Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-22dnz Total loading time: 0 Render date: 2024-04-25T07:21:38.125Z Has data issue: false hasContentIssue false

5 - Community Ecology of Salt Marshes

from Part I - Marsh Function

Published online by Cambridge University Press:  19 June 2021

Duncan M. FitzGerald
Affiliation:
Boston University
Zoe J. Hughes
Affiliation:
Boston University
Get access

Summary

Salt marshes have been useful study systems for community ecologists. They are amenable to experimental manipulation, and the simplicity and strong abiotic gradients of salt marshes lead to clear patterns and experimental outcomes. Many early ecologists believed that salt marsh ecosystems were primarily controlled by bottom-up factors (i.e., that nutrients, salinity, and other abiotic factors were the primary factors regulating productivity, and that productivity in turn regulated ecosystem trophic structure). More recently, many ecologists have argued that consumers have an important role in structuring salt marsh ecosystems through “top-down” processes. A simple conceptual approach, which we take here, is to think of salt marsh communities as being structured by bottom-up, top-down, and non-trophic processes.

Type
Chapter
Information
Salt Marshes
Function, Dynamics, and Stresses
, pp. 82 - 112
Publisher: Cambridge University Press
Print publication year: 2021

Access options

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

References

Ainouche, M. L., Baumel, A., Salmon, A., and Yannic, G. 2003. Hybridization, polyploidy and speciation in Spartina (Poaceae). New Phytologist, 161: 165172.CrossRefGoogle Scholar
Ainouche, M. L., Fortune, P. M., Salmon, A., Parisod, C., Grandbastien, M.-A., Fukunaga, K., Ricou, M., and Misset, M.-T. 2009. Hybridization, polyploidy and invasion: lessons from Spartina (Poaceae). Biological Invasions, 11: 11591173.Google Scholar
Alber, M., Swenson, E. M., Adamowicz, S. C., and Mendelssohn, I. A. 2008. Salt marsh dieback: an overview of recent events in the US. Estuarine, Coastal and Shelf Science, 80: 111.CrossRefGoogle Scholar
Alberti, J., Escapa, M., Daleo, P., Casariego, A. and Iribarne, O. 2010a. Crab bioturbation and herbivory reduce pre- and post-germination success of Sarcocornia perennis in bare patches of SW Atlantic salt marshes. Marine Ecology Progress Series, 400: 5561.CrossRefGoogle Scholar
Alberti, J., Escapa, M., Daleo, P., Iribarne, O., Silliman, B., and Bertness, M. 2007a. Local and geographic variation in grazing intensity by herbivorous crabs in SW Atlantic salt marshes. Marine Ecology Progress Series, 349: 235243.CrossRefGoogle Scholar
Alberti, J., Méndez Casariego, A., Daleo, P., Fanjul, E., Silliman, B. R., Bertness, M. D., and Iribarne, O. 2010b. Abiotic stress mediates top-down and bottom-up control in a Southwestern Atlantic salt marsh. Oecologia, 163: 181191.CrossRefGoogle Scholar
Alberti, J., Montemayor, D., Alvarez, F., Casariego, A. M., Luppi, T., Canepuccia, A., Isacch, J. P., and Iribarne, O. 2007b. Changes in rainfall pattern affect crab herbivory rates in a SW Atlantic salt marsh. Journal of Experimental Marine Biology and Ecology, 353: 126133.CrossRefGoogle Scholar
Altieri, A. H., Bertness, M. D., Coverdale, T. C., Herrmann, N. C., and Angelini, C. 2012. A trophic cascade triggers collapse of a salt-marsh ecosystem with intensive recreational fishing. Ecology, 93: 14021410.Google Scholar
Angelini, C., Griffin, J. N., Van de Koppel, , Lamers, J. L. P. M., Smolders, A. J. P., Derksen-Hooijberg, M., Van der Heide, T. and Silliman, B. R. 2016. A keystone mutualism underpins resilience of a coastal ecosystem to drought. Nature Communications, 7: 12473.CrossRefGoogle ScholarPubMed
Angelini, C., and Silliman, B. R. 2012. Patch size-dependent community recovery after massive disturbance. Ecology, 93: 101110.CrossRefGoogle ScholarPubMed
Armitage, A. R., and Fong, P. 2004. Upward cascading effects of nutrients: shifts in a benthic microalgal community and a negative herbivore response. Oecologia, 139: 560567.Google Scholar
Baldwin, A. H., and Mendelssohn, I. A. 1998. Response of two oligohaline marsh communities to lethal and nonlethal disturbance. Oecologia, 116: 543555.Google Scholar
Basan, P. B., and Frey, R. W. 1977. Actual-palaeontology and neoichnology of salt marshes near Sapelo Island, Georgia. Geological Journal Special Issue, 9: 4170.Google Scholar
Bazely, D. R., and Jefferies, R. L. 1986. Changes in the composition and standing crop of salt-marsh communities in response to the removal of a grazer. Journal of Ecology, 74: 693706.Google Scholar
Beeftink, W. G. 1977. The coastal salt marshes of western and northern Europe: an ecological and phytosociological approach. In Chapman, V. J., ed., Wet Coastal Ecosystems. Elsevier Scientific Publishing Company, Amsterdam, pp. 109155.Google Scholar
Bernik, B. M., Li, H., and Blum, M. J. 2016. Genetic variability of Spartina alterniflora intentionally introduced to China. Biological Invasions, 18: 14851498.Google Scholar
Bertness, M. D. 1984. Ribbed mussels and Spartina alterniflora production in a New England salt marsh. Ecology, 65: 17941807.Google Scholar
Bertness, M. D. 1985. Fiddler crab regulation of Spartina alterniflora production on a New England salt marsh. Ecology, 66: 10421055.CrossRefGoogle Scholar
Bertness, M. D. 1991a. Interspecific interactions among high marsh perennials in a New England salt marsh. Ecology, 72: 125137.Google Scholar
Bertness, M. D. 1991b. Zonation of Spartina patens and Spartina alterniflora in a New England salt marsh. Ecology, 72: 138148.Google Scholar
Bertness, M. D., Brisson, C. P. Coverdale, T. C. Bevil, M. C. Crotty, S. M., and Suglia, E. R. 2014. Experimental predator removal causes rapid salt marsh die-off. Ecology Letters, 17: 830835.Google Scholar
Bertness, M. D., and Callaway, R. 1994. Positive interactions in communities. Trends in Ecology and Evolution, 9: 191193.CrossRefGoogle ScholarPubMed
Bertness, M. D., Crain, C., Holdredge, C., and Sala, N. 2008. Eutrophication and consumer control of New England salt marsh primary productivity. Conservation Biology, 22: 131139.CrossRefGoogle ScholarPubMed
Bertness, M. D., and Ellison, A. M. 1987. Determinants of pattern in a New England salt marsh plant community. Ecological Monographs, 57: 129147.CrossRefGoogle Scholar
