunities for ecological research. Front. Ecol. Evol . 2 :87. doi: 10.3389/fevo.2014.00087

Received: 12 October 2014; Accepted: 21 December 2014; Published online: 12 January 2015.

Reviewed by:

Copyright © 2015 Birkhofer, Diehl, Andersson, Ekroos, Früh-Müller, Machnikowski, Mader, Nilsson, Sasaki, Rundlöf, Wolters and Smith. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY) . The use, distribution or reproduction in other forums is permitted, provided the original author(s) or licensor are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Klaus Birkhofer, Department of Biology, Biodiversity and Conservation Science, Lund University, Sölvegatan 37, 22362 Lund, Sweden e-mail: [email protected]

† These authors have contributed equally to this work.

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Published: 12 October 2020

Increasing decision relevance of ecosystem service science

  • Lisa Mandle   ORCID: orcid.org/0000-0002-1420-8529 1 ,
  • Analisa Shields-Estrada   ORCID: orcid.org/0000-0003-0790-3643 2 ,
  • Rebecca Chaplin-Kramer   ORCID: orcid.org/0000-0002-1539-5231 1 ,
  • Matthew G. E. Mitchell   ORCID: orcid.org/0000-0002-3490-9314 3 , 4 ,
  • Leah L. Bremer 5 , 6 ,
  • Jesse D. Gourevitch 7 , 8 ,
  • Peter Hawthorne 9 ,
  • Justin A. Johnson 10 ,
  • Brian E. Robinson   ORCID: orcid.org/0000-0002-8972-8318 11 ,
  • Jeffrey R. Smith   ORCID: orcid.org/0000-0001-5768-216X 12 ,
  • Laura J. Sonter   ORCID: orcid.org/0000-0002-6590-3986 7 , 13 , 14 ,
  • Gregory M. Verutes   ORCID: orcid.org/0000-0002-7667-7902 1 , 15 ,
  • Adrian L. Vogl   ORCID: orcid.org/0000-0001-9369-1071 1 ,
  • Gretchen C. Daily   ORCID: orcid.org/0000-0003-1443-1111 1 , 12 &
  • Taylor H. Ricketts   ORCID: orcid.org/0000-0001-9688-7977 7 , 8  

Nature Sustainability volume  4 ,  pages 161–169 ( 2021 ) Cite this article

4891 Accesses

133 Citations

130 Altmetric

Metrics details

  • Ecological modelling
  • Ecosystem services
  • Environmental economics

The ecosystem service (ES) community aspires to illuminate how nature contributes to human well-being, and thereby elevate consideration of nature in decision making. So far, however, policy impact of ES research has been limited. To understand why, we identify five key elements of ES research that help inform decisions by connecting the supply of ES to those who benefit from them. Our structured review of the ES literature reveals that only 13% of assessments included the full ES chain from place to value. Only 7% of assessments considered the distribution of ES benefits explicitly across demographic or other beneficiary groups (for example, private landowners versus the broader public), although disaggregation across regions or spatial units was more common (44%). Finally, crucial mediating factors that affect who benefits and how (for example, the vulnerability of beneficiaries or the availability of substitutes for ES) were considered in only 35% of assessments. Our results suggest that increasing the decision relevance of ES research requires more effectively predicting the impacts of specific decisions on the value and distribution of ES across beneficiary groups. Such efforts will need to integrate ecological models with socioeconomic and cultural dimensions of ES more closely than does the current ES literature.

This is a preview of subscription content, access via your institution

Access options

Access Nature and 54 other Nature Portfolio journals

Get Nature+, our best-value online-access subscription

24,99 € / 30 days

cancel any time

Subscribe to this journal

Receive 12 digital issues and online access to articles

111,21 € per year

only 9,27 € per issue

Buy this article

  • Purchase on SpringerLink
  • Instant access to full article PDF

Prices may be subject to local taxes which are calculated during checkout

essay on ecosystem service

Similar content being viewed by others

essay on ecosystem service

Diverse values of nature for sustainability

essay on ecosystem service

Including stewardship in ecosystem health assessment

essay on ecosystem service

Synergies and complementarities between ecosystem accounting and the Red List of Ecosystems

Data availability.

