Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-dnltx Total loading time: 0 Render date: 2024-04-24T00:38:08.267Z Has data issue: false hasContentIssue false

3 - Salt Marsh Formation

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

Historical records show a massive decline in salt marsh area (Pendleton et al. 2012), > 50% in many locations, such as sites in Australia (Saintilan and Williams, 2000; Rogers et al. 2006), the British Isles (Baily and Pearson, 2007), and New England, USA (Bertness et al. 2002). These losses are mainly fueled by an underappreciation of the large contributions of salt marsh to maintaining healthy and productive estuaries. Prior to the middle twentieth century, the value of salt marsh primarily depended on its potential for reclamation. Davis (1910) proclaimed that “…[salt marshes] are conspicuous, being generally unutilized for any purpose except for making a small amount of inferior hay, hence they are practically desert places, except where land values are sufficiently high to make it worth while to raise the surface above high tide level for building purposes, or to dike out the tides.” We now view salt marsh as a valuable estuarine habitat because it provides coastal protection from waves (Shepard et al. 2011), erosion control (Neumeier and Ciavola, 2004), water purification (Sousa et al. 2008), fish and bird habitat (Peterson and Turner, 1994; Van Eerden et al. 2005), carbon sequestration (Mcleod et al. 2011), and tourism/recreation (Barbier et al. 2011; Altieri et al. 2012). Salt marsh is also a coastal depositional environment that can accrete vertically over millennial time scales at rates equal to, or greater than, sea-level rise (Gehrels et al. 1996; Ouyang and Lee, 2014). The relatively high accretion rates and resistance of salt marshes to erosion (Mudd et al. 2010) make them valuable sites for preserving records of sea level (van de Plassche et al. 1998; Engelhart et al. 2011; Kemp et al. 2017), storms (Donnelly et al. 2001; Boldt et al. 2010; de Groot et al. 2011), and tsunamis (Morton et al. 2007; Komatsubara et al. 2008) in their sediments. Salt marsh loss and associated services have been pervasive globally, mainly due to the direct (grazing, ditching, pollution, etc.) and indirect (climate change) effects of human activities, resulting in the recent emphasis on restoration, conservation, and management (Lotze et al. 2006; Airoldi and Beck, 2007; Gedan et al. 2009). Although recent focus has been on better understanding of those mechanisms and processes that are related to salt marsh degradation, reviewing salt marsh formation and the different modes of salt marsh expansion will aid efforts aimed at preserving and increasing salt marsh habitat area and extracting climate and tectonic information from their sedimentary records.

Type
Chapter
Information
Salt Marshes
Function, Dynamics, and Stresses
, pp. 31 - 52
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

Adam, P. 1990. Saltmarsh Ecology. Cambridge University Press, Cambridge; New York.CrossRefGoogle Scholar
Adam, P. 2002. Saltmarshes in a time of change. Environmental Conservation, 29: 3961.Google Scholar
Adams, D. A. 1963. Factors influencing vascular plant zonation in North Carolina Salt Marshes. Ecology, 44: 445456.Google Scholar
Airoldi, L., and Beck, M. W. 2007. Loss, status and trends for coastal marine habitats of Europe. Oceanography and Marine Biology: An Annual Review, 45: 345405.Google Scholar
