Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-2lccl Total loading time: 0 Render date: 2024-04-26T05:57:27.828Z Has data issue: false hasContentIssue false

8 - Salt Marsh Ecogeomorphic Processes and Dynamics

from Part II - Marsh Dynamics

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 are considered some of the most biologically diverse and ecologically important regions on Earth, containing thousands of species of robust salt-tolerant plants, crabs, fish, mollusks, zooplankton, algae, and bacteria. Isolated between topographic headlands, laterally continuous behind protective barriers, or associated with extensive delta landscapes, salt marshes are regulated by a variety of physical forces such as waves, tides, rivers, and storm surges, but they are also impacted by climatic variations in temperature and precipitation, riverine flooding, local tectonics, and subsidence (i.e., a deltaic process that describes the lowering of the land surface). Biological forces also play important roles in controlling salt marsh landscapes as many species shape geomorphic development. As these landscapes form and evolve, there exist significant interactions between biology, hydrology, and geology; thus it is impossible to consider salt marsh geomorphology – i.e., how the landscape changes over time – without taking into account these principal interactions.

Type
Chapter
Information
Salt Marshes
Function, Dynamics, and Stresses
, pp. 178 - 224
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

Addino, M. S., Montemayor, D. I., Escapa, M., Alvarez, M. F., Valiñas, M. S., Lomovasky, B. J., and Iribarne, O. 2015. Effect of Spartina alterniflora Loisel, 1807 on growth of the stout razor clam Tagelus plebeius (Lightfoot, 1786) in a SW Atlantic estuary. Journal of Experimental Marine Biology and Ecology, 463: 135142.Google Scholar
Alber, M., Swenson, E. M., Adamowicz, S. C., and Mendelssohn, I. 2008. Salt marsh dieback: An overview of recent events in the US. Estuarine, Coastal and Shelf Science, 80: 111.Google Scholar
Alberti, J., Escapa, M., Daleo, P., Iribarne, O., Silliman, B., and Bertness, M. 2007. Local and geographic variation in grazing intensity by herbivorous crabs in SW Atlantic salt marshes. Marine Ecology Progress Series, 349: 235243.Google Scholar
Alberti, J., Escapa, M., Iribarne, O., Silliman, B., and Bertness, M. 2008. Crab herbivory regulates plant facilitative and competitive processes in Argentinean marshes. Ecology, 89: 155164.Google Scholar
Allen, E., and Curran, H. A., 1974. Biogenic sedimentary structures produced by crabs in lagoon margin and salt marsh environments near Beaufort, North Carolina. Journal of Sedimentary Research, 44: 538548.Google Scholar
Allen, J. R. L. 2000. Morphodynamics of Holocene salt marshes: A review sketch from the Atlantic and Southern North Sea coasts of Europe. Quaternary Science Reviews, 19: 11551231.Google Scholar
Altieri, A. H., Silliman, B. R., and Bertness, M. D. 2007. Hierarchical organization via a facilitation cascade in intertidal cordgrass bed communities. The American Naturalist, 169: 195206Google Scholar
Anderson, M. E., Smith, J. M., and McKay, S. K. 2011. Wave dissipation by vegetation. USACOE Technical Report AD1003881.Google Scholar
Angelini, C., Griffin, J. N., Van de Koppel, J., Lamers, 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, doi:10.1038/ncomms12473Google Scholar
Angelini, C., Heide, T., Griffin, J. N., Morton, J. P., Derksen-Hooijberg, M., Lamers, L. P. M., Smolders, A. J. P., and Silliman, B. R. 2015. Foundation species’ overlap enhances biodiversity and multifunctionality from the patch to landscape scale in southeastern United States salt marshes. Proceedings of the Royal Society of London B, 282: 20150421.Google Scholar
Augustin, L. N., Irish, J. L., and Lynett, P. 2009. Laboratory and numerical studies of wave damping by emergent and near-emergent wetland vegetation. Coastal Engineering, 56: 332340.CrossRefGoogle Scholar
Austen, I., Andersen, T. J., and Edelvang, K. 1999. The influence of benthic diatoms and invertebrates on the erodibility of an intertidal mudflat, the Danish Wadden Sea. Estuarine Coastal and Shelf Science, 49: 99111.Google Scholar
Bartholdy, J. 2012. Salt marsh sedimentation. In: Davis, R. A., and Dalrymple, R. W., eds, Principals of Tidal Sedimentology, pp. 151–185.Google Scholar
Bayne, B. L. 2017. Biology of oysters. Developments in Aquaculture and Fisheries Science. Vol. 41: 2844.Google Scholar
Benner, R., Fogel, M. L., and Sprague, E. K. 1991. Diagenesis of belowground biomass of Spartina alterniflora in salt‐marsh sediments. Limnology and Oceanography, 36: 13581374.Google Scholar
Bertness, M. D. 1984a. Habitat and community modification by an introduced herbivorous snail. Ecology, 65: 370381.Google Scholar
Bertness, M. D. 1984b. 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.Google Scholar
Bertness, M. D., Gough, L., and Shumway, S. 1992. Salt tolerances and the distribution of fugitive salt marsh plants. Ecology, 73: 18421851.Google Scholar
Bertness, M. D., Holdredge, C., and Altieri, A. H. 2009. Substrate mediates consumer control of salt marsh cordgrass on Cape Cod, New England. Ecology, 90: 21082117.Google Scholar
Bertness, M. D., and Miller, T. 1984. The distribution and dynamics of Uca pugnax (Smith) burrows in a New England salt marsh. Journal of Experimental Marine Biology and Ecology, 83: 211237.Google Scholar
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. 10.1002/ecs2.1795Google Scholar
Blum, L. K., and Davey, E. 2013. Below the saltmarsh surface: visualization of plant roots by computer-aided tomography. Oceanography, 26: 8587.Google Scholar
Boorman, L. A, Garbutt, A., and Barratt, D. 1998. The role of vegetation in determining patterns of the accretion of salt marsh sediment. In: Black, K. S., Paterson, D. M., and Cramp, A., eds, Sedimentary Processes in the Intertidal Zone. Geological Society (London), Special Publication No 139, pp. 389399.Google Scholar
Botto, F., and Iribarne, O. 2000. Contrasting effects of two burrowing crabs (Chasmagnathus granulata and Uca uruguayensis) on sediment composition and transport in estuarine environments. Estuarine, Coastal and Shelf Science, 51: 141151.Google Scholar
Botto, F., Iribarne, O., Gutierres, J., Bava, J., Gagliardini, A., and Valiela, I. 2006. Ecological importance of passive deposition of organic matter into burrows of the SW Atlantic crab Chasmagnathus granulatus. Marine Ecology Progress Series, 312: 201210.Google Scholar
Boudreau, B. P., and Imboden, D. M. 1987. Mathematics of tracer mixing in sediments III: The theory of nonlocal mixing within sediments. American Journal of Science, 287: 693719.Google Scholar
