Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-45l2p Total loading time: 0 Render date: 2024-04-29T01:24:17.368Z Has data issue: false hasContentIssue false

6 - Emergence of Cross-Scale Structural and Functional Processes in Ecosystem Science

Published online by Cambridge University Press:  25 February 2021

Robert G. Woodmansee
Affiliation:
Colorado State University
John C. Moore
Affiliation:
Colorado State University
Dennis S. Ojima
Affiliation:
Colorado State University
Laurie Richards
Affiliation:
Colorado State University
Get access

Summary

Fundamental knowledge about the processes that control the functioning of the biophysical workings of ecosystems has expanded exponentially since the late 1960s. Scientists, then, had only primitive knowledge about C, N, P, S, and H2O cycles; plant, animal, and soil microbialinteractions and dynamics; and land, atmosphere, and water interactions. With the advent of systems ecology paradigm (SEP) and the explosion of technologies supporting field and laboratory research, scientists throughout the world were able to assemble the knowledge base known today as ecosystem science. This chapter describes, through the eyes of scientists associated with the Natural Resource Ecology Laboratory (NREL) at Colorado State University (CSU), the evolution of the SEP in discovering how biophysical systems at small scales (ecological sites, landscapes) function as systems. The NREL and CSU are epicenters of the development of ecosystem science. Later, that knowledge, including humans as components of ecosystems, has been applied to small regions, regions, and the globe. Many research results that have formed the foundation for ecosystem science and management of natural resources, terrestrial environments, and its waters are described in this chapter. Throughout are direct and implicit references to the vital collaborations with the global network of ecosystem scientists.

Type
Chapter
Information
Natural Resource Management Reimagined
Using the Systems Ecology Paradigm
, pp. 140 - 201
Publisher: Cambridge University Press
Print publication year: 2021

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Aldridge, C. L., and Boyce, M. S. (2007). Linking occurrence and fitness to persistence: Habitat-based approach for endangered Greater Sage-grouse. Ecological Applications, 17, 508–26.CrossRefGoogle ScholarPubMed
Aldridge, C. L., Saher, D. J., Childers, T. M., Stahlnecker, K. E., and Bowen, Z. H. (2012). Crucial nesting habitat for Gunnison Sage-grouse: A spatially explicit hierarchical approach. Journal of Wildlife Management, 76, 391406.CrossRefGoogle Scholar
Anderson, D. W., Heil, R. D., Cole, C. V., and Deutsch, P. C. (1983). Identification and Characterization of Ecosystems at Different Integrative Levels. Special Publication. Athens, GA: University of Georgia, Agriculture Experiment Stations.Google Scholar
Archer, S., and Detling, J. K. (1986). Evaluation of potential herbivore mediation of plant water status in a North American mixed-grass prairie. Oikos, 47, 287–91.CrossRefGoogle Scholar
Archer, S., and Smeins, F. E. (1991). Ecosystem-level processes. In Grazing Management: An Ecological Perspective, ed. Heitschmidt, R. K. and Stuth, J. W.. Portland, OR: Timber Press, 109–39.Google Scholar
Augustine, D. J., Matchett, M. R., Toombs, T. B., Cully, J. F., Johnson, T. L., and Sidle, G. (2008). Spatiotemporal dynamics of black-tailed prairie dog colonies affected by plague. Landscape Ecology, 23, 255–67.Google Scholar
Bailey, D. W., Gross, J. E., Laca, E. A., et al. (1996). Mechanisms that result in large herbivore grazing distribution patterns. Journal of Range Management, 49, 386400.CrossRefGoogle Scholar
Baisan, C. H., and Swetnam, T. W. (1990). Fire history on a desert mountain range – Rincon Mountain Wilderness, Arizona, USA. Canadian Journal of Forest Research-Revue Canadienne De Recherche Forestiere, 20, 1559–69.CrossRefGoogle Scholar
Baker, W. L. (1989). A review of models of landscape change. Landscape Ecology, 2, 112–34.CrossRefGoogle Scholar
Baron, J. S., Barber, M. C., Adams, M., et al. (2014). The effects of atmospheric nitrogen deposition on terrestrial and freshwater biodiversity. In Nitrogen Deposition, Critical Loads, and Biodiversity, ed. Sutton, M. K., Mason, L., Sheppard, H., et al. Dordrecht: Springer, 465–80.Google Scholar
Baron, J. S., Poff, N. L., Angermeieret, P. L., et al. (2002). Meeting ecological and societal needs for freshwater. Ecological Applications, 12, 1247–60.Google Scholar
Baron, J. S., Rueth, H. M., Wolfe, A. M., et al. (2000). Ecosystem responses to nitrogen deposition in the Colorado Front Range. Ecosystems, 3, 352–68.CrossRefGoogle Scholar
Bedunah, D. J., and Sosebee, R. E. eds. (1995). Wildland Plants: Physiological Ecology and Developmental Morphology. Denver, CO: Society for Range Management.Google Scholar
Bell, C. W., Tissue, D. T., Loik, M. E., et al. (2014). Soil microbial and nutrient responses to 7 years of seasonally altered precipitation in a Chihuahuan Desert grassland. Global Change Biology, 20, 1657–73.CrossRefGoogle Scholar
Bestelmeyer, B. T., Goolsby, D. P., Archer, S. R. (2011). Spatial perspectives in state-and-transition models: A missing link to land management? Journal of Applied Ecology, 48, 746–57.CrossRefGoogle Scholar
Bhola, N., Ogutu, J. O., Piepho, H. -P., et al. (2012). Comparative changes in density and demography of large herbivores in the Masai Mara Reserve and its surrounding human-dominated pastoral ranches in Kenya. Biodiversity and Conservation, 21, 1509–30.CrossRefGoogle Scholar
Binkley, D., Adams, M., Fredericksen, T., Laclau, J. P., Makinen, H. H., and Prescott, C. (2015). Editors note: Clarity of ideas and terminology in forest ecology and management. Forest Ecology and Management, 349, 13.Google Scholar
Binkley, D., and Fisher, R. (2012). Ecology and Management of Forest Soils. New York: John Wiley and Sons.Google Scholar
Bolin, B., and Cook, R. B. ed. (1983). The Major Biogeochemical Cycles and Their Interactions (SCOPE Report 21). Chichester, published on behalf of the Scientific Committee on Problems of the Environment (SCOPE) of the International Council of Scientific Unions (ICSU). New York: John Wiley.Google Scholar