Allen, G. P., and Posamentier, H. W. 1993. Sequence stratigraphy and facies model of an incised valley fill; the Gironde Estuary, France. Journal of Sedimentary Research, 63: 378391.Google Scholar
Allen, J., and Rae, J. 1987. Late Flandrian shoreline oscillations in the Severn Estuary: a geomorphological and stratigraphical reconnaissance. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 315: 185230.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.CrossRefGoogle Scholar
Allen, J. R. L., and Haslett, S. K. 2012. Salt-marsh evolution at Northwick and Aust warths, Severn Estuary, UK: a case of constrained autocyclicity. Atlantic Geology, 50: 117.Google 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.CrossRefGoogle ScholarPubMed
Amos, C. L., Feeney, T., Sutherland, T. F., and Luternauer, J. L. 1997. The stability of fine-grained sediments from the Fraser River Delta. Estuarine, Coastal and Shelf Science, 45: 507524.Google Scholar
Anderson, J. B., Wallace, D. J., Simms, A. R., Rodriguez, A. B., Weight, R. W. R., and Taha, Z. P. 2016. Recycling sediments between source and sink during a eustatic cycle: Systems of late Quaternary northwestern Gulf of Mexico Basin. Earth-Science Reviews, 153: 111138.Google Scholar
Bahattacharya, J. P. 2006. Deltas. In: Facies Models Revisited. Eds Posamentier, H. W. and Walker, R. G.., Society for Sedimentary Geology, Tulsa, pp. 237292.CrossRefGoogle Scholar
Baily, B., and Pearson, A. W. 2007. Change detection mapping and analysis of salt marsh areas of Central Southern England from Hurst Castle Spit to Pagham Harbour. Journal of Coastal Research, 23: 15491564.Google Scholar
Bakker, J., Esselink, P., Dijkema, K., Van Duin, W., and De Jong, D. 2002. Restoration of salt marshes in the Netherlands. Hydrobiologia, 478: 2951.CrossRefGoogle Scholar
Barbier, E. B., Hacker, S. D., Kennedy, C., Koch, E. W., Stier, A. C., and Silliman, B. R. 2011. The value of estuarine and coastal ecosystem services. Ecological Monographs, 81: 169193.CrossRefGoogle Scholar
Belknap, D. F., and Kraft, J. C. 1985. Influence of antecedent geology on stratigraphic preservation potential and evolution of Delaware’s barrier systems. Marine Geology, 63: 235262.CrossRefGoogle Scholar
Belknap, D. F., Kraft, J. C., and Dunn, R. K. 1994. Transgressive valley-fill lithosomes: Delaware and Maine. In: Incised-Valley Systems: Origin and Sedimentary Sequences. Eds Dalrymple, R. W., Boyd, R. and Zaitlin, B. A.., SEPM, Special Publication 51, SEPM, Tulsa, pp. 303320.Google Scholar
Bertness, M. D., Ewanchuk, P. J., and Silliman, B. R. 2002. Anthropogenic modification of New England salt marsh landscapes. Proceedings of the National Academy of Sciences of the USA, 99: 13951398.CrossRefGoogle ScholarPubMed
Blum, M. D., and Roberts, H. H. 2009. Drowning of the Mississippi Delta due to insufficient sediment supply and global sea-level rise. Nature Geoscience, 2: 488491.CrossRefGoogle Scholar
Boldt, K. V., Lane, P., Woodruff, J. D., and Donnelly, J. P. 2010. Calibrating a sedimentary record of overwash from Southeastern New England using modeled historic hurricane surges. Marine Geology, 275: 127139.CrossRefGoogle Scholar
Bouma, T. J., van Belzen, J., Balke, T., van Dalen, J., Klaassen, P., Hartog, A. M., Callaghan, D. P., et al. 2016. Short-term mudflat dynamics drive long-term cyclic salt marsh dynamics. Limnology and Oceanography, 61: 22612275.CrossRefGoogle Scholar
Broome, S. W., Seneca, E. D., and Woodhouse, W. W. 1988. Tidal salt marsh restoration. Aquatic Botany, 32: 122.CrossRefGoogle Scholar