Bouma, T. J., Friedrichs, M., Van Wesenbeeck, B. K., Temmerman, S., Graf, G., and Herman, P. M. J. 2009. Density-dependent linkage of scale-dependent feedbacks : A flume study on the intertidal macrophyte Spartina anglica. Oikos, 118: 260268.CrossRefGoogle Scholar
Cadee, G. C. 2001. Sediment dynamics by bioturbating organisms. In: Ecological Comparisons of Sedimentary Shores, Springer-Verlag, Berlin, pp. 127147.Google Scholar
Chapman, V. J. 1960. Salt Marshes and Salt Deserts of the World. Leonard Hill: London.Google Scholar
Chmura, G. L., and Kosters, E. C. 1994. Storm deposition and 137Cs accumulation in fine-grained marsh sediments of the Mississippi Delta Plain. Estuarine Coastal and Shelf Science, 39: 3344.Google Scholar
Christiansen, T., Wiberg, P. L., and Milligan, T. G. 2000. Flow and sediment transport on a tidal salt marsh surface. Estuarine, Coastal and Shelf Science, 50: 315331.Google Scholar
Coverdale, T. C., Altieri, A. H. and Bertness, M. D. 2012. Belowground herbivory increases vulnerability of New England salt marshes to die-off. Ecology, 93: 20852094.CrossRefGoogle ScholarPubMed
Collins, L. M., Collins, J. N., and Leopold, L. B. 1987. Geomorphic processes of an estuarine marsh: preliminary hypotheses. In: Gardiner, V., ed., International Geomorphology, Part 1, Wiley, Chichester, pp. 10491072.Google Scholar
Costanza, R., Pérez-Maqueo, O., Martinez, M. L., Sutton, P., Anderson, S. J., and Mulder, K. 2008. The value of coastal wetlands for hurricane protection. AMBIO: A Journal of the Human Environment, 37: 241248.Google Scholar
Crawford, F. 2018. Geomorphology of shell ridges and their effect on the stabilization of Biloxi marshes, east Louisiana. MS thesis, University of New Orleans.Google Scholar
Cuadrado, D. G., Perillo, G. M. E. and Vitale, A. J. 2014. Modern microbial mats in siliciclastic tidal flats: Evolution, structure and the role of hydrodynamics. Marine Geology, 352: 367380Google Scholar
Currin, C. A., Chappell, W. S., and Deaton, A. 2010. Developing alternative shoreline armoring strategies: The living shoreline approach in North Carolina. In: Shipman, H., Dethier, M. N., Gelfenbaum, G., Fresh, K. L., and Dinicola, R. S., eds., Puget Sound Shorelines and the Impacts of Armoring – Proceedings of a State of the Science Workshop, May 2009, Reston, Virginia: U.S. Geological Survey, Scientific Investigations Report 2010-5254, pp. 91102.Google Scholar
Daborn, G. R., Amos, C. L., Brylinsky, M., Christian, H., Drapeau, G, Faas, R. W., Grant, J., et al. 1993. An ecological cascade effect: migratory birds affect stability of intertidal sediments. Limnology and Oceanography, 38: 225231.Google Scholar
Da Cunha Lana, P., and Guiss, C. 1992. Macrofauna-plant-biomass interactions in a euhaline salt marsh in Paranagua Bay (SE Brazil). Marine Ecology Progress Series, 80: 5764.Google Scholar
Darby, F. A., and Turner, R. E. 2008a. Effects of eutrophication on salt marsh root and rhizome biomass accumulation. Marine Ecology Progress Series, 363: 6370.Google Scholar
Darby, F. A., and Turner, R. E. 2008b. Below- and aboveground biomass of Spartina alterniflora: response to nutrient addition in a Louisiana Salt Marsh. Estuaries and Coasts, 31: 326334.CrossRefGoogle Scholar
Davenport, T. M., Seitz, R. D., Knick, K. E., and Jackson, N. 2018. Living shorelines support nearshore benthic communities in Upper and Lower Chesapeake Bay. Estuaries and Coasts, 41: 197206.Google Scholar
Davey, E., Wigand, C., Johnson, R., Sundberg, K., Morris, J., Roman, C. T., Davey, E. et al. 2011. Use of computed tomography imaging for quantifying coarse roots, rhizomes, peat, and particle densities in marsh soils. Ecological Applications, 21: 21562171.Google Scholar
Day, J. W. Jr, Kemp, G. P., Reed, D. J., Cahoon, D. R., Boumans, R. M., Suhayda, J. M., and Gambrell, R.. 2011. Vegetation death and rapid loss of surface elevation in two contrasting Mississippi delta salt marshes: the role of sedimentation, autocompaction and sea level rise. Ecological Engineering, 37: 228240.Google Scholar
Deegan, L. A., Johnson, D. S., Warren, R. S., Peterson, B. J., Fleeger, J. W., Fagherazzi, S., and Wollheim, W. 2012. Coastal eutrophication as a driver of salt marsh loss. Nature, 490: 388394.Google Scholar
Delaune, R. D., Baumann, R. H., and Gosselink, J. G. 1983. Relationships among vertical accretion, coastal submergence, and erosion in a Louisiana Gulf Coast marsh. Journal of Sedimentary Research, 53: 147157.Google Scholar
Delaune, R. D., Nyman, J. A., and Patrick, W. H. 1994. Peat collapse, ponding and wetland loss in a rapidly submerging coastal marsh. Journal of Coastal Research, 10: 10211030.Google Scholar
Delaune, R. D., Patrick, W. H., and Buresh, R. J. 1978. Sedimentation rates determined by 137Cs dating in a rapidly accreting salt marsh. Nature, 275: 532533.Google Scholar
Dionne, J.-C. 1985. Tidal marsh erosion by geese, St. Lawrence Estuary, Québec. Géographie Physique et Quaternaire, 39: 99105.Google Scholar
Dupraz, C. and Visscher, P. T. 2005. Microbial lithification in marine stromatolites and hypersaline mats. Trends in Microbiology, 13: 429438.CrossRefGoogle ScholarPubMed
Dyer, K. R. 1988. Fine sediment particle transport in estuaries. In: Dronkers, J., and van Leussen, W., eds, Physical Processes in Estuaries. Springer, Berlin, Heidelberg, pp. 295310.Google Scholar
Ellison, A. M., Bertness, M. D., and Miller, T. 1986. Seasonal patterns in the belowground biomass of Spartina alterniflora (Gramineae) across a tidal gradient. American Journal of Botany, 73: 15481554.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
Escapa, M., Perillo, G. M. E., and Iribarne, O. 2008. Sediment dynamics modulated by burrowing crab activities in contrasting SW Atlantic intertidal habitats. Estuarine, Coastal and Shelf Science, 80: 365373.Google Scholar
Escapa, C. M., Perillo, G. M. E., Iribarne, O. 2015. Biogeomorphically driven salt pan formation in Sarcocornia-dominated salt-marshes. Geomorphology, 228: 147157.Google Scholar
Fagherazzi, S., FitzGerald, D. M., Fulweiler, R. W., Hughes, Z., Wiberg, P. L., McGlathery, K. J., Morris, J. T., Tolhurst, T. J., Deegan, L. A. and Johnson, D. S. 2013a. Ecogeomorphology of salt marshes. Treatise on Geomorphology, 12: 182212.Google Scholar
Fagherazzi, S., Wiberg, P. L., Temmerman, S., Struyf, E., Zhao, Y., and Raymond, P. A. 2013b. Fluxes of water, sediments, and biogeochemical compounds in salt marshes. Ecological Processes, 2: 116.Google Scholar
Fanjul, E., Grela, M. A., and Iribarne, O. 2007. Effects of the dominant SW Atlantic intertidal burrowing crab Chasmagnathus granulatus on sediment chemistry and nutrient distribution. Marine Ecology Progress Series, 341: 177190.CrossRefGoogle Scholar
