Skip to main content Accessibility help
×
Hostname: page-component-7c8c6479df-94d59 Total loading time: 0 Render date: 2024-03-28T18:01:13.630Z Has data issue: false hasContentIssue false

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

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Salt Marshes
Function, Dynamics, and Stresses
, pp. 7 - 154
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

References

Adam, P. 1990. Saltmarsh Ecology. Cambridge University Press, Cambridge, UK.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
Aman, J., and Grimes, K. W. 2016. Measuring impacts on invasive European Green Crabs on Maine Salt Marshes: a novel approach: Report to the Maine Outdoor Heritage Fund.Google Scholar
Argow, B. A., and FitzGerald, D. M. 2006. Winter processes on northern salt marshes: evaluating the impact of in-situ peat compaction due to ice loading, Wells, ME. Estuarine, Coastal and Shelf Science, 69: 360369.CrossRefGoogle Scholar
Argow, B. A., Hughes, Z. J., and FitzGerald, D. M. 2011. Ice raft formation, sediment load, and theoretical potential for ice-rafted sediment influx on northern coastal wetlands. Continental Shelf Research, 31: 12941395.Google Scholar
Barras, J., Beville, S., Britsch, D., Hartley, S., Hawes, S., Johnston, J., Reed, D., Roy, K., Sapkota, S., and Suhaayda, J. 2004. Historical and projected coastal Louisiana land changes: 1978–2050: U.S. Geological Survey Open-File Report OFR 03-334.Google Scholar
Belknap, D. F. 1999. Sea-level rise and Gulf of Maine salt marshes. Gulf of Maine NEWS, Regional Association for Research on the Gulf of Maine, Spring, 1999: 1, 810.Google Scholar
Belknap, D. F. 2003. Salt marshes. In: Middleton, G., ed., Encyclopedia of Sediments and Sedimentary Rocks. Kluwer Academic Publishers, Dordrecht, pp. 586588.Google Scholar
Belknap, D. F., Andersen, B. G., Anderson, R. S., Anderson, W. A., Borns, H. W. Jr., Jacobson, G. Jr., et al. 1987. Late Quaternary sea-level changes in Maine. In: Nummedal, D., Pilkey, O. H. Jr. and Howard, J. D., eds., Sea-Level Fluctuation and Coastal Evolution, Society of Economic Paleontologists and Mineralogists Special Publication, No. 41, pp. 7185.CrossRefGoogle Scholar
Belknap, D. F., Gontz, A. M., and Kelley, J. T. 2005. Paleodeltas and preservation potential on a paraglacial coast – evolution of eastern Penobscot Bay, Maine. Chapter 16. In: FitzGerald, D. M. and Knight, J., eds., High Resolution Morphodynamics and Sedimentary Evolution of Estuaries. Springer, Dordrecht, pp. 335360.CrossRefGoogle Scholar
Belknap, D. F., Kelley, J. T., FitzGerald, D. M., and Buynevich, I. 2004. Quaternary Sea-level Changes and Coastal Evolution in Eastern and Central Coastal Maine, Field Trip Guidebook, International Geological Correlation Program #495, Quaternary Land-Ocean Interactions: Driving Mechanisms and Coastal Responses, Conference and Field Trip, Bar Harbor, ME, October 14–17, 2004, Dept. Earth Sciences, UniMaine, Orono.Google Scholar
Belknap, D. F., Kelley, J. T., and Gontz, A. M. 2002. Evolution of the glaciated shelf and coastline of the northern Gulf of Maine, USA. Journal of Coastal Research Special Issue, 36: 3755.CrossRefGoogle Scholar
Belknap, D. F., and Kraft, J. C. 1977. Holocene relative sea-level changes and coastal stratigraphic units on the northwest flank of the Baltimore Canyon Trough geosyncline. Journal of Sedimentary Petrology, 47: 610629.Google Scholar
Belknap, D. F., and Kraft, J. C. 1981. Preservation potential of transgressive coastal lithosomes on the U.S. Atlantic Shelf. Marine Geology, 42: 429442.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: Boyd, R., Zaitlin, B. A. and Dalrymple, R., eds., Incised Valley Fill Systems, SEPM Special Pub. 51: 303320.Google Scholar
Belknap, D. F., and Wilson, K. R. 2014. Invasive green crab impacts on salt marshes in Maine – sudden increase in erosion potential. Geological Society of America Abstracts with Programs, 46, no. 1, Abstract 55-9: 104.Google Scholar
Belknap, D. F., and Wilson, K. R. 2015. Effects of invasive Green Crabs on salt marshes in Maine. Geological Society of America Abstracts with Programs, 47, no. 1, Abstract 65-8: 127128.Google Scholar
Bertness, M. D. 1992. The ecology of a New England salt marsh. American Scientist, 80: 260268.Google Scholar
Bertness, M. D. 2007. Atlantic Shorelines: Natural History and Ecology. Princeton University Press.Google Scholar
Bertness, M. D., and Ellison, A. M. 1987. Determinants of pattern in a New England salt marsh plant community. Ecological Monographs, 57: 129147.Google Scholar
Bloom, A. L. 1964. Peat accumulation and compaction in a Connecticut coastal marsh. Journal of Sedimentary Petrology, 34: 599603.Google Scholar
Boumans, R. M., and Day, J. W. Jr. 1993. High precision measurement of surface elevation in shallow coastal areas using a sediment-erosion table. Estuaries, 16: 375380.Google Scholar
Boyd, B., and Sommerfield, C. K. 2017. Detection of fallout 241Am in U.S. Atlantic salt marsh soils. Estuarine, Coastal and Shelf Science, 196: 373378.Google Scholar
Cahoon, D. R., Lynch, J. C., and Powell, A. N. 1996. Marsh vertical accretion in a Southern California estuary U.S.A. Estuarine, Coastal and Shelf Science, 43: 1932.Google Scholar
Cahoon, D. R., and Reed, D. J. 1995. Relationships among marsh surface topography, hydroperiod, and soil accretion in a deteriorating Louisiana salt marsh. Journal of Coastal Research, 11: 357369.Google Scholar
Cahoon, D. R., Reed, D. J., and Day, J. W. Jr.. 1995. Estimating shallow subsidence in microtidal salt marshes of the southeastern United States: Kaye and Barghoorn revisited. Marine Geology, 128: 19.Google Scholar
Cahoon, D. R., and Turner, R. E. 1989. Accretion and canal impacts in a rapidly subsiding wetland II. Feldspar marker horizon technique. Estuaries, 12: 260268.Google Scholar
Chapman, V. J. 1960. Salt Marshes and Salt Deserts of the World, Interscience Publishes, Inc, New York.Google Scholar
Chmura, G. L., Anisfled, S. C., Cahoon, D. R., and Lynch, J. C. 2003. Global carbon sequestration in tidal, saline wetland soils. Global Biogeochemical Cycles, 17: 11111133.CrossRefGoogle Scholar
Chmura, G. L., Helmer, L. L., Beecher, C. B., and Sunderland, E. M. 2001. Historical rates of salt marsh accretion on the outer Bay of Fundy. Canadian Journal of Earth Sciences, 38: 10811092.Google Scholar
Clark, J. S. 1986. Late-Holocene vegetation and coastal processes at a Long Island tidal marsh. Journal of Ecology, 74: 561578.Google Scholar
Curray, J. R. 1964. Transgressions and regressions. In: Miller, R. L., ed., Papers in Marine Geology, MacMillan, New York, pp. 175203.Google Scholar
Daly, J. F., Belknap, D. F., Kelley, J. T., and Bell, T. 2007. Late Holocene sea-level change around Newfoundland. Canadian Journal of Earth Sciences, 44: 14531465.Google Scholar
Darby, F. A., and Turner, R. E.. 2008. Effects of eutrophication to salt marsh roots, rhizomes, and soils. Marine Ecology Progress Series, 363: 6370.CrossRefGoogle Scholar
Davidson-Arnott, R. G. D., van Proosdij, D. V. Ollerhead, J., and Schostak, L. 2002. Hydrodynamics and sedimentation in salt marshes: examples from a macrotidal marsh, Bay of Fundy. Geomorphology, 48: 209231.Google Scholar
Day, J. D., Britsch, L. D., Hawes, S., Shaffer, G. P., Reed, D. J., and Cahoon, D. 2000. Pattern and process of land loss in the Mississippi Delta: a spatial and temporal analysis of wetland habitat change. Estuaries, 23: 425438.CrossRefGoogle Scholar
Day, J. D., 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.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 Petrology, 53: 147157.Google Scholar
DeLaune, R. D., Nyman, J. A., and Patrick, W. H. Jr. 1994. Peat collapse, ponding and wetland loss in a rapidly submerging coastal marsh. Journal of Coastal Research, 10: 10211030.Google Scholar
DeLaune, R. D., Whitcomb, J. H., Patrick, W. H. Jr., Pardue, J. H., and Pezeshki, S. R. 1989. Accretion and canal impacts in a rapidly subsiding wetland I. 137Cs and 210Pb techniques. Estuaries, 12: 247259.Google Scholar
Dionne, M., Short, F. T., and Burdick, D. M. 1999. Fish utilization of restored, created, and reference salt-marsh habitat in the Gulf of Maine. American Fisheries Society Symposium, 22: 84404.Google Scholar
Donnelly, J. P., Bryant, S. S., Butler, J., Dowling, J., Fan, L., Hausmann, N., Newby, P., Shuman, B., Stern, J., and Webb, T. III. 2001. 700 yr sedimentary record of intense hurricane landfalls in southern New England. Geological Society of America Bulletin, 113: 714727.2.0.CO;2>CrossRefGoogle Scholar
Engelhart, S. E., and Horton, B. P. 2012. Holocene sea level database for the Atlantic coast of the United States. Quaternary Science Reviews, 54: 1225.Google Scholar
Engelhart, S. E., Horton, B. P., and Kemp, A. C. 2011. Holocene sea levels along the United States’ Atlantic coast. Oceanography, 24: 7079.Google Scholar
FitzGerald, D. M., Buynevich, I., and Argow, B. 2006. Model of tidal inlet and barrier island dynamics in a regime of accelerated sea-level rise. Journal of Coastal Research, Special Issue, 39: 789795.Google Scholar
FitzGerald, D. 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
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
Frey, R. W., and Basan, P. B. 1985. Coastal salt marshes. In: Davis, R. A. Jr., ed., Coastal Sedimentary Environments, Springer-Verlag, New York, pp. 225301.Google Scholar
Frey, R. W., and Howard, J. D. 1969. A profile of biogenic sedimentary structures in a Holocene barrier island-salt marsh complex, Georgia. Transactions of the Gulf Coast Association Geological Society, 19: 427444.Google Scholar
Gedan, K. B., and Silliman, B. R. 2009. Patterns of salt marsh loss within coastal regions of North America. In: Silliman, B., Grosholz, E., and Bertness, M.D., eds., Human Impacts on Salt Marshes: A Global Perspective, University of California Press, Los Angeles, CA, pp. 253265.Google Scholar
Gehrels, W. R. 1994. Determining relative sea-level change from salt-marsh foraminifera and plant zones on the coast of Maine, USA. Journal of Coastal Research, 10: 9901009.Google Scholar
Gehrels, W. R. 2000. Using foraminiferal transfer functions to produce high-resolution sea-level records from salt-marsh deposits, Maine, USA. The Holocene, 10: 367376.CrossRefGoogle Scholar
Gehrels, W. R., Belknap, D. F., and Kelley, J. T., 1996. Integrated high-precision analyses of Holocene relative sea-level changes: lessons from the coast of Maine. Geological Society of America Bulletin, 108: 10731088.Google Scholar
Gehrels, W. R., Kirby, J. R., Prokoph, A., Newnham, R. W., Achterberg, E. P., Evans, H., Black, S., and Scott, D. B. 2005. Onset of rapid sea-level rise in the western Atlantic Ocean. Quaternary Science Reviews, 24: 20832100.Google Scholar
Gehrels, W. R., Milne, G. A., Kirby, J. R., Patterson, R. T., and Belknap, D. F. 2004. Late Holocene sea-level changes and isostatic crustal movements in Atlantic Canada. Quaternary International, Special Issue – International Geological Correlation Program, Project 437 “Late Quaternary Highstands,” Barbados, 120: 7989.Google Scholar
Gehrels, W. R., and van de Plassche, O. 1991. Origin of the paleovalley system underlying Hammock River Marsh, Clinton, Connecticut. Journal of Coastal Research, Special Issue, 11: 7383.Google Scholar
Goodman, J. E., Wood, M. E., and Gehrels, W. R. 2007. A 17-yr record of sediment accumulation in the salt marshes of Maine (USA). Marine Geology, 242: 109121.CrossRefGoogle Scholar
Harrison, E. Z., and Bloom, A. L. 1977. Sedimentation rates on tidal salt marshes in Connecticut. Journal of Sedimentary Petrology, 47: 14841490.Google Scholar
Hartig, E. K., Gornitz, V., Kolker, A., Mushacke, F., and Fallon, D. 2002. Anthropogenic and climate-change impacts on salt marshes of Jamaica Bay, New York City. Wetlands, 22: 7189.CrossRefGoogle Scholar
Hayes, M. O., and Kana, T. W. 1976. Terrigenous clastic depositional environments, Technical Report No. 11-CRD Coastal Research Division, Department of Geology, University of South Carolina, Columbia.Google Scholar
Hine, A. C., Belknap, D. F., Hutton, J. G., Osking, E. B., and Evans, M. W. 1988. Recent geologic history and modern sedimentary processes along an incipient, low-energy, epicontinental-sea coastline: northwest Florida. Journal of Sedimentary Petrology, 58: 567579.Google Scholar
Hladik, C., and Alber, M. 2012. Accuracy assessment and correction of a LIDAR-derived salt marsh digital elevation model. Remote Sensing of Environment, 121: 224235.CrossRefGoogle Scholar
Horton, B. P., Edwards, R. J., and Lloyd, J. M. 1999. Foraminiferal-based transfer function: implications for sea-level studies. Journal of Foraminiferal Research, 29, 117129.CrossRefGoogle Scholar
Hussey, A. M. II. 1959. Age of intertidal tree stumps at Wells Beach and Kennebunk Beach, Maine. Journal of Sedimentary Petrology, 29: 464465.Google Scholar
Jacobson, H. A. 1988. Historical development of the saltmarsh at Wells, Maine. Earth Surface Processes and Landforms, 13: 475486.Google Scholar
