Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-nr4z6 Total loading time: 0 Render date: 2024-06-08T15:30:49.418Z Has data issue: false hasContentIssue false

Part III - Dealing with climate change effects

Published online by Cambridge University Press:  22 March 2019

David J. Gibson
Affiliation:
Southern Illinois University, Carbondale
Jonathan A. Newman
Affiliation:
University of Guelph, Ontario
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2019

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

15.7 References

Galvin, KA. Transitions: pastoralists living with change. Annual Review of Anthropology. 2009;38:185–98.CrossRefGoogle Scholar
Boserup, E. The conditions of agricultural growth: the economics of agrarian change under population pressure. London: Aldine Publishing Company; 1965.Google Scholar
McIntire, J, Bourzat, D, Pingali, P. Crop–livestock interaction in sub-Saharan Africa. Washington, DC: The World Bank; 1992.Google Scholar
Venables, AJ, Limão, N. Geographical disadvantage. Policy Research Working Paper 2256. Washington, DC: The World Bank; 1999.Google Scholar
Steffen, W, Broadgate, W, Deutsch, L, Gaffney, O, Ludwig, C. The trajectory of the Anthropocene: the great acceleration. The Anthropocene Review. 2015;2(1):8198.CrossRefGoogle Scholar
Homewood, K, Lambin, EF, Coast, E, Kariuki, A, Kikula, I, Kivelia, J, et al. Long-term changes in Serengeti–Mara wildebeest and land cover: pastoralism, population, or policies? Proceedings of the National Academy of Sciences of the USA. 2001;98(22):12,544–9.CrossRefGoogle ScholarPubMed
Steinfeld, H, Gerber, P, Wassenaar, T, Castel, V, de Haan, C. Livestock’s long shadow: environmental issues and options. Rome: Food & Agriculture Organization; 2006.Google Scholar
Herrero, M, Addison, J, Bedelian, C, Carabine, E, Havlík, P, Henderson, B, et al. Climate change and pastoralism: impacts, consequences and adaptation. Revue Scientifique et Technique, Office International des Epizooties. 2016;35(2):417–33.Google ScholarPubMed
Hobbs, NT, Galvin, KA, Stokes, CJ, Lackett, JM, Ash, AJ, Boone, RB, et al. Fragmentation of rangelands: implications for humans, animals, and landscapes. Global Environmental Change. 2008;18(4):776–85.Google Scholar
Thornton, PK, van de Steeg, J, Notenbaert, A, Herrero, M. The impacts of climate change on livestock and livestock systems in developing countries: a review of what we know and what we need to know. Agricultural Systems. 2009;101(3):113–27.Google Scholar
Reid, RS, Fernández-Giménez, ME, Galvin, KA. Dynamics and resilience of rangelands and pastoral peoples around the globe. Annual Review of Environment and Resources. 2014;39:217–42.Google Scholar
Porter, JR, Xie, L, Challinor, AJ, Cochrane, K, Howden, SM, Iqbal, MM, et al. Food security and food production systems. In: Field, CB, Barros, VR, editors. Climate change 2014: impacts, adaptation and vulnerability part A: global and sectoral aspects. Cambridge: Cambridge University Press; 2014.Google Scholar
Boone, RB, Conant, RT, Sircely, J, Thornton, PK, Herrero, M. Climate change impacts on selected global rangeland ecosystem services. Global Change Biology. 2018;24(3):1382–93.CrossRefGoogle ScholarPubMed
Havlík, P, Leclère, D, Valin, H, Herrero, M, Schmid, E, Soussana, JF, et al. Global climate change, food supply and livestock production systems: a bioeconomic analysis. In: Elbehri, A, editor. Climate change and food systems: global assessments and implications for food security and trade. Rome: Food & Agriculture Organization; 2015.Google Scholar
Homewood, K, Kristjanson, P, Trench, P. Staying Maasai? Livelihoods, conservation and development in East African rangelands. New York, NY: Springer; 2009.Google Scholar
Thornton, PK, Herrero, M. Adapting to climate change in the mixed crop and livestock farming systems in sub-Saharan Africa. Nature Climate Change. 2015;5(9):830.CrossRefGoogle Scholar
Reid, RS, Galvin, KA, Kruska, RS. Fragmentation in semi-arid and arid landscapes. In: Galvin, KA, Reid, RS, Behnke, R, Hobbs, NT, editors. Consequences for human and natural systems. Dordrecht: Springer; 2008. pp. 124.Google Scholar
Turner, MD, McPeak, JG, Ayantunde, A. The role of livestock mobility in the livelihood strategies of rural peoples in semi-arid West Africa. Human Ecology. 2014;42(2):231–47.Google Scholar
Megersa, B, Markemann, A, Angassa, A, Ogutu, JO, Piepho, H-P, Zárate, AV. Livestock diversification: an adaptive strategy to climate and rangeland ecosystem changes in southern Ethiopia. Human Ecology. 2014;42(4):509–20.Google Scholar
Lennox, E. Double exposure to climate change and globalization in a Peruvian highland community. Society & Natural Resources. 2015;28(7):781–96.Google Scholar
Shackleton, R, Shackleton, C, Shackleton, S, Gambiza, J. Deagrarianisation and forest revegetation in a biodiversity hotspot on the Wild Coast, South Africa. PLoS ONE. 2013;8(10):e76939.Google Scholar
Liao, C, Fei, D. Sedentarization as constrained adaptation: evidence from pastoral regions in far northwestern China. Human Ecology. 2017;45(1):2335.Google Scholar
López-i-Gelats, F, Paco, JC, Huayra, RH, Robles, ODS, Peña, ECQ, Filella, JB. Adaptation strategies of Andean pastoralist households to both climate and non-climate changes. Human Ecology. 2015;43(2):267–82.CrossRefGoogle Scholar
Namgay, K, Millar, JE, Black, RS, Samdup, T. Changes in transhumant agro-pastoralism in Bhutan: a disappearing livelihood? Human Ecology. 2014;42(5):779–92.Google Scholar
Xu, J, Yang, Y, Li, Z, Tashi, N, Sharma, R, Fang, J. Understanding land use, livelihoods, and health transitions among Tibetan nomads: a case from Gangga Township, Dingri County, Tibetan Autonomous Region of China. EcoHealth. 2008;5(2):104.Google Scholar
Schmidt, M, Pearson, O. Pastoral livelihoods under pressure: ecological, political and socioeconomic transitions in Afar (Ethiopia). Journal of Arid Environments. 2016;124:2230.Google Scholar
Galvin, KA, Beeton, TA, Boone, RB, BurnSilver, SB. Nutritional status of Maasai pastoralists under change. Human Ecology. 2015;43(3):411–24.Google Scholar
