Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-zzh7m Total loading time: 0 Render date: 2024-04-29T13:08:26.677Z Has data issue: false hasContentIssue false

Part III - Marsh Response to Stress

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. 335 - 481
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

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
Anisfeld, S. C., and Hill, T. D. 2011. Fertilization effects on elevation change and belowground carbon balance in a Long Island Sound tidal marsh. Estuaries and Coasts, doi:10.1007/s12237-011-9440-4.Google Scholar
Ball, A. S., and Drake, B. G. 1998. Stimulation of soil respiration by carbon dioxide enrichment of marsh vegetation. Soil Biology and Biochemistry, 30:12031205.Google Scholar
Baustian, M. M., Stagg, C. L., Perry, C. L., Moss, L. C., Carruthers, T. J. B., and Allison, M. 2017. Relationships between salinity and short-term soil carbon accumulations rates from marsh types across a landscape in the Mississippi River Delta. Wetlands, 37:313324.Google Scholar
Bowes, G. 1993. Facing the inevitable: plants and increasing atmospheric CO2. Annual Review of Plant Physiology & Plant Molecular Biology, 44:309332.Google Scholar
Broome, S. W., Mendelssohn, I. A., and McKee, K. L. 1995. Relative growth of Spartina patens (Ait.) Muhl. and Scirpus olneyi Gray occurring in a mixed stand as affected by salinity and flooding depth. Wetlands, 15(1):2030.Google Scholar
Brown, C. E., Pezeshki, S. R., and DeLaune, R. D. 2006. The effects of salinity and soil drying on nutrient uptake and growth of Spartina alterniflora in a simulated tidal system. Environmental and Experimental Botany, 58:140148.Google Scholar
Carey, J. C., Moran, S. B., Kelly, R. P., Kolker, A. S., and Fulweiler, R. W. 2017. The declining role of organic matter in New England salt marshes. Estuaries and Coasts, 40:626639.Google Scholar
Carniello, L., Defina, A. and D’Alpaos, L. 2009. Morphological evolution of the Venice lagoon: evidence from the past and trend for the future. Journal of Geophysical Research, 114:F04002.Google Scholar
Centritto, M. 2002. The effects of elevated [CO2] and water availability on growth and physiology of peach (Prunus persica) plants. Plant Biosystems, 136:177188.Google Scholar
Centritto, M., Lucas, M. E., and Jarvis, P. G. 2002. Gas exchange, biomass, whole-plant water-use efficiency and water uptake of peach (Prunus persica) seedlings in response to elevated carbon dioxide concentration and water availability. Tree Physiology, 22:699706.Google Scholar
Charles, H., and Dukes, J. S. 2009. Effects of warming and altered precipitation on plant and nutrient dynamics of a New England salt marsh. Ecological Applications, 19(7):17581773.Google Scholar
Cherry, J. A., McKee, K. L., and Grace, M. B. 2009. Elevated CO2 enhances biological contributions to elevation change in coastal wetlands by offsetting stressors associated with sea-level rise. Journal of Ecology, 97:6777.Google Scholar
Cleland, E. E., Chuine, I., Menzel, A., Mooney, H. A., and Schwartz, M. D. 2007. Shifting plant phenology in response to global change. Trends in Ecology and Evolution, 22:357365.Google Scholar
Coldren, G. A., Barreto, C. R., Wykoff, D. D., Morrissey, E. M., Langley, J. A., Feller, I. C., and Chapman, S. K. 2016. Chronic warming stimulates growth of marsh grasses more than mangroves in a coastal wetland ecotone. Ecology, 97(11):31673175.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(3):12201230.Google Scholar
Criddle, R. S., Hansen, L. D., Breidenbach, R. W. Ward, M. R., and Huffaker, R. C.. 1989. Effects of NaCl on metabolic heat evolution rates by barley roots. Plant Physiology 90(1):5358.Google Scholar
Cronk, J. K., and Fennessy, M. S.. 2001. Wetland Plants: Biology and Ecology. CRC Press, Boca Raton, Florida.Google Scholar
Crosby, S. C., Angermeyer, A., Adler, J. M., Bertness, M. D., Deegan, L. A., Sibinga, N. and Leslie, H. M. 2017. Spartina alterniflora biomass allocation and temperature: implications for salt marsh persistence with sea-level rise. Estuaries and Coasts, 40:213223.Google Scholar
Curtis, P.S., Balduman, L. M., Drake, B. G., and Whigman, D. F. 1990. Elevated atmospheric CO2 effects on belowground processes in C3 and C4 estuarine marsh communities. Ecology, 71(5):20012006.Google Scholar
Darby, F. A., and Turner, R. E. 2008. Effects of eutrophication on salt marsh root and rhizome biomass accumulation. Marine Ecology Progress Series, 363:6370.Google Scholar