Bruno, J. F. 2000. Facilitation of cobble beach plant communities through habitat modification by Spartina alterniflora. Ecology, 81: 11791192.CrossRefGoogle Scholar
Cahoon, D. R., White, D. A., and Lynch, J. C. 2011. Sediment infilling and wetland formation dynamics in an active crevasse splay of the Mississippi River delta. Geomorphology, 131: 5768.CrossRefGoogle Scholar
Canuel, E. A., Lerberg, E. J., Dickhut, R. M., Kuehl, S. A., Bianchi, T. S., and Wakeham, S. G. 2009. Changes in sediment and organic carbon accumulation in a highly-disturbed ecosystem: the Sacramento-San Joaquin River Delta California, USA. Marine Pollution Bulletin, 59: 154–63.Google Scholar
Chapman, V. J. 1960. Salt Marshes and Salt Deserts of the World. L. Hill, London.Google Scholar
Chung, C. H., Zhuo, R. Z., and Xu, G. W. 2004. Creation of Spartina plantations for reclaiming Dongtai, China, tidal flats and offshore sands. Ecological Engineering, 23: 135150.CrossRefGoogle Scholar
Craft, C. 2000. Co-development of wetland soils and benthic invertebrate communities following salt marsh creation. Wetlands Ecology and Management, 8: 197207.Google Scholar
Craft, C., Broome, S., and Campbell, C. 2002. Fifteen years of vegetation and soil development after brackish-water marsh creation. Restoration Ecology, 10: 248258.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.CrossRefGoogle Scholar
Currin, C. A., Delano, P. C., and Valdes-Weaver, L. M. 2008. Utilization of a citizen monitoring protocol to assess the structure and function of natural and stabilized fringing salt marshes in North Carolina. Wetlands Ecology Management, 16: 97118.CrossRefGoogle Scholar
Dalrymple, R. W., Zaitlin, B. A., and Boyd, R. 1992. Estuarine facies models: conceptual basis and stratigraphic implications. Journal of Sedimentary Petrology, 62: 11301146.Google Scholar
Davis, C. A. 1910. Salt marsh formation near Boston and its geological significance. Economic Geology, 5: 623639.Google Scholar
Davis, R. A., and Clifton, H. E. 1987. Sea-level change and the preservation potential of wave-dominated and tide-dominated coastal sequences. In: Sea-level Fluctuation and Coastal Evolution. Eds Nummedal, D., Pilkey, O. H. Jr., and Howard, J. D.., Special Publications of SEPM 41, Tulsa, pp. 167178.CrossRefGoogle Scholar
Day, J. W., Boesch, D. F., Clairain, E. J., Kemp, G. P., Laska, S. B., Mitsch, W. J., Orth, K., et al. 2007. Restoration of the Mississippi delta: lessons from Hurricanes Katrina and Rita. Science, 315: 16791684.CrossRefGoogle ScholarPubMed
de Groot, A. V., Veeneklaas, R. M., and Bakker, J. P. 2011. Sand in the salt marsh: Contribution of high-energy conditions to salt-marsh accretion. Marine Geology, 282: 240254.Google Scholar
Dijkema, K. S. 1997. Impact prognosis for salt marshes from subsidence by gas extraction in the Wadden Sea. Journal of Coastal Research, 13: 12941304.Google Scholar
Donnelly, J. P., Roll, S., Wengren, M., Butler, J., Lederer, R., and Webb, I. I. I. T. 2001. Sedimentary evidence of intense hurricane strikes from New Jersey. Geology, 29: 615618.Google Scholar
Engelhart, S. E., Horton, B. P., and Kemp, A. C. 2011. Holocene sea level changes along the United States’ Atlantic Coast. Oceanography, 24: 7079.CrossRefGoogle 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.CrossRefGoogle Scholar
Engels, J. G., Rink, F., and Jensen, K. 2011. Stress tolerance and biotic interactions determine plant zonation patterns in estuarine marshes during seedling emergence and early establishment. Journal of Ecology, 99: 277287.Google Scholar