Lkhagvadorj, D, Hauck, M, Dulamsuren, C, Tsogtbaatar, J. Pastoral nomadism in the forest-steppe of the Mongolian Altai under a changing economy and a warming climate. Journal of Arid Environments. 2013;88:82–9.CrossRefGoogle Scholar
Ogalleh, SA, Vogl, CR, Eitzinger, J, Hauser, M. Local perceptions and responses to climate change and variability: the case of Laikipia District, Kenya. Sustainability. 2012;4(12):3302–25.CrossRefGoogle Scholar
Rufino, MC, Thornton, PK, Mutie, I, Jones, PG, Van Wijk, MT, Herrero, M. Transitions in agro-pastoralist systems of East Africa: impacts on food security and poverty. Agriculture, Ecosystems & Environment. 2013;179:215–30.Google Scholar
Antwi-Agyei, P, Stringer, LC, Dougill, AJ. Livelihood adaptations to climate variability: insights from farming households in Ghana. Regional Environmental Change. 2014;14(4):1615–26.Google Scholar
Ash, A, Thornton, P, Stokes, C, Togtohyn, C. Is proactive adaptation to climate change necessary in grazed rangelands? Rangeland Ecology & Management. 2012;65(6):563–8.Google Scholar
de Leeuw, J, Osano, P, Said, MY, Ayantunde, AA, Dube, S, Neely, C, et al. Pastoral farming systems and food security in Sub-Saharan Africa. Priorities for science and policy. In: Dixon, J, Garrity, DP, Boffa, J-M, Williams, TO, Amede, T, Auricht, C, et al., editors. Farming systems and food security in Africa: priorities for science and policy under global change. London: Routledge; 2017.Google Scholar
Thornton, PK, Boone, RB, Ramirez-Villegas, J. Climate change impacts on livestock. CCAFS Working Paper no. 120. Copenhagen: CGIAR Research Program on Climate Change, Agriculture and Food Security (CCAFS); 2015.Google Scholar
Asongu, SA, Nwachukwu, JC. The mobile phone in the diffusion of knowledge for institutional quality in sub-Saharan Africa. World Development. 2016;86:133–47.Google Scholar
Rogelj, J, Schaeffer, M, Friedlingstein, P, Gillett, NP, Van Vuuren, DP, Riahi, K, et al. Differences between carbon budget estimates unravelled. Nature Climate Change. 2016;6(3):245.CrossRefGoogle Scholar
Doherty, RM, Sitch, S, Smith, B, Lewis, SL, Thornton, PK. Implications of future climate and atmospheric CO2 content for regional biogeochemistry, biogeography and ecosystem services across East Africa. Global Change Biology. 2010;16(2):617–40.Google Scholar
Jones, PG, Thornton, PK. Croppers to livestock keepers: livelihood transitions to 2050 in Africa due to climate change. Environmental Science & Policy. 2009;12(4):427–37.CrossRefGoogle Scholar
Weindl, I, Lotze-Campen, H, Popp, A, Müller, C, Havlík, P, Herrero, M, et al. Livestock in a changing climate: production system transitions as an adaptation strategy for agriculture. Environmental Research Letters. 2015;10(9):094021.CrossRefGoogle Scholar
Havlík, P, Valin, H, Herrero, M, Obersteiner, M, Schmid, E, Rufino, MC, et al. Climate change mitigation through livestock system transitions. Proceedings of the National Academy of Sciences of the USA. 2014;111(10):3709–14.CrossRefGoogle ScholarPubMed
Lieffering, M, Newton, PCD, Vibart, R, Li, FY. Exploring climate change impacts and adaptations of extensive pastoral agriculture systems by combining biophysical simulation and farm system models. Agricultural Systems. 2016;144:7786.Google Scholar
Thornton, PK, Boone, RB, Galvin, KA, BurnSilver, SB, Waithaka, MM, Kuyiah, J, et al. Coping strategies in livestock-dependent households in east and southern Africa: a synthesis of four case studies. Human Ecology. 2007;35(4):461–76.CrossRefGoogle Scholar
Puig, CJ, Greiner, R, Huchery, C, Perkins, I, Bowen, L, Collier, N, et al. Beyond cattle: potential futures of the pastoral industry in the Northern Territory. The Rangeland Journal. 2011;33(2):181–94.CrossRefGoogle Scholar
Murakami, D, Yamagata, Y. Estimation of gridded population and GDP scenarios with spatially explicit statistical downscaling. arXiv preprint arXiv:161009041. 2016.Google Scholar
Nelson, A. Estimated travel time to the nearest city of 50,000 or more people in year 2000. Global Environment Monitoring Unit–Joint Research Centre of the European Commission, Ispra, Italy. Available at http://biovaljrceceuropaeu/products/gam/ (accessed 11/07/2017). 2008.Google Scholar
Jurković, RS, Pasarić, Z. Spatial variability of annual precipitation using globally gridded data sets from 1951 to 2000. International Journal of Climatology. 2013;33(3):690–8.Google Scholar
Schilling, J, Akuno, M, Scheffran, J, Weinzierl, T. On raids and relations: climate change, pastoral conflict and adaptation in northwestern Kenya. In: Bronkhorst, S, Urmilla, B, editors. Climate change and conflict: where to for conflict sensitive climate adaptation in Africa? Berlin: Wissenschaftsverlag; 2014. pp. 241–68.Google Scholar
Lipper, L, Thornton, P, Campbell, BM, Baedeker, T, Braimoh, A, Bwalya, M, et al. Climate-smart agriculture for food security. Nature Climate Change. 2014;4(12):1068.Google Scholar
Butler, CK, Gates, S. African range wars: climate, conflict, and property rights. Journal of Peace Research. 2012;49(1):2334.CrossRefGoogle Scholar
Davis, J, Lopez-Carr, D. Migration, remittances and smallholder decision-making: implications for land use and livelihood change in Central America. Land Use Policy. 2014;36:319–29.Google Scholar
Jensen, ND, Barrett, CB, Mude, AG. Index insurance quality and basis risk: evidence from northern Kenya. American Journal of Agricultural Economics. 2016;98(5):1450–69.Google Scholar
Köhler-Rollefson, I. Innovations and diverse livelihood pathways: alternative livelihoods, livelihood diversification and societal transformation in pastoral communities. Revue Scientifique et Technique–Office International des Epizooties. 2016;35(2):611–8.Google Scholar
Greiner, R, Puig, J, Huchery, C, Collier, N, Garnett, ST. Scenario modelling to support industry strategic planning and decision making. Environmental Modelling & Software. 2014;55:120–31.Google Scholar
Vervoort, JM, Bendor, R, Kelliher, A, Strik, O, Helfgott, AER. Scenarios and the art of worldmaking. Futures. 2015;74:6270.CrossRefGoogle Scholar