Farron, S. J. 2018. Morphodynamic responses of salt marshes to sea-level rise: upland expansion, drainage evolution, and biological feedbacks. PhD Dissertation, Boston University, 159 p.Google Scholar
FitzGerald, Duncan M., Fenster, M.S., Argow, B.A., and Buynevich, I.V. 2008. Coastal impacts due to sea-level rise. Annual Review of Earth and Planetary Sciences, 36: 601647.Google Scholar
Ford, Mark A., and Grace, James B.. 1998. Effects of vertebrate herbivores on soil processes, plant biomass, litter accumulation and soil elevation changes in a coastal marsh. Journal of Ecology, 86: 974982.Google Scholar
French, J. R., and Spencer, T. 1993. Dynamics of sedimentation in a tide-dominated backbarrier salt marsh, Norfolk, UK. Marine Geology, 110: 315331.Google Scholar
French, J. R., and Stoddart, D. R. 1992. Hydrodynamics of Salt Marsh Creek Systems: Implications for Marsh Morphological Development and Material Exchange. Earth Surface Processes and Landforms 17: 235252.Google Scholar
Friedrichs, C. T., and Perry, J. E. 2001. Tidal salt marsh morphodynamics: A synthesis. Journal of Coastal Research, Special issue, no. 27: 7–37.Google Scholar
Gauthier, G., Giroux, J. F., Reed, , Béchet, A., and Bélanger, L. 2005. Interactions between land use, habitat use, and population increase in Greater Snow Geese: What are the consequences for natural wetlands? Global Change Biology, 11: 856868.Google Scholar
Gedan, K. B., Kirwan, M. L., Wolanski, E., Barbier, E. B., and Silliman, B. R. 2011. The present and future role of coastal wetland vegetation in protecting shorelines: answering recent challenges to the paradigm. Climatic Change, 106: 729.Google Scholar
Genoni, G. P. 1991. Increased burrowing by fiddler crabs Uca rapax (Smith) (Decapoda : Ocypodidae) in response to low food supply. Journal of Experimental Marine Biology and Ecology, 147: 267285.Google Scholar
Giosan, L., Syvitski, J., Constantinescu, S., and Day, J. 2014. Climate change: protect the world’s deltas. Nature, 516: 3133.Google Scholar
Gleason, M. L., Elmer, D. A., Pien, N. C., Fisher, J. S. 1979. Effects of stem density upon sediment retention by salt marsh cord grass, Spartina alterniflora Loisel. Estuaries, 2: 271273.Google Scholar
Grant, J., Bathmann, U. V., and Mills, E. L. 1986. The interaction between benthic diatom films and sediment transport. Estuarine Coastal Shelf Science, 23: 225238.Google Scholar
Gribsholt, B., Kostka, J. E., and Kristensen, E. 2003. Impact of fiddler crabs and plant roots on sediment biogeochemistry in a Georgia saltmarsh. Marine Ecology Progress Series, 259: 237251.CrossRefGoogle Scholar
Gross, M. F., Hardisky, M. A., Wolf, P. L., and Klemas, V. 1991. Relationship between aboveground and belowground biomass of Spartina alterniflora (Smooth Cordgrass). Estuaries, 14: 180191.Google Scholar
Gutiérrez, J. L., and Iribarne, O. 1999. Role of Holocene beds of the stout razor clam Tagelus plebeius in structuring present benthic communities. Marine Ecology Progress Series, 185: 213228.Google Scholar
Gutiérrez, J. L., Jones, C. G., Groffman, P. M., Findlay, S. E. G, Iribarne, O., Ribeiro, P. D., and Bruschetti, C. M. 2006. The contribution of crab burrow excavation to carbon availability in surficial salt-marsh sediments. Ecosystems, 9: 647658.Google Scholar
Gutiérrez, J. L., Jones, C. G., Strayer, D. L., and Iribarne, O. 2003. Mollusks as ecosystem engineers : The role of shell production in aquatic habitats. Oikos, 101: 7990.Google Scholar
Hannaford, J.Pinn, E. H., and Diaz, A. 2006. The impact of sika deer grazing on the vegetation and infauna of Arne saltmarsh. Marine Pollution Bulletin, 53: 5662.Google Scholar
Hatton, R. S., DeLaune, R. D., and Patrick, W. H. Jr. 1983. Sedimentation, accretion, and subsidence in marshes of Barataria Basin, Louisiana. Limnology and Oceanography, 28: 494502.Google Scholar
Hazelden, J., and Boorman, L. A. 2001. Soils and “managed retreat” in South East England. Soil Use and Management, 17: 150154.Google Scholar
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.Google Scholar
Holm, G. O. 2006. Nutrient constraints on plant community production and organic matter accumulation of subtropical floating marshes. PhD dissertation, Louisiana State University, Baton Rouge, Louisiana.Google Scholar
Hopkinson, C. S., Gosselink, J. G., and Parrondo, R. T. 1980. Production of coastal louisiana marsh plants calculated from phenometric techniques Ecology, 61: 10911098.Google Scholar
Howe, A. J., Rodriguez, J. F., and Saco, P. M. 2009. Surface evolution and carbon sequestration in disturbed and undisturbed wetland soils of the Hunter estuary, southeast Australia. Estuarine Coastal Shelf Science, 84: 7583.Google Scholar
Howes, B. L., Goehringer, D. D., and Macey, J. W. H. 1986 Factors controlling the growth form of Spartina alterniflora: feedbacks between above-ground production, sediment oxidation, nitrogen and salinity. Journal of Ecology, 74: 881898.Google Scholar
Howes, B. L., Howarth, R. W., Teal, J. M., and Valiela, I. 1981. Oxidation-reduction potentials in a saltmarsh: spatial patterns and interactions with primary production. Limnology and Oceanography, 26: 350360.Google Scholar
Howes, N. C., FitzGerald, D. M., Hughes, Z. J., Georgiou, I. Y., Kulp, M. A., Miner, M. D., Smith, J. M., and Barras, J. A. 2010. Hurricane-induced failure of low salinity wetlands. Proceedings of the National Academy of Sciences of the USA, 107: 1401414019.Google Scholar
Hu, K., Chen, Q., and Wang, H. 2015. A numerical study of vegetation impact on reducing storm surge by wetlands in a semi-enclosed estuary. Coastal Engineering, 95: 6676.Google Scholar
Hughes, Z. J., FitzGerald, D. M., Wilson, C. A., Pennings, S. C. Wiçski, K., and Mahadevan, A. 2009. Rapid headward erosion of marsh creeks in response to relative sea level rise. Geophysical Research Letters, 36: 15.Google Scholar
Hyun, J., Smith, A. C., and Kostka, J. E. 2007. Relative contributions of sulfate- and iron(III) reduction to organic matter mineralization and process controls in contrasting habitats of the Georgia saltmarsh. Applied Geochemistry, 22: 26372651.Google Scholar
Iribarne, O., Bortolus, A., Botto, F. 1997. Between-habitat differences in burrow characteristics and trophic modes in the south western Atlantic burrowing crab Chasmagnathus granulata. Marine Ecology Progress Series, 155: 137145.Google Scholar
Jaramillo, E., and Lunecke, K. 1988. The role of sediments in the distribution of Uca pugilator (Bosc) and Uca pugnax (Smith) (Crustacea, Brachyura) in a salt marsh at Cape Cod. Meeresforschung, 32: 4652.Google Scholar