Fagherazzi, S. 2013. The ephemeral life of a salt marsh. Geology, 41: 943944.Google Scholar
Fagherazzi, S., Carniello, L., D’Alpaos, L., and Defina, A. 2006. Critical bifurcation of shallow microtidal landforms in tidal flats and salt marshes. Proceedings of the National Academy of Sciences of the USA, 103: 83378341.CrossRefGoogle ScholarPubMed
Fagherazzi, S., Kirwan, M. L., Mudd, S. M., Guntenspergen, G. R., Temmerman, S., D’Alpaos, A., van de Koppel, , et al. 2012. Numerical models of salt marsh evolution: Ecological, geomorphic, and climatic factors. Reviews of Geophysics, 50: RG1002.CrossRefGoogle Scholar
Feagin, R. A., Martinez, M. L., Mendoza-Gonzalez, G., and Costanza, R. 2010. Salt marsh zonal migration and ecosystem service change in response to global sea level rise: a case study from an urban region. Ecology and Society, 15(4): 14.CrossRefGoogle Scholar
Fisher, J. J. 1962. Geomorphic Expression of Former Inlets along the Outer Banks of North Carolina, University of North Carolina at Chapel Hill.Google Scholar
Flowers, T. J., and Colmer, T. D. 2008. Salinity tolerance in halophytes. New Phytologist, 179: 945963.CrossRefGoogle ScholarPubMed
Ford, M. A., Cahoon, D. R., and Lynch, J. C. 1999. Restoring marsh elevation in a rapidly subsiding salt marsh by thin-layer deposition of dredged material. Ecological Engineering, 12: 189205.CrossRefGoogle Scholar
Galloway, W. E. 1975. Process framework for describing the morphologic and stratigraphic evolution of deltaic depositional systems. In: Deltas Models for Exploration, Ed Broussard, M. L.., Houston Geological Society, Houston, pp. 8798.Google Scholar
Gardner, L. R., and Porter, D. E. 2001. Stratigraphy and geologic history of a southeastern salt marsh basin, North Inlet, South Carolina, USA. Wetlands Ecology and Management, 9: 371385.CrossRefGoogle Scholar
Gedan, K. B., Silliman, B. R., and Bertness, M. D. 2009. Centuries of human-driven change in salt marsh ecosystems. Annual Review of Marine Science, 1: 117141.CrossRefGoogle ScholarPubMed
Gehrels, R. W., Belknap, D. F., and Kelley, J. T. 1996. Integrated high-precision analyses of Holocene relative sea-level changes: lessons from the coast of Maine. GSA Bulletin, 108: 10731088.2.3.CO;2>CrossRefGoogle Scholar
Godfrey, P. J., and Godfrey, M. M. 1974. The role of overwash and inlet dynamics in the formation of salt marshes on North Carolina barrier islands. In: Ecology of Halophytes. Eds Reimold, R. J. and Queen, W. H.., Academic Press, Inc., New York, pp. 407427.Google Scholar
Graham, S. A., and Mendelssohn, I. A. 2013. Functional assessment of differential sediment slurry applications in a deteriorating brackish marsh. Ecological Engineering, 51: 264274.Google Scholar
Gunnell, J. R., Rodriguez, A. B., and McKee, B. A. 2013. How a marsh is built from the bottom up. Geology, 41: 859862.Google Scholar
Jalowska, A. M., McKee, B. A., Laceby, J. P., and Rodriguez, A. B. 2017. Tracing the sources, fate, and recycling of fine sediments across a river-delta interface. Catena, 154: 95106.Google Scholar
Jalowska, A. M., Rodriguez, A. B., and McKee, B. A. 2015. Responses of the Roanoke Bayhead Delta to variations in sea level rise and sediment supply during the Holocene and Anthropocene. Anthropocene, 9: 4155.Google Scholar
James, L. A. 2013. Legacy sediment: definitions and processes of episodically produced anthropogenic sediment. Anthropocene, 2: 1626.Google Scholar