Katz, L. C. 1980. Effects of burrowing by the fiddler crab Uca pugnax (Smith). Estuarine and Coastal 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 the autocompaction of peat. Geological Society of America Bulletin, 75: 6368.Google Scholar
Kearney, M. S., Grace, R. E., and Stevenson, J. C. 1988. Marsh loss in Nanticoke Estuary, Chesapeake Bay. Geographical Review, 78: 205220.Google Scholar
Kearney, M. S., Rogers, A. S., Townshend, J. R. G., Rizzo, E., Stutzer, D., Stevenson, J. C., and Sundborg, K., 2002. Landsat imagery shows decline of coastal marshes in Chesapeake and Delaware Bays. Eos, 83: 173, 177–178.CrossRefGoogle Scholar
Kearney, M. S., and Stevenson, J. C. 1991. Island land loss and marsh vertical accretion rate evidence for historical sea-level changes in Chesapeake Bay. Journal of Coastal Research, 7: 403415.Google Scholar
Kelley, J. T., Almquist-Jacobson, H., Jacobson, G. H. Jr., Gehrels, W. R., and Schneider, Z. 1992. The geologic and vegetative development of tidal marshes at Wells, Maine, USA. Research Report to the Wells National Estuarine Research Reserve and the National Oceanic and Atmospheric Administration.Google Scholar
Kelley, J. T., Belknap, D. F., and Claesson, S. 2010. Drowned coastal deposits with associated archaeological remains from a sea-level “slowstand,” Northwestern Gulf of Maine, USA. Geology, 38: 695698Google Scholar
Kelley, J. T., Belknap, D. F., Kelley, A. R., and Claesson, S. H. 2013. A model for drowned terrestrial habitats with associated archeological remains in the northwestern Gulf of Maine, USA. Marine Geology, 338: 116.CrossRefGoogle Scholar
Kelley, J. T., Belknap, D. F., Jacobson, G. L. Jr., and Jacobson, H. A. 1988. The morphology and origin of salt marshes along the glaciated coastline of Maine, USA. Journal of Coastal Research, 4: 649665.Google Scholar
Kelley, J. T., and Hay, B. W. B. 1986. Bunganuc Bluffs, Day 3, Stop 6. In: Kelley, J. T. and Kelley, A. R., eds. Coastal Processes and Quaternary Stratigraphy Northern and Central Coastal Maine, Society of Economic Paleontologists and Mineralogists Eastern Section Field Trip Guidebook, pp. 66–74.Google Scholar
Kennish, M. J. 2001. Salt marsh systems in the U.S.: a review of anthropogenic impacts. Journal of Coastal Research, 17: 731748.Google Scholar
Kirwan, M. L., Murray, A. B., Donnelly, J. P., and Corbett, D. R. 2011. Rapid wetland expansion during European settlement and its implication for marsh survival under modern sediment delivery rates. Geology, 39: 507510.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
Kraft, J. C. 1971. Sedimentary facies patterns and geologic history of a Holocene marine transgression. Geological Society of America Bulletin, 82: 21312158.Google Scholar
Kraft, J. C., Allen, E. A., Belknap, D. F., John, C. J., and Maurmeyer, E. M. 1976. Delaware's Changing Shorelines. Technical Report #1, Delaware Coastal Zone Management Program, Dover.Google Scholar
Kraft, J. C., Allen, E. A., Belknap, D. F., John, C. J. and Maurmeyer, E. M. 1979. Processes and morphologic evolution of an estuarine and coastal barrier system, In: Leatherman, S. P., ed., Barrier Islands, Academic Press, New York, pp. 149183.Google Scholar
Leatherman, S. P. 1979. Migration of Assateague Island, Maryland, by inlet and overwash processes. Geology, 7: 104107.Google Scholar
Leonard, L. A., and Luther, M. E. 1995. Flow hydrodynamics in tidal marsh canopies. Limnology and Oceanography, 40: 14741484.Google Scholar
Letzsch, S. W., and Frey, R. W. 1980. Deposition and erosion in a Holocene salt marsh, Sapelo Island, Georgia. Journal of Sedimentary Petrology, 50: 529542.Google Scholar
Meredith, W. H., Saveikis, D. E., and Stachecki, C. J. 1985. Guidelines for “Open Marsh Water Management” in Delaware’s salt marshes – objectives, system designs, and installation. Wetlands, 5: 119133.Google Scholar
Moller, I., Kudella, M., Rupprecht, F., Spencer, T., Paul, M., van Wesenbeeck, B., Wolters, G., et al. Wave attenuation over coastal salt marshes under storm surge conditions. 2014. Nature Geoscience, 7: 727731.CrossRefGoogle Scholar
Morris, J. T., Porter, D., Neet, M., Noble, P. A., Schmidt, L., Lapine, L. A., and Jensen, J. R. 2005. Integrating LIDAR elevation data, multispectral imagery and neural network modeling for marsh characterization. International Journal of Remote Sensing, 26: 52215234.CrossRefGoogle Scholar
Mudd, S. M. 2011. The life and death of salt marshes in response to anthropogenic disturbance of sediment supply. Geology, 39: 511512.Google Scholar
Mudge, B. F. 1858. The salt marsh formations of Lynn. Proceedings of Essex Institute, 2: 117119.Google Scholar
National Park Service – Cape Cod National Seashore, 2017. Crab-driven vegetation losses: www.nps.gov/caco/learn/nature/crab-driven-vegetation-losses.htmGoogle Scholar
Neuendorf, K. K. E, Mehl, J. P. Jr., and Jackson, J. A. 2005. Glossary of Geology 5th Edn., American Geological Institute, Alexandria, VA.Google Scholar
Niering, W. A., and Warren, R. S. 1980. Vegetation patterns and processes in New England salt marshes. Bioscience, 30: 301307.CrossRefGoogle Scholar
Nikitina, D. L., Kemp, A. C., Horton, B. P., Vane, C. H., van de Plassche, O., and Engelhardt, S. E. 2014. Storm erosion during the past 2000 years along the north shore of Delaware Bay, USA. Geomorphology, 208: 160172.Google Scholar
Orson, R., Panageotou, W., Leatherman, S. P. 1985. Response of tidal salt marshes of the U.S. Atlantic and Gulf coasts to rising sea levels. Journal of Coastal Research, 1: 29–7.Google Scholar
Orson, R. A., Warren, R. S., and Niering, W. A. 1987. Development of a tidal marsh in a New England river valley. Estuaries, 10: 2027.Google Scholar
Orson, R. A., Warren, R. S., Niering, W. A., and Van Patten, P., eds. 1998. Research in New England Marsh-Estuarine Ecosystems, Directions and Priorities into the Next Millennium: Summary of a Sea Grant Workshop, May 15–17, 1997, 61 pp., Connecticut College, New London, CT. Connecticut Sea Grant College Program, Groton, CT: 5-11.Google Scholar
Parkinson, R. W., Craft, C., DeLaune, R. D., Donoghue, J. F., Kearney, M., Meeder, J. F., Morris, J., and Turner, R. E. 2017. Marsh vulnerability to sea-level rise. Nature Climate Change, 7: 756.Google Scholar
Rampino, M. R., and Sanders, J. E. 1980. Holocene transgression in south-central Long Island, New York. Journal of Sedimentary Petrology, 50: 10631080.Google Scholar
Redfield, A. C. 1965. Ontogeny of a salt marsh 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. Proceedings of the National Academy of Sciences, 48: 17281735.Google Scholar
Reed, D. J. 1989. Patterns of sediment deposition in subsiding coastal salt marshes, Terrebone Bay, Louisiana: the role of winter storms. Estuaries, 12: 222227.Google Scholar
Reed, D. J. 1990. The impact of sea-level rise on coastal salt marshes. Progress in Physical Geography, 14: 465481.Google Scholar
Reed, D. J. 1995. The response of coastal marshes to sea-level rise: survival or submergence? Earth Surface Processes and Landforms, 20: 3948.Google Scholar
Reed, D. J. 2002. Sea-level rise and coastal marsh sustainability: geological and ecological factors in the Mississippi delta. Geomorphology, 48: 233243.Google Scholar
Roberts, M. F. 1979. The Tidemarsh Guide, E. P. Dutton, New York.Google Scholar
Rogers, K., and Woodroffe, C. D. 2014. Tidal flats and salt marshes. In: Masselink, G., and Gehrels, R., eds., Global Environments and Global Change, John Wiley and Sons, Ltd., Chichester, UK, pp. 227250.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, 45: 717727.Google Scholar
SDNHM (San Diego Natural History Museum). 2006. www.sdplantatlas.org/NameChanges.aspx, Genus Scirpus is now Schoenoplectus.Google Scholar
Schwimmer, R. A. 2001. Rates and processes of marsh shoreline erosion in Rehoboth Bay, Delaware, USA. Journal of Coastal Research, 17: 672683.Google Scholar
Scott, D. B., and Greenberg, D. A. 1983. Relative sea-level rise and tidal development in the Fundy tidal system. Canadian Journal of Earth Sciences, 20: 15541564.Google Scholar
Scott, D. B., and Medioli, F. S. 1978. Vertical zonations of marsh foraminifera as accurate indicators of former sea levels. Nature, 272: 528531.Google Scholar
Sepanik, J. M., and McBride, R. A. 2015. Increasing rate of salt-marsh loss in a barrier-island system: Parramore and Cedar Islands, Virginia, from 1957 to 2012, Section 1.6: pp. 392–401 of: McBride, R. A. Fenster, M. S., Seminack, C. T., Richardson, T. M., Sepanik, J. M., Hanley, J. T., Bundick, J. A. and Tedder, E., Holocene barrier-island geology and morphodynamics of the Maryland and Virginia open-ocean coasts: Fenwick, Assateague, Chincoteague, Wallops, Cedar and Parramore Islands, in Brezinski, D. K., Halka, J. P., and Ortt, R. A., Jr., eds., Tripping from the Fall Line: Field Excursions for the GSA Annual Meeting, Baltimore, 2015: Geological Society of America Field Guide 40, Boulder, CO: 309–424.Google Scholar
Shaler, N. S. 1885. Preliminary report on sea-coast swamps of the Eastern United States: U.S. Geological Survey 6th Annual Report, 1885: pp. 353–398.Google Scholar
Shepard, F. P., 1960, Gulf coast barriers. In: Shepard, F. P., Phleger, F. B., and von Andel, T. H., eds., Recent Sediments, Northwest Gulf of Mexico, American Association of Petroleum Geologists, Tulsa, Oklahoma, pp. 5681.Google Scholar
Silliman, B. R., Grosholz, E. D., and Bertness, M. D., (eds.). 2009. Human Impacts on Salt Marshes: a global perspective. University of California Press, Berkeley, CA.Google Scholar
Silliman, B. R., Van der Kopple, J., Bertness, M. D., Stanton, I. E., and Mendelssohn, I. A. 2005. Drought, snails, and large-scale dieoff of southern U.S. salt marshes: Ecology, 310: 18031806.Google Scholar
Smith, D. C., and Bridges, A. E., 1982. Salt marsh dikes (dykes) as a factor in eastern Maine agriculture. Maine Historical Society Quarterly, 21: 219226.Google Scholar
Smith, D. C., Konrad, V., Koularis, H., Borns, H. W. Jr., and Hawes, E. 1989. Salt marshes as a factor in the agriculture of northeastern North America. Agricultural History, 63: 270294.Google Scholar
Snow, J. O. 1980. Secrets of a Salt Marsh. Guy Gannett Pub. Co, Portland, ME.Google Scholar
Stea, R. R., Fader, G. B. J., Scott, D. B., and Wu, P. 2001. Glaciation and relative sea-level change in Maritime Canada. In: Weddle, T. K., and Retelle, M. J., eds., Deglacial History and Relative Sea-Level Changes Northern New England and Adjacent Canada, Geological Society of America Special Paper 351: 3549.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
Stumpf, R. P. 1983. The process of sedimentation on the surface of a salt marsh. Estuarine, Coastal and Shelf Science, 17: 495508.Google Scholar
Swift, D. J. P. 1968. Coastal erosion and transgressive stratigraphy. Journal of Geology, 77: 444456.Google Scholar
Swift, D. J. P. 1975. Barrier island genesis: evidence from the central Atlantic shelf, eastern U.S.A. Sedimentary Geology, 14: 143.Google Scholar
Swisher, M. L. 1982. The rates and causes of shore erosion around a coastal lagoon, Rehoboth Bay, Delaware: M.S. thesis, Dept. Geology, University of Delaware, Newark.Google Scholar
Syvitski, J. P. M., and Saito, Y. 2007. Morphodynamics of deltas under the influence of humans. Global and Planetary Change, 57: 261282.Google Scholar
Tiner, R. W. 2009. Field Guide to Tidal Wetland Plants of the Northeastern United States and Neighboring Canada. University of Massachusetts Press, Amherst, MA.Google Scholar
Turner, R. E., Baustain, J. J., Swenson, E. M., and Spicer, J. S. 2006. Wetland sedimentation from Hurricanes Katrina and Rita. Science, 314: 449452.Google Scholar
Turner, R. E., Howes, B. L., Teal, J. M., Milan, C. S., Swenson, E. M., and Goehringer-Toner, D. 2009. Salt marshes and eutrophication: an unsustainable outcome. Limnology and Oceanography, 54: 16341642.Google Scholar
van de Plascche, O. 1986. Sea-level Research: a Manual for the Collection and Evaluation of Data. Geo Books, Norwich, England.Google Scholar
van de Plassche, O. 1991. Late Holocene sea-level fluctuations on the shore of Connecticut inferred from transgressive and regressive overlap boundaries in salt-marsh deposits: Origin of the paleovalley system underlying Hammock River Marsh, Clinton, Connecticut. Journal of Coastal Research, Special Issue 11: 159179.Google Scholar
Wang, C., Meneti, M., Stoll, M.-P., Feola, A., Belluco, E., and Marani, M. 2009. Separation of ground and low vegetations signatures in LiDAR measurements of salt-marsh environments. IEEE Transactions on Geoscience and Remote Sensing, 47: 20142023.Google Scholar
Ward, L. G., Zaprowski, B. J., Trainer, K. D., and Davis, P. T. 2008. Stratigraphy, pollen history and geochronology of tidal marshes in a Gulf of Maine estuarine system: climatic and relative sea level impacts. Marine Geology, 256: 117.Google Scholar
Wilson, K. R., Kelley, J. T., Croitoru, A., Dionne, M., Belknap, D. F., and Steneck, R. S. 2009. Stratigraphic and ecophysical characterizations of salt pools: dynamic features of the Webhannet Estuary salt marsh, Wells, Maine, USA. Estuaries and Coasts, 32: 855870.Google Scholar
Wilson, K. R., 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
Wood, M. E., Kelley, J. T., and Belknap, D. F. 1989. Pattern of sediment accumulation in the tidal marshes of Maine. Estuaries, 12: 237246.Google Scholar