Boone, R. B. (2007). Effects of fragmentation on cattle in African savannas under variable precipitation. Landscape Ecology, 22, 1355–69.CrossRefGoogle Scholar
Boone, R. B. (2017). Evolutionary computation in zoology and ecology. Current Zoology, 63, 675–86.Google Scholar
Boone, R. B., BurnSilver, S. B., Thornton, P. K., Worden, J. S., and Galvin, K. A. (2005). Quantifying declines in livestock due to subdivision. Rangeland Ecology & Management, 58, 523–32.Google Scholar
Boone, R. B., Conant, R. T., Sircely, J., Thornton, P. K., and Herrero, M. (2018). Climate change impacts on selected global rangeland ecosystem services. Global Change Biology, 24, 1382–93.CrossRefGoogle ScholarPubMed
Boone, R. B., Coughenour, M. B., Galvin, K. A., and Ellis, J. E. (2002). Addressing management questions for Ngorongoro Conservation Area using the Savanna Modeling System. African Journal of Ecology, 40, 138–50.Google Scholar
Boone, R. B., Galvin, K. A. (2014). Simulation as an approach to social–ecological integration, with an emphasis on agent-based modeling. In Understanding Society and Natural Resources: Forging New Strands of Integration across the Social Sciences, ed. Manfredo, M., et al. New York: Springer, 179202.Google Scholar
Boone, R. B., Johnson, C. M., and Johnson, L. B. (2006). Simulating wood frog movement in central Minnesota, USA using a diffusion model. Ecological Modelling, 198, 255–62.Google Scholar
Boone, R. B., and Krohn, W. B. (2002a). An introduction to modeling tools and accuracy assessment. In Predicting Species Occurrences: Issues of Accuracy and Scale, ed. Scott, J. M., Haglung, J. H., Morrison, M. L., et al. Washington, DC: Island Press, 265–70.Google Scholar
Boone, R. B., and Krohn, W. B. (2000b). Predicting broad-scale occurrences of vertebrates in patchy landscapes. Landscape Ecology, 15, 6374.CrossRefGoogle Scholar
Boone, R. B., and Krohn, W. B. (2000c). Relationship between avian range limits and plant transition zones in Maine. Journal of Biogeography, 27, 471–82.CrossRefGoogle Scholar
Boone, R. B., Lackett, J. M., Galvin, K. A., Ojima, D. S., Tucker, C. J. III (2007). Links and broken chains: Evidence of human-caused changes in land cover in remotely sensed images. Environmental Science & Policy, 10, 135–49.CrossRefGoogle Scholar
Bormann, F. H., and Likens, G. E. (1979). Pattern and Process in a Forest Ecosystem: Disturbance, Development, and the Steady State Based on the Hubbard Brook Ecosystem Study. New York: Springer.CrossRefGoogle Scholar
Briske, D. D., Fuhlendorf, S. D., and Smeins, F. E. (2005). State-and-transition models, thresholds, and rangeland health: A synthesis of ecological concepts and perspectives. Rangeland Ecology & Management, 58, 110.2.0.CO;2>CrossRefGoogle Scholar
Briske, D. D., and Heitschmidt, R. K. (1991). An ecological perspective. In Grazing Management: An Ecological Perspective, ed. Heitschmidt, R. K. and Stuth, J. W.. Portland, OR: Timber Press.Google Scholar
Brokaw, N., and Busing, R. T. (2000). Niche versus chance and tree diversity in forest gaps. Trends in Ecology and Evolution, 15, 183–8.CrossRefGoogle ScholarPubMed
Brown, J. H. (1995). Macroecology. Chicago: University of Chicago Press.Google Scholar
Brown, L. F., and Trlica, M. J. (1977). Interacting effects of soil water, temperature and irradiance on CO2 exchange rates of two dominant grasses of the shortgrass prairie. Journal of Applied Ecology, 14, 197204.Google Scholar
Burke, I. C., Lauenroth, W. K., and Wessman, C. A. 1998. Progress in understanding biogeochemistry at regional to global scales. In Successes, Limitations, and Challenges in Ecosystem Science, ed. Groffman, P. and Pace, M.. New York: Springer Verlag.Google Scholar
Burke, I. C., Mosier, A. R., Hook, P. B., et al. (2008). Organic matter and nutrient dynamics of shortgrass steppe. In Ecology of the Shortgrass Steppe: A Long-Term Perspective, ed. Lauenroth, W. K. and Burke, I. C.. Oxford: Oxford University Press.Google Scholar
BurnSilver, S., Boone, R. B., and Galvin, K. A. (2003). Linking pastoralists to a heterogeneous landscape: The case of four Maasai group ranches in Kajiado District, Kenya. In Linking Household and Remotely Sensed Data: Methodological and Practical Problems, ed. Fox, J., Mishra, V., Rindfuss, R., and Walsh, S.. Boston: Kluwer Academic Publishing, 173–99.Google Scholar
Carpenter, S. R., Mooney, H. A., Agard, J., et al. (2009). Science for managing ecosystem services: Beyond the Millennium Ecosystem Assessment. Proceedings of the National Academy of Science USA, 106, 1305–12.Google Scholar
Ceballos, G., Ehrlich, P. R., Barnosky, A. D. A., et al. (2015). Accelerated modern human-induced species losses: Entering the sixth mass extinction. Science Advances, 1(5), e1400253.CrossRefGoogle ScholarPubMed
Chapin, F. S., Chapin, M. C., Matson, P. A., and Vitousek, P. (2011). Principles of Terrestrial Ecosystem Ecology. New York: Springer.Google Scholar
Christensen, N. L. (2014). An historical perspective on forest succession and its relevance to ecosystem restoration and conservation practice in North America. Forest Ecology and Management, 330, 312–22.CrossRefGoogle Scholar
Clark, F. E. (1977). Internal cycling of 15nitrogen in shortgrass prairie. Ecology, 58, 1322–33.Google Scholar
Clark, F. E. (1981). The nitrogen cycle: Viewed with poetic licence. In Terrestrial Nitrogen Cycles: Processes, Ecosystem Strategies, and Management Impacts, ed. Clark, F. E. and Rosswall, T.. Ecological Bulletin, 33. Stockholm: Swedish Natural Science Research Council (NRF).Google Scholar
Clark, F. E. and Rosswall, T., ed. (1981). Terrestrial Nitrogen Cycles, Processes, Ecosystem, Strategies and Management Impacts. Ecological Bulletin, 33. Stockholm: Swedish Natural Science Research Council (NRF).Google Scholar
Clark, J. S. (1989). Ecological disturbance as a renewal process: Theory and application to fire history. Oikos, 56, 1730.CrossRefGoogle Scholar
Clements, F. E. (1916). Plant Succession: An Analysis of the Development of Vegetation. Publication No. 242. Washington, DC: Carnegie Institute of Washington.CrossRefGoogle Scholar