Jervey, M. T. 1988. Quantitative geological modeling of siliciclastic rock sequences and their seismic expression. In: Sea-Level Changes: An Integrated Approach. Eds Wilgus, C. K., Hastings, B. S., Ross, C. A., Posamentier, H. W., Van Wagoner, J. C., and Kendall, C. G. S. C.. Special Publication 42, SEPM, Tulsa, pp. 4769.Google Scholar
Johnson, D. W. 1919. Shore Processes and Shoreline Development. John Wiley & Sons, Incorporated, Boston.Google Scholar
Kelley, J. T., Belknap, D. F., Jacobson, G. L., and Heather, A. J. 1988. The morphology and origin of salt marshes along the glaciated coastline of Maine, USA. Journal of Coastal Research, 4: 649666.Google Scholar
Kemp, A. C., Horton, B. P., Corbett, D. R., Culver, S. J., Edwards, R. J., and van de Plassche, O. 2017. The relative utility of foraminifera and diatoms for reconstructing late Holocene sea-level change in North Carolina, USA. Quaternary Research, 71: 921.Google Scholar
Kennish, M. J. 2001. Coastal salt marsh systems in the U.S.: A review of anthropogenic impacts. Journal of Coastal Research, 17: 731748.Google Scholar
Kirwan, M. L., Guntenspergen, G. R., D’Alpaos, A., Morris, J. T., Mudd, S. M., and Temmerman, S. 2010. Limits on the adaptability of coastal marshes to rising sea level. Geophysical Research Letters, 37: L23401.Google Scholar
Kirwan, M. L., and Megonigal, J. P. 2013. Tidal wetland stability in the face of human impacts and sea-level rise. Nature, 504: 53.Google Scholar
Kirwan, M. L., Walters, D. C., Reay, W. G., and Carr, J. A. 2016. Sea level driven marsh expansion in a coupled model of marsh erosion and migration. Geophysical Research Letters, 43: 43664373.Google Scholar
Komatsubara, J., Fujiwara, O., Takada, K., Sawai, Y., Aung, T. T., and Kamataki, T. 2008. Historical tsunamis and storms recorded in a coastal lowland, Shizuoka Prefecture, along the Pacific Coast of Japan. Sedimentology, 55: 17031716.Google Scholar
Kraft, J. C. 1971. Sedimentary facies patterns and geologic history of a Holocene marine transgression. Geological Society of America Bulletin, 82: 21312158.CrossRefGoogle Scholar
Kraft, J. C., Yi, H. L., and Khalequzzaman, M. 1992. Geologic and human factors in the decline of the tidal salt marsh lithosome: the Delaware estuary and Atlantic coastal zone. Sedimentary Geology, 80: 233246.Google Scholar
Leonardi, N., and Fagherazzi, S. 2015. Local variability in erosional resistance affects large scale morphodynamic response of salt marshes to wind waves and extreme events. Geophysical Research Letters, 42: 58725879.Google Scholar
Lotze, H. K., Lenihan, H. S., Bourque, B. J., Bradbury, R. H., Cooke, R. G., Kay, M. C., Kidwell, S. M., et al. 2006. Depletion, degradation, and recovery potential of estuaries and coastal seas. Science, 312: 18061809.Google Scholar
Marani, M., D’Alpaos, A., Lanzoni, S., Carniello, L., and Rinaldo, A. 2010. The importance of being coupled: stable states and catastrophic shifts in tidal biomorphodynamics. Journal of Geophysical Research: Earth Surface, 115: F04004, doi:10.1029/2009JF001600.CrossRefGoogle Scholar
Mariotti, G., and Fagherazzi, S. 2013. Critical width of tidal flats triggers marsh collapse in the absence of sea-level rise. Proceedings of the National Academy of Sciences of the USA, 110: 53535356.Google Scholar
Mattheus, C. R., Rodriguez, A. B., and McKee, B. A. 2009. Direct connectivity between upstream and downstream promotes rapid response of lower coastal-plain rivers to land-use change. Geophysical Research Letters, 36: L20401, doi:10.1029/2009GL039995.Google Scholar