Davidson, E. A., and Janssens, I. A. 2006. Temperature sensitivity of soil carbon decomposition and feedbacks to climate change. Nature, 440:165173.Google Scholar
Davis, J., Currin, C., and Morris, J. T. 2017. Impacts of fertilization and tidal inundation on elevation change in microtidal, low relief salt marshes. Estuaries and Coasts, 40:16771687.Google Scholar
Day, J. W., Christian, R. R., Boesch, D. M., Yáñez-Arancibia, A., Morris, J., Twilley, R. R., Naylor, L. et al. 2008. Consequences of climate change on the ecogeomorphology of coastal wetlands. Estuaries and Coasts, 31:477491.Google Scholar
Day, J., Kemp, P., Reed, D., Cahoon, D., Boumans, R., Suhayda, J., and Gambrell, R. 2011. Vegetation death and rapid loss of surface elevation in two contrasting Mississippi delta salt marshes: the role of sedimentation, autocompaction and sea-level rise. Ecological Engineering, 37:229240.Google Scholar
Deegan, L., 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
DeLaune, R. D., Jugsujinda, A., Peterson, G., and Patrick, W. 2003. Impact of Mississippi River freshwater reintroduction on enhancing marsh accretionary processes in a Louisiana estuary. Estuarine, Coastal and Shelf Science, 58:653662.Google Scholar
DeLaune, R. D., and Pezeshki, S.R. 1994. The influence of subsidence and saltwater intrusion on coastal marsh stability: Louisiana Gulf Coast, U.S.A. Journal of Coastal Research, 12:7789.Google Scholar
DeLaune, R. D., Pezeshki, S., and Jugsujinda, A. 2005. Impact of Mississippi River freshwater reintroduction on Spartina patens marshes: responses to nutrient input and lowering of salinity. Wetlands, 25:155161.Google Scholar
De Leeuw, J., Olff, H., and Bakker, J. P. 1990. Year-to-year variation in peak above-ground biomass of six salt-marsh angiosperm communities as related to rainfall deficit and inundation frequency. Aquatic Botany, 36:139151.Google Scholar
Donnelly, J. P., and Bertness, M. D. 2001. Rapid shoreward encroachment of salt marsh cordgrass in response to accelerated sea-level rise. Proceedings of the National Academy of Sciences of the USA 98:1421814223.Google Scholar
Drake, B. G. 1992. A field-study of the effects of elevated CO2 on ecosystem processes in a Chesapeake Bay wetland. Australian Journal of Botany, 40:579595.Google Scholar
Dunn, R., Thomas, S. M., Keys, A. J., and Long, S. P. 1987. A comparison of the growth of C4 grass Spartina anglica with the C3 grass Lolium perenne at different temperatures. Journal of Experimental Botany, 38(188):433441.CrossRefGoogle Scholar
Erickson, J. E., Megonigal, J. P., Peresta, G., and Drake, B. G. 2007. Salinity and sea level mediate elevated CO2 effects on C3-C4 plant interactions and tissue nitrogen in a Chesapeake Bay tidal wetland. Global Change Biology, 13:202215.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.Google Scholar
Farquhar, G. D., von Caemmerer, S., and Berry, J. A. 1980. A biochemical model of photosynthetic CO2 assimilation in leaves of C3 species. Planta, 149:7890.Google Scholar
French, J. 2006. Tidal marsh sedimentation and resilience to environmental change: exploratory modelling of tidal, sea-level and sediment supply forcing in predominantly allochthonous systems. Marine Geology, 235:119136.Google Scholar
Gedan, K. B., Altieri, A. H., and Bertness, M. D. 2011. Uncertain future of New England salt marshes. Marine Ecology Progress Series, 434:229237.Google Scholar
Gedan, K. B., and Bertness, M.D. 2009. Experimental warming causes rapid loss of plant diversity in New England salt marshes. Ecology Letters, 12:842848.Google Scholar
Giosan, L., Syvitski, J., Constantinescu, S. D., and Day, J. 2014. Protect the world’s deltas. Nature, 516:3133.CrossRefGoogle ScholarPubMed
Giurgevich, J. R, and Dunn, E.L. 1979. Seasonal patterns of CO2 and water vapor exchange of the tall and short height forms of Spartina alterniflora Loisel in a Georgia salt marsh. Oecologia, 43:139156.Google Scholar
Gray, A. J., and Mogg, R. J. 2001. Climate impacts on pioneer saltmarsh plants. Climate Research, 18:105112.CrossRefGoogle Scholar
Greenway, H., and Munns, R. 1980. Mechanisms of salt tolerance in nonhalophytes. Annual Review of Plant Physiology, 31(1):149190.CrossRefGoogle Scholar
Grime, J. P. 1988. The C-S-R model of primary plant strategies–origins, implications and tests. pp. 371393. In: Plant Evolutionary Biology, Jain, S. K. (ed.), Chapman & Hall, London.CrossRefGoogle Scholar