Bertness, M. D., and Ewanchuk, P. J. 2002. Latitudinal and climate-driven variation in the strength and nature of biological interactions in New England salt marshes. Oecologia, 132: 392401.Google Scholar
Bertness, M. D., Gough, L., and Shumway, S. W. 1992a. Salt tolerances and the distribution of fugitive salt marsh species. Ecology, 73: 18421851.Google Scholar
Bertness, M. D., and Hacker, S. D. 1994. Physical stress and positive associations among marsh plants. American Naturalist, 144: 363372.Google Scholar
Bertness, M. D., and Pennings, S. C. 2000. Spatial variation in process and pattern in salt marsh plant communities in Eastern North America. Pages 39–57 in Weinstein, M. P. and Kreeger, D. A., eds., Concepts and Controversies in Tidal Marsh Ecology. Kluwer Academic Publishers, Dordrecht.Google Scholar
Bertness, M. D., and Shumway, S. W. 1992. Consumer driven pollen limitation of seed production in marsh grasses. American Journal of Botany, 79: 288293.CrossRefGoogle Scholar
Bertness, M. D., and Shumway, S. W. 1993. Competition and facilitation in marsh plants. American Naturalist, 142: 718724.Google Scholar
Bertness, M. D., Wikler, K., and Chatkupt, T. 1992b. Flood tolerance and the distribution of Iva frutescens across New England salt marshes. Oecologia, 91: 171178.CrossRefGoogle ScholarPubMed
Bilkovic, D. M., Mitchell, M. M., Isdell, R. E., Schliep, M., and Smyth, A. R. 2017. Mutualism between ribbed mussels and cordgrass enhances salt marsh nitrogen removal. Ecosphere, 8: e01795.CrossRefGoogle Scholar
Blakeslee, A. M. H., Altman, I., Miller, A. W., Byers, J. E., Hamer, C. E., and Ruiz, G. M. 2012. Parasites and invasions: a biogeographic examination of parasites and hosts in native and introduced ranges. Journal of Biogeography, 39: 609622.Google Scholar
Blum, M. J., Bando, K. J., Katz, M., and Strong, D. R. 2007. Geographic structure, genetic diversity and source tracking of Spartina alterniflora. Journal of Biogeography, 34: 20552069.CrossRefGoogle Scholar
Boesch, D. F., and Turner, R. E. 1984. Dependence of fishery species on salt marshes – the role of food and refuge. Estuaries, 7: 460468.CrossRefGoogle Scholar
Bortolus, A., and Iribarne, O. 1999. Effects of the SW Atlantic burrowing crab Chasmagnathus granulata on a Spartina salt marsh. Marine Ecology, Progress Series, 178: 7988.Google Scholar
Bradley, P. M., and Dunn, E. L. 1989. Effects of sulfide on the growth of three salt marsh halophytes of the southeastern United States. American Journal of Botany, 76: 17071713.CrossRefGoogle Scholar
Brewer, J. S., and Bertness, M. D. 1996. Disturbance and intraspecific variation in the clonal morphology of salt marsh perennials. Oikos, 77: 107116.Google Scholar
Brewer, J. S., Levine, J. M., and Bertness, M. D. 1998. Interactive effects of elevation and burial with wrack on plant community structure in some Rhode Island salt marshes. Journal of Ecology, 86: 125136.Google Scholar
Byers, J. E. 2000. Competition between two estuarine snails: implications for invasions of exotic species. Ecology, 81: 12251239.CrossRefGoogle Scholar
Byers, J. E., Rogers, T. L., Grabowski, J. H., Hughes, A. R., Piehler, M. F., and Kimbro, D. L. 2014. Host and parasite recruitment correlated at a regional scale. Oecologia, 174: 731738.Google Scholar
Callaway, J. C., Sullivan, G., and Zedler, J. B. 2003. Species-rich plantings increase biomass and nitrogen accumulation in a wetland restoration experiment. Ecological Applications, 13: 16261639.Google Scholar
Callaway, R. M. 1994. Facilitative and interfering effects of Arthrocnemum subterminale on winter annuals. Ecology, 75: 681686.Google Scholar
Callaway, R. M., and Pennings, S. C. 1998. Impact of a parasitic plant on the zonation of two salt marsh perennials. Oecologia, 114: 100105.Google Scholar
Cavanaugh, K. C., Kellner, J. R., Forde, A. J., Gruner, D. S., Parker, J. D., Rodriguez, W., and Feller, I. C. 2014. Poleward expansion of mangroves is a threshold response to decreased frequency of extreme cold events. Proceedings of the National Academy of Science, USA, 111: 723727.CrossRefGoogle ScholarPubMed
Cebrian, J. 1999. Patterns in the fate of production in plant communities. American Naturalist, 154: 449468.CrossRefGoogle ScholarPubMed
Chapman, V. J. 1974. Salt marshes and salt deserts of the world. In: Reimold, R. J. and Queen, W. H., editors. Ecology of Halophytes. Academic Press, New York, pp. 319.Google Scholar
Coverdale, T. C., Herrmann, N. C., Altieri, A. H., and Bertness, M. D.. 2013. Latent impacts: the role of historical human activity in coastal habitat loss. Frontiers in Ecology and the Environment, 11: 6974.CrossRefGoogle Scholar
Craft, C. 2007. Freshwater input structures soil properties, vertical accretion, and nutrient accumulation of Georgia and U.S. tidal marshes. Limnology and Oceanography, 52: 12201230.Google Scholar
Crain, C. M., and Bertness, M. D. 2006. Ecosystem engineering across environmental gradients: implications for conservation and management. Bioscience, 56: 211218.CrossRefGoogle Scholar
Crain, C. M., Silliman, B. R., Bertness, S. L., and Bertness, M. D. 2004. Physical and biotic drivers of plant distribution across estuarine salinity gradients. Ecology, 85: 25392549.Google Scholar
Cresswell, W., Lind, J., and Quinn, J. L. 2010. Predator-hunting success and prey vulnerability: quantifying the spatial scale over which lethal and non-lethal effects of predation occur. Journal of Animal Ecology, 79: 556562.Google Scholar
Crichton, O. W. 1960. Marsh crab: intertidal tunnel-maker and grass-eater. Estuarine Bulletin, 5: 310.Google Scholar
Crotty, S. M., Sharp, S. J., Bersoza, A. C., Prince, K. D., Cronk, K., Johnson, E., E., and Angelini, C. 2018. Foundation species patch configuration mediates salt marsh biodiversity, stability and multifunctionality. Ecology Letters, 21: 16811692.CrossRefGoogle ScholarPubMed
Currin, C. A., Newell, S. Y., and Paerl, H. W. 1995. The role of standing dead Spartina alterniflora and benthic microalgae in salt marsh food webs: considerations based on multiple stable isotope analysis. Marine Ecology Progress Series, 121: 99116.Google Scholar
Dai, T., and Wiegert, R. G. 1996a. Estimation of the primary productivity of Spartina alterniflora using a canopy model. Ecography, 19: 410423.Google Scholar