McKee, L. J., Ganju, N. K., and Schoellhamer, D. H. 2006. Estimates of suspended sediment entering San Francisco Bay from the Sacramento and San Joaquin Delta, San Francisco Bay, California. Journal of Hydrology, 323: 335352.CrossRefGoogle Scholar
McLeod, E., Chmura, G. L., Bouillon, S., Salm, R., Björk, M., Duarte, C. M., Lovelock, C. E., et al. 2011. A blueprint for blue carbon: toward an improved understanding of the role of vegetated coastal habitats in sequestering CO2. Frontiers in Ecology and the Environment, 9: 552560.Google Scholar
Möller, 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.CrossRefGoogle Scholar
Morales, J. A. 1997. Evolution and facies architecture of the mesotidal Guadiana River delta S.W. Spain-Portugal. Marine Geology, 138: 127148.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.CrossRefGoogle Scholar
Morton, R. A., Gelfenbaum, G., and Jaffe, B. E. 2007. Physical criteria for distinguishing sandy tsunami and storm deposits using modern examples. Sedimentary Geology, 200: 184207.Google Scholar
Mudd, S. M., D’Alpaos, A., and Morris, J. T. 2010. How does vegetation affect sedimentation on tidal marshes? Investigating particle capture and hydrodynamic controls on biologically mediated sedimentation. Journal of Geophysical Research, 115: F03029, doi:10.1029/2009JF001566.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
Nichols, M. M. 1989. Sediment accumulation rates and relative sea-level rise in lagoons. Marine Geology, 88: 201219.Google Scholar
Odum, W. E. 1988. Comparative ecology of tidal freshwater and salt marshes. Annual Review of Ecology and Systematics, 19: 147176.Google Scholar
Olariu, C., and Bhattacharya, J. P. 2006. Terminal distributary channels and delta front architecture of river-dominated delta systems. Journal of Sedimentary Research, 76: 212233.Google Scholar
Olliver, E. A., and Edmonds, D. A. 2017. Defining the ecogeomorphic succession of land building for freshwater, intertidal wetlands in Wax Lake Delta, Louisiana. Estuarine, Coastal and Shelf Science, 196: 4557.CrossRefGoogle Scholar
Ouyang, X., and Lee, S. Y. 2014. Updated estimates of carbon accumulation rates in coastal marsh sediments. Biogeosciences, 11: 50575071.Google Scholar
Pendleton, L., Donato, D. C., Murray, B. C., Crooks, S., Jenkins, W. A., Sifleet, S., Craft, C., et al. 2012. Estimating global “blue carbon” emissions from conversion and degradation of vegetated coastal ecosystems. PLoS One, 7: e43542.Google Scholar
Penland, S., Boyd, R., and Suter, J. R. 1988. Transgressive depositional systems of the Mississippi Delta plain; a model for barrier shoreline and shelf sand development. Journal of Sedimentary Research, 58: 932949.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.CrossRefGoogle Scholar
Pethick, J. S. 1981. Long-term accretion rates on tidal salt marshes. Journal of Sedimentary Research, 51: 571577.CrossRefGoogle Scholar
Phleger, C. F. 1971. Effect of salinity on growth of a salt marsh grass. Ecology, 52: 908911.CrossRefGoogle Scholar
Raabe, E. A., and Stumpf, R. P. 2015. Expansion of tidal marsh in response to sea-level rise: Gulf Coast of Florida, USA. Estuaries and Coasts, 39: 145157.CrossRefGoogle Scholar
Redfield, A. C. 1965. Ontogeny of a salt marsh estuary. Science, 147: 5055.CrossRefGoogle ScholarPubMed
Reed, D. J. 2002. Sea-level rise and coastal marsh sustainability: geological and ecological factors in the Mississippi delta plain. Geomorphology, 48: 233243.Google Scholar