Woodwell, G. M., Rich, P. H., and Hall, C. A. S. 1973. Carbon in estuaries. In: Woodwell, G. M., and Pecan, E., eds., Carbon and the Biosphere, U.S. Atomic Energy Commission, Springfield, VA, USA, pp. 221–240.Google Scholar
Yelverton, G. F., and Hackney, C. T. 1986. Flux of dissolved organic carbon and pore water through the substrate of a Spartina alterniflora marsh in North Carolina. Estuarine, Coastal, and Shelf Science, 22: 255267.Google Scholar

References

Adam, P. 1990. Saltmarsh Ecology. Cambridge University Press, Cambridge; New York.Google 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.Google 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.Google Scholar
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.Google 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.CrossRefGoogle 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.Google 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.Google 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.Google Scholar
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.Google 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.Google 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.Google Scholar
Broome, S. W., Seneca, E. D., and Woodhouse, W. W. 1988. Tidal salt marsh restoration. Aquatic Botany, 32: 122.Google Scholar
Bruno, J. F. 2000. Facilitation of cobble beach plant communities through habitat modification by Spartina alterniflora. Ecology, 81: 11791192.Google 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.Google 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.Google Scholar
Craft, C. 2000. Co-development of wetland soils and benthic invertebrate communities following salt marsh creation. Wetlands Ecology and Management, 8: 197207.CrossRefGoogle 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.Google 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
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.Google 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.Google 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.Google Scholar
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.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.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.Google 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.Google 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.Google Scholar
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.Google 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.Google 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.Google Scholar
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.Google 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.Google 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.Google 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.Google 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.Google 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.Google 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.Google 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.Google Scholar
Pethick, J. S. 1981. Long-term accretion rates on tidal salt marshes. Journal of Sedimentary Research, 51: 571577.Google Scholar
Phleger, C. F. 1971. Effect of salinity on growth of a salt marsh grass. Ecology, 52: 908911.Google 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.Google Scholar
Redfield, A. C. 1965. Ontogeny of a salt marsh estuary. Science, 147: 5055.Google Scholar
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.CrossRefGoogle 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.Google 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.Google 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

References

Ahmad, M. F., Dong, P., Mamat, M., Nik, W. B. W., and Mohd, M. H. 2011. The critical shear stresses for sand and mud mixture. Applied Mathematical Sciences, 5: 5371.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
Amos, C. L., Bergamasco, A., Umgiesser, G., Cappucci, S., Cloutier, D., Denat, L., Flind, M., Bonardi, M., and Cristante, S. 2004. The stability of tidal flats in Venice Lagoon–the results of in-situ measurements using two benthic, annular flumes. Journal of Marine Systems, 51: 211241.Google Scholar
Amos, C. L., Umgiesser, G., Ferrarin, C., Thompson, C. E. L. C. , Whitehouse, R. J. S., Sutherland, T. F., and Bergamasco, A. 2010. The erosion rates of cohesive sediments in Venice lagoon, Italy. Continental Shelf Research, 30: 859870.Google Scholar
Balke, T., Bouma, T. J., Horstman, E. M., Webb, E. L., Erftemeijer, P. L. A., and Herman, P. M. J. 2011. Windows of opportunity: thresholds to mangrove seedling establishment on tidal flats. Marine Ecology Progress Series, 440: 19.Google Scholar
Balke, T., Klaassen, P. C., Garbutt, A., Van der Wal, D., Herman, P. M. J., and Bouma, T. J. 2012. Conditional outcome of ecosystem engineering: A case study on tussocks of the salt marsh pioneer Spartina anglica. Geomorphology, 153154: 232238.Google Scholar
Baptist, M. J., Babovic, V., Rodríguez-Uthurburu, J., Keijzer, M., Uittenbogaard, R. E., Mynett, A., and, Verwey, A. 2007. On inducing equations for vegetation resistance. Journal of Hydraulic Research, 45: 435450.Google Scholar
Bayliss-Smith, T. P., Healey, R., Lailey, R., Spencer, T., and, Stoddart, D. R. 1979. Tidal flows in salt marsh creeks. Estuarine and Coastal Marine Science, 9: 235255.Google Scholar
Beeftink, W. G., and Rozema, J. 1993. The nature and functioning of salt marshes. In: Pollution of the North Sea, Salomons, W, Bayne, B. L., Duursma, E. K., and Forstner, U. (eds). Springer, Berlin, Heidelberg, pp. 5987.Google Scholar
Belliard, J.-P., Toffolon, M., Carniello, L., and D’Alpaos, A. 2015. An ecogeomorphic model of tidal channel initiation and elaboration in progressive marsh accretional contexts. Journal of Geophysical Research: Earth Surface, 120: 10401064.Google Scholar
Bendoni, M., Francalanci, S., Cappietti, L., and Solari, L. 2014. On salt marshes retreat: Experiments and modeling toppling failures induced by wind waves. Journal of Geophysical Research: Earth Surface, 119: 603620.Google Scholar
Bendoni, M., Mel, R., Lanzoni, S., Francalanci, S., and Oumeraci, H. 2016. Insights into lateral marsh retreat mechanism through localized field measurements. Water Resources Research, 52: 14461464.Google Scholar
Boon, J. D. I. 1975. Tidal discharge asymmetry in a salt marsh drainage system. Limnology and Oceanography, 20: 7180.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.Google Scholar
Brivio, L., Ghinassi, M., D’Alpaos, A., Finotello, A., Fontana, A., Roner, M., and Howes, N. 2016. Aggradation and lateral migration shaping geometry of a tidal point bar: An example from salt marshes of the Northern Venice Lagoon (Italy). Sedimentary Geology, 343: 141155.Google Scholar
Callaghan, D. P., Bouma, T. J., Klaassen, P., van der Wal, D., Stive, M. J. F., and Herman, P. M. J. 2010. Hydrodynamic forcing on salt-marsh development: Distinguishing the relative importance of waves and tidal flows. Estuarine, Coastal and Shelf Science, 89: 7388.Google Scholar
Callaway, J. C., and Josselyn, M. N. 1992. The introduction and spread of smooth cordgrass Spartina alterniflora in South San Francisco Bay. Estuaries, 15: 218226.Google Scholar
Carniello, L., D’Alpaos, A., and Defina, A. 2011. Modeling wind waves and tidal flows in shallow micro-tidal basins. Estuarine, Coastal and Shelf Science, 92: 263276.Google Scholar
Carniello, L., Defina, A., and D’Alpaos, L. 2012. Modeling sand-mud transport induced by tidal currents and wind waves in shallow microtidal basins: Application to the Venice Lagoon (Italy). Estuarine, Coastal and Shelf Science, 102103: 105115.Google Scholar
Chen, Y., Li, Y., Cai, T., Thompson, C., and Li, Y. 2016. A comparison of biohydrodynamic interaction within mangrove and saltmarsh boundaries. Earth Surface Processes and Landforms, 41: 19671979.Google Scholar
Chen, Z., Ortiz, A., Zong, L., and Nepf, H. 2012. The wake structure behind a porous obstruction and its implications for deposition near a finite patch of emergent vegetation. Water Resources Research, 48: 112.Google Scholar
Coco, G., Zhou, Z., van Maanen, B., Olabarrieta, M., Tinoco, R., and Townend, I. H. 2013. Morphodynamics of tidal networks: Advances and challenges. Marine Geology, 346: 116.Google Scholar
Costanza, R., d'Arge, R., de Groot, R., Farber, S., Grasso, M., Hannon, B., Limburg, K., et al. 1997. The value of the world’s ecosystem services and natural capital. Nature, 387: 253260.Google Scholar
D’Alpaos, A., Ghinassi, M., Finotello, A., Brivio, L., Bellucci, L. G. L. G., and Marani, M. 2017. Tidal meander migration and dynamics: A case study from the Venice Lagoon. Marine and Petroleum Geology, 87: 8090.Google Scholar
D’Alpaos, A., Lanzoni, S., Marani, M., Bonometto, A., Cecconi, G., and Rinaldo, A. 2007a. Spontaneous tidal network formation within a constructed salt marsh: Observations and morphodynamic modelling. Geomorphology, 91: 186197. DOI: 10.1016/j.geomorph.2007.04.013Google Scholar
D’Alpaos, A., Lanzoni, S., Marani, M., Fagherazzi, S., and Rinaldo, A. 2005. Tidal network ontogeny: Channel initiation and early development. Journal of Geophysical Research: Earth Surface, 110: 114.Google Scholar
D’Alpaos, A., Lanzoni, S., Marani, M., and Rinaldo, A. 2007b. Landscape evolution in tidal embayments: Modeling the interplay of erosion, sedimentation, and vegetation dynamics. Journal of Geophysical Research: Earth Surface, 112: 117.Google Scholar
D’Alpaos, A., Lanzoni, S., Marani, M., and Rinaldo, A. 2009. On the O’Brien–Jarrett–Marchi law. Rendiconti Lincei, 20: 225236.Google Scholar
D’Alpaos, A., Lanzoni, S., Marani, M., and Rinaldo, A. 2010. On the tidal prism-channel area relations. Journal of Geophysical Research: Earth Surface, 115: 113.Google Scholar
D’Alpaos, A., Lanzoni, S., Mudd, S. M., and Fagherazzi, S. 2006. Modeling the influence of hydroperiod and vegetation on the cross-sectional formation of tidal channels. Estuarine, Coastal and Shelf Science, 69: 311324.Google Scholar
D’Alpaos, A., and Marani, M. 2016. Reading the signatures of biologic-geomorphic feedbacks in salt-marsh landscapes. Advances in Water Resources, 93: 265275.Google Scholar
Defina, A., Carniello, L., Fagherazzi, S., and D’Alpaos, L. 2007. Self-organization of shallow basins in tidal flats and salt marshes. Journal of Geophysical Research: Earth Surface, 112: 111.Google Scholar
Di Silvio, G., Dall’Angelo, C., Bonaldo, D., Fasolato, G., Dall’Angelo, C., Bonaldo, D., and Fasolato, G. 2010. Long term model of planimetric and bathymetric evolution of a tidal lagoon. Continental Shelf Research, 30: 894903.Google Scholar
Dronkers, J. 2016. Dynamic of Coastal System. 2nd edn. World Scientific, Singapore.Google Scholar
Dronkers, J. J. 1964. Tidal Computations in Rivers and Coastal Waters. North Holland, Amsterdam.Google Scholar
Fagherazzi, S., Bortoluzzi, A., Dietrich, W. E., Adami, A., Lanzoni, S., Marani, M., and Rinaldo, A. 1999. Tidal networks 1. Automatic network extraction and preliminary scaling features from digital terrain maps. Water Resources Research, 35: 38913904.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 United States of America, 103: 83378341.Google Scholar
Fagherazzi, S., and Furbish, D. J. 2001. On the shape and widening of salt marsh creeks. Journal of Geophysical Research, 106: 991.Google Scholar
Fagherazzi, S., Gabet, E. J., and, Furbish, D. J. 2004. The effect of bidirectional flow on tidal channel planforms. Earth Surface Processes and Landforms, 29: 295309.Google Scholar
Fagherazzi, S., Hannion, M., and D’Odorico, P. 2008. Geomorphic structure of tidal hydrodynamics in salt marsh creeks. Water Resources Research, 44: 112.Google Scholar
Fagherazzi, S., Kirwan, M. L., Mudd, S. M., Guntenspergen, G. R., Temmerman, S., D’Alpaos, A., van de Koppel, J. et al. 2012. Numerical models of salt marsh evolution: Ecological, geomorphic, and climatic factors. Reviews of Geophysics, 50: 128.Google Scholar
Fagherazzi, S., and Sun, T. 2004. A stochastic model for the formation of channel networks in tidal marshes. Geophysical Research Letters, 31: 14.Google Scholar
Fagherazzi, S., Wiberg, P. L., Temmerman, S., Struyf, E., Zhao, Y., and Raymond, P. A. 2013. Fluxes of water, sediments, and biogeochemical compounds in salt marshes. Ecological Processes, 2: 116.Google Scholar
Finotello, A., Lanzoni, S., Ghinassi, M., Marani, M., Rinaldo, A., and D’Alpaos, A., 2018. Field migration rates of tidal meanders recapitulate fluvial morphodynamics. Proceedings of the National Academy of Sciences of the United States of America, 115: 14631468.Google Scholar
Folkard, A. M. 2011. Flow regimes in gaps within stands of flexible vegetation: Laboratory flume simulations. Environmental Fluid Mechanics, 11: 289306.Google Scholar
Francalanci, S., Bendoni, M., Rinaldi, M., and Solari, L. 2013. Ecomorphodynamic evolution of salt marshes: Experimental observations of bank retreat processes. Geomorphology, 195: 5365.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. 1995. Stability shear stress and equilibrium cross-sectional of sheltered tidal channels. Journal of Coastal Research, 11: 10621074.Google Scholar
