Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-qsmjn Total loading time: 0 Render date: 2024-04-24T00:36:28.876Z Has data issue: false hasContentIssue false

Part II - Species traits, functional groups, and evolutionary change

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

9.7 References

Bradley, BA, Blumenthal, DM, Wilcove, DS, Ziska, LH. Predicting plant invasions in an era of global change. Trends in Ecology & Evolution. 2010;25(5):310–8.CrossRefGoogle Scholar
Bowler, DE, Hof, C, Haase, P, Kröncke, I, Schweiger, O, Adrian, R, et al. Cross-realm assessment of climate change impacts on species’ abundance trends. Nature Ecology & Evolution. 2017;1:0067.Google Scholar
Blumenthal, DM, Resco, V, Morgan, JA, Williams, DG, Lecain, DR, Hardy, EM, et al. Invasive forb benefits from water savings by native plants and carbon fertilization under elevated CO2 and warming. New Phytologist. 2013;200(4):1156–65.CrossRefGoogle ScholarPubMed
Hulme, PE. Climate change and biological invasions: evidence, expectations, and response options. Biological Reviews. 2017;92(3):1297–313.Google Scholar
Liu, Y, Oduor, AMO, Zhang, Z, Manea, A, Tooth, IM, Leishman, MR, et al. Do invasive alien plants benefit more from global environmental change than native plants? Global Change Biology. 2017;23(8):3363–70.Google Scholar
Bellard, C, Thuiller, W, Leroy, B, Genovesi, P, Bakkenes, M, Courchamp, F. Will climate change promote future invasions? Global Change Biology. 2013;19:3740–8.Google Scholar
Bradley, BA. Regional analysis of the impacts of climate change on cheatgrass invasion shows potential risk and opportunity. Global Change Biology. 2009;15(1):196208.Google Scholar
Catford, JA, Baumgartner, JB, Vesk, PA, White, M, Buckley, YM, McCarthy, MA. Disentangling the four demographic dimensions of species invasiveness. Journal of Ecology. 2016;104(6):1745–58.Google Scholar
Seabloom, EW, Borer, ET, Buckley, Y, Cleland, EE, Davies, K, Firn, J, et al. Predicting invasion in grassland ecosystems: is exotic dominance the real embarrassment of richness? Global Change Biology. 2013;19(12):3677–87.CrossRefGoogle ScholarPubMed
Catford, JA, Vesk, PA, Richardson, DM, Pyšek, P. Quantifying levels of biological invasion: towards the objective classification of invaded and invasible ecosystems. Global Change Biology. 2012;18(1):4462.Google Scholar
Fridley, JD, Sax, DF. The imbalance of nature: revisiting a Darwinian framework for invasion biology. Global Ecology and Biogeography. 2014;23(11):1157–66.Google Scholar
Invasive Species Specialist Group (ISSG). The Global Invasive Species Database (IUCN) 2015 [available from: www.iucngisd.org/gisd/].Google Scholar
Seabloom, EW, Borer, ET, Buckley, YM, Cleland, EE, Davies, KF, Firn, J, et al. Plant species’ origin predicts dominance and response to nutrient enrichment and herbivores in global grasslands. Nature Communications. 2015;6.Google Scholar
Catford, JA, Jansson, R. Drowned, buried and carried away: effects of plant traits on the distribution of native and alien species in riparian ecosystems. New Phytologist. 2014;204(1):1936.Google Scholar
Catford, JA, Jansson, R, Nilsson, C. Reducing redundancy in invasion ecology by integrating hypotheses into a single theoretical framework. Diversity & Distributions. 2009;15(1):2240.Google Scholar
Mitchell, RM, Bakker, JD, Vincent, JB, Davies, GM. Relative importance of abiotic, biotic, and disturbance drivers of plant community structure in the sagebrush steppe. Ecological Applications. 2017;27(3):756–68.Google Scholar
Moles, AT, Flores-Moreno, H, Bonser, SP, Warton, DI, Helm, A, Warman, L, et al. Invasions: the trail behind, the path ahead, and a test of a disturbing idea. Journal of Ecology. 2012;100(1):116–27.Google Scholar
Davis, MA, Grime, JP, Thompson, K. Fluctuating resources in plant communities: a general theory of invasibility. Journal of Ecology. 2000;88:528–34.CrossRefGoogle Scholar
MacDougall, AS, Turkington, R. Are invasive species the drivers or passengers of change in degraded ecosystems? Ecology. 2005;86(1):4255.Google Scholar
Blumenthal, DM, Kray, JA. Climate change, plant traits, and invasion in natural and agricultural ecosystems. In: Ziska, L, Dukes, JS, editors. Invasive species and global climate change. Wallingford: CABI Press; 2014. pp. 6280.Google Scholar
Catford, JA, Downes, BJ, Gippel, CJ, Vesk, PA. Flow regulation reduces native plant cover and facilitates exotic invasion in riparian wetlands. Journal of Applied Ecology. 2011;48(2):432–42.Google Scholar
Catford, JA, Morris, WK, Vesk, PA, Gippel, CJ, Downes, BJ. Species and environmental characteristics point to flow regulation and drought as drivers of riparian plant invasion. Diversity and Distributions. 2014;20(9):1084–96.Google Scholar
Dawson, SK, Warton, DI, Kingsford, RT, Berney, P, Keith, DA, Catford, JA. Plant traits of propagule banks and standing vegetation reveal flooding alleviates impacts of agriculture on wetland restoration. Journal of Applied Ecology. 2017;54(6):1907–18.Google Scholar
Tecco, PA, Díaz, S, Cabido, M, Urcelay, C. Functional traits of alien plants across contrasting climatic and land-use regimes: do aliens join the locals or try harder than them? Journal of Ecology. 2010;98(1):1727.Google Scholar
Bradley, BA, Early, R, Sorte, CJB. Space to invade? Comparative range infilling and potential range of invasive and native plants. Global Ecology and Biogeography. 2015;24(3):348–59.Google Scholar
Higgins, SI, Richardson, DM. Invasive plants have broader physiological niches. Proceedings of the National Academy of Sciences of the USA. 2014;111(29):10,610–4.Google Scholar
Rejmánek, M. Invasiveness. In: Simberloff, D, Rejmánek, M, editors. Encyclopedia of biological invasions. Berkeley, CA: University of California Press; 2011. pp. 379–85.Google Scholar
Wainwright, CE, Wolkovich, EM, Cleland, EE. Seasonal priority effects: implications for invasion and restoration in a semi-arid system. Journal of Applied Ecology. 2012;49(1):234–41.CrossRefGoogle Scholar
Wolkovich, EM, Cleland, EE. The phenology of plant invasions: a community ecology perspective. Frontiers in Ecology and the Environment. 2011;9(5):287–94.Google Scholar
Wolkovich, EM, Cleland, EE. Phenological niches and the future of invaded ecosystems with climate change. AoB Plants. 2014;6:plu013.Google Scholar
van Kleunen, M, Weber, E, Fischer, M. A meta-analysis of trait differences between invasive and non-invasive plant species. Ecology Letters. 2010;13(2):235–45.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
Hellmann, JJ, Byers, JE, Bierwagen, BG, Dukes, JS. Five potential consequences of climate change for invasive species. Conservation Biology. 2008;22(3):534–43.Google Scholar
Petitpierre, B, McDougall, K, Seipel, T, Broennimann, O, Guisan, A, Kueffer, C. Will climate change increase the risk of plant invasions into mountains? Ecological Applications. 2016;26(2):530–44.CrossRefGoogle ScholarPubMed
Dukes, JS, Mooney, HA. Does global change increase the success of biological invaders? Trends in Ecology & Evolution. 1999;14(4):135–9.Google Scholar
White, TA, Campbell, BD, Kemp, PD, Hunt, CL. Impacts of extreme climatic events on competition during grassland invasions. Global Change Biology. 2001;7(1):113.Google Scholar
Bremond, L, Boom, A, Favier, C. Neotropical C3/C4 grass distributions – present, past and future. Global Change Biology. 2012;18(7):2324–34.Google Scholar
Walther, G-R, Roques, A, Hulme, PE, Sykes, MT, Pyšek, P, Kühn, I, et al. Alien species in a warmer world: risks and opportunities. Trends in Ecology & Evolution. 2009;24(12):686–93.Google Scholar
Berg, RY. Plant distribution as seen from plant dispersal: general principles and basic modes of plant dispersal. In: Kubitzki, K, editor. Dispersal and distribution: an international symposium. Hamburg: Paul Parey; 1983. pp. 1336.Google Scholar
Nathan, R, Schurr, FM, Spiegel, O, Steinitz, O, Trakhtenbrot, A, Tsoar, A. Mechanisms of long-distance seed dispersal. Trends in Ecology & Evolution. 2008;23(11):638–47.Google Scholar
Bates, AE, Pecl, GT, Frusher, S, Hobday, AJ, Wernberg, T, Smale, DA, et al. Defining and observing stages of climate-mediated range shifts in marine systems. Global Environmental Change – Human Policy Dimensions. 2014;26:2738.Google Scholar
Sheppard, CS, Alexander, JM, Billeter, R. The invasion of plant communities following extreme weather events under ambient and elevated temperature. Plant Ecology. 2012;213:1289–301.Google Scholar
Diez, JM, D’Antonio, CM, Dukes, JS, Grosholz, ED, Olden, JD, Sorte, CJB, et al. Will extreme climatic events facilitate biological invasions? Frontiers in Ecology and the Environment. 2012;10(5):249–57.Google Scholar
Catford, JA, Daehler, CC, Murphy, HT, Sheppard, AW, Hardesty, BD, Westcott, DA, et al. The intermediate disturbance hypothesis and plant invasions: implications for species richness and management. Perspectives in Plant Ecology, Evolution and Systematics. 2012;14:231–41.Google Scholar
Grime, JP. Evidence for the existence of three primary strategies in plants and its relevance to ecological and evolutionary theory. The American Naturalist. 1977;111(982):1169–94.Google Scholar
Tilman, D. Resource competition and community structure. Princeton, NJ: Princeton University Press; 1982.Google Scholar
Mata, TM, Haddad, NM, Holyoak, M. How invader traits interact with resident communities and resource availability to determine invasion success. Oikos. 2013;122(1):149–60.Google Scholar
Seabloom, EW, Harpole, WS, Reichman, OJ, Tilman, D. Invasion, competitive dominance, and resource use by exotic and native California grassland species. Proceedings of the National Academy of Sciences of the USA. 2003;100(23):13,384–9.Google Scholar
Kane, JM, Meinhardt, KA, Chang, T, Cardall, BL, Michalet, R, Whitham, TG. Drought-induced mortality of a foundation species (Juniperus monosperma) promotes positive afterlife effects in understory vegetation. Plant Ecology. 2011;212(5):733–41.Google Scholar
Liu, Y, Oduor, AMO, Zhang, Z, Manea, A, Tooth, IM, Leishman, MR, et al. Do invasive alien plants benefit more from global environmental change than native plants? Global Change Biology. 2017;23(8):3363–70.CrossRefGoogle ScholarPubMed
Blumenthal, D, Mitchell, CE, Pyšek, P, Jarošík, V. Synergy between pathogen release and resource availability in plant invasion. Proceedings of the National Academy of Sciences of the USA. 2009;106(19):7899–904.Google Scholar
Richardson, DM, Pyšek, P, Rejmánek, M, Barbour, MG, Panetta, FD, West, CJ. Naturalization and invasion of alien plants: concepts and definitions. Diversity and Distributions. 2000;6:93107.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(1):6387.Google Scholar
Rossiter, NA, Setterfield, SA, Douglas, MM, Hutley, LB. Testing the grass–fire cycle: alien grass invasion in the tropical savannas of northern Australia. Diversity & Distributions. 2003;9(3):169–76.Google Scholar
D’Antonio, CM. Fire, plant invasions, and global changes. In: Mooney, H, Hobbs, RJ, editors. Invasive species in a changing world. Washington, DC: Island Press; 2000. pp. 6594.Google Scholar
Kueffer, C. Transdisciplinary research is needed to predict plant invasions in an era of global change. Trends in Ecology & Evolution. 2010;25(11):619–20.Google Scholar
Tilman, D, Hill, J, Lehman, C. Carbon-negative biofuels from low-input high-diversity grassland biomass. Science. 2006;314(5805):1598–600.Google Scholar
Hager, HA, Sinasac, SE, Gedalof, Ze, Newman, JA. Predicting potential global distributions of two Miscanthus grasses: implications for horticulture, biofuel production, and biological invasions. PLoS ONE. 2014;9(6):e100032.Google Scholar
Parrish, DJ, Fike, JH. The biology and agronomy of switchgrass for biofuels. Critical Reviews in Plant Sciences. 2005;24(5–6):423–59.Google Scholar
Barney, JN, DiTomaso, JM. Bioclimatic predictions of habitat suitability for the biofuel switchgrass in North America under current and future climate scenarios. Biomass and Bioenergy. 2010;34(1):124–33.Google Scholar
Driscoll, D, Catford, J. Invasive plants: new pasture plants pose weed risk. Nature. 2014;516(7529):37.Google Scholar
Driscoll, DA, Catford, JA, Barney, JN, Hulme, PE, Inderjit, , Martin, TG, et al. New pasture plants intensify invasive species risk. Proceedings of the National Academy of Sciences of the USA. 2014; 111(46):16,622–7.Google Scholar
Bradley, BA, Blumenthal, DM, Early, R, Grosholz, ED, Lawler, JJ, Miller, LP, et al. Global change, global trade, and the next wave of plant invasions. Frontiers in Ecology and the Environment. 2012;10(1):20–8.Google Scholar
Haeuser, E, Dawson, W, van Kleunen, M. The effects of climate warming and disturbance on the colonization potential of ornamental alien plant species. Journal of Ecology. 2017;105(6):1698–708.Google Scholar
Rinella, MJ, Maxwell, BD, Fay, PK, Weaver, T, Sheley, RL. Control effort exacerbates invasive-species problem. Ecological Applications. 2009;19(1):155–62.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
Blumenthal, DM, Kray, JA, Ortmans, W, Ziska, LH, Pendall, E. Cheatgrass is favored by warming but not CO2 enrichment in a semi-arid grassland. Global Change Biology. 2016;22(9):3026–38.Google Scholar
Adler, PB, Leiker, J, Levine, JM. Direct and indirect effects of climate change on a prairie plant community. PLoS ONE. 2009;4(9):e6887.Google Scholar
Bansal, S, Sheley, RL. Annual grass invasion in sagebrush steppe: the relative importance of climate, soil properties and biotic interactions. Oecologia. 2016;181(2):543–57.Google Scholar
Alexander, JM, Diez, JM, Levine, JM. Novel competitors shape species’ responses to climate change. Nature. 2015;525(7570):515–8.Google Scholar