Jones, C. G., Lawton, J. H., and Shachak, M. 1994. Organisms as ecosystem engineers. Oikos 69: 373386.Google Scholar
Jordan, T., and Valiela, I. 1982. A nitrogen budget of the ribbed mussel, Geukensia demissa, and its significance in nitrogen flow in a New England salt marsh. Limnology and Oceanography, 27: 7590.Google Scholar
Julien, A. 2018. Quantifying the demographics, habitat characteristics, and foundation species role of the ribbed mussel (Geukensia demissa) in South Carolina salt marshes. Masters thesis, Dept of Biology, College of Charleston, South Carolina.Google Scholar
Katrak, G., Dittmann, S., and Seurant, L. 2008. Spatial variation in burrow morphology of the mud shore crab Helograpsus haswellianus (Brachyura, Grapsidae) in South Australian saltmarshes. Marine and Freshwater Research, 59, 902911.Google Scholar
Katz, L. C. 1980. Effects of burrowing by the fiddler crab, Uca pugnax (Smith). Estuarine Coastal and Marine Science, 11: 233237.Google Scholar
Kaye, C. A., and Barghoorn, E. S. 1964. Late Quaternary sea-level change and crustal rise at Boston Massachusetts, with notes on autocompaction of peat. GSA Bulletin, 75: 6380.Google Scholar
Kennish, M. J. 2001. Coastal salt marsh systems in the US: a review of anthropogenic impacts. Journal of Coastal Research, 17: 731748.Google Scholar
Kesel, R. H., Yodis, E. G., and McCraw, D. J. 1992. An approximation of the sediment budget of the Lower Mississippi River prior to major human modification. Earth Surface Processes and Landforms, 17: 711722.Google Scholar
Keusenkothen, M. A. 2002. The effects of deer trampling in a salt marsh. MS thesis, East Carolina University.Google Scholar
Keusenkothen, M. A., and Christian, R. R. 2004. Responses of salt marshes to disturbance in an ecogeomorphological context, with a case study of trampling by deer. In: Fagherazzi, S., Marani, M., and Blum, L., eds., The Ecogeomorphology of Tidal Marshes, Volume 59. John Wiley, Hoboken, NJ, pp. 203230.Google Scholar
King, G. M., Klug, M. J., Wiegert, R. G., and Chalmers, A. G. 1982. Relation of soil water movement and sulfide concentration of Spartina alterniflora production in a Georgia Salt Marsh. Science, 218: 6163.Google Scholar
Kirchner, J. W., Dietrich, W. E., Iseya, F., and Ikeda, H. 1990. The variability of critical shear stress, friction angle, and grain protrusion in water-worked sediments. Sedimentology, 37: 647672.Google Scholar
Kirwan, M. L., and Guntenspergen, G. R. 2010. The influence of tidal range on the stability of coastal marshland. Journal of Geophysical Research, 115: F02009, doi:10.1029/2009JF001400.Google Scholar
Kirwan, M. L., Guntenspergen, G. R., D’Alpaos, A., Morris, J., Mudd, S. M., and Temmerman, S. 2010. Limits on the adaptability of coastal marshes to rising sea level. Geophysical Research Letters, 37(23). https://doi.org/10.1029/2010GL045489Google Scholar
Kirwan, M. L., and Megonigal, J. P. 2013. Tidal wetland stability in the face of human impacts and sea-level rise. Nature, 504: 5360.Google Scholar
Kirwan, M. L., and Temmerman, S. 2009. Coastal marsh response to historical and future sea-level acceleration. Quaternary Science Reviews, 28: 18011808.Google Scholar
Kirwan, M. L., Temmerman, S., Skeehan, E. E., Guntenspergen, G. R., and Fagherazzi, S. 2016. Overestimation of marsh vulnerability to sea level rise. Nature Climate Change, 6: 253260.Google Scholar
Knutson, P. L., Brochu, R. A., and See, W. N. 1982. Wave damping in Spartina alterniflora marshes. Wetlands, 2: 87104.Google Scholar
Kobayashi, N., Raichle, A., and Asano, T. 1993. Wave attenuation by vegetation. Journal of Waterway, Port, Coastal, and Ocean Engineering Technical Report, 119(1). https://doi.org/10.1061/(ASCE)0733-950X(1993)119:1(30)Google Scholar
Koch, E. W., Barbier, E. D., Silliman, B. R., Reed, D. J., Perillo, G. M. E., Hacker, S. D., Granek, E. F., et al. 2009. Non-linearity in ecosystem services: temporal and spatial variability in coastal protection. Frontiers in Ecology and the Environment, 7: 2937.Google Scholar
Kokot, R. R. 2004. Erosión en la costa patagónica por cambio climático. Revista de la Asociación Geológica Argentina, 59: 715726.Google Scholar
Koo, B. J., Kwon, K. K., and Hyun, J. H. 2005. The sediment-water interface increment due to the complex burrows of macrofauna in a tidal flat. Ocean Science Journal, 40: 221227.Google Scholar
Kostka, J. E., Gribsholt, B., Petrie, E., Dalton, D., Skelton, H., and Kristensen, E. 2002. The rates and pathways of carbon oxidation in bioturbated saltmarsh sediments. Limnology and Oceanography, 47: 230240.Google Scholar
Kraeuter, J. N. 1976. Biodeposition by salt-marsh invertebrates. Marine Biology, 35: 215223.Google Scholar
Laegdsgaard, P. 2006. Ecology, disturbance and restoration of coastal saltmarsh in Australia: A Review. Wetlands Ecology and Management, 14: 379399.Google Scholar
Langley, J. A., McKee, K. L., Cahoon, D. R., Cherry, J. A. and Megonigal, P. 2009. Elevated CO2 stimulates marsh elevation gain, counterbalancing sea-level rise. Proceedings of the National Academy of Sciences of the USA, 106: 61826186.Google Scholar
Lanuru, M. 2008. Measuring critical erosion shear stress of intertidal sediments with EROMES erosion device. Torani Journal of Marine Science and Fisheries, 18: 390397.Google Scholar
Leeder, M. R. 1998. Lyell’s Principles of Geology: Foundations of sedimentology. Geological Society of London, Special Publication, 143: 95110.Google Scholar
Leonard, L. A., and Luther, M. E. 1995. Flow hydrodynamics in tidal marsh cano-pies. Limnology and Oceanography 18: 14741484.Google Scholar
Letzsch, W. S., and Frey, R. W. 1980. Deposition and erosion in a Holocene salt marsh, Sapelo Island, Georgia. Journal of Sedimentary Research, 50: 529542.Google Scholar
Li, H., and Yang, S. L. 2009. Trapping effect of tidal marsh vegetation on suspended sediment, Yangtze Delta. Journal of Coastal Research, 25: 915924.Google Scholar
Lightbody, A. F., and Nepf, H. M. 2006. Prediction of velocity profiles and longitudinal dispersion in salt marsh vegetation. Limnology and Oceanography, 51: 218228.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
Madsen, K. N., Nilsson, P., and Sunback, K. 1993. The influence of benthic microalgae on the stability of a subtidal sediment. Journal of Experimental Marine Biology and Ecology, 170: 159177.Google Scholar
Madsen, J. D., Chambers, P. A., James, W. F., Koch, E. W., and Westlake, D. F. 2001. The interaction between water movement, sediment dynamics and submersed macrophytes. Hydrobiologia, 444: 7184.Google Scholar