Friedrichs, C. T., and Perry, J. E. 2001. Tidal salt marsh morphodynamics: A synthesis. Journal of Coastal Research, SI: 737.Google Scholar
Gabet, E. J. 1998. Lateral migration and bank erosion in a saltmarsh tidal channel in San Francisco Bay, California. Estuaries 21: 745753.Google Scholar
Garofalo, D. 1980. The influence of wetland vegetation on tidal stream channel migration and morphology. Estuaries, 3: 258270.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
Ghinassi, M., D'alpaos, A., Gasparotto, A., Carniello, L., Brivio, L., Finotello, A., Roner, M. et al. 2018. Morphodynamic evolution and stratal architecture of translating tidal point bars: Inferences from the northern Venice Lagoon (Italy). Sedimentology, 65: 13541377.Google Scholar
Hartig, E. K., Gornitz, V., Kolker, A., Mushacke, F., and Fallon, D. 2002. Anthropogenic and climate-change impacts on salt marshes of Jamaica Bay, New York City. Wetlands, 22: 7189.Google Scholar
Horton, R. E. 1945. Erosional development of streams and their drainage basins; Hydrophysical approach to quantitative morphology. Geological Society of America Bulletin, 56: 151180.Google Scholar
Hu, K., Chen, Q., and Wang, H. 2015a. A numerical study of vegetation impact on reducing storm surge by wetlands in a semi-enclosed estuary. Coastal Engineering, 95: 6676.Google Scholar
Hu, X., and Chen, C. T. 2005. Refraction of water waves by periodic cylinder arrays. Physical Review Letters, 95: 14.Google Scholar
Hu, Z., van Belzen, J., van der Wal, D., Balke, T., Wang, Z. B., Stive, M. and Bouma, T. J. 2015b. Windows of opportunity for salt marsh vegetation establishment on bare tidal flats: The importance of temporal and spatial variability in hydrodynamic forcing. Journal of Geophysical Research: Biogeosciences, 120: 14501469.Google Scholar
Hughes, Z. J. 2012. Tidal channels on tidal flats and marshes. In: Principles of Tidal Sedimentology, Davis, R. A. and Dalrymple, R. W. (eds). Springer, Dordrecht, pp. 269300.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
Jarrett, J. T. 1976. Tidal prism-inlet area relationships. Joural of Waterways and Harbors, 95: 4352.Google Scholar
Julian, J. P., and Torres, R. 2006. Hydraulic erosion of cohesive riverbanks. Geomorphology, 76: 193206.Google Scholar
Kearney, W. S., and Fagherazzi, S. 2016. Salt marsh vegetation promotes efficient tidal channel networks. Nature Communications, 7: 17.Google Scholar
Kleinhans, M. G., Schuurman, F., Bakx, W., and Markies, H. 2009. Meandering channel dynamics in highly cohesive sediment on an intertidal mud flat in the Westerschelde estuary, the Netherlands. Geomorphology, 105: 261276.Google Scholar
Lanzoni, S., and Seminara, G. 1998. On tide propagation in convergent estuaries. Journal of Geophysical Research: Oceans, 103: 3079330812.Google Scholar
Lanzoni, S., and Seminara, G. 2002. Long-term evolution and morphodynamic equilibrium of tidal channels. Journal of Geophysical Research, 107: 113.Google Scholar
Leonard, L. A., and Croft, A. L. 2006. The effect of standing biomass on flow velocity and turbulence in Spartina alterniflora canopies. Estuarine, Coastal and Shelf Science, 69: 325336.Google Scholar
Leonard, L. A., and Luther, M. E. 1995. Flow hydrodynamics in tidal marsh canopies. Limnology and Oceanography, 40: 14741484.Google Scholar
Leonardi, N., Defne, Z., Ganju, N. K., and Fagherazzi, S. 2016a. Salt marsh erosion rates and boundary features in a shallow Bay. Journal of Geophysical Research: Earth Surface, 121: 18611875.Google Scholar
Leonardi, N., and Fagherazzi, S. 2014. How waves shape salt marshes. Geology, 42: 887890.Google Scholar
Leonardi, N., Ganju, N. K., and Fagherazzi, S. 2016b. A linear relationship between wave power and erosion determines salt-marsh resilience to violent storms and hurricanes. Proceedings of the National Academy of Sciences, 113: 6468.Google Scholar
Leopold, L. B., Collins, J. N., and Collins, L. M. 1993. Hydrology of some tidal channels in estuarine marshland near San Francisco. Catena, 20: 469493.Google Scholar
Da Lio, C., D’Alpaos, A., and Marani, M. 2013. The secret gardener: vegetation and the emergence of biogeomorphic patterns in tidal environments. Philosophical Transactions. Series A, Mathematical, Physical, and Engineering Sciences, 371: 20120367.Google Scholar
López, F., and García, M. H. 2001. Mean flow and turbulence structure of open-channel flow through non-emergent vegetation. Journal of Hydraulic Engineering, 127: 392402.Google Scholar
Marani, M., Belluco, E., D’Alpaos, A., Defina, A., Lanzoni, S., and Rinaldo, A. 2003. On the drainage density of tidal networks. Water Resources Research, 39: 111.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: 115.Google Scholar
Marani, M., D’Alpaos, A., Lanzoni, S., and Santalucia, M. 2011. Understanding and predicting wave erosion of marsh edges. Geophysical Research Letters, 38: 15.Google Scholar
Marani, M., Lanzoni, S., Zandolin, D., Seminara, G., and Rinaldo, A. 2002. Tidal meanders. Water Resources Research, 38: 714.Google Scholar
Marchi, E. 1990. Sulla stabilità delle bocche lagunari a marea. Rendiconti Lincei, 1: 137150.Google Scholar
Mariotti, G. 2018. Marsh channel morphological response to sea level rise and sediment supply. Estuarine, Coastal and Shelf Science, 209: 89101.Google Scholar
Mitsch, W. J., and Gosselink, J. G. 2000. The value of wetlands: importance of scale and landscape setting. Ecological Economics, 35: 2533.Google Scholar
Möller, I., Spencer, T., French, J. R., Leggett, D. J., and Dixon, M. 1999. Wave transformation over saltmarshes: A field and numerical modelling study from North Norfolk, England. Estuarine, Coastal and Shelf Science, 49: 411426.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.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: Earth Surface, 115: 114.Google Scholar
Mudd, S. M., Fagherazzi, S., Morris, J. T., and Furbish, D. J. 2004. Flow, sedimentation, and biomass production on a vegetated salt marsh in South Carolina: Toward a predictive model of marsh morphologic and ecologic evolution. In: The Ecogeomorphology of Tidal Marshes, Coastal and Estuarine Studies n. 59, Fagherazzi, S., Marani, M., and Blum, L. K. (eds). American Geophysical Union, Washington, D.C., pp. 165188.Google Scholar
Myrick, R. M., and Leopold, L. B. 1963. Hydraulic geometry of a small tidal estuary. United States Geological Survey Professional Paper 422: 118.Google Scholar
Nepf, H. M. 1999. Drag, turbulence, and diffusion in flow through emergent vegetation. Water Resources Research, 35: 479489.Google Scholar
Nepf, H. M. 2012. Hydrodynamics of vegetated channels. Journal of Hydraulic Research, 50: 262279.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
Nichols, M. M., Johnson, G. H., and Peebles, P. C. 1991. Modern sediments and facies model for a microtidal coastal plain estuary, the James Estuary, Virginia. Journal of Sedimentary Petrology, 61: 883899.Google Scholar
Nikora, N., and Nikora, V. 2007. A viscous drag concept for flow resistance in vegetated channels. Proceedings of the 32nd IAHR Congress, Venice.Google Scholar
O’Brien, M. P. 1969. Equilibrium flow areas of inlets on sandy coasts. Journal of Waterways and Harbors, 95: 4352.Google Scholar
Van Oyen, T., Carniello, L., D’Alpaos, A., Temmerman, S., Troch, P., and Lanzoni, S. 2014. An approximate solution to the flow field on vegetated intertidal platforms: Applicability and limitations. Journal of Geophysical Research F: Earth Surface, 119: 16821703.Google Scholar
Van Oyen, T., Lanzoni, S., D’Alpaos, A., Temmerman, S., Troch, P., and Carniello, L. 2012. A simplified model for frictionally dominated tidal flows. Geophysical Research Letters, 39: 16.Google Scholar
Pestrong, R. 1972. Tidal-flat sedimentation at cooley landing, Southwest San Francisco bay. Sedimentary Geology, 8: 251288.Google Scholar
Pethick, J. 1992. Saltmarsh geomorphology. In: Saltmarshes: Morphodynamics, Conservation and Engineering Significance, Allen, J. R. L., and Pye, K. (eds). Cambridge University Press, Cambridge, pp. 4162.Google Scholar
Pethick, J. S. 1969. Drainage in salt marshes. In: The Coastline of England and Wales. 3rd edn. Steers, J. R. (ed.). Cambridge University Press: Cambridge, pp. 752730.Google Scholar
Pethick, J. S. 1980. Velocity surges and asymmetry in tidal channels. Estuarine and Coastal Marine Science, 11: 331345.Google Scholar
Pye, K., and French, P. 1993. Erosion & Accretion Processes on British Salt Marshes. Cambridge Environmental Research Consultants.Google Scholar
Redfield, A. C. 1972. Development of a New England salt marsh. Ecological Monographs, 42: 201237.Google Scholar
Rietkerk, M., and van de Koppel, J. 2008. Regular pattern formation in real ecosystems. Trends in Ecology and Evolution, 23: 169175.Google Scholar
Rigon, R., Rinaldo, A., and Rodriguez-Iturbe, I. 1994. On landscape self-organization. Journal of Geophysical Research: Solid Earth, 99: 1197111993.Google Scholar
Rinaldo, A., Dietrich, W. E., Rigon, R., Vogel, G. K., and Rodriguez-Iturbe, I. 1995. Geomorphological signatures of varying climate. Nature, 374: 632635.Google Scholar
Rinaldo, A., Fagherazzi, S., Lanzoni, S., Marani, M., and Dietrich, W. E. 1999a. Tidal networks 2. Watershed delineation and comparative network morphology. Water Resources Research, 35: 39053917.Google Scholar
Rinaldo, A., Fagherazzi, S., Lanzoni, S., Marani, M., and Dietrich, W. E. 1999b. Tidal networks 3. Landscape-forming discharges and studies in empirical geomorphic relationships. Water Resources Research, 35: 39193929.Google Scholar
Rinaldo, A., Rodriguez-Iturbe, I., Rigon, R., Ijjasz-Vasquez, E., and Bras, R. L. 1993. Self-organized fractal river networks. Physical Review Letters, 70: 822825.Google Scholar
Rupprecht, F., Möller, I., Paul, M., Kudella, M., Spencer, T., van Wesenbeeck, B. K., Wolters, G., et al. 2017. Vegetation-wave interactions in salt marshes under storm surge conditions. Ecological Engineering, 100: 301315.Google Scholar
Salehi, M., and Strom, K. 2012. Measurement of critical shear stress for mud mixtures in the San Jacinto estuary under different wave and current combinations. Continental Shelf Research, 47: 7892.Google Scholar
Schwarz, C., Ye, Q., Wal, D., Zhang, L., Bouma, T., Ysebaert, T., and Herman, P. 2014. Impacts of salt marsh plants on tidal channels initiation and inheritance. Journal of Geophysical Research: Earth Surface, 119: 385400.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
Shi, Z., Hamilton, L. J., and Wolanski, E. 2000. Near-bed currents and suspended sediment transport in saltmarsh canopies. Journal of Coastal Research, 16: 909914.Google Scholar
Silinski, A., Heuner, M., Schoelynck, J., Puijalon, S., Schröder, U., Fuchs, E., Troch, P., et al. 2015. Effects of wind waves versus ship waves on tidal marsh plants: A flume study on different life stages of Scirpus maritimus. PLOS ONE, 10: 116.Google Scholar
Soulsby, R. L. 1997. Dynamics of Marine Sands. Thomas Telford Publications, London.Google Scholar
Soulsby, R. L., and Clarke, S. 2005. Bed shear-stresses under combined waves and currents on smooth and rough beds. Hydraulics Research Report, 1905: TR 137.Google Scholar
Steel, T. J., and Pye, K. 1997. The development of salt marsh creek networks: Evidence from the UK. Canadian Coastal Conference, pp. 1–16.Google Scholar
Stefanon, L., Carniello, L., D’Alpaos, A., Lanzoni, S., D’Alpaos, A., and Lanzoni, S. 2010. Experimental analysis of tidal network growth and development. Continental Shelf Research 30: 950962.Google Scholar
Stefanon, L., Carniello, L., D’Alpaos, A., and Rinaldo, A. 2012. Signatures of sea level changes on tidal geomorphology: Experiments on network incision and retreat. Geophysical Research Letters, 39: 16.Google Scholar
Strahler, A. N. 1957. Quantitative analysis of watershed geomorphology. Eos, Transactions American Geophysical Union, 38: 913920.Google Scholar
Tambroni, N., Luchi, R., and Seminara, G. 2017. Can tide dominance be inferred from the point bar pattern of tidal meandering channels? Journal of Geophysical Research: Earth Surface, 122: 121.Google Scholar
Tanino, Y., and Nepf, H. M. 2008. Lateral dispersion in random cylinder arrays at high Reynolds number. Journal of Fluid Mechanics, 600: 339371.Google Scholar
Tanino, Y., and Nepf, H. M. 2009. Laboratory investigation of lateral dispersion within dense arrays of randomly distributed cylinders at transitional Reynolds number. Physics of Fluids, 21: 113.Google Scholar
Temmerman, S., Bouma, T. J., Govers, G., Wang, Z. B., De Vries, M. B., Herman, P. M. J., De Vries, M. B., and Herman, P. M. J. 2005. Impact of vegetation on flow routing and sedimentation patterns: Three-dimensional modeling for a tidal marsh. Journal of Geophysical Research: Earth Surface, 110: 118.Google Scholar
Temmerman, S., Bouma, T. J., Van de Koppel, J., Van der Wal, D., De Vries, M. B., and Herman, P. M. J. 2007. Vegetation causes channel erosion in a tidal landscape. Geology, 35: 631634.Google Scholar