References

Weaver, JE. North American prairie. Lincoln, NB: Johnson Publishing Company; 1954.Google Scholar
Weaver, JE, Albertson, FW. Effects of the great drought on the prairies of Iowa, Nebraska, and Kansas. Ecology. 1936;17:567639.Google Scholar

10.5 References

Ehrlén, J, Morris, WF. Predicting changes in the distribution and abundance of species under environmental change. Ecology Letters. 2015;18(3):303–14.CrossRefGoogle ScholarPubMed
Merow, C, Latimer, AM, Wilson, AM, McMahon, SM, Rebelo, AG, Silander, JA. On using integral projection models to generate demographically driven predictions of species’ distributions: development and validation using sparse data. Ecography. 2014;37(12):1167–83.Google Scholar
Villellas, J, Doak, DF, García, MB, Morris, WF. Demographic compensation among populations: what is it, how does it arise and what are its implications? Ecology Letters. 2015;18(11):1139–52.Google Scholar
Nicolè, F, Dahlgren, JP, Vivat, A, Till‐Bottraud, I, Ehrlén, J. Interdependent effects of habitat quality and climate on population growth of an endangered plant. Journal of Ecology. 2011;99(5):1211–8.Google Scholar
Gibson, DJ. Grasses and grassland ecology. Oxford: Oxford University Press; 2009.Google Scholar
Ehrlén, J, Morris, WF, Euler, T, Dahlgren, JP. Advancing environmentally explicit structured population models of plants. Journal of Ecology. 2016;104(2):292305.Google Scholar
Adler, PB, HilleRisLambers, J. The influence of climate and species composition on the population dynamics of ten prairie forbs. Ecology. 2008;89(11):3049–60.Google Scholar
Caswell, H. Matrix population models. 2nd edn. Sunderland, MA: Sinauer Associates; 2001.Google Scholar
Easterling, MR, Ellner, SP, Dixon, PM. Size‐specific sensitivity: applying a new structured population model. Ecology. 2000;81(3):694708.Google Scholar
Griffith, AB, Salguero‐Gómez, R, Merow, C, McMahon, S. Demography beyond the population. Journal of Ecology. 2016;104(2):271–80.Google Scholar
Merow, C, Dahlgren, JP, Metcalf, CJE, Childs, DZ, Evans, MEK, Jongejans, E, et al. Advancing population ecology with integral projection models: a practical guide. Methods in Ecology and Evolution. 2014;5(2):99110.Google Scholar
Birch, LC. Experimental background to the study of the distribution and abundance of insects: I. The influence of temperature, moisture and food on the innate capacity for increase of three grain beetles. Ecology. 1953;34(4):698711.Google Scholar
Holt, RD. Bringing the Hutchinsonian niche into the 21st century: ecological and evolutionary perspectives. Proceedings of the National Academy of Sciences of the USA. 2009;106 (Suppl. 2):19,659–65.Google Scholar
Pulliam, HR. On the relationship between niche and distribution. Ecology Letters. 2000;3(4):349–61.Google Scholar
Sheth, SN, Angert, AL. Artificial selection reveals high genetic variation in phenology at the trailing edge of a species range. The American Naturalist. 2016;187(2):182–93.Google Scholar
Martorell, C. Detecting and managing an overgrazing–drought synergism in the threatened Echeveria longissima (Crassulaceae): the role of retrospective demographic analysis. Population Ecology. 2007;49(2):115–25.Google Scholar
Williams, AL, Wills, KE, Janes, JK, Vander Schoor, JK, Newton, PC, Hovenden, MJ. Warming and free‐air CO2 enrichment alter demographics in four co‐occurring grassland species. New Phytologist. 2007;176(2):365–74.Google Scholar
Toräng, P, Ehrlén, J, Ågren, J. Linking environmental and demographic data to predict future population viability of a perennial herb. Oecologia. 2010;163(1):99109.Google Scholar
Dalgleish, HJ, Koons, DN, Hooten, MB, Moffet, CA, Adler, PB. Climate influences the demography of three dominant sagebrush steppe plants. Ecology. 2011;92(1):7585.Google Scholar
Evju, M, Halvorsen, R, Rydgren, K, Austrheim, G, Mysterud, A. Effects of sheep grazing and temporal variability on population dynamics of the clonal herb Geranium sylvaticum in an alpine habitat. Plant Ecology. 2011;212(8):1299–312.Google Scholar
Adler, PB, Dalgleish, HJ, Ellner, SP. Forecasting plant community impacts of climate variability and change: when do competitive interactions matter? Journal of Ecology. 2012;100(2):478–87.Google Scholar
Sletvold, N, Dahlgren, JP, Øien, DI, Moen, A, Ehrlén, J. Climate warming alters effects of management on population viability of threatened species: results from a 30‐year experimental study on a rare orchid. Global Change Biology. 2013;19(9):2729–38.Google Scholar
Raghu, S, Osunkoya, OO, Perrett, C, Pichancourt, J-B. Historical demography of Lantana camara L. reveals clues about the influence of land use and weather in the management of this widespread invasive species. Basic and Applied Ecology. 2014;15(7):565–72.Google Scholar
Compagnoni, A, Adler, PB. Warming, competition, and Bromus tectorum population growth across an elevation gradient. Ecosphere. 2014;5(9):134.Google Scholar
Louthan, AM, Doak, DF, Goheen, JR, Palmer, TM, Pringle, RM. Mechanisms of plant–plant interactions: concealment from herbivores is more important than abiotic-stress mediation in an African savannah. Proceedings of the Royal Society of London B: Biological Sciences. 2014;281(1780):20132647.Google Scholar
Prevéy, JS, Seastedt, TR. Effects of precipitation change and neighboring plants on population dynamics of Bromus tectorum. Oecologia. 2015;179(3):765–75.Google Scholar
Dahlgren, JP, Bengtsson, K, Ehrlén, J. The demography of climate‐driven and density‐regulated population dynamics in a perennial plant. Ecology. 2016;97(4):899907.Google Scholar
Chu, C, Kleinhesselink, AR, Havstad, KM, McClaran, MP, Peters, DP, Vermeire, LT, et al. Direct effects dominate responses to climate perturbations in grassland plant communities. Nature Communications. 2016;7:11766.Google Scholar