Clements, F. E. (1935). Experimental ecology in the public service. Ecology, 16, 342–63.CrossRefGoogle Scholar
Cole, C. V., and Heil, R. D. (1981). Phosphorus effects on terrestrial nitrogen cycling. In Terrestrial Nitrogen Cycles: Processes, Ecosystem, Strategies and Management Impacts, ed. Clark, F. E. and Rosswall, T.. Ecological Bulletin, 33. Stockholm: Swedish Natural Science Research Council, 363–74.Google Scholar
Coleman, D. C., Swift, D. M., and Mitchell, J. E. (2004). From the frontier to the biosphere: A brief history of the USIBP Grasslands Biome program and its impacts on scientific research in North America. Rangelands, 26, 815.Google Scholar
Coleman, D. C., and Wall, D. H. (2014). Soil fauna: Occurrence, biodiversity, and roles in ecosystem function. In Soil Microbiology, Ecology and Biochemistry, ed. Paul, E. A.. London: Academic Press.Google Scholar
Conant, R. T., Cerri, C. E. P., Osborne, B. B., and Paustian, K. (2016). Grassland management impacts on soil carbon stocks: A new synthesis. Ecological Applications, 27, 662–8.Google Scholar
Conant, R. T., Paustian, K., and Elliot, E. T. (2001). Grassland management and conversion into grassland: Effects on soil carbon. Ecological Applications, 11, 343–55.Google Scholar
Connell, J. H. (1978). Diversity in tropical rain forests and coral reefs. Science, 199, 1302–10.Google Scholar
Connell, J. H, and Slatyer, R. O. (1977). Mechanisms of succession in natural communities and their role in community stability and organization. American Naturalist, 111, 1119–44.CrossRefGoogle Scholar
Coppock, D. L., and Detling, J. K. (1986). Alteration of bison/prairie dog grazing interaction by prescribed burning. Journal of Wildlife Management, 50, 452–5.CrossRefGoogle Scholar
Coppock, D. L., Detling, J. K., Ellis, J. E., and Dyer, M. I. (1983a). Plant–herbivore interactions in a North American mixed-grass prairie: I. Effects of black-tailed prairie dogs on seasonal aboveground plant biomass and nutrient dynamics and plant species diversity. Oecologia, 5, 19.CrossRefGoogle Scholar
Coppock, D. L., Detling, J. K., Ellis, J. E., and Dyer, M. I. (1983b). Plant–herbivore interactions in a North American mixed-grass prairie: II. Responses of bison to modification of vegetation by prairie dogs. Oecologia, 56, 1015.Google Scholar
Cotrufo, M. F., Soong, J. L., Horton, A. J., et al. (2015). Soil organic matter formation from biochemical and physical pathways of litter mass loss. Nature Geosciences, 8(10), 14.Google Scholar
Cotrufo, M. F., Wallenstein, M., Boot, M. C., Denef, K., and Paul, E. A. (2013). The Microbial Efficiency-Matrix Stabilization (MEMS) framework integrates plant litter decomposition with soil organic matter stabilization: Do labile plant inputs form stable soil organic matter? Global Change Biology, 19, 988–95.Google Scholar
Coughenour, M. B. (1991). Spatial components of plant–herbivore interactions in pastoral, ranching, and native ungulate ecosystems. Journal of Rangeland Management, 44, 530–42.CrossRefGoogle Scholar
Coughenour, M. B., Dodd, J. L., Coleman, D. C., and Lauenroth, W. K. (1979). Partitioning of carbon and SO2 sulfur in a native grassland. Oecologia, 42, 229–40.CrossRefGoogle Scholar
Coughenour, M. B., Ellis, J. E., Swift, D. M., et al. (1985). Energy extraction and use in a nomadic pastoral ecosystem. Science, 230, 619–25.Google Scholar
Covino, T. P., and McGlynn, B. L. (2007). Stream gains and losses across a mountain-to-valley transition: Impacts on watershed hydrology and stream water chemistry. Water Resources Research, 43: W10431.CrossRefGoogle Scholar
Coyne, P. I., Trlica, M. J., and Owensby, C. E. (1995). Carbon and nitrogen dynamics in range plants. In Wildland Plants: Physiological Ecology and Developmental Morphology, ed. Bedunah, D. J. and Sosebee, R. E.. Denver, CO: Society for Range Management: 59167.Google Scholar
Currie, W. S. (2011). Units of nature or processes across scales? The ecosystem concept at age 75. New Phytologist, 190, 2134.Google Scholar
Curtis, J. T. (1959). The Vegetation of Wisconsin. Madison: University of Wisconsin Press.Google Scholar
Davis, M. B., and Shaw, R. G. (2001). Range shifts and adaptive responses to climate change. Science, 292, 673–9.Google Scholar
Day, T. A., and Detling, J. K. (1994). Water relations of Agropyron smithii and Bouteloua gracilis and community evapotranspiration following long-term grazing by prairie dogs. American Midland Naturalist, 132, 381–92.CrossRefGoogle Scholar
Day, T. A., and Detling, J. K. (1990a). Grassland patch dynamics and herbivore grazing preference following urine deposition. Ecology, 71, 180–8.Google Scholar
Day, T. A., and Detling, J. K. (1990b). Changes in grass leaf water relations following urine deposition. American Midland Naturalist, 123, 171–8.CrossRefGoogle Scholar
Del Grosso, S., Ojima, D. S., Parton, W. J., et al. (2009). Global scale DAYCENT model analysis of greenhouse gas emissions and mitigation strategies for cropped soils. Global and Planetary Change, 67, 44–5.CrossRefGoogle Scholar
Del Grosso, S., Parton, W., Stohlgren, T., et al. (2008). Global potential net primary production predicted from vegetation class, precipitation, and temperature. Ecology, 89, 2117–26.Google Scholar
Detling, J. K. (1988). Grasslands and savannas: Regulation of energy flow and nutrient cycling by herbivores. In Concepts of Ecosystem Ecology, ed. Pomeroy, L. R. and Alberts, J. J.. Ecological Studies, 67. New York: Springer Verlag, 131–48.Google Scholar
Detling, J. K. (1998). Mammalian herbivores: Ecosystem-level effects in two grassland national parks. Wildlife Society Bulletin, 26, 438–48.Google Scholar
Detling, J. K., Parton, W. J., and Hunt, H. W. (1978). An empirical model for estimating CO2 exchange of Bouteloua gracilis (H.B.K.) Lag. in the shortgrass prairie. Oecologia, 33, 137–47.Google Scholar
Drury, W. H., and Nisbet, I. C. T. (1973). Succession. Journal of the Arnold Arboretum, 54, 331–68.CrossRefGoogle Scholar
Egler, F. E. (1954). Vegetation science concepts: 1. Initial floristic composition, a factor in old-field vegetation development. Vegetatio, 4, 412–17.CrossRefGoogle Scholar
Egler, F. E. (1975). Plight of the Right of Way Domain. Mt. Kisco, NY: Futura Press.Google Scholar