Temmerman, S., Meire, P., Bouma, T. J., Herman, P. M. J., Ysebaert, T., and De Vriend, H. J. 2013. Ecosystem-based coastal defence in the face of global change. Nature, 504: 7983.Google Scholar
Temmerman, S., De Vries, M. B., and Bouma, T. J. 2012. Coastal marsh die-off and reduced attenuation of coastal floods: A model analysis. Global and Planetary Change, 9293: 267274.Google Scholar
Tonelli, M., Fagherazzi, S., and Petti, M. 2010. Modeling wave impact on salt marsh boundaries. Journal of Geophysical Research: Oceans, 115: 117.Google Scholar
Torres, R., and Styles, R. 2007. Effects of topographic structure on salt marsh currents. Journal of Geophysical Research: Earth Surface, 112: F02023.Google Scholar
Townend, I. H. 2010. An exploration of equilibrium in Venice Lagoon using an idealised form model. Continental Shelf Research, 30: 984999.Google Scholar
Tucker, G. E., Catani, F., Rinaldo, A., and Bras, R. L. 2001. Statistical analysis of drainage density from digital terrain data. Geomorphology, 36: 187202.Google Scholar
Valentine, K., Mariotti, G., and Fagherazzi, S. 2014. Repeated erosion of cohesive sediments with biofilms. Advances in Geosciences, 39: 914.Google Scholar
Vandenbruwaene, W., Bouma, T. J., Meire, P., and Temmerman, S. 2013. Bio-geomorphic effects on tidal channel evolution: Impact of vegetation establishment and tidal prism change. Earth Surface Processes and Landforms, 38: 122132.Google Scholar
Vandenbruwaene, W., Temmerman, S., Bouma, T. J., Klaassen, P. C., de Vries, M. B., Callaghan, D. P., van Steeg, P. et al. 2011. Flow interaction with dynamic vegetation patches: Implications for biogeomorphic evolution of a tidal landscape. Journal of Geophysical Research: Earth Surface, 116: 113.Google Scholar
Wamsley, T. V., Cialone, M. A., Smith, J. M., Atkinson, J. H., and Rosati, J. D. 2010. The potential of wetlands in reducing storm surge. Ocean Engineering, 37: 5968.Google Scholar
van der Wegen, M., Wang, Z. B., Savenije, H. H. G., and Roelvink, J. A. 2008. Long-term morphodynamic evoluation and energy dissipation in a coastal plain, tidal embayment. Journal of Geophysical Research: Earth Surface, 113: 122.Google Scholar
van Wesenbeeck, B. K., van De, K. oppel, J., Herman, P. M. J., and Bouma, T. J. 2008. Does scale-dependent feedback explain spatial complexity in salt-marsh ecosystems? Oikos, 117: 152159.Google Scholar
White, B. L., and Nepf, H. M. 2007. Shear instability and coherent structures in shallow flow adjacent to a porous layer. Journal of Fluid Mechanics, 593: 132.Google Scholar
White, B. L., and Nepf, H. M. 2008. A vortex-based model of velocity and shear stress in a partially vegetated shallow channel. Water Resources Research, 44: 115.Google Scholar
Yang, S. L. 1998. The role of scirpus marsh in attenuation of hydrodynamics and retention of fine sediment in the Yangtze estuary. Estuarine, Coastal and Shelf Science, 47: 227233.Google Scholar

References

Ainouche, M. L., Baumel, A., Salmon, A., and Yannic, G. 2003. Hybridization, polyploidy and speciation in Spartina (Poaceae). New Phytologist, 161: 165172.Google 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.Google 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.Google 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.Google 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.Google 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.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.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.Google Scholar
Angelini, C., and Silliman, B. R. 2012. Patch size-dependent community recovery after massive disturbance. Ecology, 93: 101110.Google Scholar
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.Google 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.Google Scholar
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.Google Scholar
Bertness, M. D., and Ellison, A. M. 1987. Determinants of pattern in a New England salt marsh plant community. Ecological Monographs, 57: 129147.Google 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.Google 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.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.Google 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.Google 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.Google 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.Google 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.Google 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.Google Scholar
Cebrian, J. 1999. Patterns in the fate of production in plant communities. American Naturalist, 154: 449468.Google Scholar
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.Google 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.Google 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.Google Scholar
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.Google 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.Google 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.Google 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.Google 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.Google 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.Google Scholar
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.Google 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.Google 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.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
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.Google Scholar
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.Google 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.Google 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

References

Alexandre, A., Meunier, J. D., Colin, F., and Koud, J. M. 1997. Plant impact on the biogeochemical cycle of silicon and related weathering processesGeochimica et Cosmochimica Acta61: 677682.Google Scholar
Algar, C., and Vallino, J. 2014. Predicting microbial nitrate reduction pathways in coastal sediments. Aquatic Microbial Ecology, 71: 223238. Doi: 10.3354/ame01678Google Scholar
Anderson, I., Tibias, C., Neikirk, B., and Wetzel, R. 1997. Development of a process-based N mass balance model for a Virginia Spartina alterniflora salt marsh: implication for net DIN flux. Marine Ecology Progress Series, 159: 1327.Google Scholar
Anderson, K. A., and Downing, J. A. 2006. Dry and wet atmospheric deposition of nitrogen, phosphorus and silicon in an agricultural regionWater, Air, & Soil Pollution176: 351374.Google Scholar
Argow, B., Hughes, Z., and FitzGerald, D. 2011. Ice raft formation, sediment load, and theoretical potential for ice-rafted sediment influx on northern coastal wetlands. Continental Shelf Research, 31: 12941305.Google Scholar
Armbrust, E. V. 2009. The life of diatoms in the world’s oceansNature459: 185192.Google Scholar
Armstrong, W. 1978. Root aeration in the wetland condition. In: Hook, D. D. and Crawforth, E. M. M., eds., Plant Life in Anaerobic Environments. Ann Arbor Science Publishers, pp. 269298.Google Scholar
Baker, A. R., French, M., and Linge, K. L. 2006. Trends in aerosol nutrient solubility along a west–east transect of the Saharan dust plumeGeophysical Research Letters33: 7.Google Scholar
Bettez, N., Duncan, J., Groffman, P., Band, L., O’Neil-Dynne, , Haushal, J. S., Belt, K., and Law, N. 2015. Climate variation overwhelms efforts to reduce N delivery to coastal waters. Ecosystems, 18: 13191331.Google Scholar
Bettez, N. and Groffman, P. 2013. N deposition in and near an urban ecosystem. Environmental Science and Technology, 47: 60476051.Google Scholar
Bettez, N., Marino, R., Howarth, R., and Davidson, E. 2013. Roads as N deposition hot spots. Biogeochemistry, 114: 149163.Google Scholar
Bluth, G. J. and Kump, L. R. 1994. Lithologic and climatologic controls of river chemistryGeochimica et Cosmochimica Acta58: 23412359.Google Scholar
Bollinger, M. S., and Moore, W. 1993. Evaluation of salt marsh hydrology using radium as a tracer. Geochimica et Cosmochimica Acta, 57: 22032212.Google Scholar
Bouwman, A., Van Drecht, G., Knoop, J., Beusen, A., and Meinardi, C. 2005. Exploring changes in river nitrogen export to the world’s oceans. Global Biogeochemical Cycles, 19: 114.Google Scholar
Boynton, W. R., Hagy, J. D., Cornwell, J. C., Kemp, W. M., Green, S. M., Owens, M. S., Baker, J. E., and Larsen, R. K. 2008. Nutrient budgets and management actions in the Patuxent River Estuary, Maryland. Estuaries and Coasts, 31: 623651.Google Scholar
Boynton, W. R., and Nixon, S. W. 2013. Budget analysis of estuarine ecosystems. In: Day, J. W., Crump, B. C., Kemp, W. M., and Yanez-Arancibia, A., eds., Estuarine Ecology. 2nd edn. John Wiley and Sons, Hoboken, NJ, pp. 443464.Google Scholar
Bricker, S. B., Longstaff, B., Dennison, W., Jones, A., Boicourt, K., Wicks, C. and Woerner, J. 2008. Effects of nutrient enrichment in the nation’s estuaries: a decade of change. Harmful Algae, 8: 2132.Google Scholar
Buchsbaum, R. N., Deegan, L. A., Horowitz, J., Garritt, R. H., Giblin, A. E., Ludlam, J. P., and Shull, D. H. 2008. Effects of regular salt marsh haying on marsh plants, algae, invertebrates and birds at Plum Island Sound, Massachusetts. Wetlands Ecological Management, 17: 469487.Google Scholar
Buresh, R. J., DeLaune, R., and Patrick, W. H. Jr 1980. Nitrogen and phosphorus distribution and utilization by Spartina alterniflora in a Louisiana Gulf Marsh. Estuaries, 3: 111121.Google Scholar
Caraco, N. F., Cole, J. J., and Likens, G. E. 1990. A comparison of phosphorus immobilization in sediments of freshwater and coastal marine systems. Biogeochemistry, 9: 227290.Google Scholar
Carey, J. C., and Fulweiler, R. W., 2012a. Human activities directly alter watershed dissolved silica fluxesBiogeochemistry111: 125138.Google Scholar
Carey, J. C., and Fulweiler, R. W. 2012b. The terrestrial silica pumpPLOS ONE7, p.e52932.Google Scholar
Carey, J. C., and Fulweiler, R. W. 2013a. Nitrogen enrichment increases net silica accumulation in a temperate salt marshLimnology and Oceanography58: 99111.Google Scholar
Carey, J. C., and Fulweiler, R. W. 2013b. Watershed land use alters riverine silica cyclingBiogeochemistry113: 525544.Google Scholar
Carey, J. C., and Fulweiler, R. W. 2014a. Silica uptake by Spartina – evidence of multiple modes of accumulation from salt marshes around the world. Frontiers in Plant Science, 5: 186.Google Scholar
Carey, J. C., and Fulweiler, R. W. 2014b. Salt marsh tidal exchange increases residence time of silica in estuariesLimnology and Oceanography59: 12031212.Google Scholar
Carey, J. C., Moran, S. B. Kelly, R. P., Kolker, A. S., and Fulweiler, R. W. 2015. The declining role of organic matter in New England salt marshes. Estuaries and Coasts, 40 : 626639.Google Scholar
Castañeda-Moya, E., Twilley, R. R.Rivera-Monroy, V. H.Zhang, K.Davis, S. E. III, and Ross, M.  2010. Sediment and nutrient deposition associated with hurricane Wilma in mangroves of the Florida Coastal Everglades. Estuaries and Coasts, 33: 4558.Google Scholar
Cavatorta, J., Johnston, M., Hopkinson, C., and Valentine, V. 2003. Patterns of sedimentation in a salt marsh – dominated estuary. Biological Bulletin, 205: 239241.Google Scholar
Chalmers, A., Wiegert, R., and Wolff, P. 1985. Carbon balance in a salt marsh: interactions of diffusive export, tidal deposition and rainfall-caused erosion. Estuarine Coastal and Shelf Science, 21: 757771.Google Scholar
Chambers, M., Harvey, J. W., and Odum, W. E. 1992. Ammonium and phosphate dynamics in a Virginia salt marsh. Estuaries, 15: 349359.Google Scholar
Charette, M. A. 2007. Hydrologic forcing of submarine groundwater discharge: insight from a seasonal study of radium isotopes in a groundwater-dominated salt marsh estuary. Limnology and Oceanography, 52: 230239.Google Scholar
Charette, M. A., and Sholkovitz, E. R. 2002. Oxidative precipitation of groundwater‐derived ferrous iron in the subterranean estuary of a coastal bay. Geophysical Research Letters, 29: 851854.Google Scholar
Charette, M. A., Splivallo, R., Herbold, C., Bollinger, M. A., and Moore, W. S. 2003. Salt marsh submarine groundwater discharge as traced by radium isotopes. Marine Chemistry, 84: 113121.Google Scholar
Chen, S., Torres, R. Bizimis, M., and Wirth, E. 2012. Salt marsh sediment and metal fluxes in response to rainfall. Limnology and Oceanography: Fluids and Environments, 2: 5466.Google Scholar
Chen, Y. C., and Windom, H. L. 1997. Sediment manganese and biogenic silica as geochemical indicators in estuarine salt marshes of coastal Georgia, USAEnvironmental Geochemistry and Health19: 0. https://doi.org/10.1023/A:1018434018126.Google Scholar
Childers, D. L., Boyer, J. N., Davis, S. E., Madden, C. J. Rudnick, D. T., and Skalar, F. H. 2006. Relating precipitation and water management to nutrient concentrations in the oligotrophic “upside‐down” estuaries of the Florida Everglades. Limnology and Oceanography, 51: 602–16.Google Scholar
Childers, D. L., Davis, S., Twilley, R., and Rivera-Monroy, V. 1999. Wetland-water column interactions and the biogeochemistry of estuary-watershed coupling around the Gulf of Mexico. In: Bianchi, T., Pennock, J., Twilley, R. eds. Biogeochemistry of Gulf of Mexico Estuaries. Wiley, Hoboken, NJ, pp. 211235.Google Scholar