16.8 References

Thornton, PK, van de Steeg, J, Notenbaert, A, Herrero, M. The impacts of climate change on livestock and livestock systems in developing countries: a review of what we know and what we need to know. Agricultural Systems. 2009;101(3):113–27.CrossRefGoogle Scholar
Herrero, M, Addison, J, Bedelian, C, Carabine, E, Havlík, P, Henderson, B, et al. Climate change and pastoralism: impacts, consequences and adaptation. Revue Scientifique et Technique. 2016;35(2):417–33.Google Scholar
Seo, SN, Mendelsohn, R. An analysis of crop choice: adapting to climate change in South American farms. Ecological Economics. 2008;67(1):109–16.Google Scholar
Netting, RMc. Smallholders, householders: farm families and the ecology of intensive, sustainable agriculture. Stanford, CA: StanfordUniversity Press; 1993.Google Scholar
Dove, MR. Southeast Asian grasslands: understanding a folk landscape. New York, NY: New York Botanical Gardens Press; 2008.Google Scholar
MacDonald, GE. Cogongrass (Imperata cylindrica) – biology, ecology, and management. Critical Reviews in Plant Sciences. 2004;23(5):367–80.Google Scholar
Bartlett, HH. Fire in relation to primitive agriculture and grazing in the tropics; annotated bibliography. Ann Arbor, MI: University of Michigan Botanical Gardens; Vol. 1, 1955.Google Scholar
Dove, MR. Perception of volcanic eruption as agent of change on Merapi volcano, Central Java. Journal of Volcanology and Geothermal Research. 2008;172(3–4):329–37.Google Scholar
Th Pigeaud, TG. Java in the fourteenth century. Vol. IV Commentaries and recapitulations. The Hague: Martinus Nijhoff; 1962.Google Scholar
Wolff, JU. The place of plant names in reconstructing Proto Austronesian. In: Pawley, AK, Ross, MD, editors. Austronesian terminologies: continuity and change. Canberra: Department of Linguistics, Research School of Pacific and Asian Studies, Australian National University; 1994: pp. 511–40.Google Scholar
Hadiwidjojo, GP. Alang-Alang, kumitir. Paper presented at the Radya Pustaka Museum, 28 December 1956. Surakarta, Central Java, Indonesia.Google Scholar
Carpenter, C. Brides and bride-dressers in contemporary Java [dissertation]. Ithaca, NY: Cornell University; 1987.Google Scholar
Oxford English Dictionary. ‘swidden, n.’. Oxford: Oxford University Press (available from: http://bit.ly/2qyMSWN).Google Scholar
Conklin, HC. Shifting cultivation and succession to grassland climax. Proceedings of the Ninth Pacific Science Congress of the Pacific Science Association; 1959.Google Scholar
Dove, MR. Peasant versus government perception and use of the environment: a case-study of Banjarese ecology and river basin development in South Kalimantan. Journal of Southeast Asian Studies. 1986;17(1):113–36.Google Scholar
Dove, MR. The practical reason of weeds in Indonesia: peasant vs. state views of Imperata and Chromolaena imper. Human Ecology. 1986;14(2):163–90.Google Scholar
Clements, FE. Plant succession: an analysis of the development of vegetation. Washington, DC: Carnegie Institution of Washington; 1916.Google Scholar
Worster, D. The ecology of order and chaos. Environmental History Review. 1990;14(1/2):118.CrossRefGoogle Scholar
Laris, P, Caillault, S, Dadashi, S, Jo, A. The human ecology and geography of burning in an unstable savanna environment. Journal of Ethnobiology. 2015;35(1):111–39.Google Scholar
Wharton, CH. Man, fire and wild cattle in Southeast Asia. In: Proceedings of the Annual Tall Timbers Fire Ecology Conference; 1968;8:107–67; Tallahassee, FL: Tall Timbers Research Station.Google Scholar
Sherman, G. What ‘green desert’? The ecology of Batak grassland farming. Indonesia. 1980;29:113–48.CrossRefGoogle Scholar
Lefebvre, H. The production of space (trans. Nicholson-Smith D). Oxford: Blackwell; 1991.Google Scholar
Harvey, D. Justice, nature and the geography of difference. Cambridge, MA: Blackwell; 1996.Google Scholar
Fan, M, Li, Y, Li, W. Solving one problem by creating a bigger one: the consequences of ecological resettlement for grassland restoration and poverty alleviation in Northwestern China. Land Use Policy. 2015;42:124–30.Google Scholar
Holm, LG, Plucknett, DL, Pancho, JV, Herberger, JP. The world’s worst weeds. Honolulu, HI: University Press of Hawaii for the East-West Center; 1977.Google Scholar
Bryson, CT, Carter, R. Cogongrass, Imperata cylindrica, in the United States. Weed alert! Weed Technology. 1993;7(4):1005–9.Google Scholar
Bradley, BA, Wilcove, DS, Oppenheimer, M. Climate change increases risk of plant invasion in the Eastern United States. Biological Invasions. 2010;12(6):1855–72.Google Scholar
Geertz, C. Agricultural involution: the process of ecological change in Indonesia. Berkeley, CA: University of California Press; 1966.Google Scholar
Oxford English Dictionary. ‘weed, n.1’. Oxford: Oxford University Press (updated 15 April 2018; available from: http://bit.ly/2H4jwWF).Google Scholar
Dove, M. The banana tree at the gate: a history of marginal peoples and global markets in Borneo. New Haven, CT: Yale University Press; 2011.Google Scholar
Lippincott, CL. Effects of Imperata cylindrica (L.) Beauv. (Cogongrass) invasion on fire regime in Florida Sandhill (USA). Natural Areas Journal. 2000;20(2):140–9.Google Scholar
Sunderlin, WD, Larson, AM, Duchelle, AE, Resosudarmo, IAP, Huynh, TB, Awono, A, et al. How are REDD+ proponents addressing tenure problems? Evidence from Brazil, Cameroon, Tanzania, Indonesia, and Vietnam. World Development. 2014;55:3752.Google Scholar
Hsiang, SM, Burke, M, Miguel, E. Quantifying the influence of climate on human conflict. Science. 2013;341(6151):1235367.CrossRefGoogle ScholarPubMed
Byerlee, D, Falcon, WP, Naylor, R. The tropical oil crop revolution: food, feed, fuel, and forests. Oxford: Oxford University Press; 2016.CrossRefGoogle Scholar
Mayer, AL, Khalyani, AH. Grass trumps trees with fire. Science. 2011;334(6053):188–9.CrossRefGoogle ScholarPubMed
Staver, AC, Archibald, S, Levin, SA. The global extent and determinants of savanna and forest as alternative biome states. Science. 2011;334(6053):230–2.Google Scholar
Andela, N, Morton, DC, Giglio, L, Chen, Y, van der Werf, GR, Kasibhatla, PS, et al. A human-driven decline in global burned area. Science. 2017;356(6345):1356–62.Google Scholar
Barnes, J. Cultivating the Nile: the everyday politics of water in Egypt. Durham, NC: Duke University Press; 2014.Google Scholar
Li, A, Yarime, M. Polarization and clustering in scientific debates and problem framing: network analysis of the science–policy interface for grassland management in China. Ecology and Society. 2017;22(3).Google Scholar
Moore, FC, Mankin, JS, Becker, A. Challenges in integrating the climate and social sciences for studies of climate change impacts and adaptation. In: Barnes, J, Dove, MR, editors. Climate cultures: anthropological perspectives on climate change. New Haven, CT: Yale University Press; 2015: pp. 169–95.Google Scholar
Demeritt, D. The construction of global warming and the politics of science. Annals of the Association of American Geographers. 2001;91(2):307–37.Google Scholar
Hulme, M. Reducing the future to climate: a story of climate determinism and reductionism. Osiris. 2011;26(1):245–66.Google Scholar