Dai, T., and Wiegert, R. G. 1996b. Ramet population dynamics and net aerial primary productivity of Spartina alterniflora. Ecology, 77: 276288.Google Scholar
Daleo, P., Alberti, J., Bruschetti, C. M., Pascual, J., Iribarne, O., and Silliman, B. R. 2015. Physical stress modifies top-down and bottom-up forcing on plant growth and reproduction in a coastal ecosystem. Ecology, 96: 21472156.CrossRefGoogle Scholar
Daleo, P., Alberti, J., Canepuccia, A., Escapa, M., Fanjul, E., Silliman, B. R., Bertness, M. D., and Iribarne, O. 2008. Mychorrhizal fungi determine salt-marsh plant zonation depending on nutrient supply. Journal of Ecology, 96: 431437.CrossRefGoogle Scholar
Daleo, P., Fanjul, E., Casariego, A. M., Silliman, B. R., Bertness, M. D., and Iribarne, O. 2007. Ecosystem engineers activate mycorrhizal mutualism in salt marshes. Ecology Letters 10: 902908.Google Scholar
Daleo, P., and Iribarne, O. 2009. Beyond competition: the stress-gradient hypothesis tested in plant–herbivore interactions. Ecology, 90: 23682374.Google Scholar
Darley, W. M., Montague, C. L., Plumley, F. G., Sage, W. W., and Psalidas, A. T. 1981. Factors limiting edaphic algal biomass and productivity in a Georgia salt marsh. Journal of Phycology, 17: 122128.CrossRefGoogle Scholar
Davidson, A., Griffin, J. N., Angelini, C., Coleman, F., Atkins, R. L., and Silliman, B. R. 2015. Non-consumptive predator effects intensify grazer-plant interactions by driving vertical habitat shifts. Marine Ecology Progress Series, 537: 4958.Google Scholar
de Bettencourt, A. M. M., Neves, R. J. J., Lança, M. J., Batista, P. J., and Alves, M. J. 1994. Uncertainties in import/export studies and the outwelling theory. An analysis with the support of hydrodynamic modelling. In Mitsch, W. J., ed., Global Wetlands: Old world and new. Elsevier Science B. V., Amsterdam, pp. 235256.Google Scholar
Deegan, L. A., Johnson, D. S., Warren, R. S., Peterson, B. J., Fleeger, J. W., Fagherazzi, S., and Wollheim, W. M. 2012. Coastal eutrophication as a driver of salt marsh loss. Nature, 490: 388392.Google Scholar
Denno, R. F. 1980. Ecotope differentiation in a guild of sap-feeding insects on the salt marsh grass, Spartina patens. Ecology, 61: 702714.Google Scholar
Denno, R. F., Gratton, C., Dobel, H., and Finke, D. L. 2003. Predation risk affects relative strength of top-down and bottom-up impacts on insect herbivores. Ecology, 84: 10321044.CrossRefGoogle Scholar
Denno, R. F., Lewis, D., and Gratton, C. 2005. Spatial variation in the relative strength of top-down and bottom-up forces: causes and consequences for phytophagous insect populations. Annales Zoologici Fennici, 42: 295311.Google Scholar
Denno, R. F., Peterson, M. A., Gratton, C., Cheng, J., Langellotto, G. A., Huberty, A. F., and Finke, D. L. 2000. Feeding-induced changes in plant quality mediate interspecific competition between sap-feeding herbivores. Ecology, 81: 18141827.CrossRefGoogle Scholar
Denno, R. F., and Roderick, G. K. 1992. Density-related dispersal in planthoppers: effects of interspecific crowding. Ecology, 73: 13231334.Google Scholar
Denno, R. F., Roderick, G. K., Peterson, M. A., Huberty, A. F., Dobel, H. G., Eubanks, M. D., Losey, J. E., and Langellotto, G. A. 1996. Habitat persistence underlies intraspecific variation in the dispersal strategies of planthoppers. Ecological Monographs, 66: 389408.Google Scholar
Diaz-Ferguson, E., Robinson, J. D., Silliman, B., and Wares, J. P. 2010. Comparative phylogeography of North American Atlantic salt marsh communities. Estuaries and Coasts, 33: 828839.Google Scholar
Döbel, H. G., Denno, R. F., and Coddington, J. A. 1990. Spider (Araneae) community structure in an intertidal salt marsh: effects of vegetation structure and tidal flooding. Environmental Entomology, 19: 13561370.Google Scholar
Donnelly, J. P., Bryant, S. S., Butler, J., Dowling, J., Fan, L., Hausmann, N., Newby, P., et al. 2001a. 700 yr sedimentary record of intense hurricane landfalls in southern New England. Geological Society of America Bulletin, 113: 714727.Google Scholar
Donnelly, J. P., Roll, S., Wengren, M., Butler, J., Lederer, R., and Webb, T. III 2001b. Sedimentary evidence of intense hurricane strikes from New Jersey. Geology, 29: 615618.Google Scholar
Ellison, A. M. 1987. Effects of competition, disturbance, and herbivory on Salicornia europaea. Ecology, 68: 576586.Google Scholar
Ellison, A. M. 1991. Ecology of case-bearing moths (Lepidoptera: coleophoridae) in a New England salt marsh. Environmental Entomology, 20: 857864.Google Scholar
Elschot, K., Vermeulen, A., Vandenbruwaene, W., Bakker, J. P., Bouma, T. J., Stahl, J., Castelijns, H., and Temmerman, S. 2017. Top-down vs. bottom-up control on vegetation composition in a tidal marsh depends on scale. PLOS ONE, 12: e0169960.Google Scholar
Engels, J. G., and Jensen, K. 2010. Role of biotic interactions and physical factors in determining the distribution of marsh species along an estuarine salinity gradient. Oikos, 119: 679685.Google Scholar
Escapa, M., Minkoff, D. R., Perillo, G. M. E., and Iribarne, O. 2007. Direct and indirect effects of burrowing crab Chasmagnathus granulatus activities on erosion of southwest Atlantic Sarcocornia-dominated marshes. Limnology and Oceanography, 52: 23402349.Google Scholar
Ewanchuk, P. J., and Bertness, M. D. 2003. Recovery of a northern New England salt marsh plant community from winter icing. Oecologia, 136: 616626.Google Scholar
Fariña, J. M., He, Q., Silliman, B. R., and Bertness, M. D. 2017. Biogeography of salt marsh plant zonation on the Pacific coast of South America. Journal of Biogeography, 45: 238247.Google Scholar
Fariña, J. M., Silliman, B. R., and Bertness, M. D. 2009. Can conservation biologists rely on established community structure rules to manage novel systems?…Not in salt marshes. Ecological Applications, 19: 413422.Google Scholar
Feher, L. C., Osland, M. J., Griffith, K. T., Grace, J. B., Howard, R. J., Stagg, C. L., Enwright, N. M., et al. 2017. Linear and nonlinear effects of temperature and precipitation on ecosystem properties in tidal saline wetlands. Ecosphere, 8: e01956.Google Scholar
Finke, D. L., and Denno, R. F. 2004. Predator diversity dampens trophic cascades. Nature, 429: 407410.Google Scholar
Finke, D. L., and Denno, R. F. 2006. Spatial refuge from intraguild predation: implications for prey suppression and trophic cascades. Oecologi, 149: 265275.CrossRefGoogle ScholarPubMed