Childers, D. L., Day, J. W. Jr., and McKellar, H. N. Jr 2000. Twenty more years of marsh and estuarine flux studies: Revisiting Nixon (1980). In: Weinstein, M. P, and Kreeger, D. A., eds., Concepts and Controversies in Tidal Marsh Ecology. pp. 391–423.Google Scholar
Chmura, G. L., Kellman, L., van Ardenne, L., and Guntenspergen, G. R. 2016. Greenhouse gas fluxes from salt marshes exposed to chronic nutrient enrichment. PLOS ONE, 11: e0149937. https: //doi.org/10.1371/journal.pone.0149937Google Scholar
Coelho, J. P., Flindt, M. R., Jensen, H. S., Lillebo, A. I., and Pardal, M. A. 2004. Phosphorus speciation and availability in intertidal sediments of a temperate estuary: relationship to eutrophication and annual P fluxes. Estuarine, Coastal and Shelf Science, 61: 583590.Google Scholar
Colman, J. A. and Masterson, J. P. 2008. Transient simulations of nitrogen load for a coastal aquifer and embayment, Cape Cod, MA. Environmental Science and Technology, 42: 207213.Google Scholar
Colman, S. M. and Bratton, J. F. 2003. Anthropogenically induced changes in sediment and biogenic silica fluxes in Chesapeake BayGeology31: 7174.Google Scholar
Compton, J., Mallinson, D., Glenn, C. R., Fillippelli, G., Follmi, K., Shields, G., and Zanin, Y. 2000. Variations in the global phosphorus cycle. In: Glenn, C. R., Prevot-Lucas, L., and Lucas, J., eds., Marine Authigenesis: from Global to Microbial, Tulsa, SEPM (Society for Sedimentary Geology), pp. 2133.Google Scholar
Conley, D. J. 2002. Terrestrial ecosystems and the global biogeochemical silica cycleGlobal Biogeochemical Cycles16(4). DOI: 10.1029/2002GB001894Google Scholar
Cornelis, J. T., Delvaux, B., Georg, R. B., Lucas, Y., Ranger, J., and Opfergelt, S. 2011. Tracing the origin of dissolved silicon transferred from various soil-plant systems towards rivers: a reviewBiogeosciences8: 89112.Google Scholar
Cornu, S., Lucas, Y., Ambrosi, J. P., and Desjardins, T. 1998. Transfer of dissolved Al, Fe and Si in two Amazonian forest environments in BrazilEuropean Journal of Soil Science49: 377384.Google 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
Craft, C., Broome, S. W., and Seneca, E. D. 1988. Nitrogen, phosphorus, and organic carbon pools in a natural and transplanted marsh. Estuaries, 11: 272280.Google Scholar
Craft, C., Seneca, D., and Broome, S. W. 1993. Vertical accretion in microtidal regularly and irregularly flooded estuarine marshes. Estuarine and Coastal Shelf Science, 37: 371386.Google Scholar
Currin, C. A., Joye, S. B., and Paerl, H. W. 1996. Diel rates of N2-fixation and denitrification in a transplanted Spartina alterniflora marsh: implications for N-flux dynamics. Estuarine, Coastal, and Shelf Science, 42: 597616.Google Scholar
Currin, C. A., and Paerl, H. W. 1998a. Environmental and physiological controls on diel patterns of N2 fixation in epiphytic cyanobacterial communities. Microbial Ecology, 35: 3435.Google Scholar
Currin, C. A., and Paerl, H. W. 1998b. Epiphytic nitrogen fixation associated with standing dead shoots of smooth cordgrass, Spartina alterniflora. Estuaries, 21: 108117.Google Scholar
Cutter, G. A. and Velinsky, D. J. 1988. Temporal variations in sedimentary sulfur in a Delaware salt marsh. Marine Chemistry, 23: 311327.Google Scholar
Dacey, J. W. H. and Howes, B. L. 1984. Water uptake by roots controls water table movement and sediment oxidation in short Spartina marsh. Science, 224: 487489.Google Scholar
Dalsgaard, T., Thamdrup, B., and Canfield, D. 2005. Anaerobic ammonium oxidation (anammox) in the marine environemtn. Research in Microbiology, 156: 457464.Google Scholar
Dame, R. F., Spurrier, J. D., Williams, T. M., Kjerfve, B., Zingmark, R. G., Wolaver, T. G., Chrzanowski, T. H., McKellar, H. N., and Vernberg, F. J. 1991. Annual material processing by a salt marsh-estuarine basin in South Carolina, USA. Marine Ecology Progress Series, 72: 153166.Google Scholar
Dankers, N., Birsbergen, M., Zegers, K., Laane, R., and van der Loeff, M. R. 1984. Transportation of water, particulate and dissolved organic and inorganic matter between a salt marsh and the Ems-Dollard estuary, the Netherlands. Estuarine and Coastal Shelf Science, 19: 143165.Google Scholar
Darby, F. A., and Turner, R. E. 2008a. Below- and aboveground biomass of Spartina alterniflora: Response to nutrient addition in a Louisiana salt marsh. Estuaries and Coasts, 31: 326334.Google Scholar
Darby, F. A., and Turner, R. E. 2008b. Below- and aboveground Spartina alterniflora production in a Louisiana salt  marsh. Estuaries and Coasts, 31: 223231.Google Scholar
Davidson, E., Savage, K., Bettez, N., Marino, R., and Howarth, R. 2009. N in runoff from residential roads in a coastal area. Water, Air, and Soil Pollution, 210: 313.Google Scholar
De Bakker, N. V. J., Hemminga, M. A., and Van Soelen, J. 1999. The relationship between silicon availability, and growth and silicon concentration of the salt marsh halophyte Spartina anglicaPlant and Soil215: 1927.Google Scholar
DeLaune, R. D., Reddy, C., and Patrick, W. 1981. Accumulation of plant nutrients and heavy metals through sedimentation processes and accretion in a Louisana salt marsh. Estuaries, 4: 328334.Google Scholar
Delaune, R., and Jugsujinda, A. 2003. Denitrification potential in a Louisiana wetland receiving diverted Mississippi River water. Chemistry and Ecology, 19: 411418.Google Scholar
Desplanques, V., Cary, L., Mouret, J. C., Trolard, F., Bourrié, G., Grauby, O., and Meunier, J. D. 2006. Silicon transfers in a rice field in Camargue (France)Journal of Geochemical Exploration88: 190193.Google Scholar
Drake, D., Peterson, B. J., Galván, K. A, Deegan, L. A., Hopkinson, C., Johnson, J. M., Koop-Jakobsen, K., Lemay, L. E., and Picard, C. 2009. Salt marsh ecosystem biogeochemical responses to nutrient enrichment: a paired 15N tracer study. Ecology, 90: 2535–46.Google Scholar
Drees, L. R., Wilding, L. P., Smeck, N. E., and Senkayi, A. L. 1989. Silica in soils: quartz and disorders polymorphs, in: Minerals in Soil Environments, Dixon, J. B., and Weed, S. B., eds., Soil Science Society of America, Madison, pp. 914974.Google Scholar
Eisenreich, S. J., Emmling, P. J., and Beeton, A. M. 1977. Atmospheric loading of phosphorus and other chemicals to Lake MichiganJournal of Great Lakes Research3: 291304.Google Scholar
Epstein, E., 1994. The anomaly of silicon in plant biologyProceedings of the National Academy of Sciences91: 1117.Google Scholar
Epstein, E., 1999. SiliconAnnual Review of Plant Biology50: 641664.Google Scholar
Fagherazzi, S., Kirwan, M., Mudd, S., Guntenspergen, G., Temmerman, S., D’Alapaos, A., van de Koppel, , et al. 2012. Numerical models of salt marsh evolution: ecological, geomorphic and climatic factors. Reviews of Geophysics, Doi: 10.1029/2011RG000359.Google Scholar
Fagherazzi, S., Mariotti, G., Wiberg, P., and McGlathery, K. 2013. Marsh collapse does not require sea level rise. Oceanography, 26: 7077.Google Scholar
Forbrich, I., Giblin, A. E., and Hopkinson, C. S. 2018. Constraining marsh carbon budgets using long‐term C burial and contemporary atmospheric CO2 fluxes. Journal of Geophysical Research, 123: 867878.Google Scholar
Fowler, D., Coyle, M., Skiba, U., Sutton, M. A., Cape, J. N., Reis, S., Sheppard, L. J., et al. 2013. The global nitrogen cycle in the twenty-first century. Philosophical Transactions of the Royal Society, Series B. 368: 20130164. Doi: 10.1089/rstb.2013.0164.Google Scholar
Fowler, D., Smith, R., Muller, J., Cape, J., Sutton, M., Erishman, J., and Fagerli, H. 2007. Long term trends in sulphur and nitrogen deposition in Europe and the cause of non-linearities. Water Air Soil Pollut: Focus, 7: 4147.Google Scholar
Fulweiler, R. W. and Nixon, S. W. 2005. Terrestrial vegetation and the seasonal cycle of dissolved silica in a southern New England coastal riverBiogeochemistry74: 115130.Google Scholar
Gaillardet, J., Dupré, B., Louvat, P., and Allegre, C. J. 1999. Global silicate weathering and CO2 consumption rates deduced from the chemistry of large riversChemical Geology159: 330.Google Scholar
Gallagher, J. L., Reimhold, R. J., Linthurst, R. A., and Pfeiffer, W. T. 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
Galloway, J. N., Townsend, A. R., Erisman, J. W., Bekunda, M., Cai, Z., Freney, J. R., Martinelli, L. A., Seitzinger, S. P., and Sutton, M. A. 2008. Transformation of the nitrogen cycle: Recent trends, questions, and potential solutions. Science, 320: 889892.Google Scholar
Ganju, N., Defne, Z., Kirwan, M., Fagherazzi, S., D’Alpaos, A., and Carniello, L. 2017. Spatially integrative metrics reveal hidden vulnerability of microtidal salt marshes. Nature Communications, 8: 14156, doi: 10.1038/ncomms14156.Google Scholar
Gardener, L. R. 1990. Simulation of the diagenesis of carbon, sulfur and dissolved oxygen in salt marsh sediments. Ecological Monographs, 60: 91111.Google Scholar
Gardner, L. R., and Reeves, H. W., 2002. Spatial patterns in soil water fluxes along a forest-marsh transect in the southeastern United StatesAquatic Sciences-Research Across Boundaries64: 141155.Google Scholar
Gardner, W., and Gaines, E. 2008. Estimating pore water drainage from marsh soils using rainfall and well records. Estuarine Coastal and Shelf Science, 79: 5158.Google Scholar
Giblin, A. E. 1988. Pyrite formation during early diagenesis. Geomicrobiology Journal, 6: 7797.Google Scholar
Giblin, A. E., and Howarth, R. W. 1984. Porewater evidence for a dynamic sedimentary iron cycle in salt marshes. Limnology and Oceanography, 29: 4763.Google Scholar
Graham, W. F., and Duce, R. A. 1979. Atmospheric pathways of the phosphorus cycle. Geochimica et Cosmochimica Acta, 43: 11951208.Google Scholar
Gribsholt, B., Boschker, H. T. S., Struyf, E., Andersson, M., Tramper, A., De Brabandere, L., van Damme, S., Brion, N., et al. 2005. N processing in a tidal freshwater marsh: a whole ecosystem 15N labeling study. Limnology and Oceanography, 50: 19451959.Google Scholar
Hansel, C. M., Lentini, C. L., Tang, Y., Johnston, D. T., Wankel, S. D., and Jardine, P. M. 2015. Dominance of sulfur-fueled iron oxide reduction in low-sulfate freshwater sediments. The ISME Journal, 9: 24002412.Google Scholar
Hartzell, J. L. and Jordan, T. E. 2012. Shifts in the relative availability of phosphorus and nitrogen along estuarine salinity gradients. Biogeochemistry, 107: 489500.Google Scholar
Hartzell, J. L., Jordan, T. E., and Cornwell, J. C. 2010. Phosphorus burial in sediments along the salinity gradient of the Patuxent River, a sub estuary of the Chesapeake Bay (USA). Estuaries and Coasts, 33: 92106.Google Scholar
Hartzell, J. L., Jordan, T. E., and Cornwell, J. C. 2017. Phosphorus sequestration in sediments along the salinity gradients of Chesapeake Bay sub-estuaries. Estuaries and Coasts, 40: 16071625.Google Scholar
Harrison, K. G., 2000. Role of increased marine silica input on paleo‐pCO2 levelsPaleoceanography15: 292298.Google Scholar
Harvey, J. W., Germann, P. F., and Odum, W. E. 1987. Geomorphological control of subsurface hydrology in the creekbank zone of tidal marshesEstuarine, Coastal and Shelf Science25: 677691.Google Scholar
HELCOM 2017. Atmospheric deposition of heavy metals on the Baltic Sea. HELCOM Baltic Sea Environment Fact Sheets. Online 27.06.2018. www.helcom.fi/baltic-sea-trends/environment-fact-sheets/.Google Scholar
Hodson, M. J., White, P. J., Mead, A., and Broadley, M. R. 2005. Phylogenetic variation in the silicon composition of plantsAnnals of Botany96: 10271046.Google Scholar
Hopkinson, C., and Giblin, A.. 2008. Salt marsh N cycling. In: Capone, R., Bronk, D., Mulholland, M., and Carpenter, E., eds., Nitrogen in the Marine Environment, 2nd edn. Elsevier, Amsterdam, pp. 9911036.Google Scholar
Hopkinson, C. S. 1992. The effects of system coupling on patterns of wetland ecosystem development. Estuaries, 15: 549562.Google Scholar
Hopkinson, C. S., Cai, W.-J., and Hu, X. 2012. Carbon sequestration in wetland dominated coastal systems – a global sink of rapidly diminishing magnitude. Current Opinions in Environmental Sustainability, 4: 19.Google Scholar
Hopkinson, C. S. and Vallino, J. 1995. The nature of watershed perturbations and their influence on estuarine metabolism. Estuaries, 18: 598621.Google Scholar
Hou, L., Liu, M., Yang, Y., Ou, D., Lin, X., and Chen, H., 2010. Biogenic silica in intertidal marsh plants and associated sediments of the Yangtze EstuaryJournal of Environmental Sciences22: 374380.Google Scholar
Howarth, R. W., Anderson, D., Church, T., Greening, H., Hopkinson, C., Huber, W. Marcus, N. et al. Committee on the Causes and Management of Coastal Eutrophication. 2000. Clean Coastal Waters – Understanding and reducing the effects of nutrient pollution. Ocean Studies Board and Water Science and Technology Board, Commission on Geosciences, Environment, and Resources, National Research Council. National Academy of Sciences, Washington, DC.Google Scholar
Howarth, R. W., Anderson, D., Cloern, J., Elfring, C., Hopkinson, C., Lapointe, B. Malone, T. et al. 2000. Nutrient pollution of coastal rivers, bays and seas. Issues in Ecology, 7: 115.Google Scholar