17.5 References

National Research Council. Rangeland health: new methods to classify, inventory, and monitor rangelands. Washington, DC: The National Academies Press; 1994.Google Scholar
Tongway, DJ, Hindley, NL. Landscape function analysis manual: procedures for monitoring and assessing landscapes with special reference to minesites and rangelands. Canberra; CSIRO Sustainable Ecosystems; 2004.Google Scholar
Harris, RB. Rangeland degradation on the Qinghai-Tibetan plateau: a review of the evidence of its magnitude and causes. Journal of Arid Environments. 2010;74(1):112.CrossRefGoogle Scholar
Eldridge, DJ, Koen, TB. Detecting environmental change in eastern Australia: rangeland health in the semi-arid woodlands. Science of the Total Environment. 2003;310(1–3):211–9.Google Scholar
Damdinsuren, B, Herrick, JE, Pyke, DA, Bestelmeyer, BT, Havstad, KM. Is rangeland health relevant to Mongolia? Rangelands. 2008;30(4):25–9.Google Scholar
Bestelmeyer, BT, Duniway, MC, James, DK, Burkett, LM, Havstad, KM. A test of critical thresholds and their indicators in a desertification-prone ecosystem: more resilience than we thought. Ecology Letters. 2013;16(3):339–45.Google Scholar
Pellant, M, Shaver, P, Pyke, D, Herrick, J. Interpreting indicators of rangeland health. Version 4.0. Interagency Technical Reference 1734–6. Denver, CO: Bureau of Land Management; 2005.Google Scholar
O’Brien, RA, Johnson, CM, Wilson, AM, Elsbernd, Cv. Indicators of rangeland health and functionality in the Intermountain West. General Technical Report – Rocky Mountain Research Station, USDA Forest Service. 2003 (RMRS-GTR-104).Google Scholar
Herrick, JE, Van Zee, JW, Havstad, KM, Burkett, LM, Whitford, WG. Monitoring manual for grassland, shrubland and savanna ecosystems. Las Cruces, NM: USDA-ARS Jornada Experimental Range: University of Arizona Press; 2005.Google Scholar
Herrick, JE, Lessard, VC, Spaeth, KE, Shaver, PL, Dayton, RS, Pyke, DA, et al. National ecosystem assessments supported by scientific and local knowledge. Frontiers in Ecology and the Environment. 2010;8(8):403–8.Google Scholar
Toevs, GR, Karl, JW, Taylor, JJ, Spurrier, CS, Karl, MS, Bobo, MR, et al. Consistent indicators and methods and a scalable sample design to meet assessment, inventory, and monitoring information needs across scales. Rangelands. 2011;33(4):1420.Google Scholar
Bestelmeyer, BT, Tugel, AJ, Peacock, GL, Robinett, DG, Sbaver, PL, Brown, JR, et al. State-and-transition models for heterogeneous landscapes: a strategy for development and application. Rangeland Ecology & Management. 2009;62(1):115.Google Scholar
Bestelmeyer, BT, Ellison, AM, Fraser, WR, Gorman, KB, Holbrook, SJ, Laney, CM, et al. Analysis of abrupt transitions in ecological systems. Ecosphere. 2011;2(12):art129.Google Scholar
Duniway, MC, Nauman, TW, Johanson, JK, Green, S, Miller, ME, Williamson, JC, et al. Generalizing ecological site concepts of the Colorado plateau for landscape-level applications. Rangelands. 2016;38(6):342–9.Google Scholar
Miller, ME, Belote, RT, Bowker, MA, Garman, SL. Alternative states of a semiarid grassland ecosystem: implications for ecosystem services. Ecosphere. 2011;2(5):118.Google Scholar
Polley, HW, Bailey, DW, Nowak, RS, Stafford-Smith, M. Ecological consequences of climate change on rangelands. In: Briske, DD, editor. Rangeland systems: processes, management and challenges. Cham: Springer International Publishing; 2017: pp. 229–60.Google Scholar
Briske, DD, Joyce, LA, Polley, HW, Brown, JR, Wolter, K, Morgan, JA, et al. Climate-change adaptation on rangelands: linking regional exposure with diverse adaptive capacity. Frontiers in Ecology and the Environment. 2015;13(5):249–56.Google Scholar
Collins, M, Knutti, R, Arblaster, J, Dufresne, J-L, Fichefet, T, Friedlingstein, P, et al. Chapter 12 – Long-term climate change: projections, commitments and irreversibility. In: IPCC, editor. Climate change 2013: the physical science basis IPCC Working Group I Contribution to AR5. Cambridge: Cambridge University Press; 2013.Google Scholar
Schlaepfer, DR, Bradford, JB, Lauenroth, WK, Munson, SM, Tietjen, B, Hall, SA, et al. Climate change reduces extent of temperate drylands and intensifies drought in deep soils. Nature Communications. 2017;8:14196.Google Scholar
McCollum, DW, Tanaka, JA, Morgan, JA, Mitchell, JE, Fox, WE, Maczko, KA, et al. Climate change effects on rangelands and rangeland management: affirming the need for monitoring. Ecosystem Health and Sustainability. 2017;3(3):e01264-n/a.CrossRefGoogle Scholar
Diffenbaugh, NS, Singh, D, Mankin, JS, Horton, DE, Swain, DL, Touma, D, et al. Quantifying the influence of global warming on unprecedented extreme climate events. Proceedings of the National Academy of Sciences. 2017;114(19):4881–6.Google Scholar
Polade, SD, Pierce, DW, Cayan, DR, Gershunov, A, Dettinger, MD. The key role of dry days in changing regional climate and precipitation regimes. Scientific Reports. 2014;4:4364.Google Scholar
Pohl, B, Macron, C, Monerie, P-A. Fewer rainy days and more extreme rainfall by the end of the century in Southern Africa. Scientific Reports. 2017;7:46466.CrossRefGoogle ScholarPubMed
Sala, OE, Lauenroth, WK, Gollucio, RA, Smith, TM, Shugart, HH, Woodward, FI. Plant functional types in temperate semiarid regions. Cambridge.: Cambridge University Press; 1997: pp. 217–33.Google Scholar
Lauenroth, WK, Schlaepfer, DR, Bradford, JB. Ecohydrology of dry regions: storage versus pulse soil water dynamics. Ecosystems. 2014;17(8):1469–79.CrossRefGoogle Scholar
Andrews, CA, Bradford, JB, Norris, J, Thomas, L, Swan, M, Palumbo, J, et al. Describing past and future soil moisture in shallow loamy pinyon–juniper woodlands communities at Mesa Verde National Park. Project brief. Fort Collins, CO: National Park Service; 2018.Google Scholar
Andrews, CA, Bradford, JB, Norris, J, Thomas, L, Swan, M, Palumbo, J, et al. Describing past and future soil moisture in sandy loam grassland communities at Petrifted Forest National Park. Project brief. Fort Collins, CO: National Park Service; 2018.Google Scholar
Smith, MD, Knapp, AK, Collins, SL. A framework for assessing ecosystem dynamics in response to chronic resource alterations induced by global change. Ecology. 2009;90(12):3279–89.Google Scholar