Marani, M., D’Alpaos, A., Lanzoni, S., Carniello, L., and Rinaldo, A. 2007. Biologically-controlled multiple equilibria of tidal landforms and the fate of the Venice lagoon. Geophysical Research Letters, 34: 15.Google Scholar
Mariotti, G., and Carr, J. 2014. Dual role of salt marsh retreat: Long‐term loss and short‐term resilience. Water Resources Research, 50: 29632974.CrossRefGoogle Scholar
Mariotti, G., and Fagherazzi, S. 2010. A numerical model for the coupled long‐term evolution of salt marshes and tidal flats. Journal of Geophysical Research Earth Surface, 115 (F1). https://doi.org/10.1029/2009JF00132.Google Scholar
Mariotti, G., and Fagherazzi, S. 2013. A two-point dynamic model for the coupled evolution of channels and tidal flats. Journal of Geophysical Research, 118: 13871399.Google Scholar
Mariotti, G., Kearney, W., and Fagherazzi, S. 2016. Soil creep in salt marshes. Geology, 44: 459462.Google Scholar
McCraith, B. J., Gardner, L. R., Wethey, D. S., and Moore, W. S. 2003. The effect of fiddler crab burrowing on sediment mixing and radionuclide profiles along a topographic gradient in a southeastern salt marsh. Journal of Marine Research, 61: 359390.Google Scholar
Mehta, A. J. 1996. Interaction between fluid mud and water waves. In: Singh, V. P., and Hager, W. H., Environmental Hydraulics. Kluwer, Dordretcht, pp. 153187.Google Scholar
Melo, W. D., Perillo, G. M. E., Perillo, M. M., Schilizzi, R., and Piccolo, M. C. 2013. Late Pleistocene-Holocene deltas in the southern Buenos Aires Province, Argentina. In: Young, G., and Perillo, G. M. E., eds., Deltas: Landforms, Ecosystems and Human Activities. IAHS Press, Wallingford, UK. 358: 187195.Google Scholar
Melo, W. D., Schillizzi, R., Perillo, G. M. E. y Piccolo, M. C. 2003. Influencia del área continental pampeana sobre el origen y la morfología del estuario de Bahía Blanca. Revista de la Asociación Argentina de Sedimentología, 10: 6572.Google Scholar
Mendelssohn, I. A., McKee, K. L., and Patrick, W. H. Jr 1981. Oxygen deficiency in Spartina alterniflora roots: metabolic adaptation to anoxia. Science, 214: 439441.Google Scholar
Mendelssohn, I. A., and Seneca, E. D. 1980. The influence of soil drainage on the growth of salt marsh cordgrass Spartina alterniflora in North Carolina. Estuarine and Coastal Marine Science, 11: 2740.Google Scholar
Meyer, D. L., Townsend, E. C., and Thayer, G. W. 1997. Stabilization and erosion control value of oyster cultch for intertidal marsh. Restoration Ecology, 5: 9399.Google Scholar
Millette, T., Argow, B., Marcano, E., Hayward, C., Hopkinson, C., and Valentine, V. 2010. Salt marsh geomorphological analyses via integration of multitemporal multispectralremote sensing with LIDAR and GIS. Journal of Coastal Research, 26: 809816.Google Scholar
Minkoff, D. R., Escapa, M., Ferramola, F. E., Maraschín, S. D., Pierini, J. O., Perillo, G. M. E., and Delrieux, C. 2006. Effects of crab–halophytic plant interactions on creek growth in a S. W. Atlantic salt marsh: a cellular automata model. Estuarine, Coastal and Shelf Science, 69: 403413.Google Scholar
Minkoff, D. R., Escapa, C. M., Ferramola, F. E., and Perillo, G. M. E. 2005. Erosive processes due to physical – biological interactions based in a cellular automata model. Latin American Journal of Sedimentology and Basin Analysis, 12: 2534.Google Scholar
Mitsch, W. J., and Gosselink, J. G. 2000. Wetlands. 3rd edition. John Wiley, New York.Google Scholar
Molina, L. M., Valiñas, M. S., Pratolongo, P., Elias, R., and Perillo, G. M. E. 2017. Effect of Micropogonias furnieri on the stability of the sediment of salt marshes – an issue to be resolved. Estuaries and Coasts, 40: 17951807.Google Scholar
Moller, I., Kudella, M., Rupprecht, F., Spencer, T., Paul, M., van Wesenbeeck, B. K., Wolters, G., et al. 2014. Wave attenuation over coastal salt marshes under storm surge conditions. Nature Geoscience, 7: 727731.Google Scholar
Moller, I., and Spencer, T. 2002. Wave dissipation over macro-tidal saltmarshes: Effects of marsh edge typology and vegetation change. Journal of Coastal Research: Special Issue 36 – International Coastal Symposium (ICS 2002): 506521.Google Scholar
Montague, C. L. 1980. A natural history of temperate Western Atlantic fiddler crabs (Genus Uca) with reference to their impact on the salt marsh. Contributions in Marine Science, 23: 2555.Google Scholar
Montague, C. L. 1982. The influence of fiddler crab burrowing on metabolic processes in saltmarsh sediments. In: Kennedy, V. S., ed., Estuarine Comparisons. Academic Press, San Francisco, pp. 283301.Google Scholar
Montgomery, D. R., and Dietrich, W. E. 1988. Where do channels begin? Nature, 336: 232234.Google Scholar
Morris, J. M., 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
Mouton, E C., and Felder, D. L. 1996. Burrow distributions and population estimates for the fiddler crabs Uca spinicarpa and Uca longisignalis in a Gulf of Mexico salt marsh. Estuaries and Coasts, 19: 5161.Google Scholar
Murray, J. M. H., Meadows, A., and Meadows, P. S. 2002. Biogeomorphological implications of microscale interactions between sediment geotechnics and marine benthos: A review. Geomorphology, 47: 1530.Google Scholar
Nelson, J., Wilson, R., Coleman, F., Koenig, C., DeVries, D., Gardner, C., and Chanton, J. 2012. Flux by fin: fish-mediated carbon and nutrient flux in the northeastern Gulf of Mexico. Marine Biology, 159: 365372.Google Scholar
Nepf, H. M. 1999. Drag, turbulence, and diffusion in flow through emergent vegetation. Water Resources Research, 35: 479489.Google Scholar
Neumeier, U., and Amos, C. L. 2006. The influence of vegetation on turbulence and flow velocities in European salt‐marshes. Sedimentology, 53: 259277.Google Scholar
Neumeier, U., and Ciavola, P. 2004. Flow resistance and associated sedimentary processes in a Spartina maritima salt-marsh. Journal of Coastal Research, 20: 435447.Google Scholar
Niering, W., and Warren, R. S. 1980. Vegetation patterns and processes in New England Salt Marshes. BioScience, 30: 301307.Google Scholar
NOAA. 2018. Green Infrastructure Effectiveness Database: https://coast.noaa.gov/digitalcoast/training/gi-database.htmlGoogle Scholar
Nyman, J. A., Crozier, C., and DeLaune, R. D. 1995a. Roles and patterns of hurricane sedimentation in an estuarine marsh landscape. Estuaries, Coastal and Shelf Science, 40: 665679.Google Scholar
Nyman, J. A., Delaune, R. D., Patrick, W. H. Jr. 1990. Wetland soil formation in the rapidly subsiding Mississippi River Deltaic Plain: mineral and organic matter relationships. Estuarine, Coastal and Shelf Science, 31: 5769.Google Scholar
Nyman, J. A., DeLaune, R. D., Pezeshki, S. R., and Patrick, W. H. Jr. 1995b. Organic matter cycling and marsh stability in a rapidly submerging estuarine marsh. Estuaries, 18: 207218.Google Scholar