The data that support the findings of this study are available from the Stanford Digital Repository at https://purl.stanford.edu/pt786dv3952 . Source data are provided with this paper.

IPBES Summary for Policymakers. In Global Assessment Report on Biodiversity and Ecosystem Services (eds Díaz, S. et al.) (IPBES Secretariat, 2019).

Schaefer, M., Goldman, E., Bartuska, A. M., Sutton-Grier, A. & Lubchenco, J. Nature as capital: advancing and incorporating ecosystem services in United States federal policies and programs. Proc. Natl Acad. Sci. USA 112 , 7383–7389 (2015).

Article   CAS   Google Scholar  

Mastrángelo, M. E. et al. Key knowledge gaps to achieve global sustainability goals. Nat. Sustain . https://doi.org/10.1038/s41893-019-0412-1 (2019).

Olander, L. et al. So you want your research to be relevant? Building the bridge between ecosystem services research and practice. Ecosyst. Serv. 26 , 170–182 (2017).

Article   Google Scholar  

Polasky, S., Tallis, H. & Reyers, B. Setting the bar: standards for ecosystem services. Proc. Natl Acad. Sci. USA https://doi.org/10.1073/pnas.1406490112 (2015).

Rieb, J. et al. When, where and how nature matters for ecosystem services: challenges for the next generation of ecosystem service models. BioScience 67 , 820–833 (2017).

Natural Capital Protocol (Natural Capital Coalition, 2016).

Mandle, L., Ouyang, Z., Salzman, J. & Daily, G. C. Green Growth that Works: Natural Capital Policy and Finance Mechanisms from the World (Island Press, 2019).

Transforming our World: The 2030 Agenda for Sustainable Development (UN, 2015).

Díaz, S. et al. Assessing nature’s contributions to people: recognizing culture, and diverse sources of knowledge, can improve assessments. Science 359 , 270–272 (2018).

Arkema, K. K. et al. Embedding ecosystem services in coastal planning leads to better outcomes for people and nature. Proc. Natl Acad. Sci. USA 112 , 7390–7395 (2015).

Van Wensem, J. et al. Identifying and assessing the application of ecosystem services approaches in environmental policies and decision making. Integr. Environ. Assess. Manag. 13 , 41–51 (2017).

Ricketts, T. H. & Lonsdorf, E. Mapping the margin: comparing marginal values of tropical forest remnants for pollination services. Ecol. Appl. 23 , 1113–1123 (2013).

Mandle, L., Tallis, H., Sotomayor, L. & Vogl, A. L. Who loses? Tracking ecosystem service redistribution from road development and mitigation in the Peruvian Amazon. Front. Ecol. Environ. 13 , 309–315 (2015).

Wieland, R., Ravensbergen, S., Gregr, E. J., Satterfield, T. & Chan, K. M. A. Debunking trickle-down ecosystem services: the fallacy of omnipotent, homogeneous beneficiaries. Ecol. Econ. 121 , 175–180 (2016).

Polasky, S. & Segerson, K. Integrating ecology and economics in the study of ecosystem services: some lessons learned. Annu. Rev. Resour. Econ. 1 , 409–434 (2009).

Keeler, B. L. et al. Linking water quality and well-being for improved assessment and valuation of ecosystem services. Proc. Natl Acad. Sci. USA 109 , 18619–18624 (2012).

Vogl, A. L. et al. Valuing investments in sustainable land management in the Upper Tana River basin, Kenya. J. Environ. Manag. 195 , 78–91 (2017).

Arkema, K., Guannel, G. & Verutes, G. Coastal habitats shield people and property from sea-level rise and storms. Nat. Clim. Change 3 , 913–918 (2013).

Plummer, M. L. Assessing benefit transfer for the valuation of ecosystem services. Front. Ecol. Environ. 7 , 38–45 (2009).

Tallis, H., Polasky, S., Lozano, J. S. & Wolny, S. in Inclusive Wealth Report 2012: Measuring Progress Toward Sustainability 195–214 (Cambridge Univ. Press, 2012).