Howarth, R. W. and Hobbie, J. E. 1982. The regulation of decomposition and heterotophic microbial activity in salt marsh soils: a review. In: Estuarine Comparisons, pp. 183207. Kennedy, V. S., ed., Academic Press, Orlando, FL, pp. 183–220.Google Scholar
Howarth, R. W., Jensen, H. S., Marino, R., and Postma, H. 1995. Transport to and processing of P in near-shore and oceanic waters. In: Tiessen, H., ed., Phosphorus and the Global Environment. John Wiley and Sons Ltd, Hoboken, NJ, pp. 17.Google Scholar
Howarth, R. W., Swaney, D., Billen, G., Carnier, J., Hong, B., Humborg, C., Johnes, P., Morth, C., and Marino, R. 2012. Nitrogen fluxes from the landscape are controlled by net anthropogenic nitrogen inputs and by climate. Frontiers in Ecology and Environment, 10: 3743.Google Scholar
Howarth, R. W. and Teal, J. 1979. Sulfate reduction in a New England salt marsh. Limnology and Oceanography, 24: 9991013.Google Scholar
Howes, B. L., Howarth, R. W., Teal, J. M., and Valiela, I. 1981. Oxidation-reduction potentials in a salt marsh: Spatial patterns and interactions with primary production. Limnology and Oceanography, 26: 350360.Google Scholar
Hyfield, E., Day, J., Cable, J., and Justic, D. 2008. The impacts of re-introducing Mississippi River water on the hydrologic budget and nutrient inpus of a deltaic estuary. Ecological Engineering, 32: 347359.Google Scholar
Iler, R. J. K. 1979. The Chemistry of Silica: Solubility, Polymerization, Colloid and Surface Properties, and Biochemistry. Wiley, New York, N.Y.Google Scholar
Jacobs, S., Struyf, E., Maris, T., and Meire, P., 2008. Spatiotemporal aspects of silica buffering in restored tidal marshesEstuarine, Coastal and Shelf Science80: 4252.Google Scholar
Johnson, T. C. and Eisenreich, S. J., 1979. Silica in Lake Superior: mass balance considerations and a model for dynamic response to eutrophicationGeochimica et Cosmochimica Acta43: 7792.Google Scholar
Jordan, T. E., Cornwell, J. C., Boynton, W. R., and Anderson, J. T. 2008. Changes in the phosphorus biogeochemistry along and estuarine salinity gradient: The iron conveyer belt. Limnology and Oceanography, 53: 172184.Google Scholar
Joye, S. B. and Paerl, H. W. 1994. Nitrogen cycling in microbial mats: rates and patterns of denitrification and nitrogen fixation. Marine Biology, 119: 285295Google Scholar
Kaplan, W., Valiela, I., and Teal, J. M. 1979. Denitrification in a Massachusetts salt marsh ecosystem. Limnology and Oceanography, 26: 350360.Google Scholar
Karstens, S., Buczko, U., and Glatzel, S. 2015. Phosphorus storage and mobilization in coastal Phragmites wetlands: Influence of local-scale hydrodynamics. Estuarine, Coastal and Shelf Science, 164: 124133.Google Scholar
Kinney, E. L. and Valiela, I. 2013. Changes in the δ15N in salt marsh sediments in a long-term fertilization study. Marine Ecology Progress Series, 477: 4152.Google Scholar
Koch, M. Maltby, E., Oliver, G., and Bakker, S. 1992. Factors controlling denitrification rates in tidal mudflats and fringing salt marshes in South-west England. Estuarine, Coastal and Shelf Science, 34: 471485.Google Scholar
Koop-Jacobsen, K., and Giblin, A. 2009. Anammox in tidal marsh sediments: the role of salinity, nitrogen loading, and marsh vegetation. Estuaries and Coasts, 32: 238245.Google Scholar
Koop-Jakobsen, K., and Giblin, A. E. 2010. The effect of increased nitrate loading on nitrate reduction via denitrification and DNRA in salt marsh sediments Limnol. Oceanography, 55: 789802.Google Scholar
Koretsky, C. M., VanCappellen, P., DiChistina, T. J., Kostka, J. E., Lowe, K. L., Moore, C. M. Roychoudhury, A. N., and Viollier, E. 2005. Salt marsh pore water chemistry does not correlate with microbial community structure. Estuarine Coastal and Shelf Science, 62: 233251.Google Scholar
Kostka, J. E., and Luther, G. W. III 1995. Seasonal cycling of Fe in salt marsh sediments. Biogeochemistry, 29: 159181.Google Scholar
Lane, R. R., Day, J. W., Justic, D., Reyes, E., Marx, B., Day, J. N., and Hyfield, E. 2004. Changes in stoichiometric Si, N and P ratios of Mississippi River water diverted through coastal wetlands to the Gulf of MexicoEstuarine, Coastal and Shelf Science60: 110.Google Scholar
Lanning, F. C., and Eleuterius, L. N. 1981. Silica and ash in several marsh plantsGulf and Caribbean Research7: 4752.Google Scholar
Lanning, F. C., and Eleuterius, L. N. 1983. Silica and ash in tissues of some coastal plantsAnnals of Botany51: 835850.Google Scholar
Lanning, F. C., and Eleuterius, L. N. 1985. Silica and ash in tissues of some plants growing in the coastal area of Mississippi, USAAnnals of Botany56: 157172.Google Scholar
Lanning, F. C. and Eleuterius, L. N. 1989. Silica deposition in some C3 and C4 species of grasses, sedges and composites in the USA. Annals of Botany, 64: 395410.Google Scholar
LeMay, L. 2007. The impact of drainage ditches on salt marsh flow patterns, sedimentation and morphology: Rowley River, Massachusetts. MS thesis. The College of William and Mary, Williamsburg, Virginia, USA.Google Scholar
Lettrich, M. 2011. Nitrogen advection and denitrification loss in southeastern North Carolina salt marshes. MS thesis, University of North Carolina Wilmington.Google Scholar
Lillebø, A. I., Coelho, J. P., Flindt, M. R., Jensen, H. S., Marques, J. C., Pedersen, J. B., and Pardal, M. A. 2007. Spartina maritima influence on the dynamics of the phosphorous sedimentary cycle in a warm temperate estuary (Mondego Estuary, Portugal). Hydrobiologia, 587: 195204.Google Scholar
Lillebø, A. I., Neto, J. M., Flindt, M. R., Jensen, H. S., Marques, J. C., and Pardal, M. A. 2004. Phosphorous dynamics in a temperate intertidal estuary. Estuarine, Coastal and Shelf Science, 61: 101109.Google Scholar
Lloret, J., and Valiela, I. 2016. Unprecedented decrease in deposition of nitrogen oxides over North America: the relative effects of emission controls and prevailing air-mass trajectories. Biogeochemistry, 129: 165180.Google Scholar
Loomis, M. J., and Craft, C. B. 2010. Carbon sequestration and nutrient (nitrogen, phosphorus) accumulation in river-dominated tidal marshes, Georgia, USA. Soil Society of America Journal, 74: 10281036.Google Scholar
Maavara, T., Parsons, C. T., Ridenour, C., Stojanovic, S., Durr, H. H., Powley, H. R. and Van Cappellen, P. 2015. Global phosphorus retention by river damming. Proceedings of the National Academy of Sciences, 122: 1560315608.Google Scholar
Maguire, T. J. 2017. Anthropogenic perturbations to the biogeochemical cycle of silicon. PhD dissertation, Boston University, Boston, MA.Google Scholar
Mahowald, N., Jickells, T. D., Baker, A. R., Artaxo, P., Benitex-Nelson, C. R., Bergametti, G., Bond, T. C., et al. 2008. Global distribution of atmospheric phosphorus sources, concentrations and deposition rates and anthropogenic impacts. Global Biogeochemical Cycles, 22: 119.Google Scholar
Mariotti, G., and Fagherazzi, S., 2013. Critical width of tidal flats triggers marsh collapse in the absence of sea-level riseProceedings of the National Academy of Sciences110: 53535356.Google Scholar
Martin, R. M. and Moseman-Valtierra, S. 2015. Greenhouse gas fluxes vary between Phragmities alstralis and native vegetation zones in coastal wetlands along a salinity gradient. Wetlands, 35: 10211031.Google Scholar
Martin, R. M., Wigand, C., Elmstrom, E., Lloret, J., and Valiea, I. 2018. Long-term nutrient addition increase respiration and nitrous oxide emissions in New England salt marsh. Ecology and Evolution, 8: 49584966.Google Scholar
Mateos-Naranjo, E., Andrades-Moreno, L., and Davy, A. J. 2013. Silicon alleviates deleterious effects of high salinity on the halophytic grass Spartina densifloraPlant Physiology and Biochemistry63: 115121.Google Scholar
Mateos-Naranjo, E., Gallé, A., Florez-Sarasa, I., Perdomo, J. A., Galmés, J., Ribas-Carbó, M., and Flexas, J. 2015. Assessment of the role of silicon in the Cu-tolerance of the C 4 grass Spartina densifloraJournal of Plant Physiology178: 7483.Google Scholar
McKeague, J. A., and Cline, M. G. 1963. Silica in soilsAdvances in Agronomy15: 339396.Google Scholar
McLeod, E., Chmura, G., Bouillon, S. Salm, R. Bjork, M. Duarte, C., Lovelock, C., Schlesinger, W., and Silliman, B. 2011. A blueprint for the blue carbon: toward an improved understanding of the role of vegetated coastal habitats in sequestering CO2. Frontiers Ecology Environment, 9: 552560.Google Scholar
Mendelssohn, I. A., and Morris, J. T. 2000. Eco-physiological controls on the productivity of Spartina alternilflora, Loisel. In: Concepts and Controversies in Tidal Marsh Ecology. Weinstein, M. P., and Kreeger, D. A., eds., Kluwer Academic Publishers, Dordrecht, the Netherlands, pp. 5980.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 Coastal and Marine Science, 2: 2740.Google Scholar
Merrill, J. Z. and Cornwell, J. C.. 2000. The role of oligohaline marshes in estuarine nutrient cycling. In: Concepts and Controversies in Tidal Marsh Ecology. Weinstein, M. P., and Kreeger, D. A., eds., Kluwer Academic Publishers, Dordrecht, the Netherlands, pp 425441.Google Scholar
Metson, G. S., Lin, J., Harrison, J. A., and Compton, J. E. 2017. Linking terrestrial phosphorus inputs to riverine export across the United States. Water Research, 124: 177191.Google Scholar
Meybeck, M. 2003. Global occurrence of major elements in riversTreatise on Geochemistry5: 207223.Google Scholar
Miyazako, T., Kamiya, H., Godo, T., Koyama, Y., Sato, S., Kishi, M., Fujihara, A., Tabayashi, Y., and Yamamuro, M. 2015. Long-term trends in nitrogen and phosphorus concentrations in the Hii River as influenced by atmospheric deposition from East Asia. Limnology and Oceanography, 60: 629640.Google Scholar
Moisander, P. H., Piehler, M. F., and Paerl, H. W. 2005. Diversity and activity of epiphytic nitrogen-fixers on standing dead stems of the salt marsh grass Spartina alterniflora. Aquatic Microbial Ecology, 39: 271279.Google Scholar
Morris, J. T., and Lajtha, K. 1986. Decomposition and nutrient dynamics of litter from four species of freshwater emergent macrophytes. Hydrobiologia, 131: 215223.Google Scholar
Morris, J. T., Shaffer, G. P., and Nyman, J. A. 2013. Brinson Review: Perspectives on the influence of nutrients on the sustainability of coastal wetlands. Wetlands, 33: 975988.Google Scholar
Müller, F., Struyf, E., Hartmann, J., Wanner, A., and Jensen, K. 2013. A comprehensive study of silica pools and fluxes in Wadden Sea salt marshesEstuaries and Coasts36: 11501164.Google Scholar
Newell, S. Y., Hopkinson, C. S., and Scott, L. 1992. Patterns of nitrogenase activity (acetylene reduction) associated with standing, decaying shoots of Spartina alterniflora. Estuarine, Coastal and Shelf Science, 35: 127140Google Scholar
Nixon, S. W. 1980. Between coastal marshes and coastal waters – a review of twenty years of speculation and research on the role of salt marshes in estuarine productivity and water chemistry. In: Hamilton, P. and MacDonald, K. B., eds., Estuarine Wetland Process with Emphasis on Modelling. Plenum Publishing Corporation, New York, pp. 438525.Google Scholar
Nixon, S. W. 1995. Coastal marine eutrophication: a definition, social causes, and future concerns. Ophelia, 41: 199219.Google Scholar
Nixon, S. W., Ammerman, J. W., Atkinson, L. P., Berounsky, V. M., Billen, G., Boicourt, W. C., Boynton, W. R., et al. 1996. The fate of nitrogen and phosphorus at the land-sea margin of the North Atlantic Ocean. Biogeochemistry, 35: 141180.Google Scholar
Nixon, S. W., and Oviatt, C. A. 1973. Analysis of local variation in the standing crop of Spartina alterniflora. Botanica Marine, IVI: 103–9.Google Scholar
Nixon, W., and Pilson, M. 1984. Estuarine total system metabolism and organic exchange calculated from nutrient ratios: an example from Narragansett Bay. In: The Estuary as a Filter. Kennedy, V. S., ed., Academic Press, New York, pp. 261290.Google Scholar
Noe, G. B., Childers, D. L., and Jones, R. D. 2001. Phosphorus biogeochemistry and the impact of P enrichment: Why is the everglades so unique? Ecosystems, 4: 603.Google Scholar
Norris, A. R., and Hackney, C. T. 1999. Silica content of a mesohaline tidal marsh in North CarolinaEstuarine, Coastal and Shelf Science49: 597605.Google Scholar
Nuttle, W. K. 1988. The extent of lateral water movement in the sediments of a New England salt marshWater Resources Research24: 20772085.Google Scholar
Odum, E. P. 2000. Tidal marshes as outwelling pulsing systems. In: Weinstein, M. P., and Kreeger, D. A., eds., Concepts and Controversies in Tidal Marsh Ecology. Kluwer Academic Publishers, Dordrecht, the Netherlands, pp. 37.Google Scholar
Paludan, C. and Morris, J. T. 1999. Distribution and speciation of phosphorus along a salinity gradient in intertidal marshes. Biogeochemistry, 45: 197221.Google Scholar
Packett, C. R., and Chambers, R. M. 2006. Distribution and nutrient status of haplotypes of the marsh grass Phragmites australis along the Rappahannock River in Virginia. Estuaries and Coasts, 29: 12221225.Google Scholar
Portnoy, J. W., Nowicki, B. L., Roman, C. T., and Urish, D. W. 1997. The discharge of nitrate-contaminated groundwater from a developed shoreline to a marsh-fringed estuary. Water Resources Research, 34: 30953104.Google Scholar