Egler, F. E. (1981). Untitled letter to the editor. Bulletin of the Ecological Society of America, 62, 230–2.Google Scholar
Ellis, J. E., and Galvin, K. A. (1994). Climate patterns and land-use practices in the dry zones of Africa. BioScience, 44, 340–9.CrossRefGoogle Scholar
Ellis, J. E., and Swift, D. M. (1988). Stability of African pastoral ecosystems: Alternate paradigms and implications for development. Journal of Range Management, 41, 450–9.CrossRefGoogle Scholar
Evangelista, P. H., Kumar, S., Stohlgren, T. J., et al. (2008). Modelling invasion for a habitat generalist and a specialist plant species. Diversity and Distributions, 14, 808–17.CrossRefGoogle Scholar
Evangelista, P. H., Norman, J., Berhanu, L., Kumar, S., and Alley, N. (2008). Predicting habitat suitability for the endemic mountain nyala (Tragelaphus buxtoni) in Ethiopia. Wildlife Research, 35, 409–16.Google Scholar
Evangelista, P., Stohlgren, T. J., Morisette, J. T., and Kumar, S. (2009). Mapping invasive tamarisk (Tamarix): A comparison of single-scene and time-series analyses of remotely sensed data. Remote Sensing, Ecological Status and Change by Remote Sensing special issue, 1, 519–33.Google Scholar
Evangelista, P., Swartzinski, P., and Waltermire, R. (2007). A profile of the mountain nyala (Tragelophus buxtoni). African Indaba, 5(2), special report.Google Scholar
Fassnacht, S. R., Dressler, K. A., and Bales, R. C. (2003). Snow water equivalent interpolation for the Colorado River Basin from snow telemetry (SNOTEL) data. Water Resources Research, 39(8), 1208.Google Scholar
Fassnacht, S. R., Sexstone, G. A., Kashipazha, A. H., et al. (2016). Deriving snow-cover depletion curves for different spatial scales from remote sensing and snow telemetry data. Hydrological Processes, 30, 1708–17.CrossRefGoogle Scholar
Fassnacht, S. R., Web, R. W., and Sanford, W. E. (2017). Headwater regions – physical, ecological, and social approaches to understand these areas: Introduction to special issue. Frontiers of Earth Science, 11(3), 443–6.Google Scholar
Fassnacht, S. R., Williams, M. W., and Corrao, M. V. (2009). Changes in the surface roughness of snow from millimetre to metre scales. Ecological Complexity, 6, 221–9.Google Scholar
Fedy, B. C., Doherty, K. E., Aldridge, C. L., et al. (2014). Habitat prioritization across large landscapes, multiple seasons, and novel areas: An example using Greater Sage-Grouse in Wyoming. Wildlife Monographs, 190, 139.CrossRefGoogle Scholar
Fenn, M. E., Baron, J. S., Allen, E. B., et al. (2003). Ecological effects of nitrogen deposition in the western United States. BioScience, 53, 404–20.Google Scholar
Forrester, J. W. (1968). Principles of Systems. Cambridge, MA: Wright-Allen Press.Google Scholar
Galvin, K. A., Reid, R. S., Behnke, R. H. Jr., and Hobbs, N. T. (eds) (2008). Fragmentation in Semi-arid and Arid Landscapes. Dordrecht: Springer.Google Scholar
Galvin, K. A., Thornton, P. K., Boone, R. B., and Sunderland, J.. (2004). Climate variability and impacts on East African livestock herders. African Journal of Range and Forage Sciences, 21,183–9.Google Scholar
Gao, W., Zheng, Y. F., Slusser, J. R., et al. (2004). Effects of supplementary ultraviolet-B irradiance on maize yield and qualities: A field experiment. Photochemistry and Photobiology, 80, 127–31.Google Scholar
Gleason, H. A. (1926). The individualistic concept of the plant association. Bulletin of the Torrey Botanical Club, 53, 726.CrossRefGoogle Scholar
Gleason, H. A. (1939). The individualistic concept of the plant association. American Midland Naturalist, 21, 92110.Google Scholar
Goselink, R., Klop, G., Dijkstra, J., and Bannink, A. (2014). Phosphorus metabolism in dairy cattle: A literature study on recent developments and missing links. Livestock Research, Livestock Research Report 910. Wageningen: Wageningen University and Research Centre.Google Scholar
Gough, C. (2012). Terrestrial primary production: Fuel for life. Nature Education Knowledge, 3, 28.Google Scholar
Graham, J., Jarnevich, C., Young, N., Newman, G., and Stohlgren, T. (2011). How will climate change affect the potential distribution of Eurasion tree sparrows Passer montanus in North America? Current Zoology, 57, 648–54.Google Scholar
Green, R. A., and Detling, J. K. (2000). Defoliation-induced enhancement of total aboveground nitrogen yield of grasses. Oikos, 91, 280–4.Google Scholar
Heitschmidt, R. K., and Stuth, J. W. (1991). Grazing Management: An Ecological Perspective. Portland, OR: Timber Press.Google Scholar
Henderson, B. B., Gerber, P. J., Hilinski, T. E., et al. (2015). Greenhouse gas mitigation potential of the world’s grazing lands: Modeling soil carbon and nitrogen fluxes of mitigation practices. Agriculture, Ecosystems & Environment, 207, 91100.Google Scholar
Hilbert, D. W., Swift, D. M., Detling, J. K., and Dyer, M. I. (1981). Relative growth rates and the grazing optimization hypothesis. Oecologia, 51, 1418.Google Scholar
Hobbs, N. T., Andrén, H., Persson, J., Aronsson, M., and Chapron, G. (2012). Native predators reduce harvest of reindeer by Sámi pastoralists. Ecological Applications, 22, 1640–54.Google Scholar
Hobbs, N. T., Galvin, K. A., Stokes, C. J., et al. (2008). Fragmentation of rangelands: Implications for humans, animals, and landscapes. Global Environmental Change 18, 776–85.Google Scholar
Hobbs, N. T., and Hooten, M. B.. (2015). Bayesian Models: A Statistical Primer for Ecologists. Princeton, NJ: Princeton University Press.Google Scholar
Holland, E. A., and Detling, J. K. (1990). Plant response to herbivory and belowground nitrogen cycling. Ecology, 71, 1040–9.Google Scholar
Holland, E. A., Parton, W. J., Detling, J. K., and Coppock, D. L. (1992). Physiological responses of plant populations to herbivory and their consequences for ecosystem nutrient flow. American Naturalist, 140, 685706.Google Scholar
Holling, C. S. (1973). Resilience and stability of ecological systems. Annual Review of Ecology and Systematics, 4, 123.Google Scholar
Holling, C. S. (1992). Cross-scale morphology, geometry, and dynamics of ecosystems. Ecological Monographs, 62, 447502.CrossRefGoogle Scholar