11.7 References

Sala, OE, Vivanco, L, Flombaum, P. Grassland ecology. In: Encyclopedia of biodiversity [Internet]. 2nd edn.; Burlington, USA: Elsevier Science. 2013. pp. 17.Google Scholar
O’Mara, FP. The role of grasslands in food security and climate change. Annals of Botany. 2012;110(6):1263–70.Google Scholar
Suttle, KB, Thomsen, MA, Power, ME. Species interactions reverse grassland responses to changing climate. Science. 2007;315(5812):640–2.Google Scholar
Craine, JM, Elmore, AJ, Olson, K, Tolleson, D. Climate change and cattle nutritional stress. Global Change Biology. 2010;16(10):2901–11.Google Scholar
Robinson, EA, Ryan, GD, Newman, JA. A meta‐analytical review of the effects of elevated CO2 on plant–arthropod interactions highlights the importance of interacting environmental and biological variables. New Phytologist. 2012;194(2):321–36.Google Scholar
Terborgh, J, Estes, JA. Trophic cascades: predators, prey, and the changing dynamics of nature. Washington, DC: Island Press; 2013.Google Scholar
Wand, SJE, Midgley, G, Jones, MH, Curtis, PS. Responses of wild C4 and C3 grass (Poaceae) species to elevated atmospheric CO2 concentration: a meta‐analytic test of current theories and perceptions. Global Change Biology. 1999;5(6):723–41.Google Scholar
Tyliaakis, JM, Didham, RK, Bascompte, J, Wardle, DA. Global change and species interactions in terrestrial ecosystems. Ecology Letters. 2008;11(12):1351–63.Google Scholar
Zvereva, EL, Kozlov, MV. Consequences of simultaneous elevation of carbon dioxide and temperature for plant–herbivore interactions: a meta-analysis. Global Change Biology. 2006;12(1):2741.Google Scholar
Johnson, SN, Hartley, SE. Elevated carbon dioxide and warming impact silicon and phenolic‐based defences differently in native and exotic grasses. Global Change Biology. 2017.Google Scholar
Gherardi, LA, Sala, OE. Enhanced interannual precipitation variability increases plant functional diversity that in turn ameliorates negative impact on productivity. Ecology Letters. 2015;18(12):1293–300.Google Scholar
IPCC. Climate change 2013: the physical science basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Stocker, TF, Qin, D, Plattner, G-K, Tignor, M, Allen, SK, Boschung, J, et al., editors. Cambridge: Cambridge University Press; 2013. 1535 pp.Google Scholar
Stireman, J, Dyer, LA, Janzen, DH, Singer, M, Lill, J, Marquis, RJ, et al. Climatic unpredictability and parasitism of caterpillars: implications of global warming. Proceedings of the National Academy of Sciences of the USA. 2005;102(48):17,384–7.Google Scholar
Jones, CG, Hartley, SE. A protein competition model of phenolic allocation. Oikos. 1999;86(1):2744.Google Scholar
Bezemer, TM, Jones, TH. Plant–insect herbivore interactions in elevated atmospheric CO2: quantitative analyses and guild effects. Oikos. 1998;82:212–22.Google Scholar
Stiling, P, Cornelissen, T. How does elevated carbon dioxide (CO2) affect plant–herbivore interactions? A field experiment and meta‐analysis of CO2‐mediated changes on plant chemistry and herbivore performance. Global Change Biology. 2007;13(9):1823–42.Google Scholar
Luo, Y, Hui, D, Zhang, D. Elevated CO2 stimulates net accumulations of carbon and nitrogen in land ecosystems: a meta‐analysis. Ecology. 2006;87(1):5363.Google Scholar
Gargallo‐Garriga, A, Sardans, J, Pérez‐Trujillo, M, Oravec, M, Urban, O, Jentsch, A, et al. Warming differentially influences the effects of drought on stoichiometry and metabolomics in shoots and roots. New Phytologist. 2015;207(3):591603.Google Scholar
Johnson, SN, Lopaticki, G, Hartley, SE. Elevated atmospheric CO2 triggers compensatory feeding by root herbivores on a C3 but not a C4 grass. PloS ONE. 2014;9(3):e90251.Google Scholar
Morgan, JA, LeCain, DR, Pendall, E, Blumenthal, DM, Kimball, BA, Carrillo, Y, et al. C4 grasses prosper as carbon dioxide eliminates desiccation in warmed semi-arid grassland. Nature. 2011;476(7359):202.Google Scholar
Zavaleta, ES, Shaw, MR, Chiariello, NR, Thomas, BD, Cleland, EE, Field, CB, et al. Grassland responses to three years of elevated temperature, CO2, precipitation, and N deposition. Ecological Monographs. 2003;73(4):585604.Google Scholar
Hartley, SE, Jones, CG, Couper, GC, Jones, TH. Biosynthesis of plant phenolic compounds in elevated atmospheric CO2. Global Change Biology. 2000;6(5):497506.Google Scholar
Reich, PB, Hobbie, SE, Lee, TD. Plant growth enhancement by elevated CO2 eliminated by joint water and nitrogen limitation. Nature Geoscience. 2014;7(12):920.Google Scholar
Hartley, SE, DeGabriel, JL. The ecology of herbivore‐induced silicon defences in grasses. Functional Ecology. 2016;30(8):1311–22.Google Scholar
Ryalls, JM, Hartley, SE, Johnson, SN. Impacts of silicon-based grass defences across trophic levels under both current and future atmospheric CO2 scenarios. Biology Letters. 2017;13(3):20160912.Google Scholar
McLarnon, E, McQueen-Mason, S, Lenk, I, Hartley, SE. Evidence for active uptake and deposition of Si-based defenses in tall fescue. Frontiers in Plant Science. 2017;8:1199.Google Scholar
Buitenwerf, R, Bond, W, Stevens, N, Trollope, W. Increased tree densities in South African savannas: > 50 years of data suggests CO2 as a driver. Global Change Biology. 2012;18(2):675–84.CrossRefGoogle Scholar
Cui, T, Martz, L, Guo, X. Grassland phenology response to drought in the Canadian Prairies. Remote Sensing. 2017;9(12):1258.Google Scholar
Harris, G, Thirgood, S, Hopcraft, JGC, Cromsight, JPGM, Berger, J. Global decline in aggregated migrations of large terrestrial mammals. Endangered Species Research. 2009;7:5576.Google Scholar
DeLucia, EH, Nabity, PD, Zavala, JA, Berenbaum, MR. Climate change: resetting plant–insect interactions. Plant Physiology. 2012;160(4):1677–85.Google Scholar
AbdElgawad, H, Peshev, D, Zinta, G, Van den Ende, W, Janssens, IA, Asard, H. Climate extreme effects on the chemical composition of temperate grassland species under ambient and elevated CO2: a comparison of fructan and non-fructan accumulators. PLoS ONE. 2014;9(3):e92044.Google Scholar
Massey, FP, Massey, K, Ennos, AR, Hartley, SE. Impacts of silica-based defences in grasses on the feeding preferences of sheep. Basic and Applied Ecology. 2009;10(7):622–30.Google Scholar
Reynolds, OL, Keeping, MG, Meyer, JH. Silicon‐augmented resistance of plants to herbivorous insects: a review. Annals of Applied Biology. 2009;155(2):171–86.Google Scholar
Wade, RN, Karley, AJ, Johnson, SN, Hartley, SE. Impact of predicted precipitation scenarios on multitrophic interactions. Functional Ecology. 2017;31(8):1647–58.Google Scholar
Flannigan, MD, Krawchuk, MA, de Groot, WJ, Wotton, BM, Gowman, LM. Implications of changing climate for global wildland fire. International Journal of Wildland Fire. 2009;18(5):483507.Google Scholar
Jolly, WM, Cochrane, MA, Freeborn, PH, Holden, ZA, Brown, TJ, Williamson, GJ, et al. Climate-induced variations in global wildfire danger from 1979 to 2013. Nature Communications. 2015;6:7537.Google Scholar
Bond, WJ, Keeley, JE. Fire as a global ‘herbivore’: the ecology and evolution of flammable ecosystems. Trends in Ecology & Evolution. 2005;20(7):387–94.Google Scholar
Alstad, AO, Damschen, EI, Givnish, TJ, Harrington, JA, Leach, MK, Rogers, DA, et al. The pace of plant community change is accelerating in remnant prairies. Science Advances. 2016;2(2):e1500975.Google Scholar
Archibald, S. Managing the human component of fire regimes: lessons from Africa. Philosophical Transactions of the Royal Society B: Biological Sciences. 2016;371(1696):20150346.Google Scholar
Hempson, GP, Parr, CL, Archibald, S, Anderson, TM, Mustaphi, CJC, Dobson, AP, et al. Continent‐level drivers of African pyrodiversity. Ecography. 2018;41(6):889–99.Google Scholar
Ojima, DS, Schimel, D, Parton, W, Owensby, C. Long- and short-term effects of fire on nitrogen cycling in tallgrass prairie. Biogeochemistry. 1994;24(2):6784.Google Scholar
Van de Vijver, CADM, Poot, P, Prins, HHT. Causes of increased nutrient concentrations in post-fire regrowth in an East African savanna. Plant and Soil. 1999;214(1–2):173–85.Google Scholar
Byenkya, GS, Gumisiriza, G, Kasigwa, H. Evaluation of control strategies for Cymbopogon nardus in grazing areas of Uganda. Journal of Agricultural Science and Technology. 2013;3(9B):656.Google Scholar
Bale, JS, Masters, GJ, Hodkinson, ID, Awmack, C, Bezemer, TM, Brown, VK, et al. Herbivory in global climate change research: direct effects of rising temperature on insect herbivores. Global Change Biology. 2002;8(1):116.Google Scholar
Awmack, CS, Leather, SR. Host plant quality and fecundity in herbivorous insects. Annual Review of Entomology. 2002;47(1):817–44.Google Scholar
Nord, EA, Lynch, JP. Plant phenology: a critical controller of soil resource acquisition. Journal of Experimental Botany. 2009;60(7):1927–37.Google Scholar
Mattson, WJ, Haack, RA. The role of drought in outbreaks of plant-eating insects. Bioscience. 1987;37(2):110–8.Google Scholar
de Sassi, C, Tylianakis, JM. Climate change disproportionately increases herbivore over plant or parasitoid biomass. PLoS ONE. 2012;7(7):e40557.Google Scholar
Yu, G, Shen, H, Liu, J. Impacts of climate change on historical locust outbreaks in China. Journal of Geophysical Research: Atmospheres. 2009;114(D18).Google Scholar
Johnson, SN, Staley, JT, McLeod, FA, Hartley, SE. Plant‐mediated effects of soil invertebrates and summer drought on above‐ground multitrophic interactions. Journal of Ecology. 2011;99(1):5765.Google Scholar
Pritchard, J, Griffiths, B, Hunt, E. Can the plant‐mediated impacts on aphids of elevated CO2 and drought be predicted? Global Change Biology. 2007;13(8):1616–29.Google Scholar
Hopcraft, JGC, Anderson, TM, Pérez‐Vila, S, Mayemba, E, Olff, H. Body size and the division of niche space: food and predation differentially shape the distribution of Serengeti grazers. Journal of Animal Ecology. 2012;81(1):201–13.Google Scholar
Hopcraft, JGC, Olff, H, Sinclair, A. Herbivores, resources and risks: alternating regulation along primary environmental gradients in savannas. Trends in Ecology & Evolution. 2010;25(2):119–28.Google Scholar
McNaughton, SJ. Grazing lawns: animals in herds, plant form, and coevolution. The American Naturalist. 1984;124:863–86.Google Scholar
Archibald, S. African grazing lawns – how fire, rainfall, and grazer numbers interact to affect grass community states. Journal of Wildlife Management. 2008;72(2):492501.Google Scholar
Donaldson, JE, Archibald, S, Govender, N, Pollard, D, Luhdo, Z, Parr, CL. Ecological engineering through fire–herbivory feedbacks drives the formation of savanna grazing lawns. Journal of Applied Ecology. 2018;55(1):225–35.Google Scholar
Ogutu, JO, Piepho, H-P, Dublin, HT. Reproductive seasonality in African ungulates in relation to rainfall. Wildlife Research. 2015;41(4):323–42.Google Scholar
Langvatn, R, Mysterud, A, Stenseth, NC, Yoccoz, NG. Timing and synchrony of ovulation in red deer constrained by short northern summers. The American Naturalist. 2004;163(5):763–72.Google Scholar
Mduma, SA, Sinclair, A, Hilborn, R. Food regulates the Serengeti wildebeest: a 40‐year record. Journal of Animal Ecology. 1999;68(6):1101–22.Google Scholar
Post, E, Pedersen, C. Opposing plant community responses to warming with and without herbivores. Proceedings of the National Academy of Sciences of the USA. 2008;105(34):12,353–8.Google Scholar
Hartley, SE, Gange, AC. Impacts of plant symbiotic fungi on insect herbivores: mutualism in a multitrophic context. Annual Review of Entomology. 2009;54:323–42.Google Scholar
Brosi, GB, McCulley, RL, Bush, LP, Nelson, JA, Classen, AT, Norby, RJ. Effects of multiple climate change factors on the tall fescue–fungal endophyte symbiosis: infection frequency and tissue chemistry. New Phytologist. 2011;189(3):797805.Google Scholar
Harrington, R, Woiwod, I, Sparks, T. Climate change and trophic interactions. Trends in Ecology & Evolution. 1999;14(4):146–50.Google Scholar
Beale, CM, Baker, NE, Brewer, MJ, Lennon, JJ. Protected area networks and savannah bird biodiversity in the face of climate change and land degradation. Ecology Letters. 2013;16(8):1061–8.Google Scholar
Stenseth, NC, Mysterud, A. Climate, changing phenology, and other life history traits: nonlinearity and match–mismatch to the environment. Proceedings of the National Academy of Sciences of the USA. 2002;99(21):13,379–81.Google Scholar
Woodroffe, R, Groom, R, McNutt, JW. Hot dogs: high ambient temperatures impact reproductive success in a tropical carnivore. Journal of Animal Ecology. 2017;86(6):1329–38.Google Scholar
Loarie, SR, Tambling, CJ, Asner, GP. Lion hunting behaviour and vegetation structure in an African savanna. Animal Behaviour. 2013;85(5):899906.Google Scholar
Riginos, C. Climate and the landscape of fear in an African savanna. Journal of Animal Ecology. 2015;84(1):124–33.Google Scholar
deMenocal, PB. African climate change and faunal evolution during the Pliocene–Pleistocene. Earth and Planetary Science Letters. 2004;220(1–2):324.Google Scholar
Grime, JP, Fridley, JD, Askew, AP, Thompson, K, Hodgson, JG, Bennett, CR. Long-term resistance to simulated climate change in an infertile grassland. Proceedings of the National Academy of Sciences of the USA. 2008;105(29):10,028–32.Google Scholar