McDowell, NG. Mechanisms linking drought, hydraulics, carbon metabolism, and vegetation mortality. Plant Physiology. 2011;155(3):1051–9.Google Scholar
McDowell, N, Pockman, WT, Allen, CD, Breshears, DD, Cobb, N, Kolb, T, et al. Mechanisms of plant survival and mortality during drought: why do some plants survive while others succumb to drought? New Phytologist. 2008;178(4):719–39.Google Scholar
Bradford, JB, Schlaepfer, DR, Lauenroth, WK, Yackulic, CB, Duniway, M, Hall, S, et al. Future soil moisture and temperature extremes imply expanding suitability for rainfed agriculture in temperate drylands. Scientific Reports. 2017;7(1):12923.Google Scholar
Mueller, KE, Blumenthal, DM, Pendall, E, Carrillo, Y, Dijkstra, FA, Williams, DG, et al. Impacts of warming and elevated CO2 on a semi-arid grassland are non-additive, shift with precipitation, and reverse over time. Ecology Letters. 2016;19(8):956–66.Google Scholar
Becklin, KM, Anderson, JT, Gerhart, LM, Wadgymar, SM, Wessinger, CA, Ward, JK. Examining plant physiological responses to climate change through an evolutionary lens. Plant Physiology. 2016;172(2):635.Google Scholar
Munson, SM, Long, AL. Climate drives shifts in grass reproductive phenology across the western USA. New Phytologist. 2017;213(4):1945–55.Google Scholar
Lauenroth, WK, Sala, OE. Long-term forage production of North American shortgrass steppe. Ecological Applications. 1992;2:397403.Google Scholar
Bunting, EL, Munson, SM, Villarreal, ML. Climate legacy and lag effects on dryland plant communities in the southwestern U.S. Ecological Indicators. 2017;74:216–29.Google Scholar
Sala, OE, Gherardi, LA, Reichmann, L, Jobbágy, E, Peters, D. Legacies of precipitation fluctuations on primary production: theory and data synthesis. Philosophical Transactions of the Royal Society B: Biological Sciences. 2012;367(1606):3135–44.Google Scholar
Knapp, AK, Smith, MD. Variation among biomes in temporal dynamics of aboveground primary production. Science. 2001;291:481–4.Google Scholar
Huxman, TE, Smith, MD, Fay, PA, Knapp, AK, Shaw, MR, Loik, ME, et al. Convergence across biomes to a common rain-use efficiency. Nature. 2004;429(6992):651–4.Google Scholar
Heisler-White, JL, Blair, JM, Kelly, EF, Harmoney, K, Knapp, AK. Contingent productivity responses to more extreme rainfall regimes across a grassland biome. Global Change Biology. 2009;15(12):2894–904.Google Scholar
Campbell, BD, Stafford Smith, DM. A synthesis of recent global change research on pasture and rangeland production: reduced uncertainties and their management implications. Agriculture, Ecosystems & Environment. 2000;82(1):3955.Google Scholar
Poorter, H, Navas, M-L. Plant growth and competition at elevated CO2: on winners, losers and functional groups. New Phytologist. 2003;157(2):175–98.Google Scholar
Browning, DM, Duniway, MC, Laliberte, AS, Rango, A. Hierarchical analysis of vegetation dynamics over 71 years: soil–rainfall interactions in a Chihuahuan Desert ecosystem. Ecological Applications. 2012;22(3):909–26.Google Scholar
Sankaran, M, Hanan, NP, Scholes, RJ, Ratnam, J, Augustine, DJ, Cade, BS, et al. Determinants of woody cover in African savannas. Nature. 2005;438(7069):846–9.Google Scholar
D’Antonio, CM, Vitousek, PM. Biological invasions by exotic grasses, the grass/fire cycle, and global change. Annual Review of Ecology and Systematics. 1992;23:6387.Google Scholar
Sandel, B, Dangremond, EM. Climate change and the invasion of California by grasses. Global Change Biology. 2012;18(1):277–89.Google Scholar
Schlesinger, WH. Biogeochemistry: an analysis of global change. San Diego, CA: Academic Press; 1997.Google Scholar
Munson, SM, Belnap, J, Okin, GS. Responses of wind erosion to climate-induced vegetation changes on the Colorado Plateau. Proceedings of the National Academy of Sciences. 2011;108(10):3854–9.Google Scholar
Nearing, MA, Pruski, FF, O’Neal, MR. Expected climate change impacts on soil erosion rates: a review. Journal of Soil and Water Conservation. 2004;59(1):4350.Google Scholar
Hoover, DL, Duniway, MC, Belnap, J. Testing the apparent resistance of three dominant plants to chronic drought on the Colorado Plateau. Journal of Ecology. 2017;105(1):152–62.Google Scholar
Gremer, JR, Bradford, JB, Munson, SM, Duniway, MC. Desert grassland responses to climate and soil moisture suggest divergent vulnerabilities across the southwestern United States. Global Change Biology. 2015;21(11):4049–62.Google Scholar
Nusser, SM, Goebel, JJ. The National Resources Inventory: a long-term multi-resource monitoring programme. Environmental and Ecological Statistics. 1997;4(3):181204.Google Scholar
Munson, SM, Duniway, MC, Johanson, JK. Rangeland monitoring reveals long-term plant responses to precipitation and grazing at the landscape scale. Rangeland Ecology & Management. 2016;69(1):7683.Google Scholar
Miller, ME. Broad-scale assessment of rangeland health, Grand Staircase–Escalante National Monument, USA. Rangeland Ecology & Management. 2008;61(3):249–62.Google Scholar
Bradford, JB, Betancourt, JL, Butterfield, BJ, Munson, SM, Wood, TE. Anticipatory natural resource science and management for a dynamic future. Frontiers in Ecology and the Environment. 2018;16(5):295303.Google Scholar
Toevs, GR, Taylor, JJ, Spurrier, CS, MacKinnon, WC, Bobo, MR. Assessment, Inventory, and Monitoring Strategy for Integrated Renewable Resources Management. US Department of the Interior, Bureau of Land Management, National Operations Center; 2011.Google Scholar
Hunt, ER Jr, Everitt, JH, Ritchie, JC, Moran, MS, Booth, DT, Anderson, GL, et al. Applications and research using remote sensing for rangeland management. Photogrammetric Engineering & Remote Sensing. 2003;69(6):675–93.Google Scholar
Joyce, LA, Briske, DD, Brown, JR, Polley, HW, McCarl, BA, Bailey, DW. Climate change and North American rangelands: assessment of mitigation and adaptation strategies. Rangeland Ecology & Management. 2013;66(5):512–28.Google Scholar
West, NE. History of rangeland monitoring in the U.S.A. Arid Land Research and Management. 2003;17(4):495545.Google Scholar