Homer, C., Meyer, D. K., Aldridge, C. L., and Schell, S. J. (2013). Detecting annual and seasonal changes in a sagebrush ecosystem with remote sensing-derived continuous fields. Journal of Applied Remote Sensing, 7(1), 10.1117/1.JRS 7.073508.Google Scholar
Horn, H. S. (1975). Forest succession. Scientific American, 232, 90–8.Google Scholar
Houpt, T. R. (1959). Utilization of blood urea in ruminants. American Journal of Physiology, 197, 115–20.CrossRefGoogle ScholarPubMed
Hutchinson, G. E. (1957). Concluding remarks. Cold Spring Harbor Symposium on Quantitative Biology, 22, 415–27.Google Scholar
Ingham, R. E., and Detling, J. K. (1984). Plant–herbivore interactions in a North American mixed-grass prairie: III. Soil nematode population and root biomass dynamics on a black-tailed prairie dog colony and an adjacent uncolonized area. Oecologia, 63, 307–13.Google Scholar
Ingham, R. E., and Detling, J. K. (1990). Effects of root-feeding nematodes on aboveground net primary production in a North American grassland. Plant and Soil, 121, 279–81.Google Scholar
Innis, G. S., ed. (1978). Grassland Simulation Model. Ecological Studies, 26. New York: Springer.Google Scholar
Jaramillo, V. J., and Detling, J. K. (1988). Grazing history, defoliation, and competition: Effects on shortgrass production and nitrogen accumulation. Ecology, 69, 1599–608.Google Scholar
Jaramillo, V. J., and Detling, J. K. (1992a). Small-scale heterogeneity in a semiarid North American grassland I: Tillering, N uptake and retranslocation in simulated urine patches. Journal of Applied Ecology, 29, 18.CrossRefGoogle Scholar
Jaramillo, V. J., and Detling, J. K. (1992b). Small scale heterogeneity in a semiarid North American grassland II: Cattle grazing of simulated urine patches. Journal of Applied Ecology, 29, 913.CrossRefGoogle Scholar
Johnson, E. A., and Fryer, G. I. (1989). Population dynamics in lodgepole pine–Engelmann spruce forests. Ecology, 70, 1335–45.Google Scholar
Kampf, S. K., and Burges, S. J. (2007). A framework for classifying and comparing distributed hillslope and catchment hydrologic models. Water Resources Research, 43, W05423.Google Scholar
Kampf, S. K., Strobl, B., Hammond, J., et al. (2018). Testing the waters: mobile apps for crowdsourced streamflow data. EOS 99, https://doi.org/10.1029/2018EO096355.Google Scholar
Kampf, S. K., Tyler, S. W., Ortiz, C. A., Muñoz, J. F., and Adkins, P. L. (2005). Evaporation and land surface energy budget at the Salar de Atacama, Northern Chile. Journal of Hydrology, 310, 236–52.Google Scholar
Kandylis, K. (1984). The role of sulphur in ruminant nutrition: A review. Livestock Production Science, 11, 611–24.Google Scholar
Kareiva, P., and Andersen, M. (1988). Spatial aspects of species interactions: The wedding of models and experiments. In Community Ecology, ed. Hasting, A.. New York: Springer, 3854.Google Scholar
Keane, R. E., Hessburg, P. F., Landres, P. B., and Swanson, F. J. (2009). The use of historical range and variability in landscape management. Forest Ecology and Management, 258, 1025–37.Google Scholar
Kilgore, B. M. (1973). The ecological role of fire in Sierran conifer forests: Its application to National Park management. Quaternary Research, 3, 496513.Google Scholar
Klein, J. A., Harte, J., and Zhao, X.-Q. (2007). Experimental warming, not grazing, decreases rangeland quality on the Tibetan Plateau. Ecological Applications, 17, 541–57.Google Scholar
Kumar, S., Graham, J., West, A. M., and Evangelista, P. H. (2014). Using district-level occurrences in MaxEnt for predicting the invasion potential of an exotic insect pest in India. Computers and Electronics in Agriculture, 103, 5562.CrossRefGoogle Scholar
Kumar, S., Neven, L. G., and Wee, Y. L. (2014). Evaluating correlative and mechanistic niche models for assessing the risk of pest establishment. Ecosphere, 5(7), 86.CrossRefGoogle Scholar
Kumar, S., Spaulding, S. A., Stohlgren, T. J., et al. (2009). Potential habitat distribution for the freshwater diatom Didymosphenia geminata in the continental US. Frontiers in Ecology and the Environment, 7, 415–20.Google Scholar
Lauenroth, W. K. (1979). Grassland primary production: North American grasslands in perspective. In Perspectives on Grassland Ecology, ed. French, N. R.. Ecological Studies, 32. New York: Springer Verlag, 324.Google Scholar
Lauenroth, W. K., and Burke, I. C., eds. (2008). Ecology of the Shortgrass Steppe: A Long-Term Perspective. Oxford: Oxford University Press.CrossRefGoogle Scholar
Lauenroth, W. K., Burke, I. C., and Gutmann, M. (1999). The structure and function of ecosystems in the central North American grassland region. Great Plains Research, 9, 223–59.Google Scholar
Laurnroth, W. K., Milchunas, D. G., Sala, O. E., Burke, I. C., and Morgan, J. A. (2008). Net primary production in the Shortgrass Steppe. In Ecology of the Shortgrass Steppe: A Long-Term Perspective. Lauenroth, W. K and Burke, I. C.. Oxford: Oxford University Press, 270305.Google Scholar
Lefsky, M. A., Cohen, W. B., Acker, S. A., et al. (1999). Lidar remote sensing of the canopy structure and biophysical properties of Douglas-fir western hemlock forests. Remote Sensing of Environment, 70, 339–61.Google Scholar
Levin, S. A. (1992). The problem of pattern and scale in ecology: The Robert H. MacArthur Award lecture. Ecology, 73, 1943–67.Google Scholar
Levins, R. (1969). Some demographic and genetic consequences of environmental heterogeneity for biological control. Bulletin of the Entomological Society of America, 15, 237–40.Google Scholar
Li, Z., Liu, S., Tan, Z., et al. (2014). Comparing cropland net primary production estimates from inventory, a satellite-based model, and a process-based model in the Midwest of the United States. Ecological Modelling, 277, 112.Google Scholar
Liu, J., Mooney, H., Hull, V., et al. (2015). Systems integration for global sustainability. Science, 347, 1258832.CrossRefGoogle ScholarPubMed
Loucks, O. L. (1970). Evolution of diversity, efficiency, and community stability. American Zoologist, 10, 1725.CrossRefGoogle ScholarPubMed
MacDonald, L. H. (2000). Evaluating and managing cumulative effects: Process and constraints. Environmental Management, 26, 299315.Google Scholar