Nyman, J. A., Walters, R., Delaune, R. D., Patrick, W. H. Jr. 2006. Marsh vertical accretion via vegetative growth. Estuaries Coastal and Shelf Science, 69: 370380.Google Scholar
O’Donnell, J. E. D. 2017. Living shorelines: A review of literature relevant to New England coasts. Journal of Coastal Research, 332: 435451.Google Scholar
Odum, W. 1988. Comparative ecology of tidal freshwater and salt marshes. Annual Review of Ecology and Systematics, 19: 147176.Google Scholar
Onorevole, K. M., Thompson, S. P., and Piehler, M. F. 2018. Living shorelines enhance nitrogen removal capacity over time. Ecological Engineering, 120: 238248.Google Scholar
Pan, J., Bournod, C. N., Pizani, N. V., Cuadrado, D. G., and Carmona, N. B. 2013. Characterization of microbial mats from a siliciclastic tidal flat (Bahía Blanca Estuary, Argentina). Geomicrobiology Journal, 30: 665674.Google Scholar
Pennings, S. C., and Callaway, R. 1992. Salt marsh plant zonation: the relative importance of competition and physical factors. Ecology, 73: 681690.Google Scholar
Pennings, S. C., Carefoot, T. H., Siska, E. L., Chase, M.G., and Page, T. A. 1998. Feeding preferences of a generalist salt-marsh crab: relative importance of multiple plant traits. Ecology, 79: 19681979.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
Perillo, G. M. E. 2019. Geomorphology of tidal courses and depressions. In: Perillo, G. M. E., Wolanski, E., Cahoon, D. R., and Hopkinson, C., eds., Coastal Wetlands: An Integrated Ecosystem Approach. Elsevier, Amsterdam, pp. 185210.Google Scholar
Perillo, G. M. E., Drapeau, G., Piccolo, M. C., and Chaouq, N. 1993. Tidal circulation pattern on a tidal flat, Minas Basin, Canada. Marine Geology, 112: 219236Google Scholar
Perillo, G. M. E., and Iribarne, O. 2003a. New mechanisms studied for creek formation in tidal flats: from crabs to tidal channels. EOS American Geophysical Union Transactions, 84: 15.Google Scholar
Perillo, G. M. E., and Iribarne, O. 2003b. Processes of tidal channels develop in salt and freshwater marshes. Earth Surface Processes and Landforms, 28: 14731482.Google Scholar
Perillo, G. M. E., Minkoff, D. R., and Piccolo, M. C. 2005. Novel mechanism of stream formation in coastal wetlands by crab–fish–groundwater interaction. Geo-Marine Letters, 25: 214220.Google Scholar
Pestrong, R. 1965. The development of drainage patterns on tidal marshes. Stanford University Publications. Earth Science, 10(2): 1–87.Google Scholar
Pestrong, R. 1972. Tidal-flat sedimentation at cooley landing, Southwest San Francisco bay. Sedimentary Geology, 8: 251288.Google Scholar
Peterson, G. W., and Turner, R. E. 1994. The value of salt marsh edge vs interior as a habitat for fish and decapod crustaceans in a Louisiana tidal marsh. Estuaries, 17: 235262.Google Scholar
Pethick, J. S. 1980. Velocity surges and asymmetry in tidal channels. Estuarine Coastal and Marine Science, 11: 331345.Google Scholar
Pethick, J. S. 1992. Saltmarsh geomorphology. In: Allen, J. R. L., and Pye, K., eds, Saltmarshes: Morphodynamics, Conservation and Engineering Significance, Cambridge University Press, Cambridge, UK. pp. 4162.Google Scholar
Piazza, B. P., Banks, P. D., and La Peyre, M. K. 2005. The potential for created oyster shell reefs as a sustainable shoreline protection strategy in Louisiana. Restoration Ecology, 13: 499506.Google Scholar
Postma, H. 1961. Transport and accumulation of suspended matter in the Dutch Wadden Sea. Netherlands Journal of Sea Research, 1: 148180.Google Scholar
Pratolongo, P. D., Mazzon, C., Zapperi, G., Piovan, M. J., and Brinson, M. M. 2013. Land cover changes in tidal salt marshes of the Bahía Blanca estuary (Argentina) during the past 40 years. Estuarine, Coastal and Shelf Science, 133: 2331.Google Scholar
Pratolongo, P. D., Perillo, G. M. E., and Piccolo, M. C. 2010. Combined effects of waves and marsh plants on mud deposition events at a mudflat-saltmarsh edge. Estuarine, Coastal and Shelf Sciences, 87: 207212.Google Scholar
Priestas, A. M., and Fagherazzi, S. 2011. Morphology and hydrodynamics of wave-cut gullies. Geomorphology, 131: 113.Google Scholar
Pye, K., and French, P. W. 1993. Erosion and Accretion Processes on British Saltmarshes. Volume One. Introduction: Saltmarsh Processes and Morphology. Report No. ES19. Ministry of Agriculture, Fisheries and Food. Cambridge Environmental Research Consultants, Cambridge.Google Scholar
Reddy, K. R., and Delaune, R. D. 2008. Biogeochemistry of Wetlands: Science and Applications. CRC Press, Boca Raton, FL.Google Scholar
Redfield, A. C. 1965. Ontogeny of a saltmarsh estuary. Science, 147: 5055.Google Scholar
Redfield, A. C. 1972. Development of a New England Salt Marsh. Ecological Monographs, 42: 201237.Google Scholar
Redfield, A. C., and Rubin, M. 1962. The age of salt marsh peat and its relation to recent changes in sea level at Barnstable, Massachusetts. Proceeding of the National Academy of Science of the United States of America, 48: 17281735.Google Scholar
Reed, D. J. 1988. Sediment dynamics and deposition in a retreating coastal salt marsh. Estuarine, Coastal and Shelf Science, 26: 6769.Google Scholar
Reed, D. J., Spencer, T., Murray, A., French, J. R., and Leonard, L. 1999. Marsh surface sediment deposition and the role of tidal creeks: implications for created and managed coastal marshes. Journal of Coastal Conservation, 5: 8190.Google Scholar
Rice, D. L. 1986. Early diagenesis in bioadvective sediments: Relationships between the diagenesis of beryllium-7, sediment reworking rates, and the abundance of conveyor-belt deposit-feeders. Journal of Marine Research, 44: 149184.Google Scholar
Ringold, P. 1979. Burrowing, root mat density, and the distribution of fiddler crabs in the Eastern United States. Journal of Experimental Marine Biology and Ecology, 36: 1121.Google Scholar
Robertson, T. L., and Weis, J. S. 2005. A comparison of epifaunal communities associated with the stems of salt marsh grasses Phragmites australis and Spartina alterniflora. Wetlands, 25: 17.Google Scholar
Roman, C. T., Peck, J. A., Allen, J. R., King, J. W., and Appleby, P. G. 1997. Accretion of a New England (U.S.A.) salt marsh in response to inlet migration, storms, and sea-level rise. Estuarine Coastal and Shelf Science, 46: 717727.Google Scholar
Schwimmer, R. 2001, Rates and processes of marsh shoreline erosion in Rehoboth Bay, Delaware, U.S.A. Journal of Coastal Research, 17: 672683.Google Scholar
Schwimmer, R., and Pizzuto, J. 2000. A model for the evolution of marsh shorelines. Journal of Sedimentary Research, 70: 10261035.Google Scholar