18.6 References

Ellis, EC, Ramankutty, N. Putting people in the map: anthropogenic biomes of the world. Frontiers in Ecology and the Environment. 2008;6(8):439–47.Google Scholar
Gang, CC, Zhou, W, Chen, YZ, Wang, ZQ, Sun, ZG, Li, JL, et al. Quantitative assessment of the contributions of climate change and human activities on global grassland degradation. Environmental Earth Sciences. 2014;72(11):4273–82.Google Scholar
Baer, SG, Heneghan, L, Eviner, V. Applying soil ecological knowledge to restore ecosystem services. In: Wall, DH, Bardgett, RD, Behan-Pelletier, V, Herrick, JE, Jones, HP, Ritz, K, et al., editors. Soil ecology and ecosystem services. Oxford: Oxford University Press; 2012. pp. 377–93.Google Scholar
Hobbs, RJ, Cramer, VA. Restoration ecology: interventionist approaches for restoring and maintaining ecosystem function in the face of rapid environmental change. Annual Review of Environment and Resources. 2008;33:3961.Google Scholar
Society for Ecological Restoration International Science Policy Working Group. The SER international primer on ecological restoration. Tuscon, AZ: Society for Ecological Restoration International; 2004.Google Scholar
Harris, JA, Hobbs, RJ, Higgs, E, Aronson, J. Ecological restoration and global climate change. Restoration Ecology. 2006;14(2):170–6.Google Scholar
Millennium Ecosystem Assessment. Ecosystems and human well-being: biodiversity synthesis. Washington, DC: World Resources Institute; 2005.Google Scholar
Smith, NG, Schuster, MJ, Dukes, JS. Rainfall variability and nitrogen addition synergistically reduce plant diversity in a restored tallgrass prairie. Journal of Applied Ecology. 2016;53(2):579–86.Google Scholar
White, PS, Walker, JL. Approximating nature’s variation: selecting and using reference information in restoration ecology. Restoration Ecology. 1997;5(4):338–49.Google Scholar
Papanastasis, VP. Restoration of degraded grazing lands through grazing management: can it work? Restoration Ecology. 2009;17(4):441–5.Google Scholar
Ehrenfeld, JG, Ravit, B, Elgersma, K. Feedback in the plant–soil system. Annual Review of Environment and Resources. 2005;30:75115.Google Scholar
Reinhart, KO, Callaway, RM. Soil biota and invasive plants. New Phytologist. 2006;170(3):445–57.Google Scholar
White, PS, Jentsch, A. Disturbance, succession, and community assembly in terrestrial plant communities. In: Temperton, VM, Hobbs, RJ, Nuttle, T, Hale, S, editors. Assembly rules and restoration ecology. Washington, DC: Island Press; 2004. pp. 342–66.Google Scholar
McIntyre, S, Nicholls, AO, Manning, AD. Trajectories of floristic change in grassland: landscape, land use legacy and seasonal conditions overshadow restoration actions. Applied Vegetation Science. 2017;20(4):582–93.Google Scholar
McGovern, ST, Evans, CD, Dennis, P, Walmsley, CA, Turner, A, McDonald, MA. Increased inorganic nitrogen leaching from a mountain grassland ecosystem following grazing removal: a hangover of past intensive land-use? Biogeochemistry. 2014;119(1–3):125–38.Google Scholar
Peters, DPC, Havstad, KM, Archer, SR, Sala, OE. Beyond desertification: new paradigms for dryland landscapes. Frontiers in Ecology and the Environment. 2015;13(1):412.Google Scholar
Nsikani, MM, van Wilgen, BW, Gaertner, M. Barriers to ecosystem restoration presented by soil legacy effects of invasive alien N-2-fixing woody species: implications for ecological restoration. Restoration Ecology. 2018;26(2):235–44.Google Scholar
Hawkes, CV, Wren, IF, Herman, DJ, Firestone, MK. Plant invasion alters nitrogen cycling by modifying the soil nitrifying community. Ecology Letters. 2005;8(9):976–85.Google Scholar
Mitsch, WJ, Jorgensen, SE. Ecological engineering and ecosystem restoration. Hoboken, NJ: Wiley; 2004.Google Scholar
Brudvig, LA, Barak, RS, Bauer, JT, Caughlin, TT, Laughlin, DC, Larios, L, et al. Interpreting variation to advance predictive restoration science. Journal of Applied Ecology. 2017;54(4):1018–27.Google Scholar
Hobbs, RJ, Harris, JA. Restoration ecology: repairing the Earth’s ecosystems in the new millennium. Restoration Ecology. 2001;9(2):239–46.Google Scholar
Peters, DPC, Yao, J, Sala, OE, Anderson, JP. Directional climate change and potential reversal of desertification in arid and semiarid ecosystems. Global Change Biology. 2012;18(1):151–63.Google Scholar
Hipp, AL, Larkin, DJ, Barak, RS, Bowles, ML, Cadotte, MW, Jacobi, SK, et al. Phylogeny in the service of ecological restoration. American Journal of Botany. 2015;102(5):647–8.Google Scholar
Smith, AB, Alsdurf, J, Knapp, M, Baer, SG, Johnson, LC. Phenotypic distribution models corroborate species distribution models: a shift in the role and prevalence of a dominant prairie grass in response to climate change. Global Change Biology. 2017;23(10):4365–75.Google Scholar
Beaumont, LJ, Hughes, L. Potential changes in the distributions of latitudinally restricted Australian butterfly species in response to climate change. Global Change Biology. 2002;8(10):954–71.Google Scholar
Notaro, M, Mauss, A, Williams, JW. Projected vegetation changes for the American Southwest: combined dynamic modeling and bioclimatic-envelope approach. Ecological Applications. 2012;22(4):1365–88.Google Scholar
Araujo, MB, Pearson, RG, Thuiller, W, Erhard, M. Validation of species–climate impact models under climate change. Global Change Biology. 2005;11(9):1504–13.Google Scholar
Heikkinen, RK, Luoto, M, Araujo, MB, Virkkala, R, Thuiller, W, Sykes, MT. Methods and uncertainties in bioclimatic envelope modelling under climate change. Progress in Physical Geography. 2006;30(6):751–77.Google Scholar
Booth, TH. Identifying particular areas for potential seed collections for restoration plantings under climate change. Ecological Management & Restoration. 2016;17(3):228–34.Google Scholar
Gray, MM, St Amand, P, Bello, NM, Galliart, MB, Knapp, M, Garrett, KA, et al. Ecotypes of an ecologically dominant prairie grass (Andropogon gerardii) exhibit genetic divergence across the US Midwest grasslands’ environmental gradient. Molecular Ecology. 2014;23(24):6011–28.Google Scholar
Johnson, LC, Olsen, JT, Tetreault, H, DeLaCruz, A, Bryant, J, Morgan, TJ, et al. Intraspecific variation of a dominant grass and local adaptation in reciprocal garden communities along a US Great Plains’ precipitation gradient: implications for grassland restoration with climate change. Evolutionary Applications. 2015;8(7):705–23.Google Scholar
Nixon, AE, Fisher, RJ, Stralberg, D, Bayne, EM, Farr, DR. Projected responses of North American grassland songbirds to climate change and habitat availability at their northern range limits in Alberta, Canada. Avian Conservation and Ecology. 2016;11(2).Google Scholar
Jennings, MD, Harris, GM. Climate change and ecosystem composition across large landscapes. Landscape Ecology. 2017;32(1):195207.Google Scholar
Broadhurst, LM, Lowe, A, Coates, DJ, Cunningham, SA, McDonald, M, Vesk, PA, et al. Seed supply for broadscale restoration: maximizing evolutionary potential. Evolutionary Applications. 2008;1(4):587–97.Google Scholar
Hufford, KM, Mazer, SJ. Plant ecotypes: genetic differentiation in the age of ecological restoration. Trends in Ecology & Evolution. 2003;18(3):147–55.Google Scholar
Jones, TA. Ecologically appropriate plant materials for restoration applications. Bioscience. 2013;63(3):211–9.Google Scholar
Maschinski, J, Wright, SJ, Koptur, S, Pinto-Torres, EC. When is local the best paradigm? Breeding history influences conservation reintroduction survival and population trajectories in times of extreme climate events. Biological Conservation. 2013;159:277–84.Google Scholar
McKay, JK, Christian, CE, Harrison, S, Rice, KJ. ‘How local is local?’ A review of practical and conceptual issues in the genetics of restoration. Restoration Ecology. 2005;13(3):432–40.Google Scholar
Vander Mijnsbrugge, K, Bischoff, A, Smith, B. A question of origin: where and how to collect seed for ecological restoration. Basic and Applied Ecology. 2010;11(4):300–11.Google Scholar
Daehler, CC, Strong, DR. Status, prediction and prevention of introduced cordgrass Spartina spp invasions in Pacific estuaries, USA. Biological Conservation. 1996;78(1–2):51–8.Google Scholar
Linhart, YB, Grant, MC. Evolutionary significance of local genetic differentiation in plants. Annual Review of Ecology and Systematics. 1996;27:237–77.Google Scholar
Montalvo, AM, Williams, SL, Rice, KJ, Buchmann, SL, Cory, C, Handel, SN, et al. Restoration biology: a population biology perspective. Restoration Ecology. 1997;5(4):277–90.Google Scholar
Bucharova, A, Frenzel, M, Mody, K, Parepa, M, Durka, W, Bossdorf, O. Plant ecotype affects interacting organisms across multiple trophic levels. Basic and Applied Ecology. 2016;17(8):688–95.Google Scholar