Ridge, J. T., Rodriguez, A. B., and Fodrie, F. J. 2017. Salt marsh and fringing oyster reef transgression in a shallow temperate estuary: implications for restoration, conservation and blue carbon. Estuaries and Coasts, 40: 10131027.Google Scholar
Roberts, H. H. 1997. Dynamic changes of the Holocene Mississippi River Delta Plain: the delta cycle. Journal of Coastal Research, 13: 605627.Google Scholar
Rodriguez, A. B., Anderson, J. B., Banfield, L. B., Taviani, M., Abdulah, K., and Snow, J. N. 2000. Identification of a −15m middle Wisconsin shoreline on the Texas inner continental shelf. Palaeogeography, Palaeoclimatology, Palaeoecology, 158: 2543.CrossRefGoogle Scholar
Rodriguez, A. B., Fodrie, F. J., Ridge, J. T., Lindquist, N. L., Theuerkauf, E. J., Coleman, S. E., et al. 2014. Oyster reefs can outpace sea-level rise. Nature Climate Change, 4: 493497.Google Scholar
Rodriguez, A. B., Simms, A. R., and Anderson, J. B. 2010. Bay-head deltas across the northern Gulf of Mexico back step in response to the 8.2 ka cooling event. Quaternary Science Reviews, 29: 39833993.Google Scholar
Rogers, K., Wilton, K. M., and Saintilan, N. 2006. Vegetation change and surface elevation dynamics in estuarine wetlands of southeast Australia. Estuarine, Coastal and Shelf Science, 66: 559569.Google Scholar
Saintilan, N., and Hashimoto, T. R. 1999. Mangrove-saltmarsh dynamics on a bay-head delta in the Hawkesbury River estuary, New South Wales, Australia. Hydrobiologia, 413: 95102.Google Scholar
Saintilan, N., and Williams, R. 2010. Short Note: The decline of saltmarsh in southeast Australia: Results of recent surveys. Wetlands Australia Journal, 18: 4954.Google Scholar
Schwimmer, R. A., and Pizzuto, J. E. 2000. A model for the evolution of marsh shorelines. Journal of Sedimentary Research, 70: 10261035.Google Scholar
Shennan, I., and Horton, B. 2002. Holocene land- and sea-level changes in Great Britain. Journal of Quaternary Science, 17: 511526.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: e27374.Google Scholar
Shideler, G. L. 1984. Suspended sediment responses in a wind-dominated estuary of the Texas Gulf Coast. Journal of Sedimentary Petrology, 54: 731745.Google Scholar
Simms, A. R., and Rodriguez, A. B. 2014. Where do coastlines stabilize following rapid retreat? Geophysical Research Letters, 41: 16981703.Google Scholar
Simms, A. R., and Rodriguez, A. B. 2015. The Influence of valley morphology on the rate of Bayhead Delta Progradation. Journal of Sedimentary Research, 85: 3844.Google Scholar
Simms, A. R., Rodriguez, A. B., and Anderson, J. B. 2018. Bayhead deltas and shorelines: Insights from modern and ancient examples. Sedimentary Geology, 374: 1735.Google Scholar
Singh Chauhan, P. P. 2009. Autocyclic erosion in tidal marshes. Geomorphology, 110: 4557.Google Scholar
Snow, A. A., and Vince, S. W. 1984. Plant Zonation in an Alaskan Salt Marsh: II. An experimental study of the role of edaphic conditions. Journal of Ecology, 72: 669684.Google Scholar
Sousa, A. I., Lillebø, A. I., Caçador, I., and Pardal, M. A. 2008. Contribution of Spartina maritima to the reduction of eutrophication in estuarine systems. Environmental Pollution, 156: 628635.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
Syvitski, J. P. M., Kettner, A. J., Overeem, I., Hutton, E. W. H., Hannon, M. T., Brakenridge, G. R., Day, J., et al. 2009. Sinking deltas due to human activities. Nature Geoscience, 2: 681686.Google Scholar