MacDonald, L. H., and Huffman, E. L. (2004). Post-fire soil water repellency: Persistence and soil moisture thresholds. Soil Science Society of America Journal 68: 1729–34.Google Scholar
MacDonald, L. H., Smart, A. W., and Wissmar, R. C. (1991). Monitoring guidelines to evaluate effects of forestry activities on streams in the Pacific Northwest and Alaska. EPA/910/9–91–001. Seattle, WA: US Environmental Protection Agency.Google Scholar
McGill, W. B., Hunt, H. W., Woodmansee, R. G., and Reuss, J. O. (1981). Phoenix: A model of the dynamics of carbon and nitrogen in grassland soils. In Terrestrial Nitrogen Cycles: Processes, Ecosystem, Strategies and Management Impacts, ed. Clark, F. E. and Rosswall, T.. Ecological Bulletin, 33. Stockholm: Swedish Natural Science Research Council, 49115.Google Scholar
McHale, D. N., Bunn, M. R., Pickett, S. T. A., and Twine, W. (2013). Urban ecology in a developing world: Why advanced socioecological theory needs Africa. Frontiers in Ecology and the Environment, 11, 556–64.Google Scholar
McHale, M. R., Pickett, S. T. A., Barbosa, O., Bunn, D. N., and Cadenasso, M. L. (2015). The new global urban realm: Complex connected, diffuse, and diverse social-ecological systems. Sustainability, 7, 5211–40.Google Scholar
McNaughton, S. J. (1976). Serengeti migratory wildebeest: Facilitation of energy flow by grazing. Science, 191, 92–4.CrossRefGoogle ScholarPubMed
McNaughton, S. J. (1979). Grazing as an optimization process: Grass-ungulate relationships in the Serengeti. American Naturalist, 113, 691703.Google Scholar
MEA (Millennium Ecosystem Assessment). (2005). Ecosytems and human well-being synthesis. Washington, DC: Island Press.Google Scholar
Meadows, D. H. (2008). Thinking in Systems: A Primer. White River Junction, VT: Chelsea Green Publishing.Google Scholar
Milchunas, D. G., and Lauenroth, W. K. (1992). Carbon dynamics and estimates of primary production by harvest, C-14 dilution, and C-14 turnover. Ecology, 73, 593607.Google Scholar
Milchunas, D. G., Lauenroth, W. K., Burke, I. C., and Detling, J. K. (2008). Effects of grazing on vegetation. In Ecology of the Shortgrass Steppe: A Long-Term Perspective, ed. Lauenroth, W. K. and Burke, I. C.. Oxford: Oxford University Press.Google Scholar
Milly, P. C. D., Betancourt, J., Falkenmark, M., et al. (2008). Climate change – Stationarity is dead: Whither water management? Science, 319, 573–4.Google Scholar
Mingyang, L., Yunwei, J., Kumar, S., and Stohlgen, T. J. (2008). Modeling potential habitats for alien species Dreissena polymorpha in continental USA. Acta Ecological Sinica, 28, 4253–8.Google Scholar
Moir, R. J. (1970). Implications of the N:S ratio and differential recycling. In Symposium: Sulphur in Nutrition, ed. Muth, O. H. and Oldfield, J. E.. Westport , CT: AVI Publishing Co.Google Scholar
Mosier, A. R., Parton, W. J., Martin, R. E., et al. (2008). Soil–atmosphere exchange of trace gasses in the Colorado shortgrass steppe. Ecology of the Shortgrass Steppe: A Long-Term Perspective ed. Lauenroth, W. K. and Burke, I. C.. Oxford: Oxford University Press.Google Scholar
Nash, K. L., Allen, C. R., Angeler, D. G., et al. (2014). Discontinuities, cross-scale patterns, and the organization of ecosystems. Ecology, 95, 654–67.Google Scholar
NREL. (2018). Natural Resource Ecology Laboratory. www.nrel.colostate.edu (accessed June 6, 2018).Google Scholar
Odum, E. P. (1969). The strategy of ecosystem development. Science, 164, 262–70.Google Scholar
Odum, E. P., and Odum, H. T. (1963). Fundamentals of Ecology, 2nd edn. E. P. Odum in collaboration with H. T. Odum. Philadelphia and London: W. B. Saunders.Google Scholar
Ogle, S. M., Breidt, F. J., Eve, M. D., and Paustain, K. (2003). Uncertainty in estimating land use and management impacts on soil organic matter storage for US agricultural lands between 1982 and 1997. Global Change Biology, 9, 1521–42.Google Scholar
Ogle, S. M., Breidt, F. J., and Paustain, K. (2005). Agricultural management impacts on soil organic carbon storage under moist and dry climatic conditions of temperate and tropical regions. Biogeochemistry, 72, 87121.Google Scholar
Ogutu, J. O., Piepho, H.-P., Dublin, H. T., Bhola, N. and Reid, R. S. (2009). Dynamics of Mara-Serengeti ungulates in relation to land use changes. Journal of Zoology, 278, 114.Google Scholar
Ojima, D. S., Galvin, K. A., and Turner, B. L. (1994). The global impact of land-use change. BioScience, 44, 300–4.Google Scholar
Ojima, D., Garcia, L., Elgaali, E., et al. (1999). Potential climate change impacts on water resources in the Great Plains. Journal of the American Water Resources Association, 35(6), 1443–54.Google Scholar
Ojima, D. S., Kittel, T. G. F., Rosswall, T., and Walker, B. H. (1991). Critical issues for understanding global change effects on terrestrial ecosystems. Ecological Applications, 1, 316–25.Google Scholar
Ojima, D., Mosier, A., DelGrosso, S. J., and Parton, W. J. (2000). TRAGNET analysis and synthesis of trace gas fluxes. Global and Biogeochemical Cycles, 14, 995–7.Google Scholar
Ojima, D. S., Schimel, D. S., Parton, W. J., and Owensby, C. W. (1994). Long- and short-term effects of fire on nitrogen cycling in tallgrass prairie. Biogeochemistry, 24, 6784.Google Scholar
O’Neill, R. V. (1988). Hierarchy theory and global change. In Scales and Global Change: Spatial and Temporal Variability in Biospheric and Geospheric Processes, ed. Rosswall, T., Woodmansee, R. G., and Risser, P. G.. SCOPE Series 38. Hoboken, NJ: John Wiley and Sons, 2945.Google Scholar
Painter, E. L., Detling, J. K., and Steingraeber, D. A. (1993). Plant morphology and grazing history: Relationships between native grasses and herbivores. Vegetatio, 106, 3762.CrossRefGoogle Scholar
Parton, W. J., Anderson, D. W., Cole, C. V., and Stewart, J. W. B. (1983). Simulation of soil organic matter formation and mineralization in semiarid agroecosystems. In Nutrient Cycling in Agricultural Ecosystems, ed. Lowrance, R. R., Todd, R. L., Asmussen, L. E., and Leonard, R. A.. Special Publication No. 23. Athens, GA: University of Georgia, College of Agriculture Experiment Station.Google Scholar
Parton, W. J., Hartman, M., Ojima, D., and Schimel, D. (1998). DAYCENT and its land surface submodel: Description and testing. Global and Planetary Change, 19, 3548.Google Scholar