Bucharova, A, Michalski, S, Hermann, JM, Heveling, K, Durka, W, Holzel, N, et al. Genetic differentiation and regional adaptation among seed origins used for grassland restoration: lessons from a multispecies transplant experiment. Journal of Applied Ecology. 2017;54(1):127–36.Google Scholar
Bischoff, A, Steinger, T, Muller-Scharer, H. The importance of plant provenance and genotypic diversity of seed material used for ecological restoration. Restoration Ecology. 2010;18(3):338–48.Google Scholar
Mendola, ML, Baer, SG, Johnson, LC, Maricle, BR. The role of ecotypic variation and the environment on biomass and nitrogen in a dominant prairie grass. Ecology. 2015;96(9):2433–45.Google Scholar
Edmands, S. Between a rock and a hard place: evaluating the relative risks of inbreeding and outbreeding for conservation and management. Molecular Ecology. 2007;16(3):463–75.Google Scholar
Galloway, LF, Fenster, CB. Population differentiation in an annual legume: local adaptation. Evolution. 2000;54(4):1173–81.Google Scholar
Leiss, KA, Muller-Scharer, H. Performance of reciprocally sown populations of Senecio vulgaris from ruderal and agricultural habitats. Oecologia. 2001;128(2):210–6.Google Scholar
Falk, DA, Richards, CM, Montalvo, AM, Knapp, EE. Population and ecological genetics in restoration ecology. In: Falk, DA, Palmer, MA, Zedler, JB, editors. Foundations of restoration ecology. Washington, DC: Island Press; 2006. pp. 1441.Google Scholar
Michalski, SG, Durka, W, Jentsch, A, Kreyling, J, Pompe, S, Schweiger, O, et al. Evidence for genetic differentiation and divergent selection in an autotetraploid forage grass (Arrhenatherum elatius). Theoretical and Applied Genetics. 2010;120(6):1151–62.Google Scholar
Gibson, DJ, Allstadt, AJ, Baer, SG, Geisler, M. Effects of foundation species genotypic diversity on subordinate species richness in an assembling community. Oikos. 2012;121(4):496507.Google Scholar
Gustafson, DJ, Gibson, DJ, Nickrent, DL. Conservation genetics of two co-dominant grass species in an endangered grassland ecosystem. Journal of Applied Ecology. 2004;41(2):389–97.Google Scholar
Dolan, RW, Marr, DL, Schnabel, A. Capturing genetic variation during ecological restorations: an example from Kankakee Sands in Indiana. Restoration Ecology. 2008;16(3):386–96.Google Scholar
Young, AG, Murray, BG. Genetic bottlenecks and dysgenic gene flow into re-established populations of the grassland daisy, Rutidosis leptorrhynchoides. Australian Journal of Botany. 2000;48(3):409–16.Google Scholar
Aitken, SN, Whitlock, MC. Assisted gene flow to facilitate local adaptation to climate change. Annual Review of Ecology, Evolution, and Systematics. 2013;44:367–88.Google Scholar
De Frenne, P, Coomes, DA, De Schrijver, A, Staelens, J, Alexander, JM, Bernhardt-Roemermann, M, et al. Plant movements and climate warming: intraspecific variation in growth responses to nonlocal soils. New Phytologist. 2014;202(2):431–41.Google Scholar
Liancourt, P, Spence, LA, Song, DS, Lkhagva, A, Sharkhuu, A, Boldgiv, B, et al. Plant response to climate change varies with topography, interactions with neighbors, and ecotype. Ecology. 2013;94(2):444–53.Google Scholar
Nicotra, AB, Atkin, OK, Bonser, SP, Davidson, AM, Finnegan, EJ, Mathesius, U, et al. Plant phenotypic plasticity in a changing climate. Trends in Plant Science. 2010;15(12):684–92.Google Scholar
Rice, KJ, Emery, NC. Managing microevolution: restoration in the face of global change. Frontiers in Ecology and the Environment. 2003;1(9):469–78.Google Scholar
Christensen, JH, Hewitson, B, Busuioc, A, Chen, A, Gao, X, Held, I, et al. Regional climate projections. In: Solomon, S, Qin, D, Manning, M, Chen, Z, Marquis, M, Avery, KB, et al., editors. Climate change 2007: the physical science basis contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge: Cambridge University Press; 2007.Google Scholar
Polley, HW, Briske, DD, Morgan, JA, Wolter, K, Bailey, DW, Brown, JR. Climate change and North American rangelands: trends, projections, and implications. Rangeland Ecology & Management. 2013;66(5):493511.Google Scholar
Kimball, S, Lulow, ME, Balazs, KR, Huxman, TE. Predicting drought tolerance from slope aspect preference in restored plant communities. Ecology and Evolution. 2017;7(9):3123–31.Google Scholar
Richter, S, Kipfer, T, Wohlgemuth, T, Guerrero, CC, Ghazoul, J, Moser, B. Phenotypic plasticity facilitates resistance to climate change in a highly variable environment. Oecologia. 2012;169(1):269–79.Google Scholar
Nooten, SS, Hughes, L. The power of the transplant: direct assessment of climate change impacts. Climatic Change. 2017;144(2):237–55.Google Scholar
Carter, DL, Blair, JM. Seed source affects establishment and survival for three grassland species sown into reciprocal common gardens. Ecosphere. 2012;3(11):10.Google Scholar
Wilson, LR, Gibson, DJ, Baer, SG, Johnson, LC. Plant community response to regional sources of dominant grasses in grasslands restored across a longitudinal gradient. Ecosphere. 2016;7(4):art e01329.Google Scholar
Isbell, F, Craven, D, Connolly, J, Loreau, M, Schmid, B, Beierkuhnlein, C, et al. Biodiversity increases the resistance of ecosystem productivity to climate extremes. Nature. 2015;526(7574):574–7.Google Scholar
Kreyling, J, Jentsch, A, Beierkuhnlein, C. Stochastic trajectories of succession initiated by extreme climatic events. Ecology Letters. 2011;14(8):758–64.Google Scholar
Hooper, DU, Chapin, FS, Ewel, JJ, Hector, A, Inchausti, P, Lavorel, S, et al. Effects of biodiversity on ecosystem functioning: a consensus of current knowledge. Ecological Monographs. 2005;75(1):335.Google Scholar
Funk, JL, Cleland, EE, Suding, KN, Zavaleta, ES. Restoration through reassembly: plant traits and invasion resistance. Trends in Ecology & Evolution. 2008;23(12):695703.Google Scholar
Verdu, M, Gomez-Aparicio, L, Valiente-Banuet, A. Phylogenetic relatedness as a tool in restoration ecology: a meta-analysis. Proceedings of the Royal Society B: Biological Sciences. 2012;279(1734):1761–7.Google Scholar
Barber, NA, Jones, HP, Duvall, MR, Wysocki, WP, Hansen, MJ, Gibson, DJ. Phylogenetic diversity is maintained despite richness losses over time in restored tallgrass prairie plant communities. Journal of Applied Ecology. 2017;54(1):137–44.Google Scholar
Khalil, MI, Gibson, DJ, Baer, SG. Phylogenetic diversity reveals hidden patterns related to population source and species pools during restoration. Journal of Applied Ecology. 2017;54(1):91101.Google Scholar
Khalil, MI, Gibson, DJ, Baer, SG. Functional diversity is more sensitive to biotic filters than phylogenetic diversity during community assembly. Ecosphere. 2018;9:art e02164.Google Scholar
Barak, RS, Hipp, AL, Cavender-Bares, J, Pearse, WD, Hotchkiss, SC, Lynch, EA, et al. Taking the long view: integrating recorded, archeological, paleoecological, and evolutionary data into ecological restoration. International Journal of Plant Sciences. 2016;177(1):90102.Google Scholar
Forrestel, EJ, Donoghue, MJ, Edwards, EJ, Jetz, W, du Toit, JCO, Smith, MD. Different clades and traits yield similar grassland functional responses. Proceedings of the National Academy of Sciences of the USA. 2017;114(4):705–10.Google Scholar
Venail, P, Gross, K, Oakley, TH, Narwani, A, Allan, E, Flombaum, P, et al. Species richness, but not phylogenetic diversity, influences community biomass production and temporal stability in a re-examination of 16 grassland biodiversity studies. Functional Ecology. 2015;29(5):615–26.Google Scholar
Ratajczak, Z, Nippert, JB, Briggs, JM, Blair, JM. Fire dynamics distinguish grasslands, shrublands and woodlands as alternative attractors in the Central Great Plains of North America. Journal of Ecology. 2014;102(6):1374–85.Google Scholar
Ratajczak, Z, Nippert, JB, Ocheltree, TW. Abrupt transition of mesic grassland to shrubland: evidence for thresholds, alternative attractors, and regime shifts. Ecology. 2014;95(9):2633–45.Google Scholar
Gibson, DJ, Hulbert, LC. Effects of fire, topography and year-to-year climatic variation on species composition in tallgrass prairie. Vegetatio. 1987;72(3):175–85.Google Scholar
Bowles, ML, Jones, MD. Repeated burning of eastern tallgrass prairie increases richness and diversity, stabilizing late successional vegetation. Ecological Applications. 2013;23(2):464–78.Google Scholar
Collins, SL, Calabrese, LB. Effects of fire, grazing and topographic variation on vegetation structure in tallgrass prairie. Journal of Vegetation Science. 2012;23(3):563–75.Google Scholar
Kane, K, Debinski, DM, Anderson, C, Scasta, JD, Engle, DM, Miller, JR. Using regional climate projections to guide grassland community restoration in the face of climate change. Frontiers in Plant Science. 2017;8:11.Google Scholar