Foster, W. A., and Treherne, J. E. 1976. Insects of marine saltmarshes: problems and adaptations. In: Cheng, L., ed., Marine Insects. North-Holland Publishing Company, Amsterdam, pp. 542.Google Scholar
Frey, R. W., and Basan, P. B. 1978. Coastal salt marshes. In: Davis, R. A. Jr., ed., Coastal Sedimentary Environments. Springer-Verlag, New York, pp. 101169.Google Scholar
Gabler, C. A., Osland, M. J., Grace, J. B., Stagg, C. L., Day, R. H., Hartley, S. B., Enwright, N. M., et al. 2017. Macroclimatic change expected to transform coastal wetland ecosystems this century. Nature Climate Change, 7: 142147.CrossRefGoogle Scholar
Gallagher, J. L., Reimold, R. J., Linthurst, R. A., and Pfeiffer, W. J. 1980. Aerial production, mortality, and mineral accumulation-export dynamics in Spartina alterniflora and Juncus roemerianus plant stands in a Georgia salt marsh. Ecology, 61: 303312.Google Scholar
Ganong, W. F. 1903. The vegetation of the Bay of Fundy salt and diked marshes: an ecological study. Botanical Gazette, 36: 161186, 280–302, 350–367, 429–455.Google Scholar
Gedan, K. B., Crain, C. M., and Bertness, M. D. 2009. Small-mammal herbivore control of secondary succession in New England tidal marshes. Ecology, 90: 430440.Google Scholar
Grewell, B. J. 2008a. Hemiparasites generate environmental heterogeneity and enhance species coexistence in salt marshes. Ecological Applications, 18: 12971306.Google Scholar
Grewell, B. J. 2008b. Parasite facilitates plant species coexistence in a coastal wetland. Ecology, 89: 14811488.Google Scholar
Griffin, J. N., and Silliman, B. R. 2011 Predator diversity stabilizes and strengthens trophic control of a keystone grazer. Biology Letters, 7: 7982.Google Scholar
Grosholz, E. 2010. Avoidance by grazers facilitates spread of an invasive hybrid plant. Ecology Letters, 13: 145153.Google Scholar
Guo, H., and Pennings, S. C. 2012. Mechanisms mediating plant distributions across estuarine landscapes in a low-latitude tidal estuary. Ecology, 93: 90100.Google Scholar
Hacker, S. D., and Bertness, M. D. 1995. A herbivore paradox: why salt marsh aphids live on poor-quality plants. American Naturalist, 145: 192210.CrossRefGoogle Scholar
Hacker, S. D., and Bertness, M. D. 1999. Experimental evidence for factors maintaining plant species diversity in a New England salt marsh. Ecology, 80: 20642073.Google Scholar
Hackney, C. T., and Bishop, T. D. 1981. A note on the relocation of marsh debris during a storm surge. Estuarine, Coastal and Shelf Science, 12: 621624.Google Scholar
Haines, E. B. 1976. Stable carbon isotope ratios in the biota, soils and tidal water of a Georgia salt marsh. Estuarine and Coastal Marine Science, 4: 609616.Google Scholar
Haines, E. B., and Montague, C. L. 1979. Food sources of estuarine invertebrates analyzed using 13C/12C ratios. Ecology, 60: 4856.Google Scholar
Hanley, T. C., Kimbro, D. L., and Hughes, A. R. 2017. Stress and subsidy effects of seagrass wrack duration, frequency, and magnitude on salt marsh community structure. Ecology, 98: 18841895.Google Scholar
Hardwick-Witman, M. N. 1985. Biological consequences of ice rafting in a New England salt marsh community. Journal of Experimental Marine Biology and Ecology, 87: 283298.Google Scholar
He, Q., Altieri, A. H., and Cui, B. 2015. Herbivory drives zonation of stress-tolerant marsh plants. Ecology, 96: 13181328.Google Scholar
He, Q., and Bertness, M. D. 2014. Extreme stresses, niches, and positive species interactions along stress gradients. Ecology, 95: 14371443.Google Scholar
He, Q., Bertness, M. D., and Altieri, A. H. 2013. Global shifts towards positive species interactions with increasing environmental stress. Ecology Letters, 16: 695706.Google Scholar
He, Q., Bertness, M. D., Bruno, F., Li, B., Chen, G., Coverdale, T. C., Altieri, A. H., et al. 2014. Economic development and coastal ecosystem change in China. Scientific Reports, 4: 5995.Google Scholar
He, Q., and Cui, B. 2015. Multiple mechanisms sustain a plant-animal facilitation on a coastal ecotone. Scientific Reports, 5: 8612.Google Scholar
He, Q., Cui, B., Bertness, M. D., and An, Y. 2012. Testing the importance of plant strategies on facilitation using congeners in a coastal community. Ecology, 93: 20232029.Google Scholar
He, Q., and Silliman, B. R. 2015. Biogeographic consequences of nutrient enrichment for plant-herbivore interactions in coastal wetlands. Ecology Letters, 18: 462471.Google Scholar
He, Q., and Silliman, B. R. 2016. Consumer control as a common driver of coastal vegetation worldwide. Ecological Monographs, 86: 278294.Google Scholar
He, Q., Silliman, B. R., and Cui, B. 2017a. Incorporating thresholds into understanding salinity tolerance: a study using salt-tolerant plants in salt marshes. Ecology and Evolution, 2017: 63266333.Google Scholar
He, Q., Silliman, B. R., Liu, Z., and Cui, B. 2017b. Natural enemies govern ecosystem resilience in the face of extreme droughts. Ecology Letters, 20: 194201.Google Scholar
Hensel, M. J. S., and Silliman, B. R. 2013. Consumer diversity across kingdoms supports multiple functions in a coastal ecosystem. Proceedings of the National Academy of Science, USA, 110: 2062120626.Google Scholar
Hilton, G. M., Ruxton, G. D., and Cresswell, W. 1999. Choice of foraging area with respect to predation risk in redshanks: the effects of weather and predator activity. Oikos, 87: 295302.Google Scholar
Ho, C.-K., and Pennings, S. C. 2008. Consequences of omnivory for trophic interactions on a salt marsh shrub. Ecology, 89: 17141722.CrossRefGoogle ScholarPubMed
Holdredge, C., Bertness, M. D., and Altieri, A. H. 2008. Role of crab herbivory in die-off of New England salt marshes. Conservation Biology, 23: 672679.CrossRefGoogle ScholarPubMed
Hopkinson, C. S., Gosselink, J. G., and Parrondo, R. T. 1978. Aboveground production of seven marsh plant species in coastal Louisiana. Ecology, 59: 760769.Google Scholar
Hovel, K. A., Bartholomew, A., and Lipcius, R. N. 2001. Rapidly entrainable tidal vertical migrations in the salt marsh snail Littoraria irrorata. Estuaries, 24: 808816.Google Scholar
Hughes, A. R., and Lotterhos, K. E. 2014. Genotypic diversity at multiple spatial scales in the foundation marsh species, Spartina alterniflora. Marine Ecology Progress Series, 497: 105117.Google Scholar
