Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-22dnz Total loading time: 0 Render date: 2024-04-27T02:06:45.823Z Has data issue: false hasContentIssue false

6 - The Role of Marshes in Coastal Nutrient Dynamics and Loss

from Part I - Marsh Function

Published online by Cambridge University Press:  19 June 2021

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

Summary

Sixty-five years ago, Teal’s (1962) study showed that salt marsh primary production was greater than community respiration. To explain this result, he suggested that marshes exported excess organic matter either directly as organic matter, or as organisms, to coastal waters. This concept, that marshes were “outwelling” material to the adjacent estuary and coastal oceans, was soon expanded to nutrients as well. However, the actual importance of the marsh in supplying organic matter and nutrients to adjacent coastal systems has been controversial and reviews debating the importance of outwelling from marshes have regularly appeared over the decades (Nixon 1980, Childers et al. 2000, Odum 2000, Valiela et al. 2000, Boynton and Nixon 2013). It has also been argued that in some cases the coastal ocean can act as a source of nutrients to the marsh and estuary (“inwelling”).

Type
Chapter
Information
Salt Marshes
Function, Dynamics, and Stresses
, pp. 113 - 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

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.CrossRefGoogle 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.CrossRefGoogle 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.CrossRefGoogle 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.CrossRefGoogle 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
×