19.7 References

Crutzen, PJ. The ‘anthropocene’. Journal de Physique Archives. 2002;12(10):15.Google Scholar
Crutzen, PJ. Geology of mankind. Nature. 2002;415:23.Google Scholar
Crutzen, PJ, Stoermer, EF. The ‘Anthropocene’. Global Change Newsletter, International Geosphere–Biosphere Programme (IGBP). 2000;41:17–8.Google Scholar
Ruddiman, WF, Ellis, EC, Kaplan, JO, Fuller, DQ. Defining the epoch we live in. Science. 2015;348(6230):38–9.Google Scholar
Bai, X, van der Leeuw, S, O’Brien, K, Berkhout, F, Biermann, F, Brondizio, ES, et al. Plausible and desirable futures in the Anthropocene: a new research agenda. Global Environmental Change. 2016;39:351–62.Google Scholar
Rockström, J, Steffen, W, Noone, K, Persson, Å, Chapin III, FS, Lambin, E, et al. Planetary boundaries: exploring the safe operating space for humanity. Ecology and Society. 2009;14(2):art32 (www.ecologyandsociety.org/vol14/iss2/art32/).Google Scholar
Newman, JA, Varner, G, Linquist, S. Defending biodiversity: environmental science and ethics. Cambridge: Cambridge University Press; 2017.Google Scholar
Ellis, EC, Ramankutty, N. Putting people in the map: anthropogenic biomes of the world. Frontiers in Ecology and the Environment. 2008;6:439–47.Google Scholar
Olson, DM, Dinerstein, E, Wikramanayake, ED, Burgess, ND, Powell, GVN, Underwood, EC, et al. Terrestrial ecoregions of the world: a new map of life on Earth. BioScience. 2001;51(11):933–8.Google Scholar
Sayre, NF, Davis, DK, Bestelmeyer, B, Williamson, JC. Rangelands: where anthromes meet their limits. Land. 2017;6(2):31.Google Scholar
Ellis, EC. Ecology in an anthropogenic biosphere. Ecological Monographs. 2015;85(3):287331.Google Scholar
Fuentes, A, Baynes-Rock, M. Anthropogenic landscapes, human action and the process of co-construction with other species: making anthromes in the Anthropocene. Land. 2017;6(1):15.Google Scholar
Jax, K. Ecological units: definitions and application. The Quarterly Review of Biology. 2006;81:237–58.Google Scholar
Garcia, RK, Newman, JA. Is it possible to care for ecosystems? Policy paralysis and ecosystem management. Ethics, Policy & Environment. 2016;19(2):170–82.Google Scholar
Hobbs, RJ, Higgs, ES, Hall, CM, editors. Novel ecosystems: intervening in the new ecological world order. Chichester: John Wiley & Sons; 2013.Google Scholar
Parmesan, C, Hanley, ME. Plants and climate change: complexities and surprises. Annals of Botany. 2015;116(6):849–64.Google Scholar
Jouzel, J, Masson-Delmotte, V, Cattani, O, Dreyfus, G, Falourd, S, Hoffmann, G, et al. Orbital and millennial Antarctic climate variability over the past 800,000 years. Science. 2007;317(5839):793–6.Google Scholar
Lüthi, D, Le Floch, M, Bereiter, B, Blunier, T, Barnola, J-M, Siegenthaler, U, et al. High-resolution carbon dioxide concentration record 650,000–800,000 years before present. Nature. 2008;453(7193):379.Google Scholar
Seastedt, TR, Hobbs, RJ, Suding, KN. Management of novel ecosystems: are novel approaches required? Frontiers in Ecology and the Environment. 2008;6(10):547–53.Google Scholar
Six, LJ, Bakker, JD, Bilby, RE. Vegetation dynamics in a novel ecosystem: agroforestry effects on grassland vegetation in Uruguay. Ecosphere. 2014;5(6):115.Google Scholar
Chapin III, FS, Starfield, AM. Time lags and novel ecosystems in response to transient climatic change in arctic Alaska. Climatic Change. 1997;35(4):449–61.Google Scholar
Wilcox, BP, Sorice, MG, Young, MH. Dryland ecohydrology in the Anthropocene: taking stock of human–ecological interactions. Geography Compass. 2011;5:112–27.Google Scholar
Wilson, EO. The creation: an appeal to save life on Earth. New York, NY: Norton; 2006.Google Scholar
McGill, BJ, Dornelsa, M, Gotelli, NJ, Magurran, AE. Fifteen forms of biodiversity trend in the Anthropocene. Trends in Ecology & Evolution. 2014;30(2):104–13.Google Scholar
Vellend, M, Baeten, L, Becker-Scarpitta, A, Boucher-Lalonde, V, McCune, JL, Messier, J, et al. Plant biodiversity change across scales during the Anthropocene. Annual Review of Plant Biology. 2017;68(1):563–86.Google Scholar
Ellis, EC, Antill, EC, Kreft, H. All is not loss: plant biodiversity in the Anthropocene. PLoS ONE. 2012;7(1):e30535.Google Scholar
Sutherland, WJ, Woodroof, HJ. The need for environmental horizon scanning. Trends in Ecology & Evolution. 2009;24(10):523–7.Google Scholar
Sutherland, WJ, Aveling, R, Brooks, TM, Clout, M, Dicks, LV, Fellman, L, et al. A horizon scan of global conservation issues for 2014. Trends in Ecology & Evolution. 2014;29(1):1522.Google Scholar
Rocha, JC, Peterson, GD, Biggs, R. Regime shifts in the Anthropocene: drivers, risks, and resilience. PLoS ONE. 2015;10(8): e0134639.Google Scholar
Benito-Garzón, M, Leadley, PW, Fernández-Manjarrés, JF. Assessing global biome exposure to climate change through the HoloceneAnthropocene transition. Global Ecology and Biogeography. 2014;23(2):235–44.Google Scholar
Beals, SC, Preston, DL, Wessman, CA, Seastedt, TR. Resilience of a novel ecosystem after the loss of a keystone species: plague epizootics and urban prairie dog management. Ecosphere. 2015;6(9):art157.Google Scholar
Wolkovich, EM, Cook, BI, McLauchlan, KK, Davies, TJ. Temporal ecology in the Anthropocene. Ecology Letters. 2014;17(11):1365–79.Google Scholar
Verburg, PH, Dearing, JA, Dyke, JG, Leeuw, Svd, Seitzinger, S, Steffen, W, et al. Methods and approaches to modelling the Anthropocene. Global Environmental Change. 2016;39:328–40.Google Scholar
Biermann, F, Bai, X, Bondre, N, Broadgate, W, Arthur Chen, C-T, Dube, OP, et al. Down to Earth: contextualizing the Anthropocene. Global Environmental Change. 2016;39:341–50.Google Scholar
Poesen, J. Soil erosion in the Anthropocene: research needs. Earth Surface Processes and Landforms. 2018;43(1):6484.Google Scholar
Svenning, J-C, Pedersen, PBM, Donlan, CJ, Ejrnæs, R, Faurby, S, Galetti, M, et al. Science for a wilder Anthropocene: synthesis and future directions for trophic rewilding research. Proceedings of the National Academy of Sciences of the USA. 2016;113(4):898906.Google Scholar
Zimov, SA, Zimov, NS, Tikhonov, AN, Chapin III, FS. Mammoth steppe: a high productivity phenomenon. Quarternary Science Reviews. 2012;57:2645.Google Scholar
Fernández-Giménez, ME, Venable, NH, Angerer, J, Fassnacht, SR, Reid, RS, Khishigbayar, J. Exploring linked ecological and cultural tipping points in Mongolia. Anthropocene. 2017;17:4669.Google Scholar
Clement, S, Standish, RJ. Novel ecosystems: governance and conservation in the age of the Anthropocene. Journal of Environmental Management. 2018;208:3645.Google Scholar
Seddon, N, Mace, GM, Naeem, S, Tobias, JA, Pigot, AL, Cavanagh, R, et al. Biodiversity in the Anthropocene: prospects and policy. Proceedings of the Royal Society B: Biological Sciences. 2016;283(1844).Google Scholar
Reich, PB, Hobbie, SE, Lee, TD, Pastore, MA. Unexpected reversal of C3 versus C4 grass response to elevated CO2 during a 20-year field experiment. Science. 2018;360(6386):317–20.Google Scholar
Flato, G, Marotzke, J, Abiodun, B, Braconnot, P, Chou, SC, Collins, W, et al. Evaluation of climate models. In: Stocker, TF, Qin, D, Plattner, G-K, Tignor, M, Allen, SK, Boschung, J, et al., editors. Climate change 2013: the physical science basis contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge: Cambridge University Press; 2013.Google Scholar
Hawkins, E, Sutton, R. The potential to narrow uncertainty in regional climate predictions. Bulletin of the American Meteorological Society. 2009;90(8):1095–108.Google Scholar
Newman, JA. Climate change and the fate of cereal aphids in Southern Britain. Global Change Biology. 2005;11(6):940–4.Google Scholar
Ziter, C, Robinson, EA, Newman, JA. Climate change and voltinism in Californian insect pest species: sensitivity to location, scenario and climate model choice. Global Change Biology. 2012;18(9):2771–80.Google Scholar
Langille, AB, Arteca, EM, Newman, JA. The impacts of climate change on the abundance and distribution of the Spotted Wing Drosophila (Drosophila suzukii) in the United States and Canada. PeerJ. 2017;5:e3192.Google Scholar
Minigan, JN, Hager, HA, Peregrine, AS, Newman, JA. Current and potential future distribution of the American dog tick (Dermacentor variabilis, Say) in North America. Ticks and Tick-Borne Diseases. 2018;9(2):354–62.Google Scholar
Clarke, L, Edmonds, J, Krey, V, Richels, R, Rose, S, Tavoni, M. International climate policy architectures: overview of the EMF 22 International Scenarios. Energy Economics. 2009;31:S64S81.Google Scholar
Royer, DL. CO2-forced climate thresholds during the Phanerozoic. Geochimica et Cosmochimica Acta. 2006;70(23):5665–75.Google Scholar

References

Gore, A. An inconvenient truth: The planetary emergency of global warming and what we can do about it. New York, NY: Rodale; 2006.Google Scholar
Harris, RMB, Carter, O, Gilfedder, L, Porfirio, LL, Lee, G, Bindoff, NL. Noah’s Ark conservation will not preserve threatened ecological communities under climate change. PLoS ONE. 2015;10(4):e0124014.Google Scholar
Mariotte, P, Kardol, P, editors. Grassland biodiversity and conservation in a changing world. New York, NY: Nova Publishers; 2014.Google Scholar
Acar, Z, López-Francos, A, Porqueddu, C, editors. New approaches for grassland research in a context of climate and socio-economic changes. Zaragoza: CIHEAM; 2012.Google Scholar
Parmesan, C, Burrows, MT, Duarte, CM, Poloczanska, ES, Richardson, AJ, Schoeman, DS, et al. Beyond climate change attribution in conservation and ecological research. Ecology Letters. 2013;16:5871.Google Scholar
Reich, PB, Hobbie, SE, Lee, TD, Pastore, MA. Unexpected reversal of C3 versus C4 grass response to elevated CO2 during a 20-year field experiment. Science. 2018;360(6386):317–20.Google Scholar
Parmesan, C, Hanley, ME. Plants and climate change: complexities and surprises. Annals of Botany. 2015;116(6):849–64.Google Scholar
Bardgett, RD, Gibson, DJ. Plant ecological solutions to global food security. Journal of Ecology. 2017;105(4):14.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
×