12.7 References

Bardgett, RD, Cook, R. Functional aspects of soil animal diversity in agricultural grasslands. Applied Soil Ecology. 1998;10:263–76.Google Scholar
Bardgett, RD, van der Putten, WH. Belowground biodiversity and ecosystem functioning. Nature. 2014;515:505–11.CrossRefGoogle ScholarPubMed
van Der Heijden, MGA, Bardgett, RD, van Straalen, NM. The unseen majority: soil microbes as drivers of plant diversity and productivity in terrestrial ecosystems. Ecology Letters. 2008;11:296310.Google Scholar
Bardgett, RD, Freeman, C, Ostle, NJ. Microbial contributions to climate change through carbon cycle feedbacks. The ISME Journal. 2008;2:805–14.Google Scholar
Classen, AT, Sundqvist, MK, Henning, JA, Newman, GS, Moore, JAM, Cregger, MA, et al. Direct and indirect effects of climate change on soil microbial and soil microbial–plant interactions: what lies ahead? Ecosphere. 2015;6:121.Google Scholar
Jenkinson, DS, Adams, DE, Wild, A. Model estimates of CO2 emissions from soil in response to global warming. Nature. 1991;351:304–6.Google Scholar
Cox, PM, Betts, RA, Jones, CD, Spall, SA, Totterdell, IJ. Acceleration of global warming due to carbon-cycle feedbacks in a coupled climate model. Nature. 2000;408:184–7.Google Scholar
Dorrepaal, E, Toet, S, van Logtestijn, RSP, Swart, E, van de Weg, MJ, Callaghan, TV, et al. Carbon respiration from subsurface peat accelerated by climate warming in the subarctic. Nature. 2009;460:616–79.Google Scholar
Zhou, J, Xue, K, Xie, J, Deng, Y, Wu, L, Cheng, X, et al. Microbial mediation of carbon-cycle feedbacks to climate warming. Nature Climate Change. 2012;2:106–10.Google Scholar
Blankinship, J, Niklaus, P, Hungate, B. A meta-analysis of responses of soil biota to global change. Oecologia. 2011;165:553–65.Google Scholar
Crowther, TW, Todd-Brown, KEO, Rowe, CW, Wieder, WR, Carey, JC, Machmuller, MB, et al. Quantifying global soil carbon losses in response to warming. Nature. 2016;540:104–8.Google Scholar
Schuur, EA, Vogel, JG, Crummer, KG, Lee, H, Sickman, JO, Osterkamp, TE. The effect of permafrost thaw on old carbon release and net carbon exchange from tundra. Nature. 2009;459:556–9.Google Scholar
Xue, K, Yuan, MM, Shi, ZJ, Qin, Y, Deng, Y, Cheng, L, et al. Tundra soil carbon is vulnerable to rapid microbial decomposition under climate warming. Nature Climate Change. 2016;6:595600.Google Scholar
Luo, Y, Sherry, R, Zhou, X, Wan, S. Terrestrial carbon‐cycle feedback to climate warming: experimental evidence on plant regulation and impacts of biofuel feedstock harvest. Global Change Bioenergy. 2009;1:6274.Google Scholar
White, SR, Bork, EW, Cahill, JF. Direct and indirect drivers of plant diversity responses to climate and clipping across northern temperate grassland. Ecology. 2014;95:3093–103.Google Scholar
Hoeppner, SS, Dukes, JS. Interactive responses of old-field plant growth and composition to warming and precipitation. Global Change Biology. 2012;18:1754–68.Google Scholar
Mowll, W, Blumenthal, DM, Cherwin, K, Smith, A, Symstad, AJ, Vermeire, LT, et al. Climatic controls of aboveground net primary production in semi-arid grasslands along a latitudinal gradient portend low sensitivity to warming. Oecologia. 2015;177:959–69.Google Scholar
Wookey, PA, Aerts, R, Bardgett, RD, Baptist, F, Brathen, KA, Cornelissen, JHC, et al. Ecosystem feedbacks and cascade processes: understanding their role in the responses of Arctic and alpine ecosystems to environmental change. Global Change Biology. 2009;15:1153–72.Google Scholar
Bardgett, RD, Wardle, DA. Aboveground–belowground linkages: biotic interactions, ecosystem processes, and global change. Oxford: Oxford University Press; 2010.Google Scholar
Kardol, P, Cregger, MA, Campany, CE, Classen, AT. Soil ecosystem functioning under climate change: plant species and community effects. Ecology. 2010;91:767–81.Google Scholar
Eisenhauer, N, Cesarz, S, Koller, R, Worm, K, Reich, PB. Global change belowground: impacts of elevated CO2, nitrogen, and summer drought on soil food webs and biodiversity. Global Change Biology. 2012;18:435–47.Google Scholar
Eisenhauer, N, Dobies, T, Cesarz, S, Hobbie, SE, Meyer, RJ, Worm, K, et al. Plant diversity effects on soil food webs are stronger than those of elevated CO2 and N deposition in a long-term grassland experiment. Proceedings of the National Academy of Sciences. 2013;110:6889–94.Google Scholar
Mueller, KE, Blumenthal, DM, Carrillo, Y, Cesarz, S, Ciobanu, M, Hines, J, et al. Elevated CO2 and warming shift the functional composition of soil nematode communities in a semiarid grassland. Soil Biology and Biochemistry. 2016;103:4651.Google Scholar
Reich, PB, Hobbie, SE, Lee, TD. Plant growth enhancement by elevated CO2 eliminated by joint water and nitrogen limitation. Nature Geoscience. 2014;7:920–4.Google Scholar
Pendall, E, Mosier, AR, Morgan, JA. Rhizodeposition stimulated by elevated CO2 in a semiarid grassland. New Phytologist. 2004;162:447–58.Google Scholar
Nie, M, Lu, M, Bell, J, Raut, S, Pendall, E. Altered root traits due to elevated CO2: a meta‐analysis. Global Ecology and Biogeography. 2013;22:1095–105.Google Scholar
Drigo, B, Pijl, AS, Duyts, H, Kielak, AM, Gamper, HA, Houtekamer, MJ, et al. Shifting carbon flow from roots into associated microbial communities in response to elevated atmospheric CO2. Proceedings of the National Academy of Sciences. 2010;107:10,938–42.Google Scholar
Högberg, P, Read, DJ. Towards a more plant physiological perspective on soil ecology. Trends in Ecology and Evolution. 2006;21:548–54.Google Scholar
Rillig, MC, Wright, SF, Allen, MF, Field, CB. Rise in carbon dioxide changes in soil structure. Nature. 1999;400:628.Google Scholar
Wilson, GW, Rice, CW, Rillig, MC, Springer, A, Hartnett, DC. Soil aggregation and carbon sequestration are tightly correlated with the abundance of arbuscular mycorrhizal fungi: results from long‐term field experiments. Ecology Letters. 2009;12:452–61.Google Scholar
Fontaine, S, Bardoux, G, Abbadie, L, Mariotti, A. Carbon input to soil may decrease soil carbon content. Ecology Letters. 2004;7:314–20.Google Scholar
Langley, JA, McKinley, DC, Wolf, AA, Hungate, BA, Drake, BG, Megonigal, JP. Priming depletes soil carbon and releases nitrogen in a scrub-oak ecosystem exposed to elevated CO2. Soil Biology and Biochemistry. 2009;41:5460.Google Scholar
Phillips, RP, Finzi, AC, Bernhardt, ES. Enhanced root exudation induces microbial feedbacks to N cycling in a pine forest under long-term CO2 fumigation. Ecology Letters. 2011;14:187–94.Google Scholar
Hu, S, Chapin III, FS, Firestone, MK, Field, CB, Chiariello, NR. Nitrogen limitation of microbial decomposition in a grassland under elevated CO2. Nature. 2001;409:188–91.Google Scholar
Diaz, S, Grime, JP, Harris, J, McPherson, E. Evidence of a feedback mechanism limiting plant-response to elevated carbon-dioxide. Nature. 1993;364:616–7.Google Scholar
Reichstein, M, Bahn, M, Ciais, P, Frank, D, Mahecha, MD, Seneviratne, SI, et al. Climate extremes and the carbon cycle. Nature. 2013;500:287–95.Google Scholar
Arnone III, JA, Verburg, PSJ, Johnson, DW, Larsen, JD, Jasoni, RL, Lucchesi, AJ, et al. Prolonged suppression of ecosystem carbon dioxide uptake after an anomalously warm year. Nature. 2008;455;383–6.Google Scholar
Hawkes, CV, Kivlin, SN, Rocca, JD, Huguet, V, Thomsen, MA, Suttle, KB. Fungal community responses to precipitation. Global Change Biology. 2011;17:1637–45.Google Scholar
De Vries, FT, Bloem, J, Quirk, H, Stevens, CJ, Bol, R, Bardgett, RD. Extensive management promotes plant and microbial nitrogen retention in temperate grassland. PLoS ONE. 2012;7:e51201.Google Scholar
De Vries, FT, Shade, A. Controls on soil microbial community stability under climate change. Frontiers in Microbiology. 2013;4:265.Google Scholar
Lau, JA, Lennon, JT. Rapid responses of soil microorganisms improve plant fitness in novel environments. Proceedings of the National Academy of Sciences. 2012;109:14,058–62.Google Scholar
Mariotte, P, Canarini, A, Dijkstra, FA. Stoichiometric N:P flexibility and mycorrhizal symbiosis favour plant resistance against drought. Journal of Ecology. 2017;105:958–67.Google Scholar
Rubin, RL, van Groenigen, KJ, Hungate, BA. Plant growth promoting rhizobacteria are more effective under drought: a meta-analysis. Plant and Soil. 2017;416:309–23.Google Scholar
Kaisermann, A, Vries, FT, Griffiths, RI, Bardgett, RD. Legacy effects of drought on plant–soil feedbacks and plant–plant interactions. New Phytologist. 2017;215:1413–24.Google Scholar
Meisner, A, De Deyn, GB, de Boer, W, Van der Putten, WH. Soil biotic legacy effects of extreme weather events influence plant invasiveness. Proceedings of the National Academy of Sciences. 2013;110:9835–8.Google Scholar
Schrama, M, Bardgett, RD. Grassland invasibility varies with drought effects on soil functioning. Journal of Ecology. 2016;104:1250–8.Google Scholar
De Vries, FT, Liiri, ME, Bjørnlund, L, Bowker, MA, Christensen, S, Setälä, HM, et al. Land use alters the resistance and resilience of soil food webs to drought. Nature Climate Change. 2012;2:276–80.CrossRefGoogle Scholar
Knapp, AK, Carroll, CJ, Denton, EM, La Pierre, KJ, Collins, SL, Smith, MD. Differential sensitivity to regional-scale drought in six central US grasslands. Oecologia. 2015;177:949–57.Google Scholar