Ta, T. K. O., Nguyen, V. L., Tateishi, M., Kobayashi, I., Saito, Y., and Nakamura, T. 2002. Sediment facies and Late Holocene progradation of the Mekong River Delta in Bentre Province, southern Vietnam: an example of evolution from a tide-dominated to a tide- and wave-dominated delta. Sedimentary Geology, 152: 313325.Google Scholar
Theuerkauf, E. J., and Rodriguez, A. B. 2017. Placing barrier-island transgression in a blue-carbon context. Earth’s Future, 5: 789810.Google Scholar
Theuerkauf, E. J., Stephens, J. D., Ridge, J. T., Fodrie, F. J., and Rodriguez, A. B. 2015. Carbon export from fringing saltmarsh shoreline erosion overwhelms carbon storage across a critical width threshold. Estuarine, Coastal and Shelf Science, 164: 367378.Google Scholar
Thomas, M. A., and Anderson, J. B. 1994. Sea-level controls on the facies architecture of the Trinity/Sabine incised-valley system, Texas continental shelf. In: Incised-Valley Systems: Origin and Sedimentary Sequences. Eds Dalrymple, R. W., Boyd, R., and Zaitlin, B. A.., SEPM, Special Publication 51, SEPM, Tulsa, pp. 6382.Google Scholar
Törnqvist, T. E., Gonzalez, J. L., Newsom, L., van der Borg, K., de Jong, A. F. M., and Kurnik, C. W. 2004. Deciphering Holocene sea-level history on the U.S. Gulf Coast: a high-resolution record from the Mississippi Delta. Geological Society of America Bulletin, 116: 10261039.CrossRefGoogle Scholar
van de Plassche, O., van der Borg, K., and de Jong, A. F. M. 1998. Sea level-climate correlation during the past 1400 yr. Geology, 26: 319322.Google Scholar
Van der Wal, D., Wielemaker-Van den Dool, A., and Herman, P. M. J. 2008. Spatial patterns, rates and mechanisms of saltmarsh cycles Westerschelde, the Netherlands. Estuarine, Coastal and Shelf Science, 76: 357368.Google Scholar
Van Eerden, M. R., Drent, R. H., Stahl, J., and Bakker, J. P. 2005. Connecting seas: western Palaearctic continental flyway for water birds in the perspective of changing land use and climate. Global Change Biology, 11: 894908.Google Scholar
Warren, R. S., Fell, P. E., Rozsa, R., Brawley, A. H., Orsted, A. C., Olson, E. T., Swamy, V., and Niering, W. A. 2002. Salt marsh restoration in Connecticut: 20 years of science and management. Restoration Ecology, 10: 497513.Google Scholar
Watson, E. B., and Byrne, R. 2013. Late Holocene marsh expansion in Southern San Francisco Bay, California. Estuaries and Coasts, 36: 643653.Google Scholar
White, W. A., Morton, R. A., and Holmes, C. W. 2002. A comparison of factors controlling sedimentation rates and wetland loss in fluvial-deltaic sytems, Texas Gulf coast. Geomorphology, 44: 4766.Google Scholar
Williams, K., Ewel, K. C., Stumpf, R. P., Putz, F. E., and Workman, T. W. 1999. Sea-level rise and coastal forest retreat on the west coast of Florida, USA. Ecology, 80: 20452063.CrossRefGoogle Scholar
Williams, P. B., and Orr, M. K. 2002. Physical evolution of restored breached levee salt marshes in the San Francisco Bay Estuary. Restoration Ecology, 10: 527542.Google Scholar
Xiao, D., Zhang, L., and Zhu, Z. 2010. The range expansion patterns of Spartina alterniflora on salt marshes in the Yangtze Estuary, China. Estuarine, Coastal and Shelf Science, 88: 99104.Google Scholar
Yang, S. L., Li, H., Ysebaert, T., Bouma, T. J., Zhang, W. X., Wang, Y. Y., Li, P., et al. 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
Zhang, R. S., Shen, Y. M., Lu, L. Y., Yan, S. G., Wang, Y. H., Li, J. L., and Zhang, Z. L. 2004. Formation of Spartina alterniflora salt marshes on the coast of Jiangsu Province, China. Ecological Engineering, 23: 95105.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
×