Querné, J., Ragueneau, O., and Poupart, N. 2012. In situ biogenic silica variations in the invasive salt marsh plant, Spartina alterniflora: a possible link with environmental stressPlant and Soil352: 157171.Google Scholar
Reddy, K. R., Kadlec, R. H., Flaig, E., and Gale, P. M. 1999. Phosphorus retention in streams and wetlands – a review. Critical Reviews in Environmental Science and Technology, 29: 86146.Google Scholar
Redfield, A. C. 1963. The influence of organisms on the composition of seawaterThe Sea2: 2677.Google Scholar
Redfield, G. W. 2002. Atmospheric deposition of phosphorus to the Everglades: Concepts, constraints, and published deposition rates for ecosystem management. The Scientific World Journal, 2: 18431873.Google Scholar
Reed, D. J. 1972. Patterns of sediment deposition in subsiding coastal salt marshes. Terrebone Bay, Louisiana: the role of winter storms. Estuaries, 12: 222227.Google Scholar
Reimold, R. G. 1972. The movement of phosphorus through the marsh cord grass, Spartina alterniflora. Loisel. Limnology and Oceanography, 17: 606611.Google Scholar
Rozema, J., Leendertse, P., Bakker, J., and van Wijnen, H.. 2000. Nitrogen and vegetation dynamics in European salt marshes. In: Weinstein, M. P., and Kreeger, D. A., eds., Concepts and Controversy in Tidal Marsh Ecology. Kluwer Academic Publishers, Dordrecht, the Netherlands, pp. 469494.Google Scholar
Ruttenberg, K. 1992. Development of a sequential extraction method for different forms of phosphorus in marine sediments Limnology and Oceanography, 37: 14601482.Google Scholar
Ruttenberg, K. 2014.The global phosphorus cycle. In: Holland, H. D., and Turekian, K. K., eds., Treatise on Geochemistry, 2nd edn, vol 10. Elsevier, Oxford, pp. 499558.Google Scholar
Sauer, D., Saccone, L., Conley, D. J., Herrmann, L., and Sommer, M., 2006. Review of methodologies for extracting plant-available and amorphous Si from soils and aquatic sedimentsBiogeochemistry80: 89108.Google Scholar
Seitzinger, S. P., Harrison, J. A., Dumont, E., Beusen, A. H. W., and Bouwman, A. F. 2005. Sources and delivery of carbon, nitrogen, and phosphorus to the coastal zone: An overview of the Global Nutrient Export from Watersheds (NEWS) models and their application. Global Biogeochemical Cycles 19: GB4S01, doi: 10.1029/2005GB002606.Google Scholar
Seitzinger, S. P., Mayorga, E., Bouwman, A., Kroeze, C., Beusen, C., Billen, , van Drecht, G., et al. 2010. Global river nutrient export: a scenario analysis of past and future trends. Global Biogeochemical Cycles, 24: GBOA08, doi: 10.1029/2009GB003587, 2010.Google Scholar
Simonson, R. W. 1995. Airborne dust and its significance to soilsGeoderma65: 143.Google Scholar
Scudlark, J. R., and Church, T. M. 1989. The sedimentary flux of nutrients at a Delaware salt marsh site: A geochemical perspective. Biogeochemistry, 7: 5575.Google Scholar
Smith, S. V. 1991. Stoichiometry of C: N: P fluxes in shallow-water marine ecosystems. In: Cole, J., Lovett, J., and Findlay, S., eds., Comparative Analyses of Ecosystems. Patterns, mechanisms, theories. New York, Springer-Verlag, pp. 259286.Google Scholar
Sommer, M., Kaczorek, D., Kuzyakov, Y., and Breuer, J. 2006. Silicon pools and fluxes in soils and landscapes – a reviewJournal of Plant Nutrition and Soil Science169: 310329.Google Scholar
Sorrell, B. Mendelssohn, I. A., McKee, K., and Woods, R. A. 2000. Ecophysiology of wetland plant roots: A modeling comparison of aeration in relation to species distribution. Annals of Botany, 86: 675685.Google Scholar
Statham, P. J. 2012. Nutrients in estuaries – An overview and the potential impacts of climate change. Science of the Total Environment, 434: 213227.Google Scholar
Staver, L. W. 2015. Ecosystem dynamics in tidal marshes constructed with fine grained, nutrient rich dredged material. PhD thesis. University of Maryland, College Park, MD.Google Scholar
Street‐Perrott, F. A., and Barker, P. A. 2008. Biogenic silica: a neglected component of the coupled global continental biogeochemical cycles of carbon and siliconEarth Surface Processes and Landforms33: 14361457.Google Scholar
Stribling, J. M. and Cornwell, J. C. 2001. Nitrogen, phosphorus and sulfur dynamics in a low salinity marsh system dominated by Spartina alterniflora. Wetlands, 21: 629638.Google Scholar
Stumm, W., and Sulzberger, B. 1992. The cycling of iron in natural environments: Considerations based on laboratory studies of heterogeneous redox processes. Geochimica et Cosmochimica Acta, 56: 32333257.Google Scholar
Struyf, E., Dausse, A., Van Damme, S., Bal, K., Gribsholt, B., Boschker, H. T., Middelburg, J. J., and Meire, P. 2006. Tidal marshes and biogenic silica recycling at the land‐sea interfaceLimnology and Oceanography51: 838846.Google Scholar
Struyf, E., Mörth, C. M., Humborg, C., and Conley, D. J., 2010. An enormous amorphous silica stock in boreal wetlandsJournal of Geophysical Research: Biogeosciences, 115(G4): 18.Google Scholar
Struyf, E., Smis, A., Van Damme, S., Garnier, J., Govers, G., Van Wesemael, B., Conley, D. J., et al. 2010. Historical land use change has lowered terrestrial silica mobilizationNature Communications1: 129.Google Scholar
Struyf, E., Van Damme, S., Gribsholt, B., Middelburg, J. J., and Meire, P. 2005. Biogenic silica in tidal freshwater marsh sediments and vegetation (Schelde estuary, Belgium)Marine Ecology Progress Series303: 5160.Google Scholar
Sullivan, M., and Daiber, F. 1974. Response in production of cord grass Spartina alterniflora, to inorganic nitrogen and phosphorus fertilizer. Chesapeake Science, 15: 121123.Google Scholar
Sundareshwar, P. V., and Morris, J. T. 1999. Phosphorus sorption characteristics of intertidal marsh sediments along an estuarine salinity gradient. Limnology and Oceanography, 44: 16931701.Google Scholar
Sundby, B., Vale, C., Caetano, M., and Luther, G. W. III 2003. Redox chemistry in the root zone of a salt marsh sediment in the Tagus Estuary, Portugal. Aquatic Geochemistry, 9: 257271.Google Scholar
Teal, J. 1962. Energy flow in a salt marsh ecosystem of Georgia. Ecology, 43: 614624.Google Scholar
Tegen, I., and Kohfeld, K. E. 2006. Atmospheric transport of silicon. In: Ittekkot, V., Unger, D., Humborg, C., and An, N. T., eds., The Silicon Cycle: Human Perturbations and Impacts on Aquatic Systems, Island Press, Washington pp. 8191.Google Scholar
Tipping, E., Benham, S., Boyle, J. F., Crow, P., Davies, J., Fischer, U., Guyatt, H., et al. 2014. Atmospheric deposition of phosphorus to land and freshwater. Environmental Science Processes and Impacts. DOI: 10.1039/c3em00641g.Google Scholar
Tobias, C., Anderson, I., Canuel, E., and Macko, S. 2001. N cycling through a fringing marsh-aquifer ecotone. Marine Ecology Progress Series, 210: 2539.Google Scholar
Tobias, C., and Neubauer, S. 2009. Salt marsh biogeochemistry – an overview. In: Perillo, G., Wolanski, E., Cahoon, D. R., and Brinson, M. M., eds., Coastal Wetlands: An Integrated Ecosystems Approach, Elsevier, Amsterdam, pp. 445492.Google Scholar
Tobias, C., and Neubauer, S. 2018. Salt marsh biogeochemistry – an overview. In: Wolanski, E., Perillo, G., Cahoon, D., and Hopkinson, C., eds., Coastal Wetlands: a synthesis. 2nd edn. Elsevier, Amsterdam, pp. 539596.Google Scholar
Torres, R., Mwamba, M., and Goni, M. 2003. Properties of intertidal marsh sediment mobilized by rainfall. Limnology and Oceanography, 48: 12451253.Google Scholar
Tréguer, P. J., and De La Rocha, C. L. 2013. The world ocean silica cycleAnnual Review of Marine Science5: 477501.Google Scholar
Treguer, P., Nelson, D. M., Van Bennekom, A. J., DeMaster, D. J., Leynaert, A., and Quéguiner, B. 1995. The silica balance in the world ocean: a reestimate. Science: 375–375.Google Scholar
Turner, R. E., Swenson, E. M., Milan, C. S., and Lee, J. M. 2007. Hurricane signals in salt marsh sediments: Inorganic sources and soil volume. Limnology and Oceanography, 52: 12311238.Google Scholar
Tyler, C., Mastronicola, T., and McGlathery, K. 2003. N fixation and N limitation of primary production along a natural marsh chronosequence. Oecologia, 136: 431438.Google Scholar
Valiela, I., and Teal, J. M. 1974. Nutrient limitation in salt marsh vegetation. In: Reimold, R. J., and Queen, W. H., eds., The Ecology of Halophytes. Academic Press, New York, pp. 547563.Google Scholar
Valiela, I., and Teal, J. 1979. The N budget of a salt marsh ecosystem. Nature, 280: 652656.Google Scholar
Valiela, I., and Cole, M. L. 2002. Comparative evidence that salt marshes and mangroves may protect seagrass meadows from land-derived nitrogen loads. Ecosystems, 5: 92102.Google Scholar
Valiela, I., Cole, M. L., McClelland, J., Hauxwell, J., Cebrian, J., and Joye, S. B. 2000. Role of salt marshes as part of coastal landscapes. In: Weinstein, M. P., and Kreeger, D. A., eds., Concepts and Controversies in Tidal Marsh Ecology.Kluwer Academic Publishers, Dordrecht, the Netherlands, pp. 2338.Google Scholar
Valigura, R., Alexander, R., Castro, M., Meyers, T., Paerl, H., Stacey, P., and Turner, R., eds. 2000. Nitrogen loading in coastal water bodies: an atmospheric perspective. Coastal and Estuarine Studies No. 57. AGU, Washington, DC.Google Scholar
Vallino, J. J., and Hopkinson, C. S. 1998. Estimation of dispersion and characteristics of mixing times in Plum Island Sound Estuary. Estuarine, Coastal and Shelf Science, 46: 333350.Google Scholar
Vallino, J. J., Hopkinson, C. S., and Garritt, R. H. 2005. Estimating estuarine gross production, community respiration and net ecosystem production: A nonlinear inverse technique. Ecological Modeling, 187: 281296.Google Scholar
VanZomeren, C. 2011. Fate of Mississippi River diverted nitrate on vegetated and non-vegetated coastal marshes of Breton Sounds Estuary. MS thesis. Louisiana State University.Google Scholar
Vieillard, A. M., Fulweiler, R. W., Hughes, Z. J., and Carey, J. C. 2011. The ebb and flood of silica: quantifying dissolved and biogenic silica fluxes from a temperate salt marshEstuarine, Coastal and Shelf Science95: 415423.Google Scholar
Wedepohl, K. H. 1995. The composition of the continental crustGeochimica et Cosmochimica Acta59: 12171232.Google Scholar
Weston, N. B., Porubsky, W. P., Samarkin, V. A., Erickson, M., Macavoy, S. E., and Joye, S. B. 2006. Porewater stoichiometry of terminal metabolic products, sulfate, and dissolved organic carbon and nitrogen in estuarine intertidal creek-bank sediments. Biogeochemistry, 77: 375408.Google Scholar
Weston, N. 2013. Declining sediments and rising seas: an unfortunate convergence for tidal wetlands. Estuaries and Coasts, 37: 123.Google Scholar
White, D., and Howes, B. 1994. Long-term 15N-nitrogen retention in the vegetated sediments of a New England salt marsh. Limnology and Oceanography, 39: 133140.Google Scholar
Whitney, D. M., Chalmers, A. G., Haines, E. B., Hanson, R. B., Pomeroy, L. R., and Sherr, B. 1981. The cycles of nitrogen and phosphorus. In: Pomeroy, L. R., and Wiegert, R. G., eds. The Ecology of a Salt Marsh. Springer Verlag. N.Y, pp. 163181.Google Scholar
Wilson, A. M., and Gardner, L. R. 2006. Tidally driven groundwater flow and solute exchange in a marsh: numerical simulationsWater Resources Research42:https://doi.org/10.1029/2005WR004302.Google Scholar
Woodwell, G. M., and Whitney, D. E. 1977. Flax pond ecosystems study: exchanges of phosphorus between a salt marsh and the coastal waters of Long Island Sound. Marine Biology, 41: 16.Google Scholar
Wolaver, T. G., and Spurrier, J. D. 1988. The exchange of phosphorus between a euhaline vegetated marsh and the adjacent tidal creek. Estuarine and Shelf Sciences, 26: 203214.Google Scholar
Wolaver, T. G., and Zieman, J. 1984. The role of tall and medium Spartina alterniflora zones in the processing of nutrients in tidal water. Estuarine, Coastal and Shelf Sciences, 19: 113.Google Scholar
Wolaver, T. G., Zieman, J. C., Wetzel, R. and Webb, K. L. 1983. Tidal exchange of nitrogen and phosphorus between a mesohaline vegetated marsh and the surrounding estuary in the lower Chesapeake Bay. Estuarine, Coastal and Shelf Science, 16: 321332.Google Scholar
Yang, Q.Tian, H.FriedrichsM. A. M., HopkinsonC. S., Lu, C., and  Najjar, R. G. 2015. Increased nitrogen export from eastern North America to the Atlantic Ocean due to climatic and anthropogenic changes during 1901–2008. Journal of Geophysical Research: Biogeosciences12010461068.Google Scholar
Yu, D., Yan, W., Chen, N., Peng, B., Hong, H., and Zhuo, G. 2015. Modeling increased riverine nitrogen export: source tracking and integrated watershed-coast management. Marine Pollution Bulletin, 101: 642652.Google Scholar
Zhang, J., Zhang, G. S., and Liu, S. M. 2005. Dissolved silicate in coastal marine rainwaters: Comparison between the Yellow Sea and the East China Sea on the impact and potential link with primary productionJournal of Geophysical Research: Atmospheres110(D16).Google Scholar
Zhou, J., Wu, Y., Kang, Q., and Zhang, J. 2007. Spatial variations in carbon, nitrogen, phosphorus and sulfur in the salt marsh sediments of the Yangtze Estuary in Chine. Estuarine, Coastal and Shelf Science, 71: 4759.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
×