Bardgett, RD, Manning, P, Morrien, E, De Vries, FT. Hierarchical responses of plant–soil interactions to climate change: consequences for the global carbon cycle. Journal of Ecology. 2013;101:334–43.Google Scholar
Post, E, Pedersen, C, Wilmers, CC, Forchhammer, MC. Warming, plant phenology and the spatial dimension of trophic mismatch for large herbivores. Proceedings of the Royal Society B: Biological Sciences. 2008;275:2005–13.Google Scholar
Robbirt, KM, Roberts, DL, Hutchings, MJ, Davy, AJ. Potential disruption of pollination in a sexually deceptive orchid by climatic change. Current Biology. 2014;24:2845–9.Google Scholar
Berg, MP, Kiers, ET, Driessen, G, van der Heijden, M, Kooi, BW, Kuenen, F, et al. Adapt or disperse: understanding species persistence in a changing world. Global Change Biology. 2010;16:587–98.Google Scholar
van der Putten, WH, Macel, M, Visser, ME. Predicting species distribution and abundance responses to climate change: why it is essential to include biotic interactions across trophic levels. Philosophical Transactions of the Royal Society B: Biological Sciences. 2010;365:2025–34.Google Scholar
Heide, OM. Control of flowering and reproduction in temperate grasses. New Phytologist. 1994;128:347–62.Google Scholar
Keller, F, Korner, C. The role of photoperiodism in alpine plant development. Arctic Antarctic and Alpine Research. 2003;35:361–8.Google Scholar
Pautasso, M, Doring, TF, Garbelotto, M, Pellis, L, Jeger, MJ. Impacts of climate change on plant diseases – opinions and trends. European Journal of Plant Pathology. 2012;133:295313.Google Scholar
Veresoglou, SD, Rillig, MC. Suppression of fungal and nematode plant pathogens through arbuscular mycorrhizal fungi. Biology Letters. 2012;8:214–7.Google Scholar
Pieterse, CMJ, Zamioudis, C, Berendsen, RL, Weller, DM, Van Wees, SCM, Bakker, P. Induced systemic resistance by beneficial microbes. Annual Review of Phytopathology. 2014;52:347–75.Google Scholar
Vierheilig, H, Steinkellner, S, Khaosaad, T, Garcia-Garrido, GM. The biocontrol effect of mycorrhization on soilborne fungal pathogens and the autoregulation of AM symbiosis: one mechanism, two effects? In: Varma, A, editor. Mycorrhiza: genetics and molecular biology. Berlin: Springer; 2008. pp. 307320.Google Scholar
Sikes, BA, Cottenie, K, Klironomos, JN. Plant and fungal identity determines pathogen protection of plant roots by arbuscular mycorrhizas. Journal of Ecology. 2009;97:1274–80.Google Scholar
Chomel, M, Guittonny-Larcheveque, M, Fernandez, C, Gallet, C, DesRochers, A, Pare, D, et al. Plant secondary metabolites: a key driver of litter decomposition and soil nutrient cycling. Journal of Ecology. 2016;104:1527–41.Google Scholar
Engelkes, T, Morrien, E, Verhoeven, KJF, Bezemer, TM, Biere, A, Harvey, JA, et al. Successful range-expanding plants experience less above-ground and below-ground enemy impact. Nature. 2008;456:946–8.Google Scholar
Morrien, E, Duyts, H, van der Putten, WH. Effects of native and exotic range-expanding plant species on taxonomic and functional composition of nematodes in the soil food web. Oikos. 2012;121:181–90.Google Scholar
Wilschut, RA, Geisen, S, ten Hooven, FC, van der Putten, WH. Interspecific differences in nematode control between range-expanding plant species and their congeneric natives. Soil Biology & Biochemistry. 2016;100:233–41.Google Scholar
Hampe, A, Petit, RJ. Conserving biodiversity under climate change: the rear edge matters. Ecology Letters. 2005;8:461–7.Google Scholar
Wardle, DA, Bardgett, RD, Callaway, RM, van der Putten, WH. Terrestrial ecosystem responses to species gains and losses. Science. 2011;332:1273–7.Google Scholar
Crutsinger, GM, Sanders, NJ, Classen, AT. Comparing intra- and inter-specific effects on litter decomposition in an old-field ecosystem. Basic and Applied Ecology. 2009;10:535–43.Google Scholar
Semchenko, M, Saar, S, Lepik, A. Intraspecific genetic diversity modulates plant–soil feedback and nutrient cycling. New Phytologist. 2017;216:90–8.Google Scholar
Jump, AS, Penuelas, J. Running to stand still: adaptation and the response of plants to rapid climate change. Ecology Letters. 2005;8:1010–20.Google Scholar
Moran, EV, Alexander, JM. Evolutionary responses to global change: lessons from invasive species. Ecology Letters. 2014;17:637–49.Google Scholar
Leff, JW, Jones, SE, Prober, SM, Barberán, A, Borer, ET, Firn, JL, et al. Consistent responses of soil microbial communities to elevated nutrient inputs in grasslands across the globe. Proceedings of the National Academy of Sciences. 2015;112:10,967–72.Google Scholar
Bullock, JM, Dhanjal-Adams, KL, Milne, A, Oliver, TH, Todman, LC, Whitmore, AP, et al. Resilience and food security: rethinking an ecological concept. Journal of Ecology. 2017;105:880–4.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;526:574–7.Google Scholar
Bardgett, RD, McAlister, E. The measurement of soil fungal: bacterial biomass ratios as an indicator of ecosystem self-regulation in temperate meadow grasslands. Biology and Fertility of Soils. 1999;29:282–90.Google Scholar
De Vries, FT, Hoffland, E, van Eekeren, N, Brussaard, L, Bloem, J. Fungal/bacterial ratios in grasslands with contrasting nitrogen management. Soil Biology and Biochemistry. 2006;38:2092–103.Google Scholar
Six, J, Frey, SD, Thiet, RK, Batten, KM. Bacterial and fungal contributions to carbon sequestration in agroecosystems. Soil Science Society of America Journal. 2006;70:555–69.Google Scholar
Gordon, H, Haygarth, PM, Bardgett, RD. Drying and rewetting effects on soil microbial community composition and nutrient leaching. Soil Biology and Biochemistry. 2008;40:302–11.Google Scholar
Rooney, N, McCann, K, Gellner, G, Moore, JC. Structural asymmetry and the stability of diverse food webs. Nature. 2006;442:265–9.Google Scholar
Fierer, N, Bradford, MA, Jackson, RB. Toward an ecological classification of soil bacteria. Ecology. 2007;88:1354–64.Google Scholar
Isbell, F, Adler, PR, Eisenhauer, N, Fornara, D, Kimmel, K, Kremen, C, et al. Benefits of increasing plant diversity in sustainable agroecosystems. Journal of Ecology. 2017;105:871–9.Google Scholar
De Deyn, GB, Quirk, H, Oakley, S, Ostle, NJ, Bardgett, RD. Increased plant carbon translocation linked to overyielding in grassland species mixtures. PLoS ONE. 2012;7:e45926.Google Scholar
Fornara, DA, Tilman, D. Plant functional composition influences rates of soil carbon and nitrogen accumulation. Journal of Ecology. 2008;96:314–22.Google Scholar
Lange, M, Eisenhauer, N, Sierra, CA, Bessler, H, Engels, C, Griffiths, RI, et al. Plant diversity increases soil microbial activity and soil carbon storage. Nature Communications. 2015;6:6707.Google Scholar
Gould, IJ, Quinton, JN, Weigelt, A, De Deyn, GB, Bardgett, RD. Plant diversity and root traits benefit physical properties key to soil function in grasslands. Ecology Letters. 2016;19:1140–9.Google Scholar
De Deyn, GB, Quirk, H, Yi, Z, Oakley, S, Ostle, NJ, Bardgett, RD. Vegetation composition promotes carbon and nitrogen storage in model grassland communities of contrasting soil fertility. Journal of Ecology. 2009;97:864–75.Google Scholar
Hiiesalu, I, Pärtel, M, Davison, J, Gerhold, P, Metsis, M, Moora, M et al. Species richness of arbuscular mycorrhizal fungi: associations with grassland plant richness and biomass. New Phytologist. 2014;203:233–44.Google Scholar
van der Heijden, MGA. Mycorrhizal fungi reduce nutrient loss from model grassland ecosystems. Ecology. 2010;91:1163–71.Google Scholar
Martínez-García, LB, De Deyn, GB, Pugnaire, FI, Kothamasi, D, van der Heijden, MGA. Symbiotic soil fungi enhance ecosystem resilience to climate change. Global Change Biology. 2017;23:5228–36.Google Scholar
van der Heijden, MG, Klironomos, JN, Ursic, M, Moutoglis, P. Mycorrhizal fungal diversity determines plant biodiversity, ecosystem variability and productivity. Nature. 1998;396:6972.Google Scholar
Thakur, MP, Tilman, D, Purschke, O, Ciobanu, M, Cowles, J, Isbell, F, et al. Climate warming promotes species diversity, but with greater taxonomic redundancy, in complex environments. Science Advances. 2017;3:e1700866.Google Scholar
Bardgett, RD. Plant trait-based approaches for interrogating belowground function. Biology and Environment: Proceedings of the Royal Irish Academy. 2017;117:113.Google Scholar
McSherry, ME, Ritchie, ME. Effects of grazing on grassland soil carbon: a global review. Global Change Biology. 2013;19:1347–57.Google Scholar
Klein, JA, Harte, J, Zhao, X-Q. Experimental warming causes large and rapid species loss, dampened by simulated grazing, on the Tibetan Plateau. Ecology Letters. 2004;7:1170–9.Google Scholar
Olofsson, J, Oksanen, L, Callaghan, T, Hulme, PE, Oksanen, T, Suominen, O. Herbivores inhibit climate‐driven shrub expansion on the tundra. Global Change Biology. 2009;15:2681–93.Google Scholar
Wolf, B, Zheng, X, Brüggemann, N, Chen, W, Dannenmann, M, Han, X, et al. Grazing-induced reduction of natural nitrous oxide release from continental steppe. Nature. 2010;464:881–4.Google Scholar
Paz-Ferreiro, J, Medina-Roldán, E, Ostle, NJ, McNamara, NP, Bardgett, RD. Grazing increases the temperature sensitivity of soil organic matter decomposition in a temperate grassland. Environmental Research Letters. 2012;7:014027.Google Scholar
Koerner, SE, Collins, SL. Interactive effects of grazing, drought, and fire on grassland plant communities in North America and South Africa. Ecology. 2014;95:98109.Google Scholar