Hughes, A. R., Moore, A. F. P., and Piehler, M. F. 2014. Independent and interactive effects of two facilitators on their habitat-providing host plant, Spartina alterniflora. Oikos, 123: 488499.Google Scholar
Jensen, A. 1985. The effect of cattle and sheep grazing on salt-marsh vegetation at Skallingen, Denmark. Vegetatio, 60: 3748.Google Scholar
Johnson, D. S., and Heard, R. 2017. Bottom-up control of parasites. Ecosphere, 8: e01885.Google Scholar
Keddy, P. A. 1990. Competitive hierarchies and centrifugal organization in plant communities. In: Grace, J. B., and Tilman, D., eds., Perspectives on Plant Competition. Academic Press, Inc., San Diego, pp. 265290.Google Scholar
Kimbro, D. L. 2012. Tidal regime dictates the cascading consumptive and nonconsumptive effects of multiple predators on a marsh plant. Ecology, 93: 334344.Google Scholar
Kirwan, M. L., Guntenspergen, G. R., and Morris, J. T. 2009. Latitudinal trends in Spartina alterniflora productivity and the response of coastal marshes to global change. Global Change Biology, 15: 19821989.Google Scholar
Kirwan, M. L., Murray, A. B., and Boyd, W. S. 2008. Temporary vegetation disturbance as an explanation for permanent loss of tidal wetlands. Geophysical Research Letters, 35: L05403.Google Scholar
Kneib, R. T. 1987. Predation risk and use of intertidal habitats by young fishes and shrimp. Ecology, 68: 379386.Google Scholar
Kneib, R. T. 2000. Salt marsh ecoscapes and production transfers by estuarine nekton in the southeastern United States. In: Weinstein, M. P. and Kreeger, D. A., editors. Concepts and Controversies in Tidal Marsh Ecology. Kluwer Academic Publishers, Dordrecht.Google Scholar
Kuijper, D. P. J., and Bakker, J. P. 2005. Top-down control of small herbivores on salt-marsh vegetation along a productivity gradient. Ecology, 86: 914923.Google Scholar
Kuris, A. M., Hechinger, R. F., Shaw, J. C., Whitney, K. L., Aguirre-Macedo, L., Boch, C. A., Dobson, A. P., et al. 2008. Ecosystem energetic implications of parasite and free-living biomass in three estuaries. Nature, 454: 515518.CrossRefGoogle ScholarPubMed
Kwak, T. J., and Zedler, J. B. 1997. Food web analysis of southern California coastal wetlands using multiple stable isotopes. Oecologia, 110: 262277.Google Scholar
Lafferty, K. D. 1993. Effects of parasitic castration on growth, reproduction and population dynamics of the marine snail Cerithidea californica. Marine Ecology Progress Series, 96: 229237.Google Scholar
Lafferty, K. D., Dobson, A. P., and Kuris, A. M. 2006. Parasites dominate food web links. Proceedings of the National Academy of Science, USA, 103: 1121111216.Google Scholar
Lafferty, K. D., and Morris, K. 1996. Altered behavior of parasitized killifish increases susceptibility to predation by bird final hosts. Ecology, 77: 13901397.Google Scholar
Langdon, C. J., and Newell, R. I. E. 1990. Utilization of detritus and bacteria as food sources by two bivalve suspension-feeders, the oyster Crassostrea virginica and the mussel Geukensia demissa. Marine Ecology Progress Series, 58: 299310.Google Scholar
Lee, S. C., and Silliman, B. R. 2006. Competitive displacement of a detritivorous salt marsh snail. Journal of Experimental Marine Biology and Ecology, 339: 7585.Google Scholar
Levin, P. S., Ellis, J., Petrik, R., and Hay, M. E. 2002. Indirect effects of feral horses on estuarine communities. Conservation Biology, 16: 13641371.Google Scholar
Levine, J. M., Brewer, J. S., and Bertness, M. D. 1998. Nutrients, competition and plant zonation in a New England salt marsh. Journal of Ecology, 86: 285292.Google Scholar
Lewis, D. B., and Eby, L. A. 2002. Spatially heterogeneous refugia and predation risk in intertidal salt marshes. Oikos, 96: 119129.Google Scholar
Li, H., Zhang, X., Zheng, R., Li, X., Elmer, W. H., Wolfe, L. M., and Li, B. 2014a. Indirect effects of non-native Spartina alterniflora and its fungal pathogen (Fusarium palustre) on native saltmarsh plants in China. Journal of Ecology, 102: 11121119.Google Scholar
Li, S., and Pennings, S. C. 2016. Disturbance in Georgia salt marshes: variation across space and time. Ecosphere, 7: e01487.Google Scholar
Li, S., and Pennings, S. C. 2017. Timing of disturbance affects biomass and flowering of a saltmarsh plant and attack by stem-boring herbivores. Ecosphere, 8: e01675.Google Scholar
Li, Z., Wang, W., and Zhang, Y. 2014b. Recruitment and herbivory affect spread of invasive Spartina alterniflora in China. Ecology, 95: 19721980.Google Scholar
Linthurst, R. A. 1980. An evaluation of aeration, nitrogen, pH and salinity as factors affecting Spartina alterniflora growth: a summary. In: Kennedy, V. S., ed., Estuarine Perspectives. Academic Press, New York, pp. 235247Google Scholar
Liu, W., Maung-Douglas, K., Strong, D. R., Pennings, S. C., and Zhang, Y. 2016. Geographical variation in vegetative growth and sexual reproduction of the invasive Spartina alterniflora in China. Journal of Ecology, 104: 173181.Google Scholar
Lynch, J. J., O'Neil, E., and Lay, D. W. 1947. Management significance of damage by geese and muskrats to gulf coast marshes. Journal of Wildlife Management, 11: 5076.Google Scholar
Malek, J. C., and Byers, J. E. 2017. The effects of tidal elevation on parasite heterogeneity and co-infection in the eastern oyster, Crassostrea virginica. Journal of Experimental Marine Biology and Ecology, 494: 3227.Google Scholar
Marczak, L. B., Więski, K. Denno, R. F., and Pennings, S. C. 2013. Importance of local vs. geographic variation in salt marsh plant quality for arthropod herbivore communities. Journal of Ecology, 101: 11691182.Google Scholar
Maricle, B. R., Cobos, D. R., and Campbell, C. S. 2007. Biophysical and morphological leaf adaptations to drought and salinity in salt marsh grasses. Environmental and Experimental Botany, 60: 458467.Google Scholar
Maricle, B. R., Crosier, J. J., Bussiere, B. C., and Lee, R. W. 2006. Respiratory enzyme activities correlate with anoxia tolerance in salt marsh grasses. Journal of Experimental Marine Biology and Ecology, 337: 3037.Google Scholar
Maricle, B. R., and Lee, R. W. 2007. Root respiration and oxygen flux in salt marsh grasses from different elevational zones. Marine Biology, 151: 413423.Google Scholar
McKee, K. L., Mendelssohn, I. A., and Materne, M. D. 2004. Acute salt marsh dieback in the Mississippi River deltaic plain: a drought-induced phenomenon? Global Ecology and Biogeography, 13: 6573.Google Scholar