Scordo, F., Bohn, V., Piccolo, M. C., and Perillo, G. M. 2018. Mapping and monitoring lakes intra-annual variability in semi-arid regions: a case study in Patagonian Plains (Argentina). Water, 10: 889.Google Scholar
Scyphers, S. B., Powers, S. P., Heck, K. L., and Byron, D. 2011. Oyster reefs as natural breakwaters mitigate shoreline loss and facilitate fisheries. PLOS ONE. 6(8). doi:10.1371/journal.pone.0022396.Google Scholar
Shaffer, G. P., Sasser, C. E., Gosselink, J. G., and Rejmanek, M. 1992. Vegetation dynamics in the emerging Atchafalaya Delta, Louisiana, USA. Journal of Ecology, 80: 677687.Google Scholar
Sharma, P., Gardner, L. R., Moore, W. S., and Bollinger, M. S. 1987. Sedimentation and bioturbation in a salt marsh as revealed by 210Pb, 137Cs, and 7Be studies. Limnology and Oceanography, 32: 313326.Google Scholar
Shepard, C. C., Crain, C. M., and Beck, M. W. 2011. The protective role of coastal marshes: a systematic review and meta-analysis. PLOS ONE, 6(11): e27374. https://doi.org/10.1371/journal.pone.0027374Google Scholar
Shi, B. W., Yang, S. L., Wang, Y. P., Bouma, T. J., and Zhu, Q. 2012. Relating accretion and erosion at an exposed tidal wetland to the bottom shear stress of combined current-wave action. Geomorphology, 138: 380389.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: 18031807.Google Scholar
Silliman, B. R., and Zieman, J. 2001. Top-down control of Spartina alterniflora production by periwinkle grazing in a Virginia salt marsh. Ecology, 82: 28302845.Google Scholar
Slatyer, R. A., Fok, E. S. Y., Hocking, R., and Backwell, P. R.Y. 2008. Why do fiddler crabs build chimneys? Biology Letters of the Royal Society, 4: 616618.Google Scholar
Smith, S. M. 2009. Multi-decadal changes in salt marshes of Cape Cod, Massachusetts: a photographic analysis of vegetation loss, species shifts, and geomorphic change. Northeastern Naturalist, 16: 183208.Google Scholar
Smith, J. E., Bentley, S. J., Snedden, G. A., and White, C. 2015. What role do hurricanes play in sediment delivery to subsiding river deltas? Scientific Reports, 5, Article number: 17582.Google Scholar
Smith, J. M., and Frey, R. W. 1985. Biodeposition by the ribbed mussel Geukensia demissa in a salt marsh, Sapelo Island, Georgia. Journal of Sedimentary Research, 55: 817828.Google Scholar
Soudry, D. 2000. Microbial phosphate sediment. In: Riding, R. E., and Awramik, S. M., eds., Microbial Sediments. Springer-Verlag, Berlin, pp. 127136.Google Scholar
Spalding, M. D., Ruffo, S., Lacambra, C., Meliane, I., Zeitlin Hale, L., Shepard, C. C., and Beck, M. W. 2014. The role of ecosystems in coastal protection: Adapting to climate change and coastal hazards. Ocean and Coastal Management, 90: 5057.Google Scholar
Steel, T. J., and Pye, K. 1997. The development of saltmarsh tidal creek networks: Evidence from the U.K. Proceedings of the Canadian Coastal Conference, pp. 267–280.Google Scholar
Stevenson, J. C., Ward, L. G., and Kearney, M. S. 1986. Vertical accretion in marshes with varying rates of sea-level rise. In: Wolfe, D. A., ed.,  Estuarine Variability, Academic Press, New York, pp. 241259.Google Scholar
Stolz, J. F. 2000. Structure of microbial mats and biofilms. In: Riding, R. E., and Awramik, S. M., eds., Microbial Sediments. Springer-Verlag, Berlin, pp. 18.Google Scholar
Stumpf, R. P. 1983. The process of sedimentation on the surface of a salt marsh. Estuarine, Coastal and Shelf Science, 17: 495508.Google Scholar
Syvitsky, J., Kettner, A. J., Overeem, I., Hutton, E. W., Hannon, M. T., Brakenridge, G. R., Day, J., et al. 2009. Sinking deltas due to human activities. Nature Geoscience, 2: 681686.Google Scholar
Takeda, S., and Kurihara, Y. 1987. The effects of burrowing of Helice tridens (De Haan) on the soil of a salt-marsh habitat. Journal of Experimental Marine Biology and Ecology, 113: 7989.Google Scholar
Taylor, K. L., and Grace, J. B. 1995. The effects of vertebrate herbivory on plant community structure in the coastal marshes of the Pearl River, Louisiana, USA. Wetlands, 15: 6873.Google Scholar
Teal, J. M. 1958. Distribution of fiddler crabs in Georgia salt marshes. Ecology, 39: 186193.Google Scholar
Temmerman, S., Bouma, T. J., Govers, G., Wang, Z. B., de Vries, M. B., and Herman, P. M. J. 2005. Impact of vegetation on flow routing and sedimentation patterns: Three dimensional modelling for a tidal marsh. Journal of Geophysical Research, 110: F04019, doi: 10.1029/2005JF000301.Google Scholar
Temmerman, S., Bouma, T. J., Van de Koppel, J., Van der Wal, , De Vries, D. M. B., and Herman, P. M. J. 2007. Vegetation causes channel erosion in a tidal landscape. Geology, 35: 631634.Google Scholar
Temmerman, S., Govers, G., Meire, P., and Wartel, S. 2003. Modelling long-term tidal marsh growth under changing tidal conditions and suspended sediment concentrations, Scheldt Estuary, Belgium. Marine Geology, 193: 151169.Google Scholar
Temmerman, S., Govers, G., Wartel, S. and Meire, P. 2004. Modelling estuarine variations in tidal marsh sedimentation: response to changing sea levels and suspended sediment concentrations. Marine Geology, 212: 119.Google Scholar
Thomas, C. R., and Blum, L. K. 2010. Importance of the fiddler crab Uca pugnax to salt marsh soil organic matter accumulation. Marine Ecology Progress Series, 414: 167177.Google Scholar
Tolhurst, T. J., Black, K. S., Shayler, S. A., Mather, S., Black, I., Baker, K., and Paterson, D. M. 1999. Measuring the in Situ erosion shear stress of intertidal sediments with the cohesive strength meter (CSM). Estuarine, Coastal and Shelf Science, 49: 281294.Google Scholar
Turner, R. E. 2010. Doubt and the values of an ignorance-based world view for wetland restoration: Coastal LouisianaEstuaries and Coasts32: 10541068.Google Scholar
Turner, R. E. 2011. Beneath the salt marsh canopy: loss of soil strength with increasing nutrient loads. Estuaries and Coasts, 34: 10841093.Google Scholar
Turner, R. E., Swenson, E. M., and Milan, C. S. 2002. Organic and inorganic contributions to vertical accretion in salt marsh sediments. In: Weinstein, M. P., and Kreeger, D. A., eds., Concepts and Controversies in Tidal Marsh Ecology. Springer, Dordrecht, pp. 583595.Google Scholar
Underwood, G. J. C, Paterson, D. M., and Parkes, R. J. 1995. The measurement of microbial carbohydrate exopolymers from intertidal sediments. Limnology and Oceanography, 40: 12431453.Google Scholar
Valentine, K., and Mariotti, G. 2019. Wind-driven water level fluctuations drive marsh edge erosion variability in microtidal coastal bays. Continental Shelf Research, 176: 7689.Google Scholar
Valiela, I., Teal, J. M., and Persson, N. Y. 1976. Production and dynamics of experimentally enriched salt marsh vegetation: belowground biomass. Limnology and Oceanography, 21: 245252.Google Scholar