13.8 References

Ahuja, I, de Vos, RCH, Bones, AM, Hall, RD. Plant molecular stress responses face climate change. Trends in Plant Science. 2010;15(12):664–74.Google Scholar
Franks, SJ, Hoffman, AA. Genetics of climate change adaptation. Annual Review of Genetics. 2012;46:185208.Google Scholar
Visser, ME. Keeping up with a warming world; assessing the rate of adaptation to climate change. Proceedings of the Royal Society B: Biological Sciences. 2008;275(1635):649–59.Google Scholar
Shaw, RG, Etterson, JR. Rapid climate change and the rate of adaptation: insight from experimental quantitative genetics. New Phytologist. 2012;195(4):752–65.Google Scholar
Kinnison, MT, Hairston, NG. Eco-evolutionary conservation biology: contemporary evolution and the dynamics of persistence. Functional Ecology. 2007;21(3):444–54.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
Berg, MP, Kiers, ET, Driessen, G, Van Der Heijden, M, Kooi, BW, Kuenen, F, et al. Adapt or disperse: understanding species persistence in a changing world. Global Change Biology. 2010;16(2):587–98.Google Scholar
Bossdorf, O, Richards, CL, Pigliucci, M. Epigenetics for ecologists. Ecology Letters. 2008;11(2):106–15.Google Scholar
Antonovics, J. Evolution in closely adjacent plant populations X: long-term persistence of prereproductive isolation at a mine boundary. Heredity. 2006;97(1):33–7.Google Scholar
Heap, IM. The occurrence of herbicide-resistant weeds worldwide. Pesticide Science. 1997;51(3):235–43.Google Scholar
Axelrod, DI. Rise of the grassland biome, central North America. The Botanical Review. 1985;51(2):163201.Google Scholar
McMillan, C. The role of ecotypic variation in the distribution of the central grassland of North America. Ecological Monographs. 1959;29(4):285308.Google Scholar
McMillan, C. Nature of the plant community. III. Flowering behavior within two grassland communities under reciprocal transplanting. American Journal of Botany. 1957;44(2):144–53.Google Scholar
Gustafson, DJ, Gibson, DJ, Nickrent, DL. Competitive relationships of Andropogon gerardii (Big Bluestem) from remnant and restored native populations and select cultivated varieties. Functional Ecology. 2004;18(3):451–7.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
Gustafson, DJ, Gibson, DJ, Nickrent, DL. Random amplified polymorphic DNA variation among remnant big bluestem (Andropogon gerardii Vitman) populations from Arkansas’ Grand Prairie. Molecular Ecology. 1999;8(10):1693–701.Google Scholar
Vogel, KP, Hopkins, AA, Moore, KJ, Johnson, KD, Carlson, IT. Genetic variation among Canada Wildrye accessions from Midwest USA remnant prairies for biomass yield and other traits. Crop Science. 2006;46(6):2348–53.Google Scholar
Fu, Y-B, Phan, AT, Coulman, B, Richards, KW. Genetic diversity in natural populations and corresponding seed collections of little bluestem as revealed by AFLP markers. Crop Science. 2004;44(6):2254–60.Google Scholar
Swanson, DL, Palmer, JS. Spring migration phenology of birds in the Northern Prairie region is correlated with local climate change. Journal of Field Ornithology. 2009;80(4):351–63.Google Scholar
Jones, T, Cresswell, W. The phenology mismatch hypothesis: are declines of migrant birds linked to uneven global climate change? Journal of Animal Ecology. 2010;79(1):98108.Google Scholar
Monteith, KL, Bleich, VC, Stephenson, TR, Pierce, BM, Conner, MM, Klaver, RW, et al. Timing of seasonal migration in mule deer: effects of climate, plant phenology, and life-history characteristics. Ecosphere. 2011;2(4):134.Google Scholar
Turner, BM. Epigenetic responses to environmental change and their evolutionary implications. Philosophical Transactions of the Royal Society B: Biological Sciences. 2009;364(1534):3403–18.Google Scholar
Kim, J-M, To, TK, Matsui, A, Tanoi, K, Kobayashi, NI, Matsuda, F, et al. Acetate-mediated novel survival strategy against drought in plants. Nature Plants. 2017;3:17097.Google Scholar
Bradshaw, WE, Holzafel, CM. Light, time, and the physiology of biotic response to rapid climate change in animals. Annual Review of Physiology. 2010;72:147–66.Google Scholar
Hoffmann, I. Climate change and the characterization, breeding and conservation of animal genetic resources. Animal Genetics. 2010;41:3246.Google Scholar
Bradshaw, WE, Holzapfel, CM. Genetic response to rapid climate change: it’s seasonal timing that matters. Molecular Ecology. 2008;17(1):157–66.Google Scholar
Etterson, JR, Shaw, RG. Constraint to adaptive evolution in response to global warming. Science. 2001;294(5540):151–4.Google Scholar
Hoffmann, AA, Sgro, CM. Climate change and evolutionary adaptation. Nature. 2011;470(7335):479–85.Google Scholar
Gienapp, P, Teplitsky, C, Alho, JS, Mills, JA, Merilä, J. Climate change and evolution: disentangling environmental and genetic responses. Molecular Ecology. 2008;17(1):167–78.Google Scholar
Anderson, JT, Inouye, DW, McKinney, AM, Colautti, RI, Mitchell-Olds, T. Phenotypic plasticity and adaptive evolution contribute to advancing flowering phenology in response to climate change. Proceedings of the Royal Society B: Biological Sciences. 2012;279(1743):3843–52.Google Scholar
Frei, ER, Ghazoul, J, Matter, P, Heggli, M, Pluess, AR. Plant population differentiation and climate change: responses of grassland species along an elevational gradient. Global Change Biology. 2014;20(2):441–55.Google Scholar
Dunnell, KL, Travers, SE. Shifts in the flowering phenology of the northern Great Plains: patterns over 100 years. American Journal of Botany. 2011;98(6):935–45.Google Scholar
Hyatt, LA, Rosenberg, MS, Howard, TG, Bole, G, Fang, W, Anastasia, J, et al. The distance dependence prediction of the Janzen–Connell hypothesis: a meta-analysis. Oikos. 2003;103(3):590602.Google Scholar
Diacon-Bolli, JC, Edwards, PJ, Bugmann, H, Scheidegger, C, Wagner, HH. Quantification of plant dispersal ability within and beyond a calcareous grassland. Journal of Vegetation Science. 2013;24(6):1010–9.Google Scholar
Soons, MB, Heil, GW, Nathan, R, Katul, GG. Determinants of long-distance seed dispersal by wind in grasslands. Ecology. 2004;85(11):3056–68.Google Scholar
Clobert, J, Le Galliard, J-F, Cote, J, Meylan, S, Massot, M. Informed dispersal, heterogeneity in animal dispersal syndromes and the dynamics of spatially structured populations. Ecology Letters. 2009;12(3):197209.Google Scholar
Vittoz, P, Engler, R. Seed dispersal distances: a typology based on dispersal modes and plant traits. Botanica Helvetica. 2007;117(2):109–24.Google Scholar
Rosas, CA, Engle, DM, Shaw, JH, Palmer, MW. Seed dispersal by Bison bison in a tallgrass prairie. Journal of Vegetation Science. 2008;19(6):769–78.Google Scholar
Couvreur, M, Bart, C, Verheyen, V, Hermy, M. Large herbivores as mobile links between isolated nature reserves through adhesive seed dispersal. International Association of Vegetation Science. 2004;7(2):229–36.Google Scholar
Myers, JA, Vellend, M, Gardescu, S, Marks, PL. Seed dispersal by white-tailed deer: implications for long-distance dispersal, invasion, and migration of plants in eastern North America. Oecologia. 2004;139(1):3544.Google Scholar
Boeye, J, Travis, JMJ, Stoks, R, Bonte, D. More rapid climate change promotes evolutionary rescue through selection for increased dispersal distance. Evolutionary Applications. 2013;6(2):353–64.Google Scholar
Baguette, M, Blanchet, S, Legrand, D, Stevens, VM, Turlure, C. Individual dispersal, landscape connectivity and ecological networks. Biological Reviews. 2013;88(2):310–26.Google Scholar
Lavergne, S, Mouquet, N, Thuiller, W, Ronce, O. Biodiversity and climate change: integrating evolutionary and ecological responses of species and communities. Annual Review of Ecology, Evolution, and Systematics. 2010;41(1):321–50.Google Scholar
Travis, JMJ, Delgado, M, Bocedi, G, Baguette, M, Bartoń, K, Bonte, D, et al. Dispersal and species’ responses to climate change. Oikos. 2013;122(11):1532–40.Google Scholar
Cormont, A, Malinowska, AH, Kostenko, O, Radchuk, V, Hemerik, L, WallisDeVries, MF, et al. Effect of local weather on butterfly flight behaviour, movement, and colonization: significance for dispersal under climate change. Biodiversity and Conservation. 2011;20(3):483503.Google Scholar
Bönsel, AB, Sonneck, A-G. Habitat use and dispersal characteristic by Stethophyma grossum: the role of habitat isolation and stable habitat conditions towards low dispersal. Journal of Insect Conservation. 2011;15(3):455–63.Google Scholar
Pakeman, RJ. Plant migration rates and seed dispersal mechanisms. Journal of Biogeography. 2001;28(6):795800.Google Scholar
Graham, RW, Lundelius, EL, Graham, MA, Schroeder, EK, Toomey, RS, Anderson, E, et al. Spatial response of mammals to Late Quaternary environmental fluctuations. Science. 1996;272(5268):1601–6.Google Scholar
Clark, JS, Fastie, C, Hurtt, G, Jackson, ST, Johnson, C, King, GA, et al. Reid’s paradox of rapid plant migration dispersal theory and interpretation of paleoecological records. BioScience. 1998;48(1):1324.Google Scholar
Trakhtenbrot, A, Nathan, R, Perry, G, Richardson, DM. The importance of long-distance dispersal in biodiversity conservation. Diversity and Distributions. 2005;11(2):173–81.Google Scholar
Hampe, A. Plants on the move: the role of seed dispersal and initial population establishment for climate-driven range expansions. Acta Oecologica. 2011;37(6):666–73.Google Scholar
Jump, AS, Peñuelas, J. Running to stand still: adaptation and the response of plants to rapid climate change. Ecology Letters. 2005;8(9):1010–20.Google Scholar
Parmesan, C, Yohe, G. A globally coherent fingerprint of climate change impacts across natural systems. Nature. 2003;421(6918):3742.Google Scholar
Neilson, RP, Pitelka, LF, Solomon, AM, Nathan, R, Midgley, GF, Fragoso, JMV, et al. Forecasting regional to global plant migration in response to climate change. BioScience. 2005;55(9):749–59.Google Scholar
Brown, DA, Gersmehl, PJ. Migration models for grasses in the American midcontinent. Annals of the Association of American Geographers. 1985;75(3):383–94.Google Scholar
van Dorp, D, Schippers, P, van Groenendael, JM. Migration rates of grassland plants along corridors in fragmented landscapes assessed with a cellular automation model. Landscape Ecology. 1997;12(1):3950.Google Scholar
Collingham, YC, Huntley, B. Impacts of habitat fragmentation and patch size upon migration rates. Ecological Applications. 2000;10(1):131–44.Google Scholar
Sorte, CJB. Predicting persistence in a changing climate: flow direction and limitations to redistribution. Oikos. 2013;122(2):161–70.Google Scholar
Johnson, WC, Millett, BV, Gilmanov, T, Voldseth, RA, Guntenspergen, GR, Naugle, DE. Vulnerability of northern prairie wetlands to climate change. BioScience. 2005;55(10):863–72.Google Scholar
Zimmermann, NE, Kienast, F. Predictive mapping of alpine grasslands in Switzerland: species versus community approach. Journal of Vegetation Science. 1999;10(4):469–82.Google Scholar
Erasmus, BFN, Van Jaarsveld, AS, Chown, SL, Kshatriya, M, Wessels, KJ. Vulnerability of South African animal taxa to climate change. Global Change Biology. 2002;8(7):679–93.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
Van der Putten, WH, Macel, M, Visser, ME. Predicting species distribution and abundance responses to climate change: why it is essential to include biotic interactions across trophic levels. Philosophical Transactions of the Royal Society B: Biological Sciences. 2010;365(1549):2025–34.Google Scholar
Gilman, SE, Urban, MC, Tewksbury, J, Gilchrist, GW, Holt, RD. A framework for community interactions under climate change. Trends in Ecology & Evolution. 2010;25(6):325–31.Google Scholar
HilleRisLambers, J, Harsch, MA, Ettinger, AK, Ford, KR, Theobald, EJ. How will biotic interactions influence climate change-induced range shifts? Annals of the New York Academy of Sciences. 2013;1297(1):112–25.Google Scholar
Bertin, RI. Plant phenology and distribution in relation to recent climate change. The Journal of the Torrey Botanical Society. 2008;135(1):126–46.Google Scholar
Hulme, PE. Adapting to climate change: is there scope for ecological management in the face of a global threat? Journal of Applied Ecology. 2005;42(5):784–94.Google Scholar
Higgins, PAT, Harte, J. Biophysical and biogeochemical responses to climate change depend on dispersal and migration. BioScience. 2006;56(5):407–17.Google Scholar
Prober, SM, Hilbert, DW, Ferrier, S, Dunlop, M, Gobbett, D. Combining community-level spatial modelling and expert knowledge to inform climate adaptation in temperate grassy eucalypt woodlands and related grasslands. Biodiversity and Conservation. 2012;21(7):1627–50.Google Scholar
Bliege Bird, R, Codding, BF, Kauhanen, PG, Bird, DW. Aboriginal hunting buffers climate-driven fire-size variability in Australia’s spinifex grasslands. Proceedings of the National Academy of Sciences of the USA. 2012;109(26):10,287–92.Google Scholar
Malcolm, JR, Markham, A, Neilson, RP, Garaci, M. Estimated migration rates under scenarios of global climate change. Journal of Biogeography. 2002;29(7):835–49.Google Scholar
Malcolm, JR, Liu, C, Neilson, RP, Hansen, L, Hannah, LEE. Global warming and extinctions of endemic species from biodiversity hotspots. Conservation Biology. 2006;20(2):538–48.Google Scholar
Pacifici, M, Foden, WB, Visconti, P, Watson, JEM, Butchart, SHM, Kovacs, KM, et al. Assessing species vulnerability to climate change. Nature Climate Change. 2015;5(3):215–24.Google Scholar
Davis, MB, Shaw, RG, Etterson, JR. Evolutionary responses to changing climate. Ecology. 2005;86(7):1704–14.Google Scholar
Corlett, RT, Westcott, DA. Will plant movements keep up with climate change? Trends in Ecology & Evolution. 2013;28(8):482–8.Google Scholar
Huntley, B, Barnard, P. Potential impacts of climatic change on southern African birds of fynbos and grassland biodiversity hotspots. Diversity and Distributions. 2012;18(8):769–81.Google Scholar
Marini, , Barbet-Massin, M, Lopes, LE, Jiguet, F. Predicted climate-driven bird distribution changes and forecasted conservation conflicts in a neotropical savanna. Conservation Biology. 2009;23(6):1558–67.Google Scholar
Ouled Belgacem, A, Louhaichi, M. The vulnerability of native rangeland plant species to global climate change in the west Asia and north African regions. Climatic Change. 2013;119(2):451–63.Google Scholar
Mawdsley, JR, O’Malley, R, Ojima, DS. A review of climate-change adaptation strategies for wildlife management and biodiversity conservation. Conservation Biology. 2009;23(5):1080–9.Google Scholar
Heller, NE, Zavaleta, ES. Biodiversity management in the face of climate change: a review of 22 years of recommendations. Biological Conservation. 2009;142(1):1432.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
Rowland, E, Davison, J, Graumlich, L. Approaches to evaluating climate change impacts on species: a guide to initiating the adaptation planning process. Environmental Management. 2011;47(3):322–37.Google Scholar