Parton, W. J., Schimel, D. S., Cole, C. V., and Ojima, D. S. (1987). Analysis of factors controlling soil organic matter levels in Great Plains grasslands. Soil Science Society of America Journal, 51, 1173–9.Google Scholar
Parton, W., Silver, W. L., Burke, I. C., et al. (2007). Global-scale similarities in nitrogen release patterns during long-term decomposition. Science, 315, 362–4.Google Scholar
Paul, E. A., ed. (2015). Soil Microbiology, Ecology and Biochemistry Academic Press, 4th edn. Burlington, MA: Academic Press .Google Scholar
Paul, E. A., Paustian, K., Elliott, E. T., and Cole, C. V., eds. (1997). Soil Organic Matter in Temperate Ecosystems. Boca Raton, FL: CRC Press.Google Scholar
Paustian, K., Cole, C. V., Sauerbeck, D., and Sampson, N. (1998). CO2 mitigation by agriculture: An overview. Climatic Change, 40, 135–62.Google Scholar
Peters, D. P. C., Sala, O. E., Allen, C. D, Covich, A., and Brunson, M. (2007). Cascading events in linked ecological and socioeconomic systems. Frontiers in Evolution and the Environment, 5, 221–4.Google Scholar
Phillips, S. J., and Dudik, M. (2008). Modeling of species distributions with Maxent: New extensions and a comprehensive evaluation. Ecography, 31, 161–75.Google Scholar
Phillips, S. J., Dudik, M., and Schapire, R. E. (2004). A maximum entropy approach to species distribution modeling. Proceedings of the 21st International Conference on Machine Learning. New York: ACM Press, 655–62.Google Scholar
Pickett, S. T. A., Cadenasso, M. L., Childers, D. L., McDonnell, M. J., and Zhou, W. (2016). Evolution and future of urban ecological science: Ecology in, of, and for the city. Ecosystem Health and Sustainability, 2(7), e01229.Google Scholar
Pickett, S. T. A., Cadenasso, M. L., and Meiners, S. J. (2009). Ever since Clements: From succession to vegetation dynamics and understanding to intervention. Applied Vegetation Science, 12, 921.Google Scholar
Polley, H. W., and Detling, J. K. (1989). Defoliation, nitrogen, and competition: Effects on plant growth and nitrogen nutrition. Ecology, 70, 721–7.Google Scholar
Rahel, F. J. (2002). Homogenization of freshwater faunas. Annual Review of Ecology, Evolution, and Systematics, 33, 291315.Google Scholar
Reddy, K. R., Kakani, V. G., Zhao, D., Koti, S., and Gao, W. (2004). Interactive effects of ultraviolet-B radiation and temperature on cotton physiology, growth, development and hyperspectral reflectance. Photochemistry and Photobiology, 79, 416–27.Google Scholar
Redfield, A .C. (1934). On the proportions of organic derivatives in sea water and their relation to the composition of plankton. James Johnstone Memorial Volume. Liverpool: Liverpool University Press, 176–92.Google Scholar
Reid, R. S. (2012). Savanas of Our Birth. London: University of California Press.Google Scholar
Reid, R. S., Gichohi, H., Said, M. Y., et al. (2008). Fragmentation of a peri-urban savanna, Athi-Kaputiei Plains, Kenya. In Fragmentation in Semi-Arid and Arid Landscapes, ed. Galvin, K. A., Reid, R. S., Behnke Jr, R. H.., and Hobbs, N. T.. Dordrecht: Springer, 195224.Google Scholar
Reid, R. S., and Ellis, J. E. (1995). Impacts of pastoralists on woodlands in South Turkana, Kenya: Livestock-mediated tree regeneration. Ecological Applications, 5, 978–92.Google Scholar
Risser, P. G. (1990). Landscape pattern and its effects on energy and nutrient distribution. In Changing Landscapes: An Ecological Perspective, Zonneveld, I. S. and Forman, R. R. T.. New York: Springer, 4556.Google Scholar
Robertson, G. P., and Groffman, P. M. (2015). Nitrogen transformations. In Soil Microbiology, Ecology and Biochemistry, ed. Paul, E. A.. Burlington, MA: Academic Press, 421–46.Google Scholar
Rodewald, A. D. (2003). The importance of land uses within the landscape matrix. Wildlife Society Bulletin, 31, 586–92.Google Scholar
Rollins, M. G. (2009). LANDFIRE: A nationally consistent vegetation, wildland fire, and fuel assessment. International Journal of Wildland Fire, 18, 235–49.CrossRefGoogle Scholar
Romme, W. H. (1982). Fire and landscape diversity in subalpine forests of Yellowstone National Park. Ecological Monographs, 52, 199221.Google Scholar
Romme, W. H., Whitby, T. G., Tinker, D. B., and Turner, M. G. (2016). Deterministic and stochastic processes lead to divergence in plant communities 25 years after the 1988 Yellowstone fires. Ecological Monographs, 86, 327–51.Google Scholar
Root, T. (1988). Environmental factors associated with avian distributional limits. Journal of Biogeography, 15, 489505.Google Scholar
Rosswall, T., Woodmansee, R. G., and Risser, P. G. (1988). Scales and Global Change: Spatial and Temporal Variability in Biospheric and Geospheric Processes. Scope Series (Book 38). Hoboken, NJ: John Wiley and Sons.Google Scholar
Ryan, M. G. (1991). Effects of climate change on plant respiration. Ecological Applications, 1, 157–67.Google Scholar
Ryan, M. G., and Asao, S. (2014). Phloem transport in trees. Tree Physiology, 34, 14.Google Scholar
Safranyik, L., Shrimpton, D. M., and Whitney, H. S. (1974). Management of lodgepole pine to reduce losses from the mountain pine beetle. Forestry Technical Report 1. Victoria, BC: Natural Resources Canada, Canadian Forest Service, Pacific Forestry Centre.Google Scholar
Sankaran, M., et al. (2005). Determinants of woody cover in African savannas. Nature, 438, 846–9.Google Scholar
Sauer, J. R., Niven, D. K., Hines, J. E., et al. (2017). The North American Breeding Bird Survey, Results and Analysis 1966–2015. Version 2.07.2017. Laurel, MD: USGS Patuxent Wildlife Research Center.Google Scholar
Schimel, D. S., Melillo, M., Tian, H., et al. (2000). Contribution of increasing CO2 and climate to carbon storage by ecosystems in the United States. Science, 287, 2004–6.Google Scholar
Schimel, D. S., Parton, W. J., Adamsen, F. J., et al. (1986). Role of cattle in the volatile loss of nitrogen from a shortgrass steppe. Biogeochemistry, 2, 3952.Google Scholar
Schimel, D. S., Stillwell, M. A., and Woodmansee, R. G. (1985). Biogeochemistry C, N, and P in a soil catena of the shortgrass steppe. Ecology, 66, 276–82.Google Scholar