Costanza, R. et al. Changes in the global value of ecosystem services. Glob. Environ. Change https://doi.org/10.1016/j.gloenvcha.2014.04.002 (2014).

Granek, E. F. et al. Ecosystem services as a common language for coastal ecosystem-based management. Conserv. Biol. 24 , 207–216 (2010).

Ruckelshaus, M. et al. Notes from the field: lessons learned from using ecosystem service approaches to inform real-world decisions. Ecol. Econ . https://doi.org/10.1016/j.ecolecon.2013.07.009 (2013).

Ellis, A. M., Myers, S. S. & Ricketts, T. H. Do pollinators contribute to nutritional health? PLoS ONE 10 , e114805 (2015).

Olsson, P., Folke, C. & Hughes, T. P. Navigating the Transition to Ecosystem-Based Management of the Great Barrier Reef, Australia. Proc. Natl Acad. Sci. USA 105 , 9489–9494 (2008).

Costanza, R. et al. The value of the world’s ecosystem services and natural capital. Nature 387 , 253–260 (1997).

SEEA Experimental Ecosystem Accounting Revision (System of Environmental Economic Accounting, 2020); https://go.nature.com/2sqGqFn

Aburto-Oropeza, O. et al. Mangroves in the Gulf of California increase fishery yields. Proc. Natl Acad. Sci. USA https://doi.org/10.1073/pnas.0804601105 (2008).

Keeler, B. L. et al. The social costs of nitrogen. Sci. Adv. 2 , e1600219 (2016).

Kenter, J. O. et al. What are shared and social values of ecosystems? Ecol. Econ. 111 , 86–99 (2015).

Pascual, U. et al. Valuing nature’s contributions to people: the IPBES approach. Curr. Opin. Environ. Sustain. 26–27 , 7–16 (2017).

Samberg, L. H., Gerber, J. S., Ramankutty, N., Herrero, M. & West, P. C. Subnational distribution of average farm size and smallholder contributions to global food production. Environ. Res. Lett. 11 , 124010 (2016).

Jean, N. et al. Combining satellite imagery and machine learning to predict poverty. Science 353 , 790–794 (2016).

Wolff, S., Schulp, C. J. E. & Verburg, P. H. Mapping ecosystem services demand: a review of current research and future perspectives. Ecol. Indic. 55 , 159–171 (2015).

Dawson, N. & Martin, A. Assessing the contribution of ecosystem services to human wellbeing: a disaggregated study in western Rwanda. Ecol. Econ. 117 , 62–72 (2015).

Daw, T., Brown, K., Rosendo, S. & Pomeroy, R. Applying the ecosystem services concept to poverty alleviation: the need to disaggregate human well-being. Environ. Conserv. 38 , 370–379 (2011).

Ruhl, J. B. & Salzman, J. The effects of wetland mitigation banking on people. Natl Wetl. Newsl. 28 , 7–13 (2006).

Google Scholar  

Kabisch, N. & Haase, D. Green justice or just green? Provision of urban green spaces in Berlin, Germany. Landsc. Urban Plan. 122 , 129–139 (2014).

Farley, K. A. & Bremer, L. L. ‘Water Is Life’: local perceptions of páramo grasslands and land management strategies associated with payment for ecosystem services. Ann. Am. Assoc. Geogr. 107 , 371–381 (2017).

Pascual, U. et al. Social equity matters in payments for ecosystem services. BioScience 64 , 1027–1036 (2014).

Mastrangelo, M. E. & Laterra, P. From biophysical to social-ecological trade-offs: integrating biodiversity conservation and agricultural production in the Argentine Dry Chaco. Ecol. Soc. 20 , 20 (2015).

Guerry, A. D. et al. Natural capital and ecosystem services informing decisions: from promise to practice. Proc. Natl Acad. Sci. USA 112 , 7348–7355 (2015).

Rieb, J. T. et al. When, where, and how nature matters for ecosystem services: challenges for the next generation of ecosystem service models. BioScience 67 , 820–833 (2017).

Villa, F., Bagstad, K. J., Voigt, B., Johnson, G. W. & Portela, R. A methodology for adaptable and robust ecosystem services assessment. PLoS ONE 9 , e91001 (2014).