14.7 References

Still, CJ, Berry, JA, Collatz, GJ, DeFries, RS. Global distribution of C3 and C4 vegetation: carbon cycle implications. Global Biogeochemical Cycles. 2003;17(1):6-1–6-14.Google Scholar
Del Grosso, S, Parton, W, Stohlgren, T, Zheng, D, Bachelet, D, Prince, S, et al. Global potential net primary production predicted from vegetation class, precipitation, and temperature. Ecology. 2008;89(8):2117–26.Google Scholar
Staver, AC, Bond, WJ, Stock, WD, Van Rensburg, SJ, Waldram, MS. Browsing and fire interact to suppress tree density in an African savanna. Ecological Applications. 2009;19(7):1909–19.Google Scholar
Lehmann, CER, Archibald, SA, Hoffmann, WA, Bond, WJ. Deciphering the distribution of the savanna biome. New Phytologist. 2011;191(1):197209.Google Scholar
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
Teeri, JA, Stowe, LG. Climatic patterns and the distribution of C4 grasses in North America. Oecologia. 1976;23(1):112.Google Scholar
Chazdon, RL. Ecological aspects of the distribution of C4 grasses in selected habitats of Costa Rica. Biotropica. 1978;10(4):265–9.Google Scholar
Rundel, PW. The ecological distribution of C4 and C3 grasses in the Hawaiian Islands. Oecologia. 1980;45(3):354–9.Google Scholar
Hattersley, PW. The distribution of C3 and C4 grasses in Australia in relation to climate. Oecologia. 1983;57(1–2):113–28.Google Scholar
Ehleringer, JR, Cerling, TE, Helliker, BR. C4 photosynthesis, atmospheric CO2, and climate. Oecologia. 1997;112(3):285–99.Google Scholar
Edwards, EJ, Still, CJ. Climate, phylogeny and the ecological distribution of C4 grasses. Ecology Letters. 2008;11(3):266–76.Google Scholar
Christin, P-A, Osborne, CP, Sage, RF, Arakaki, M, Edwards, EJ. C4 eudicots are not younger than C4 monocots. Journal of Experimental Botany. 2011;62(9):3171–81.Google Scholar
Pagani, M, Zachos, JC, Freeman, KH, Tipple, B, Bohaty, S. Marked decline in atmospheric carbon dioxide concentrations during the Paleogene. Science. 2005;309(5734):600–3.Google Scholar
Black, CC, Chen, TM, Brown, RH. Biochemical basis for plant competition. Weed Science. 1969;17(3):338–44.Google Scholar
Ehleringer, JR. Implications of quantum yield differences on the distributions of C3 and C4 grasses. Oecologia. 1978;31(3):255–67.Google Scholar
Atkinson, RRL, Mockford, EJ, Bennett, C, Christin, P-A, Spriggs, EL, Freckleton, RP, et al. C4 photosynthesis boosts growth by altering physiology, allocation and size. Nature Plants. 2016;2(5):16038.Google Scholar
Sage, RF. The evolution of C4 photosynthesis. New Phytologist. 2004;161(2):341–70.Google Scholar
Collatz, GJ, Berry, JA, Clark, JS. Effects of climate and atmospheric CO2 partial pressure on the global distribution of C4 grasses: present, past, and future. Oecologia. 1998;114(4):441–54.Google Scholar
Edwards, EJ, Osborne, CP, Strömberg, CAE, Smith, SA, Consortium, CG. The origins of C4 grasslands: integrating evolutionary and ecosystem science. Science. 2010;328(5978):587–91.Google Scholar
Brummitt, RK, Pando, F, Hollis, S, Brummitt, NA. Plant taxonomic database standards no. 2. World geographical scheme for recording plant distributions, 2nd edn. Pittsburgh, PA: Published for the International Working Group on Taxonomic Databases For Plant Sciences (TDWG) by the Hunt Institute for Botanical Documentation, Carnegie Mellon University; 2001.Google Scholar
Osborne, CP, Salomaa, A, Kluyver, TA, Visser, V, Kellogg, EA, Morrone, O, et al. A global database of C4 photosynthesis in grasses. New Phytologist. 2014;204(3):441–6.Google Scholar
Dixon, AP, Faber-Langendoen, D, Josse, C, Morrison, J, Loucks, CJ. Distribution mapping of world grassland types. Journal of Biogeography. 2014;41(11):2003–19.Google Scholar
Osborne, CP. Atmosphere, ecology and evolution: what drove the Miocene expansion of C4 grasslands? Journal of Ecology. 2008;96(1):3545.Google Scholar
Edwards, EJ, Smith, SA. Phylogenetic analyses reveal the shady history of C4 grasses. Proceedings of the National Academy of Sciences of the USA. 2010;107(6):2532–7.Google Scholar
Beerling, DJ, Osborne, CP. The origin of the savanna biome. Global Change Biology. 2006;12(11):2023–31.Google Scholar
Spriggs, EL, Christin, P-A, Edwards, EJ. C4 photosynthesis promoted species diversification during the Miocene grassland expansion. PLoS ONE. 2014;9(5):e97722.Google Scholar
Preston, JC, Sandve, SR. Adaptation to seasonality and the winter freeze. Frontiers in Plant Science. 2013;4:167.Google Scholar
McKeown, M, Schubert, M, Marcussen, T, Fjellheim, S, Preston, JC. Evidence for an early origin of vernalization responsiveness in temperate Pooideae grasses. Plant Physiology. 2016;172(1):416–26.Google Scholar
Visser, V, Clayton, WD, Simpson, DA, Freckleton, RP, Osborne, CP. Mechanisms driving an unusual latitudinal diversity gradient for grasses. Global Ecology and Biogeography. 2014;23(1):6175.Google Scholar
Paruelo, JM, Lauenroth, WK. Relative abundance of plant functional types in grasslands and shrublands of North America. Ecological Applications. 1996;6(4):1212–24.Google Scholar
Smith, MD, Knapp, AK. Dominant species maintain ecosystem function with non‐random species loss. Ecology Letters. 2003;6(6):509–17.Google Scholar
Epstein, HE, Lauenroth, WK, Burke, IC, Coffin, DP. Productivity patterns of C3 and C4 functional types in the US Great Plains. Ecology. 1997;78(3):722–31.Google Scholar
Winslow, JC, Hunt, ER Jr, Piper, SC. The influence of seasonal water availability on global C3 versus C4 grassland biomass and its implications for climate change research. Ecological Modelling. 2003;163(1–2):153–73.Google Scholar
Murphy, BP, Bowman, DMJS. The interdependence of fire, grass, kangaroos and Australian Aborigines: a case study from central Arnhem Land, northern Australia. Journal of Biogeography. 2007;34(2):237–50.Google Scholar
Fischer, V, Joseph, C, Tieszen, LL, Schimel, DS. Climate controls on C3 vs. C4 productivity in North American grasslands from carbon isotope composition of soil organic matter. Global Change Biology. 2008;14(5):1141–55.Google Scholar
Bremond, L, Boom, A, Favier, C. Neotropical C3/C4 grass distributions – present, past and future. Global Change Biology. 2012;18(7):2324–34.Google Scholar
Auerswald, K, Wittmer, MHOM, Bai, Y, Yang, H, Taube, F, Susenbeth, A, et al. C4 abundance in an Inner Mongolia grassland system is driven by temperature–moisture interaction, not grazing pressure. Basic and Applied Ecology. 2012;13(1):6775.Google Scholar
Griffith, DM, Anderson, TM, Osborne, CP, Strömberg, CAE, Forrestel, EJ, Still, CJ. Biogeographically distinct controls on C3 and C4 grass distributions: merging community and physiological ecology. Global Ecology and Biogeography. 2015;24(3):304–13.Google Scholar
Hatch, M, editor. Chemical energy costs for CO2 fixation by plants with differing photosynthetic pathways. Prediction and measurement of photosynthetic productivity. Trebon, Czechoslovakia: PUDOC; 1970.Google Scholar
Ehleringer, J, Björkman, O. Quantum yields for CO2 uptake in C3 and C4 plants: dependence on temperature, CO2, and O2 concentration. Plant Physiology. 1977;59(1):8690.Google Scholar
Intergovernmental Panel on Climate Change. Climate change 2014: mitigation of climate change. Cambridge: Cambridge University Press; 2015.Google Scholar
Lehmann, CER, Anderson, TM, Sankaran, M, Higgins, SI, Archibald, S, Hoffmann, WA, et al. Savanna vegetation–fire–climate relationships differ among continents. Science. 2014;343(6170):548–52.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
Pearson, RG, Dawson, TP. Predicting the impacts of climate change on the distribution of species: are bioclimate envelope models useful? Global Ecology and Biogeography. 2003;12(5):361–71.Google Scholar
Taylor, KE, Stouffer, RJ, Meehl, GA. An overview of CMIP5 and the experiment design. Bulletin of the American Meteorological Society. 2012;93(4):485–98.Google Scholar
Owensby, CE, Coyne, PI, Ham, JM, Auen, LM, Knapp, AK. Biomass production in a tallgrass prairie ecosystem exposed to ambient and elevated CO2. Ecological Applications. 1993;3(4):644–53.Google Scholar
Polley, HW, Johnson, HB, Derner, JD. Increasing CO2 from subambient to superambient concentrations alters species composition and increases above‐ground biomass in a C3/C4 grassland. New Phytologist. 2003;160(2):319–27.Google Scholar
Morgan, JA, Milchunas, DG, LeCain, DR, West, M, Mosier, AR. Carbon dioxide enrichment alters plant community structure and accelerates shrub growth in the shortgrass steppe. Proceedings of the National Academy of Sciences of the USA. 2007;104(37):14,724–9.Google Scholar
Morgan, JA, LeCain, DR, Pendall, E, Blumenthal, DM, Kimball, BA, Carrillo, Y, et al. C4 grasses prosper as carbon dioxide eliminates desiccation in warmed semi-arid grassland. Nature. 2011;476(7359):202.Google Scholar
Wittmer, MHOM, Auerswald, K, Bai, Y, Schaeufele, R, Schnyder, H. Changes in the abundance of C3/C4 species of Inner Mongolia grassland: evidence from isotopic composition of soil and vegetation. Global Change Biology. 2010;16(2):605–16.Google Scholar