Senft, R. L., Coughenour, M. B., Bailey, D. W., et al. (1987). Large herbivore foraging and ecological hierarchies. BioScience, 37, 789–99.Google Scholar
Six, J., Conant, R. T., Paul, E. A., and Paustian, K. (2002). Stabilization mechanisms of soil organic matter: Implications for C-saturation of soils. Plant and Soil, 241, 155–76.CrossRefGoogle Scholar
Six, J., Elliott, E. T., and Paustian, K. (2000). Soil microaggregate turnover and microaggregate formation: A mechanism for C sequestration under no-tillage agriculture. Soil Biology and Biochemistry, 32, 2099–103.Google Scholar
Six, J., Ogle, S. M., Conant, R. T., Mosier, A. R., and Paustian, K. (2004). The potential to mitigate global warming with no-tillage management is only realized when practiced in the long term. Global Change Biology, 10, 155–60.Google Scholar
Smith, P., Martino, D., Cai, Z., et al. (2008). Greenhouse gas mitigation in agriculture. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 363, 789813.Google Scholar
Stabach, J. A., Wittemyer, G., Boone, R. B., Reid, R. S., and Worden, J. S. (2016). Variation in habitat selection by white-bearded wildebeest across different degrees of human disturbance. Ecosphere, 7(8), e01428.Google Scholar
Stohlgren, T. J. (2007). Measuring Plant Diversity: Lessons from the Field. New York: Oxford University Press.Google Scholar
Stohlgren, T. J., Barnett, D., Flather, C., et al. (2006). Species richness and patterns of invasion in plants, birds, and fishes in the United States. Biological Invasions, 8, 427–57.Google Scholar
Stohlgren, T. J., Barnett, D., and Kartesz, J. (2003). The rich get richer: Patterns of plant invasions in the United States. Frontiers in Ecology and the Environment, 1(1), 1114.Google Scholar
Stohlgren, T. J., Binkley, D., Chong, G. W., et al. (1999). Exotic plant species invade hot spots of native plant diversity. Ecological Monographs, 69, 2546.Google Scholar
Stohlgren, T. J., Bull, K. A., and Otsuki, Y. (1998). Comparison of rangeland sampling techniques in the Central Grasslands. Journal of Range Management, 51, 164–72.Google Scholar
Stohlgren, T. J., and Schnase, J. L. (2006). Risk analysis for biological hazards: What we need to know about invasive species. Risk Analysis, 26, 163–73.Google Scholar
Stohlgren, T. J., Szalanski, A. L., Gaskin, J., et al. (2014). From hybrid swarms to swarms of hybrids. Environment and Ecology Research, 2(8), 311–18.Google Scholar
Swetnam, T. W. (1993). Fire history and climate-change in giant sequoia groves. Science, 262, 885–9.Google Scholar
Thurow, T. L. (1991). Hydrology and erosion. In Grazing Management: An Ecological Perspective, ed. Heitschmidt, R. K. and Stuth, J. W.. Portland, OR: Timber Press.Google Scholar
Turner, M. G., Romme, W. H., Gardner, R. H., O’Neill, R. V., and Kratz, T. K. (1993). A revised concept of landscape equilibrium: Disturbance and stability on scaled landscapes. Landscape Ecology, 8, 213–27.Google Scholar
Van Dyne, G. M., ed. (1969). The Ecosystem Concept in Natural Resource Management. New York: Academic Press.Google Scholar
Van Dyne, G. M., and Anway, J. C. (1976). Research program for and process of building and testing grassland ecosystem models. Journal of Range Management, 29, 114–22.Google Scholar
Wakie, T. T., Laituri, M., and Evangelista, P. H. (2016). Assessing the distribution and impacts of Prosopis juliflora through participatory approaches. Applied Geography, 66, 132–43.Google Scholar
Wall, D. H., Bradford, M. A., St. John, M. G., et al. (2008). Global decomposition experiment shows soil animal impacts on decomposition are climate-dependent. Global Change Biology, 14, 2661–77.Google Scholar
Wallenstein, M. D., and Hall, E. K. (2012). A trait-based framework for predicting when and where microbial adaptation to climate change will affect ecosystem functioning. Bioegeochemistry, 109, 3547.Google Scholar
Wann, G. T., Aldridge, G. L., and Braun, C. E. (2016). Effects of seasonal weather on breeding phenology and reproductive success of alpine ptarmigan in Colorado. PLoS ONE, 11(7), e0158913.Google Scholar
Watt, A. S. (1947). Pattern and process in the plant community. The Journal of Ecology, 35: 122.Google Scholar
Westoby, M., Walker, B., and Noymeir, I. (1989). Opportunistic management for rangelands not at equilibrium. Journal of Range Management, 42, 266–74.Google Scholar
Whicker, A. D., and Detling, J. K. (1988). Ecological consequences of prairie dog disturbances. BioScience, 38, 778–85.Google Scholar
Whittaker, R. H. (1975). Communities and Ecosystems, 2nd edn. New York: Macmillan.Google Scholar
Wiens, J. A. (1984). On understanding a non-equilibrium world: Myth and reality in community patterns and processes. In Ecological Communities: Conceptual Issues and the Evidence, ed. Strong, D. R., Simberloff, D., Abele, L. G., and Thistle, A. B.. Princeton, NJ: Princeton University Press.Google Scholar
Wiens, J. A. (1989). Spatial scaling in ecology. Functional Ecology, 3, 385–97.Google Scholar
Williams, G. J. III, and Kemp, P. R. (1978). Simultaneous measurement of leaf and root gas exchange of shortgrass prairie species. International Journal of Plant Sciences, 139, 150–7.Google Scholar
Woodmansee, R. G. (1978). Additions and losses of nitrogen in grassland ecosystems. BioScience, 28, 448–53.Google Scholar
Woodmansee, R. G. (1990). Biogeochemistry cycles and ecological hierarchies. In Changing Landscapes: An Ecological Perspective, ed. Zonneveld, I. S. and Forman, R. R. T.. New York: Springer, 5771.Google Scholar
Woodmansee, R. G, Dodd, J. L., Bowman, R. A., Clark, F. E., and Dickinson, C. E. (1978). Nitrogen budget of a shortgrass prairie ecosystem. Oecologia, 34, 363–76.Google Scholar
Woodmansee, R. G., Vallis, I., and Mott, J. J. (1981). Grassland nitrogen. In Terrestrial Nitrogen Cycles: Processes, Ecosystem Strategies, and Management Impacts, ed. Clark, F. E. and Rosswall, T.. Ecological Bulletin, 33. Stockholm: Swedish Natural Science Research Council, 443–62.Google Scholar
WSS. (2018). Web Soil Survey. USDA Natural Resource Conservasion Service. https://websoilsurvey.sc.egov.usda.gov/App/HomePage.htm (accessed June 5, 2018).Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×