Mendelssohn, I. A. 1979. Nitrogen metabolism in the height forms of Spartina alterniflora in North Carolina. Ecology, 60: 574584.Google Scholar
Mendelssohn, I. A., and Morris, J. T. 2000. Eco-physiological controls on the productivity of Spartina alterniflora Loisel. In: Weinstein, M. P. and Kreeger, D. A., eds., Concepts and Controversies in Tidal Marsh Ecology. Kluwer Academic Publishers, Dordrecht, pp. 5980.Google Scholar
Minello, T. J., Able, K. W., Weinstein, M. P., and Hays, C. G. 2003. Salt marshes as nurseries for nekton: testing hypotheses on density, growth and survival through meta-analysis. Marine Ecology Progress Series, 246: 3959.Google Scholar
Minello, T. J., and Zimmerman, R. J. 1992. Utilization of natural and transplanted Texas salt marshes by fish and decapod crustaceans. Marine Ecology Progress Series, 90: 273285.Google Scholar
Montague, C. L. 1982. The influence of fiddler crab burrows and burrowing on metabolic processes in salt marsh sediments. In: Kennedy, V. S., ed., Estuarine Comparisons. Academic Press, New York, pp. 283301.Google Scholar
Moon, D. C., Rossi, A. M., and Stiling, P. 2000. The effects of abiotically induced changes in host plant quality (and morphology) on a salt marsh planthopper and its parasitoid. Ecological Entomology, 25: 325331.Google Scholar
Moon, D. C., and Stiling, P. 2002. The effects of salinity and nutrients on a tritrophic salt-marsh system. Ecology, 83: 24652476.Google Scholar
Moon, D. C., and Stiling, P. 2004. The influence of a salinity and nutrient gradient on coastal vs. upland tritrophic complexes. Ecology, 85: 27092716.Google Scholar
Morris, J. T., Kjerfve, B., and Dean, J. M. 1990. Dependence of estuarine productivity on anomalies in mean sea level. Limnology and Oceanography, 35: 926930.Google Scholar
Morris, J. T., Sundareshwar, P. V., Nietch, C. T., Kjerfve, B., and Cahoon, D. R. 2002. Responses of coastal wetlands to rising sea level. Ecology, 83: 28692877.Google Scholar
Morris, J. T., Sundberg, K., and Hopkinson, C. S. 2013. Salt marsh primary production and its responses to relative sea level and nutrients in estuaries at Plum Island, Massachusetts, and North Inlet, South Carolina, USA. Oceanography, 26: 7884.CrossRefGoogle Scholar
Morris, R. K. A., Reach, I. S., Duffy, M. J., Collins, T. S., and Leafe, R. N. 2004. On the loss of saltmarshes in south-east England and the relationship with Nereis diversicolor. Journal of Applied Ecology, 41: 787791.Google Scholar
Naeem, S. 2002. Ecosystem consequences of biodiversity loss: the evolution of a paradigm. Ecology, 83: 15371552.Google Scholar
Naeem, S., Thompson, L. J., Lawler, S. P., Lawton, J. H., and Woodfin, R. M. 1994. Declining biodiversity can alter the performance of ecosystems. Nature, 368: 734737.Google Scholar
Nifong, J. C., and Silliman, B. R. 2013. Impacts of a large-bodied, apex predator (Alligator mississippiensis Daudin 1801) on salt marsh food webs. Journal of Experimental Marine Biology and Ecology, 440: 185191.Google Scholar
Nyman, J. A., Crozier, C. R., and DeLaune, R. D. 1995. Roles and patterns of hurricane sedimentation in an estuarine marsh landscape. Estuarine, Coastal and Shelf Science, 40: 665679.Google Scholar
O'Donnell, J. P. R., and Schalles, J. F. 2016. Examination of abiotic drivers and their influence on Spartina alterniflora biomass over a twenty-eight year period using Landsat 5 TM satellit imagery of the central Georgia coast. Remote Sensing, 8: 477.Google Scholar
Odum, W. E. 1988. Comparative ecology of tidal freshwater and salt marshes. Annual Review of Ecology and Systematics, 19: 147176.Google Scholar
Osgood, D. T., Santos, M. C. F. V., and Zieman, J. C. 1995. Sediment physico-chemistry associated with natural marsh development on a storm-deposited sand flat. Marine Ecology Progress Series, 120: 271283.Google Scholar
Osland, M. J., Enwright, N., Day, R. H., and Doyle, T. W. 2013. Winter climate change and coastal wetland foundation species: salt marshes vs. mangrove forests in the southeastern United States. Global Change Biology, 19: 14821494.Google Scholar
Pace, M. L., Shimmel, S., and Darley, W. M. 1979. The effect of grazing by a gastropod, Nassarius obsoletus, on the benthic microbial community of a salt marsh mudflat. Estuarine and Coastal Marine Science, 9: 121134.Google Scholar
Page, H. M. 1997. Importance of vascular plant and algal production to macro-invertebrate consumers in a southern California salt marsh. Estuarine, Coastal and Shelf Science, 45: 823834.Google Scholar
Paramor, O. A., and Hughes, R. G. 2004. The effects of bioturbation and herbivory by the polychaete Nereis diversicolor on loss of saltmarsh in south-east England. Journal of Applied Ecology, 41: 449463.Google Scholar
Pautzke, S. M., Mather, M. E., Finn, J. T., Deegan, L. A., and Muth, R. M. 2010. Seasonal use of a New England estuary by foraging contingents of migratory striped bass. Transactions of the American Fisheries Society, 139: 257269.Google Scholar
Penfound, W. T., and Hathaway, E. S. 1938. Plant communities in the marshlands of southeastern Louisiana. Ecological Monographs, 8: 156.Google Scholar
Pennings, S. C., and Bertness, M. D. 2001. Salt marsh communities. In Bertness, M. D., Gaines, S. D., and Hay, M. E., eds., Marine Community Ecology. Sinauer Associates, Sunderland, pp. 289316.Google Scholar
Pennings, S. C., and Callaway, R. M. 1992. Salt marsh plant zonation: the relative importance of competition and physical factors. Ecology, 73: 681690.Google Scholar
Pennings, S. C., and Callaway, R. M. 1996. Impact of a parasitic plant on the structure and dynamics of salt marsh vegetation. Ecology, 77: 14101419.Google Scholar
Pennings, S. C., Ho, C.-K., Salgado, C. S., Więski, K., Davé, N., Kunza, A. E., and Wason, E. L. 2009. Latitudinal variation in herbivore pressure in Atlantic Coast salt marshes. Ecology, 90: 183195.Google Scholar
Pennings, S. C., and Moore, D. J. 2001. Zonation of shrubs in western Atlantic salt marshes. Oecologia, 126: 587594.Google Scholar
Pennings, S. C., and Richards, C. L. 1998. Effects of wrack burial in salt-stressed habitats: Batis maritima in a southwest Atlantic salt marsh. Ecography, 21: 630638.Google Scholar
Pennings, S. C., Selig, E. R., Houser, L. T., and Bertness, M. D. 2003. Geographic variation in positive and negative interactions among salt marsh plants. Ecology, 84: 15271538.Google Scholar