Griffith, DM, Cotton, JM, Powell, RL, Sheldon, ND, Still, CJ. Multi‐century stasis in C3 and C4 grass distributions across the contiguous United States since the industrial revolution. Journal of Biogeography. 2017;44(11):2564–74.Google Scholar
Higgins, SI, Scheiter, S. Atmospheric CO2 forces abrupt vegetation shifts locally, but not globally. Nature. 2012;488(7410):209.Google Scholar
Knapp, AK, Briggs, JM, Collins, SL, Archer, SR, Bret‐Harte, MS, Ewers, BE, et al. Shrub encroachment in North American grasslands: shifts in growth form dominance rapidly alters control of ecosystem carbon inputs. Global Change Biology. 2008;14(3):615–23.Google Scholar
Van Auken, O. Causes and consequences of woody plant encroachment into western North American grasslands. Journal of Environmental Management. 2009;90(10):2931–42.Google Scholar
Silva, JF, Zambrano, A, Fariñas, MR. Increase in the woody component of seasonal savannas under different fire regimes in Calabozo, Venezuela. Journal of Biogeography. 2001;28(8):977–83.Google Scholar
Roques, KG, O’Connor, TG, Watkinson, AR. Dynamics of shrub encroachment in an African savanna: relative influences of fire, herbivory, rainfall and density dependence. Journal of Applied Ecology. 2001;38(2):268–80.Google Scholar
Sankaran, M. Fire, grazing and the dynamics of tall-grass savannas in the Kalakad-Mundanthurai Tiger Reserve, South India. Conservation and Society. 2005;3(1):425.Google Scholar
Misra, R. Indian savannas. In: Bourlière, F, editor. Tropical savannas, ecosystems of the world. Vol. 13. Amsterdam, The Netherlands: Elsevier; 1983. pp. 151–66.Google Scholar
Peng, H-Y, Li, X-Y, Li, G-Y, Zhang, Z-H, Zhang, S-Y, Li, L, et al. Shrub encroachment with increasing anthropogenic disturbance in the semiarid Inner Mongolian grasslands of China. Catena. 2013;109:3948.Google Scholar
Wigley, BJ, Bond, WJ, Hoffman, MT. Thicket expansion in a South African savanna under divergent land use: local vs. global drivers? Global Change Biology. 2010;16(3):964–76.Google Scholar
Kgope, BS, Bond, WJ, Midgley, GF. Growth responses of African savanna trees implicate atmospheric [CO2] as a driver of past and current changes in savanna tree cover. Austral Ecology. 2010;35(4):451–63.Google Scholar
Bond, WJ, Midgley, GF. A proposed CO2‐controlled mechanism of woody plant invasion in grasslands and savannas. Global Change Biology. 2000;6(8):865–9.Google Scholar
Scholes, RJ, Archer, SR. Tree–grass interactions in savannas. Annual Review of Ecology and Systematics. 1997;28(1):517–44.Google Scholar
Van Auken, OW. Shrub invasions of North American semiarid grasslands. Annual Review of Ecology and Systematics. 2000;31(1):197215.Google Scholar
Klink, CA, Joly, CA. Identification and distribution of C3 and C4 grasses in open and shaded habitats in São Paulo State, Brazil. Biotropica. 1989;21(1):30–4.Google Scholar
Bond, WJ, Midgley, GF, Woodward, FI. The importance of low atmospheric CO2 and fire in promoting the spread of grasslands and savannas. Global Change Biology. 2003;9(7):973–82.Google Scholar
Bond, WJ, Keeley, JE. Fire as a global ‘herbivore’: the ecology and evolution of flammable ecosystems. Trends in Ecology & Evolution. 2005;20(7):387–94.Google Scholar
Scheiter, S, Higgins, SI, Osborne, CP, Bradshaw, C, Lunt, D, Ripley, BS, et al. Fire and fire‐adapted vegetation promoted C4 expansion in the Late Miocene. New Phytologist. 2012;195(3):653–66.Google Scholar
Tilman, D, Wedin, D. Dynamics of nitrogen competition between successional grasses. Ecology. 1991;72(3):1038–49.Google Scholar
Keeley, JE, Pausas, JG, Rundel, PW, Bond, WJ, Bradstock, RA. Fire as an evolutionary pressure shaping plant traits. Trends in Plant Science. 2011;16(8):406–11.Google Scholar
Forrestel, EJ, Donoghue, MJ, Smith, MD. Convergent phylogenetic and functional responses to altered fire regimes in mesic savanna grasslands of North America and South Africa. New Phytologist. 2014;203(3):1000–11.Google Scholar
Bond, WJ, Woodward, FI, Midgley, GF. The global distribution of ecosystems in a world without fire. New Phytologist. 2005;165(2):525–38.Google Scholar
Keeley, JE, Rundel, PW. Fire and the Miocene expansion of C4 grasslands. Ecology Letters. 2005;8(7):683–90.Google Scholar
Hessl, AE, Ariya, U, Brown, P, Byambasuren, O, Green, T, Jacoby, G, et al. Reconstructing fire history in central Mongolia from tree-rings. International Journal of Wildland Fire. 2012;21(1):8692.Google Scholar
Boutton, TW, Cameron, GN, Smith, BN. Insect herbivory on C3 and C4 grasses. Oecologia. 1978;36(1):2132.Google Scholar
McNaughton, S. Ecology of a grazing ecosystem: the Serengeti. Ecological Monographs. 1985;55(3):259–94.Google Scholar
Stebbins, GL. Coevolution of grasses and herbivores. Annals of the Missouri Botanical Garden. 1981;68(1):7586.Google Scholar
Coughenour, MB. Graminoid responses to grazing by large herbivores: adaptations, exaptations, and interacting processes. Annals of the Missouri Botanical Garden. 1985;72(4):852–63.Google Scholar
Forrestel, EJ, Donoghue, MJ, Smith, MD. Functional differences between dominant grasses drive divergent responses to large herbivore loss in mesic savanna grasslands of North America and South Africa. Journal of Ecology. 2015;103(3):714–24.Google Scholar
Sage, RF, Christin, P-A, Edwards, EJ. The C4 plant lineages of planet Earth. Journal of Experimental Botany. 2011;62(9):3155–69.Google Scholar
Grass Phylogeny Working Group II. New grass phylogeny resolves deep evolutionary relationships and discovers C4 origins. New Phytologist. 2012;193(2):304–12.Google Scholar
Christin, P-A, Osborne, CP, Chatelet, DS, Columbus, JT, Besnard, G, Hodkinson, TR, et al. Anatomical enablers and the evolution of C4 photosynthesis in grasses. Proceedings of the National Academy of Sciences of the USA. 2013;110(4):1381–6.Google Scholar
Visser, V, Woodward, FI, Freckleton, RP, Osborne, CP. Environmental factors determining the phylogenetic structure of C4 grass communities. Journal of Biogeography. 2012;39(2):232–46.Google Scholar
Clayton, WD, Govaerts, R, Harman, KT, Williamson, H, Vorontsova, M. World checklist of Poaceae. Facilitated by the Royal Botanic Gardens, Kew. Published on the Internet 2011. Available from: http://apps.kew.org/wcsp/Google Scholar
Sage, RF. Photosynthesis: mining grasses for a better Rubisco. Nature Plants. 2016;2(12):16192.Google Scholar
Christin, P-A, Osborne, CP. The recurrent assembly of C4 photosynthesis, an evolutionary tale. Photosynthesis Research. 2013;117(1–3):163–75.Google Scholar
Hattersley, PW, editor. C4 photosynthetic pathway variation in grasses (Poaceae): its significance for arid and semi-arid lands. In: Desertified grasslands: their biology and management (Linnean Society Symposium Series No. 13). London: Academic Press; 1992.Google Scholar
Taub, DR. Climate and the US distribution of C 4 grass subfamilies and decarboxylation variants of C4 photosynthesis. American Journal of Botany. 2000;87(8):1211–5.Google Scholar
Cabido, M, Pons, E, Cantero, JJ, Lewis, JP, Anton, A. Photosynthetic pathway variation among C4 grasses along a precipitation gradient in Argentina. Journal of Biogeography. 2008;35(1):131–40.Google Scholar
Hatch, MD. C4 photosynthesis: a unique elend of modified biochemistry, anatomy and ultrastructure. Biochimica et Biophysica Acta– Reviews on Bioenergetics. 1987;895(2):81106.Google Scholar
Wang, Y, Bräutigam, A, Weber, AP, Zhu, X-G. Three distinct biochemical subtypes of C4 photosynthesis? A modelling analysis. Journal of Experimental Botany. 2014;65(13):3567–78.Google Scholar
Ghannoum, O. C4 photosynthesis and water stress. Annals of Botany. 2008;103(4):635–44.Google Scholar
Taylor, SH, Ripley, BS, Woodward, FI, Osborne, CP. Drought limitation of photosynthesis differs between C3 and C4 grass species in a comparative experiment. Plant, Cell & Environment. 2011;34(1):6575.Google Scholar
Taylor, SH, Ripley, BS, Martin, T, De‐Wet, LA, Woodward, FI, Osborne, CP. Physiological advantages of C4 grasses in the field: a comparative experiment demonstrating the importance of drought. Global Change Biology. 2014;20(6):19922003.Google Scholar
Liu, H, Osborne, CP. Water relations traits of C4 grasses depend on phylogenetic lineage, photosynthetic pathway, and habitat water availability. Journal of Experimental Botany. 2014;66(3):761–73.Google Scholar
Taylor, SH, Hulme, SP, Rees, M, Ripley, BS, Woodward FI, Osborne, CP. Ecophysiological traits in C3 and C4 grasses: a phylogenetically controlled screening experiment. New Phytologist. 2010;185(3):780–91.Google Scholar
Liu, H, Edwards, EJ, Freckleton, RP, Osborne, CP. Phylogenetic niche conservatism in C4 grasses. Oecologia. 2012;170(3):835–45.Google Scholar
Ripley, B, Visser, V, Christin, P-A, Archibald, S, Martin, T, Osborne, C. Fire ecology of C3 and C4 grasses depends on evolutionary history and frequency of burning but not photosynthetic type. Ecology. 2015;96(10):2679–91.CrossRefGoogle 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
×