van Asselen, S., Stouthamer, E., and van Asch, Th. W. J. 2009. Effects of peat compaction on delta evolution: a review on processes, responses, measuring and modeling. Earth-Science Reviews, 92: 3551.Google Scholar
van Eerdt, M. 1986. The influence of basic soil and vegetation parameters on salt marsh cliff strength. In: Gardiner, V., ed., International Geomorphology, Part 1, Wiley, Chichester, pp. 10731086.Google Scholar
van Proosdij, D., Davidson-Arnott, R. G. D., and Ollerhead, J. 2006. Controls on spatial patterns of sediment deposition across a macro-tidal salt marsh surface over single tidal cycles. Estuarine, Coastal and Shelf Science, 69: 6486Google Scholar
van Rijn, L. C., van Rossum, H., and Termes, P. 1990. Field verification of 2–D and 3–D suspended‐sediment models. Journal of Hydraulic Engineering, 116: 12701288.Google Scholar
van Wieren, S. E., and Bakker, J. P. 2008. The impact of browsing and grazing herbivores on biodiversity. In: Gordon, I. J., and Prins, H. H. T., eds. The Ecology of Browsing and Grazing. Springer, Berlin, pp. 236292.Google Scholar
Vandenbruwaene, W., Meire, P., and Temmerman, S. 2012. Formation and evolution of a tidal channel network within a constructed tidal marsh. Geomorphology, 151–152: 114125.Google Scholar
Vu, H. D., and Pennings, S. C. 2018. Predators mediate above- vs. belowground herbivory in a salt marsh crab. Ecosphere, 9(2). doi:10.1002/ecs2.2107.Google Scholar
Vu, H. D., Wieski, K., and Pennings, S. C. 2017. Ecosystem engineers drive creek formation in salt marshes. Ecology, 98: 162174.Google Scholar
Wang, J. Q., Zhang, X. D., Jiang, L. F., Bertness, M. D., Fang, C. M., Chen, J. K., Hara, T., and Li, B. 2010. Bioturbation of burrowing crabs promotes sediment turnover and carbon and nitrogen movements in an estuarine salt marsh. Ecosystems, 13: 586599.Google Scholar
Wang, J. Q., Zhang, Nie, Fu, M., Chen, C. Z., J. K., and Li, B. 2008. Exotic Spartina alterniflora provides compatible habitats for native estuarine crab Sesarma dehaani in the Yangtze River Estuary. Ecological Engineering, 34: 5764.Google Scholar
Wang, M., Gao, X., and Wang, W. 2014. Differences in burrow morphology of crabs between Spartina alterniflora marsh and mangrove habitats. Ecological Engineering, 69: 213219.Google Scholar
Wang, Y. P., Zhang, R., and Gao, S. 1999. Velocity variations in salt marsh creeks, Jiangsu, China. Journal of Coastal Research, 15: 471477.Google Scholar
Watts, C. W., Tolhurst, T. J., Black, K. S., and Whitmore, A. P. 2003. In Situ measurements of erosion shear stress and geotechnical shear strength of the intertidal sediments of the experimental managed realignment scheme at Tollesbury, Essex, UK. Estuarine, Coastal and Shelf Science, 58: 611–20.Google Scholar
Weissburg, M. 1992. Functional analysis of fiddler crab foraging: sex-specific mechanics and constraints in Uca pugnax (Smith). Journal of Experimental Marine Biology and Ecology, 156: 105124.Google Scholar
West, J. M., and Zedler, J. B. 2000. Marsh-creek connectivity: fish use of a tidal salt marsh in Southern California. Estuaries, 23: 699710.Google Scholar
Widdows, J., and Brinsley, M. 2002. Impact of biotic and abiotic processes on sediment dynamics and the consequences to the structure and functioning of the intertidal zone. Journal of Sea Research, 48: 143156.Google Scholar
Widdows, J., Brinsley, M. D., Bowley, N., and Barrett, C. 1998. A benthic annular flume for in situ measurement of suspension feeding/biodeposition rates and erosion potential of intertidal cohesive sediments. Estuarine Coastal and Shelf Science, 46: 2738.Google Scholar
Widdows, J., Pope, N., and Brinsley, M. 2008. Effect of Spartina anglica stems on nearbed hydrodynamics, sediment erodability and morphological changes on an intertidal mudflat. Marine Ecology Progress Series, 362: 4557.Google Scholar
Wigand, C., Brennan, P., Stolt, M., Holt, M., and Ryba, S. 2009. Soil respiration rates in coastal marshes subject to increasing watershed nitrogen loads in southern New England, US. Wetlands, 29: 952963.Google Scholar
Wilson, K., Kelley, J., Croitoru, A., Dionne, M., Belknap, D., and Steneck, R. 2009. Stratigraphic and ecophysical characterizations of salt pools: dynamic landforms of the Webhannet Salt Marsh, Wells, ME, USA. Estuaries and Coasts, 32: 855870.Google Scholar
Wilson, K., Kelley, J. T., Tanner, B. R., and Belknap, D. F. 2010. Probing the origins and stratigraphic signature of salt pools from north-temperate marshes in Maine, U.S.A. Journal of Coastal Research, 26: 10071026.Google Scholar
Wilson, C., and Allison, M. 2008. An equilibrium profile model for retreating marsh shorelines in southeast Louisiana. Estuarine, Coastal and Shelf Science, 80, 483494.Google Scholar
Wilson, C. A., Hughes, Z. J., and FitzGerald, D. M. 2012. The effects of crab bioturbation on mid-Atlantic saltmarsh tidal creek extension: geotechnical and geochemical changes. Estuarine, Coastal and Shelf Science, 106: 3344.Google Scholar
Wilson, C. A., Hughes, Z. J., FitzGerald, D. M., Hopkinson, C. S., Valentine, V., and Kolker, A. S. 2014. Saltmarsh pool and tidal creek morphodynamics: dynamic equilibrium of northern latitude saltmarshes? Geomorphology, 213: 99115.Google Scholar
Windham, L. 2001. Comparison of biomass production and decomposition between Phragmites australis (Common Reed) and Spartina patens (Salt Hay Grass) in brackish tidal marshes of New Jersey, USA. Wetlands, 21: 179188.Google Scholar
Xin, P., Jin, G., Li, L. and Barry, D. A. 2009. Effects of crab burrows on pore water flows in salt marshes. Advances in Water Resources, 32: 439449.Google Scholar
Yallop, M. L., Paterson, D. M., and Wellsbury, P. 2000. Interrelationships between rates of microbial production, exopolymer production, microbial biomass, and sediment stability in biofilms of intertidal sediments. Microbial Ecology, 39: 116127.Google Scholar
Yang, S. L., Li, M., Dai, S. B., Liu, Z., Zhang, J. and Ding, P. X. 2006. Drastic decrease in sediment supply from the Yangtze River and its challenge to coastal wetland management. Geophysical Research Letters, 33: 47.Google Scholar
Yang, S. L., Li, H., Ysebaert, T., Bouma, T. J., Zhang, W. X., Wang, Y. Y., Li, P., Li, M., and Ding, P. X. 2008. Spatial and temporal variations in sediment grain size in tidal wetlands, Yangtze Delta: On the role of physical and biotic controls. Estuarine Coastal and Shelf Science, 77: 657671.Google Scholar
Yapp, R. H., Johns, D., and Jones, O. T. 1917. The salt marshes of the Dovey Estuary. Journal of Ecology, 5: 65103.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
×