Pennings, S. C., and Silliman, B. R. 2005. Linking biogeography and community ecology: latitudinal variation in plant-herbivore interaction strength. Ecology, 86: 23102319.Google Scholar
Pennings, S. C., Stanton, L. E., and Brewer, J. S. 2002. Nutrient effects on the composition of salt marsh plant communities along the southern Atlantic and Gulf Coasts of the United States. Estuaries, 25: 11641173.Google Scholar
Pfeiffer, W. J., and Wiegert, R. G. 1981. Grazers on Spartina and their predators. In: Pomeroy, L. R. and Wiegert, R. G., ed., The Ecology of a Salt Marsh. Springer-Verlag, New York, pp. 87112.Google Scholar
Pung, O. J., Khan, R. N., Vives, S. P., and Walker, C. B. 2002. Prevalence, geographical distribution, and fitness effects of Microphallus turgidus (Trematoda: Microphallidae) in grass shrimp (Palaemonetes spp.) from coastal Georgia. Journal of Parasitology, 88: 8992.Google Scholar
Purer, E. A. 1942. Plant ecology of the coastal salt marshlands of San Diego county, California. Ecological Monographs, 12: 82111.Google Scholar
Rachlin, J. W., Stalter, R., Kincaid, D., and Warkentine, B. E. 2012. Parsimony analysis of East Coast salt marsh plant distributions. Northeastern Naturalist, 19: 279296.Google Scholar
Rand, T. A. 2004. Competition, facilitation, and compensation for insect herbivory in an annual salt marsh forb. Ecology, 85: 20462052.Google Scholar
Ranwell, D. S. 1961. Spartina salt marshes in southern England. I. The effects of sheep grazing at the upper limits of Spartina marsh in Bridgwater Bay. Journal of Ecology, 49: 325340.Google Scholar
Raybould, A. F., Gray, A. J., and Clarke, R. T. 1998. The long-term epidemic of Claviceps purpurea on Spartina anglica in Poole Harbour: pattern of infection, effects on seed production and the role of Fusarium heterosporum. New Phytologist, 138: 497505.Google Scholar
Redfield, A. C. 1972. Development of a New England salt marsh. Ecological Monographs, 42: 201237.Google Scholar
Reidenbaugh, T. G., and Banta, W. C. 1980. Origin and effects of Spartina wrack in a Virginia salt marsh. Gulf Research Reports, 6: 393401.Google Scholar
Rejmanek, M., Sasser, C., and Peterson, G. W. 1988. Hurricane-induced sediment deposition in a Gulf Coast marsh. Estuarine and Coastal Shelf Science, 27: 217222.Google Scholar
Richard, G. A. 1978. Seasonal and environmental variations in sediment accretion in a Long Island salt marsh. Estuaries, 1: 2935.Google Scholar
Richards, C. L., Hamrick, J. L., Donovan, L. A., and Mauricio, R. 2004. Unexpectedly high clonal diversity of two salt marsh perennials across a severe environmental gradient. Ecology Letters, 7: 11551162.Google Scholar
Rowcliffe, J. M., Watkinson, A. R., and Sutherland, W. J. 1998. Aggregative responses of brent geese on salt marsh and their impact on plant community dynamics. Oecologia, 114: 417426.Google Scholar
Rozema, J., Bijwaard, P., Prast, G., and Broekman, R. 1985. Ecophysiological adaptations of coastal halophytes from foredunes and salt marshes. Vegetatio, 62: 499521.Google Scholar
Schalles, J. F., Hladik, C. M., Lynes, A. A., and Pennings, S. C. 2013. Landscape estimates of habitat types, plant biomass, and invertebrate densities in a Georgia salt marsh. Oceanography, 26: 8897.Google Scholar
Schindler, D. E., Johnson, B. M., MacKay, N. A., Bouwes, N., and Kitchell, J. F. 1994. Crab: snail size-structured interactions and salt marsh predation gradients. Oecologia, 97: 4961.Google Scholar
Schoolmaster, D. R. Jr., and Stagg, C. L. 2018. Resource competition model predicts zonation and increasing nutrient use efficiency along a wetland salinity gradient. Ecology, 99: 670680.Google Scholar
Schrama, M., Berg, M. P., and Olff, H. 2012. Ecosystem assembly rules: the interplay of green and brown webs during salt marsh succession. Ecology, 93: 23532364.Google Scholar
Schubauer, J. P., and Hopkinson, C. S. 1984. Above- and belowground emergent macrophyte production and turnover in a coastal marsh ecosystem, Georgia. Limnology and Oceanography, 29: 10521065.Google Scholar
Seiple, W. 1981. The ecological significance of the locomoter activity rhythms of Sesarma cinereum (Bosc) and Sesarma reticulatum (Say) (Decapoda, Grapsidae). Crustaceana, 40: 515.Google Scholar
Seliskar, D. M., Gallagher, J. L., Burdick, D. M., and Mutz, L. A. 2002. The regulation of ecosystem functions by ecotypic variation in a dominant plant: a Spartina alterniflora salt-marsh case study. Journal of Ecology, 90: 111.Google Scholar
Shields, J. D., and Squyars, C. M. 2000. Mortality and hematology of blue crabs, Callinectes sapidus, experimentally infected with the parasitic dinoflagelate Hematodinium perezi. Fishery Bulletin, 98: 139152.Google Scholar
Shumway, S. W., and Bertness, M. D. 1994. Patch size effects on marsh plant secondary succession mechanisms. Ecology, 75: 564568.Google Scholar
Silliman, B. R., and Bertness, M. D. 2002. A trophic cascade regulates salt marsh primary production. Proceedings of the National Academy of Science, USA, 99: 1050010505.Google Scholar
Silliman, B. R., and Bortolus, A. 2003. Underestimation of Spartina productivity in western Atlantic marshes: marsh invertebrates eat more than just detritus. Oikos, 101: 549554.Google Scholar
Silliman, B. R., and Newell, S. Y. 2003. Fungal farming in a snail. Proceedings of the National Academy of Science, USA, 100: 1564315648.Google Scholar
Silliman, B. R., Schrack, E., He, Q., Cope, R., Santoni, A., Van der Heide, T., Jacobi, R., and Van de Koppel, J. 2015. Facilitation shifts paradigms and can amplify coastal restoration efforts. Proceedings of the National Academy of Science, USA, 112: 1429514300.Google Scholar
Silliman, B. R., Van de Koppel, J., Bertness, M. D., Stanton, L. E., and Mendelssohn, I. A. 2005. Drought, snails, and large-scale die-off of southern U.S. salt marshes. Science 310: 18031806.Google Scholar
Silliman, B. R., and Zieman, J. C. 2001. Top-down control of Spartina alterniflora production by periwinkle grazing in a Virginia salt marsh. Ecology, 82: 28302845.Google Scholar
Srivastava, D. S., and Jefferies, R. L. 1996. A positive feedback: herbivory, plant growth, salinity, and the desertification of an Arctic salt-marsh. Journal of Ecology, 84: 3142.Google Scholar