Díaz, S. et al. Assessing nature’s contributions to people. Science 359 , 270–272 (2018).

Millennium Ecosystem Assessment Ecosystems and Human Well-being: A Framework for Assessment (Island Press, 2003).

Fleiss, J. L. Measuring nominal scale agreement among many raters. Psychol. Bull. 76 , 378–382 (1971).

Gamer, M., Lemon, J., Fellows, I. & Singh, P. irr: Various Coefficients of Interrater Reliability and Agreement (2012).

Landis, J. R. & Koch, G. G. The measurement of observer agreement for categorical data. Biometrics 33 , 159–174 (1977).

Tallis, H. et al. A global system for monitoring ecosystem service change. BioScience 62 , 977–986 (2012).

Daily, G. C. et al. Ecosystem services in decision making: time to deliver. Front. Ecol. Environ. 7 , 21–28 (2009).

Download references

Acknowledgements

This research was supported in part by the Scenarios Society and Solutions Research Coordination Network (grant no. NSF-DEB-13-38809), the Social Science and Humanities Research Council of Canada (430-2014-00861) and the Natural Capital Project–Stanford. We thank W. Bierbower for help with scripts for randomly selecting and downloading papers. We are grateful to E. Bennett, D. Guswa, K. Watson and the many Natural Capital Project members who provided helpful suggestions and feedback along the way.

Author information

Authors and affiliations.

Natural Capital Project, Department of Biology and Woods Institute for the Environment, Stanford University, Stanford, CA, USA

Lisa Mandle, Rebecca Chaplin-Kramer, Gregory M. Verutes, Adrian L. Vogl & Gretchen C. Daily

Department of Integrative Biology, The University of Texas at Austin, Austin, TX, USA

Analisa Shields-Estrada

Institute for Resources, Environment and Sustainability, University of British Columbia, Vancouver, British Columbia, Canada

Matthew G. E. Mitchell

Centre for Sustainable Food Systems, Faculty of Land and Food Systems, University of British Columbia, Vancouver, British Columbia, Canada

University of Hawaiʻi Economic Research Organization, Honolulu, HI, USA

Leah L. Bremer

University of Hawaiʻi Water Resources Research Center, Honolulu, HI, USA

Gund Institute for Environment, University of Vermont, Burlington, VT, USA

Jesse D. Gourevitch, Laura J. Sonter & Taylor H. Ricketts

Rubenstein School of Environment and Natural Resources, University of Vermont, Burlington, VT, USA

Jesse D. Gourevitch & Taylor H. Ricketts

Natural Capital Project, Institute on the Environment, University of Minnesota, St Paul, MN, USA

Peter Hawthorne

Department of Applied Economics, University of Minnesota, St Paul, MN, USA

Justin A. Johnson

Department of Geography, McGill University, Montreal, Québec, Canada

Brian E. Robinson

Department of Biology, Center for Conservation Biology, Stanford University, Stanford, CA, USA

Jeffrey R. Smith & Gretchen C. Daily

Centre for Biodiversity and Conservation Science, The University of Queensland, St Lucia, Queensland, Australia

Laura J. Sonter

School of Earth and Environmental Sciences, The University of Queensland, St Lucia, Queensland, Australia

Faculty of Political and Social Sciences, Universidade de Santiago de Compostela, Santiago de Compostela, Spain

Gregory M. Verutes

You can also search for this author in PubMed   Google Scholar

Contributions

All authors were involved in study conceptualization and design, and contributed to review and editing; A.L.V., A.S.-E., R.C.-K., B.E.R., G.M.V., J.D.G., J.A.J., J.R.S., L.J.S., L.L.B., L.M., M.G.E.M. and P.L.H. contributed to data collection; A.S.-E., L.M. and G.M.V. conducted the formal analyses; L.M. led project administration and supervision; A.S.-E. led data management and visualization; and L.M. and T.H.R. led the writing of the original draft.

Corresponding author

Correspondence to Analisa Shields-Estrada .

Ethics declarations

Competing interests.

The authors declare no competing interests.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary information

Supplementary information.

Supplementary Figs. 1–3 and Appendices A and B.

Reporting Summary

Source data, source data fig. 2.