Steever, E. Z., Warren, R. S., and Niering, W. A. 1976. Tidal energy subsidy and standing crop production of Spartina alterniflora. Estuarine and Coastal Marine Science, 4: 473478.Google Scholar
Stiling, P., and Moon, D. C. 2005. Quality or quantity: the direct and indirect effects of host plants on herbivores and their natural enemies. Oecologia, 142: 413420.Google Scholar
Stiling, P. D., and Strong, D. R. 1983. Weak competition among Spartina stem borers by means of murder. Ecology, 64: 770778.Google Scholar
Stiling, P. D., and Strong, D. R. 1984. Experimental density manipulation of stem-boring insects: some evidence for interspecific competition. Ecology, 65: 16831685.Google Scholar
Stiven, A. E., and Kuenzler, E. J. 1979. The response of two salt marsh molluscs, Littorina irrorata and Geukensia demissa, to field manipulations of density and Spartina litter. Ecological Monographs, 49: 151171.Google Scholar
Sullivan, G., Callaway, J. C., and Zedler, J. B. 2007. Plant assemblage composition explains and predicts how biodiversity affects salt marsh functioning. Ecological Monographs, 77: 569590.Google Scholar
Taylor, K. L., Grace, J. B., Guntenspergen, G. R., and Foote, A. L. 1994. The interactive effects of herbivory and fire on an oligohaline marsh, little lake, Louisiana, USA. Wetlands, 14: 8287.Google Scholar
Teal, J. M. 1958. Distribution of fiddler crabs in Georgia salt marshes. Ecology, 39: 185193.Google Scholar
Teal, J. M. 1962. Energy flow in the salt marsh ecosystem of Georgia. Ecology, 43: 614624.Google Scholar
Teal, J. M., and Howes, B. L. 1996. Interannual variability of a salt-marsh ecosystem. Limnology and Oceanography, 41: 802809.Google Scholar
Tilman, D., Wedin, D., and Knops, J. 1996. Productivity and sustainability influenced by biodiversity in grassland ecosystems. Nature, 379: 718720.Google Scholar
Tobias, C., and Neubauer, S. C. 2009. Salt marsh biogeochemistry: an overview.In: Perillo, G. M. E., Wolanski, E., Cahoon, D. R., and Brinson, M. M., eds., Coastal Wetlands: An Integrated Ecosystem Approach. Elsevier, Amsterdam, pp. 445492.Google Scholar
Travis, S. E., and Grace, J. B. 2010. Predicting performance for ecological restoration: a case study using Spartina alterniflora. Ecological Applications, 20: 192204.Google Scholar
Travis, S. E., and Hester, M. W. 2005. A space-for-time substitution reveals the long-term decline in genotypic diversity of a widespread salt marsh plant, Spartina alterniflora, over a span of 1500 years. Journal of Ecology, 93: 417430.Google Scholar
Turner, M. G. 1987. Effects of grazing by feral horses, clipping, trampling, and burning on a Georgia salt marsh. Estuaries, 10: 5460.Google Scholar
Turner, R. E. 1976. Geographic variations in salt marsh macrophyte production: a review. Contributions in Marine Science, 20: 4768.Google Scholar
Valiela, I., and Rietsma, C. S. 1995. Disturbance of salt marsh vegetation by wrack mats in Great Sippewissett Marsh. Oecologia, 102: 106112.Google Scholar
Valiela, I., Teal, J. M., and Deuser, W. G. 1978. The nature of growth forms in the salt marsh grass Spartina alterniflora. American Naturalist, 112: 461470.CrossRefGoogle Scholar
van de Koppel, J., van der Wal, D., Bakker, J. P., and Herman, P. M. J. 2005. Self-organization and vegetation collapse in salt marsh ecosystems. American Naturalist, 165: E1E12.Google Scholar
Van Der Wal, R., Van Wijnen, H., Van Wijnen, S., Beucher, O., and Bos, D. 2000. On facilitation between herbivores: how brent geese profit from brown hares. Ecology, 81: 969980.Google Scholar
Voss, C. M., Christian, R. R., and Morris, J. T. 2013. Marsh macrophyte responses to inundation anticipate impacts of sea-level rise and indicate ongoing drowning of North Carolina marshes. Marine Biology, 160: 181194.Google Scholar
Vu, H., Więski, K., and Pennings, S. C. 2017. Ecosystem engineers drive creek formation in salt marshes. Ecology, 98: 162174.Google Scholar
Wang, X. Y., Shen, D. W., Jiao, J., Xu, N. N. Yu, S. Zhou, X. F., Shi, M. M., and Chen, X. Y. 2012. Genotypic diversity enhances invasive ability of Spartina alterniflora. Molecular Ecology, 21: 25422551.Google Scholar
Wiegert, R. G., Chalmers, A. G., and Randerson, P. F. 1983. Productivity gradients in salt marshes: the response of Spartina alterniflora to experimentally manipulated soil water movement. Oikos, 41: 16.Google Scholar
Więski, K., Guo, H., Craft, C. B., and Pennings, S. C. 2010. Ecosystem functions of tidal fresh, brackish and salt marshes on the Georgia coast. Estuaries and Coasts, 33: 161169.Google Scholar
Więski, K., and Pennings, S. 2014a. Latitudinal variation in resistance and tolerance to herbivory of a salt marsh shrub. Ecography, 37: 763769.Google Scholar
Więski, K., and Pennings, S. C. 2014b. Climate drivers of Spartina alterniflora saltmarsh production in Georgia, USA. Ecosystems, 17: 473484.Google Scholar
Wilson, A. M., Evans, T., Moore, W., Schutte, C. A., Joye, S. B., Hughes, A. H., and Anderson, J. L. 2015. Groundwater controls ecological zonation of salt marsh macrophytes. Ecology, 96: 840849.Google Scholar
Windham, L., and Lathrop, R. G. Jr. 1999. Effects of Phragmites australis (common reed) invasion on aboveground biomass and soil properties in brackish tidal marsh of the Mullica River, New Jersey. Estuaries, 22: 927935.Google Scholar
Worm, B., Barbier, E. B., Beaumont, N., Duffy, J. E., Folke, C., Halpern, B. S., Jackson, J. B. C., Lotze, H. K., et al. 2006. Impacts of biodiversity loss on ocean ecosystem services. Science, 314: 787790.Google Scholar
Zacheis, A. M. Y., Hupp, J. W., and Ruess, R. W. 2001. Effects of migratory geese on plant communities of an Alaskan salt marsh. Journal of Ecology, 89: 5771.Google Scholar
Zerebecki, R. A., Crutsinger, G. M., and Hughes, A. R. 2017. Spartina alterniflora genotypic identity affects plant and consumer responses in an experimental marsh community. Journal of Ecology, 105: 661673.Google Scholar
Zhang, L., Wang, B., and Qi, L. 2017. Phylogenetic relatedness, ecological strategy, and stress determine interspecific interactions within a salt marsh community. Aquatic Science, 79: 587595.Google Scholar
Zheng, S., Shao, D., Asaeda, T., Sun, T., and Luo, S. 2016. Modeling the growth dynamics of Spartina alterniflora and the effects of its control measures. Ecological Engineering, 97: 144156.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
×