Numerical Source Data.

Source Data Fig. 3

Source data fig. 4, source data fig. 5, rights and permissions.

Reprints and permissions

About this article

Cite this article.

Mandle, L., Shields-Estrada, A., Chaplin-Kramer, R. et al. Increasing decision relevance of ecosystem service science. Nat Sustain 4 , 161–169 (2021). https://doi.org/10.1038/s41893-020-00625-y

Download citation

Received : 04 March 2020

Accepted : 11 September 2020

Published : 12 October 2020

Issue Date : February 2021

DOI : https://doi.org/10.1038/s41893-020-00625-y

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

This article is cited by

Social–ecological benefits of land–sea planning at multiple scales in mesoamerica.

  • Jade M. S. Delevaux
  • Jess M. Silver
  • Katie K. Arkema

Nature Sustainability (2024)

Perceived ecosystem services differ substantially from calculated services using biophysical models

  • Peiyuan Chen
  • Yuchen Zhou

Landscape Ecology (2024)

Enhancing human well-being through cognitive and affective pathways linking landscape sensation to cultural ecosystem services

  • Clive E. Sabel

A social-ecological approach to support equitable land use decision-making

  • Maria Brück
  • Felipe Benra
  • David J. Abson

Ambio (2024)

A bibliographic review of the relationship between ecosystem services and human well-being

  • Qingxu Huang
  • Yansong Bai

Environment, Development and Sustainability (2024)

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

essay on ecosystem service

COMMENTS

  1. What Are Ecosystem Services, and How Do They Help Our Planet?

    Those services, which are often called ecosystem services, include providing resources such as food and water, maintaining habitats that support biodiversity, offering opportunities for recreation, and helping to regulate human-caused impacts like climate change.

  2. Ecosystem Services and Their Main Types Essay - IvyPanda

    Ecosystem services enable people to obtain resources from the land itself through proper care and use of resources. Four types of ecosystem services are currently distinguished: providing, regulating, cultural, and supporting (Pearce, 2020).

  3. Ecosystem services | Definition, History, & Importance

    ecosystem services, outputs, conditions, or processes of natural systems that directly or indirectly benefit humans or enhance social welfare. Ecosystem services can benefit people in many ways, either directly or as inputs into the production of other goods and services.

  4. What Are Ecosystem Services? - Science ABC

    Ecosystem services are all the processes and outputs that nature provides us with. These include provisioning services (food, water), regulating services (waste water treatment, pollution control), supporting services (shelter), and cultural services (recreation and tourism).

  5. Ecosystems, Ecosystem Services, and Biodiversity | U.S ...

    Ecosystems, which are composed of living things that interact with the physical environment, provide numerous essential benefits to people. These benefits, termed ecosystem services, encompass four primary functions: provisioning materials, such as food and fiber; regulating crit.

  6. The ecosystem service approach and its application as a tool ...

    Abstract. Ecosystem services are the benefits that natural environments supply to human beings. Due to the immense diversity of ecosystems and objectives for which their services are being assessed, there are no standard methodologies for this type of evaluation.

  7. Ecosystem services—current challenges and opportunities for ...

    In this article we identify major challenges and opportunities for ecologists involved in empirical or modeling ecosystem service research. The first challenge arises from the fact that the ecosystem service concept has not been generated in the context of managed systems.

  8. Ecosystem Services | Journal | ScienceDirect.com by Elsevier

    Ecosystem Services, associated with the Ecosystem Services Partnership (ESP), is an international, interdisciplinary journal that deals with the science, policy and practice of Ecosystem Services defined as the direct and indirect contributions of ecosystems to human wellbeing.

  9. Increasing decision relevance of ecosystem service science

    To increase the policy relevance of ecosystem service benefits research, studies need to better predict the impact of specific decisions, according to an analysis of the literature.

  10. Full article: Disentangling ‘ecosystem services’ and ‘nature ...

    The papers in this special issue adopt a conceptual framework based on the UK National Ecosystem Assessment which identifies the interaction between: environmental spaces and cultural practices, generating material benefits and non-material benefits in terms of identities, experience, and capabilities (Kenter Citation 2016). While evidence of ...