Hanson, A., Johnson, R., Wigand, C., Oczkowski, A., Davey, E., and Markham, E. 2016. Responses of Spartina alterniflora to multiple stressors: changing precipitation patterns, accelerated sea level rise, and nutrient enrichment. Estuaries and Coasts, 39:13761385.Google Scholar
Henry, M. K., and Twilley, R. R. 2013. Soil development in a coastal Louisiana wetland during a climate-induced vegetation shift from salt marsh to mangrove. Journal of Coastal Research, 29:12731283.Google Scholar
Herbert, E. R., Boon, P., Burgin, A. J., Neubauer, S. C., Franklin, R. B., Ardón, M., Hopfensperger, K. N., Lamers, L. P. M., and Gell, P. 2015. A global perspective on wetland salinization: ecological consequences of a growing threat to freshwater wetlands. Ecosphere, 6(10):206.Google Scholar
Howard, R. J., and Mendelssohn, I. A. 1999. Salinity as a constraint on growth of oligohaline marsh macrophytes. I. Species variation in stress tolerance. American Journal of Botany, 86(6):785794.Google Scholar
Hughes, A. L. H., Wilson, A. M., and Morris, J. T. 2012. Hydrologic variability in a salt marsh: assessing the links between drought and acute marsh dieback. Estuarine, Coastal and Shelf Science, 111:95106.Google Scholar
Idso, K. E., and Idso, S. B. 1994. Plant responses to atmospheric CO2 enrichment in the face of environmental constraints: a review of the past 10 years’ research. Agricultural and Forest Meteorology, 69:153203.Google Scholar
IPCC, 2014: Climate Change 2014: Synthesis Report. Contribution of Working Groups I, II and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Core Writing Team, Pachauri, R.K. and Meyer, L.A. (eds.)]. IPCC, Geneva, Switzerland.Google Scholar
Jacob, J., Greitner, C. and Drake, B. G. 1995. Acclimation of photosynthesis in relation to Rubisco and non-structural carbohydrate contents and in situ carboxylase activity in Scirpus olneyi grown at elevated CO2 in the field. Plant, Cell and Environment, 18:875884.CrossRefGoogle Scholar
Janousek, C. N., Buffington, K. J., Thorne, K. M., Guntenspergen, G. R., Takekawa, J. Y., and Dugger, B. D. 2016. Potential effects of sea-level rise on plant productivity: species-specific responses in northeast Pacific tidal marshes. Marine Ecology Progress Series, 548:111125.Google Scholar
Johnson, D. S., Warren, R. S., Deegan, L. A., and Mozdzer, T. J. 2016. Saltmarsh plant responses to eutrophication. Ecological Applications, 26(8):26492661.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
Keddy, P. A. 1990. Competitive hierarchies and centrifugal organization in plant communities. pp. 265290. In: Perspectives on Plant Competition, Grace, J.B. and Tilman, D. (eds.), Academic Press, Inc., San Diego, USA.Google Scholar
Keddy, P.A. 2011. Wetland Ecology: Principles and Conservation. Vol. 2. Cambridge University Press, Cambridge, U.K.Google Scholar
Kirwan, M. L., and Guntenspergen, G. R. 2013. Feedbacks between inundation, root production, and shoot growth in a rapidly submerging brackish marsh. Journal of Ecology, 100:764770.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., Guntenspergen, G. R., and Langley, J. A. 2014. Temperature sensitivity of organic-matter decay in tidal marshes. Biogeosciences, 11:48014808.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., Langley, J. A., Guntenspergen, G. R. and Megonigal, J. P. 2013. The impact of sea-level rise on organic matter decay rates in Chesapeake Bay brackish tidal marshes. Biogeosciences, 10:18691876.Google Scholar
Kirwan, M. L., and Megonigal, J. P. 2013. Tidal wetland stability in the face of human impacts and sea-level rise. Nature, 504:5360.Google Scholar
Kirwan, M. L., and Mudd, S. M. 2012. Response of salt-marsh carbon accumulation to climate change. Nature, 489:550553.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
Lamers, L., Govers, L., Jenssen, I., Geurts, J. Van der Welle, M., Van Katwijk, M., Van der Heide, T., Roelofs, J., and Smolders, A. 2013. Sulfide as a soil phytotoxin--a review. Frontiers in Plant Science, 4:14.Google Scholar
Langley, J. A., and Hungate, B. A. 2014. Plant community feedbacks and long-term ecosystem responses to multi-factored global change. AoB Plants, 6:plu035. doi:10.1093/aobpla/plu035Google Scholar
Langley, J. A., McKee, K. L., Cahoon, D. R., Cherry, J. A., and Megonigal, J. P. 2009. Elevated CO2 stimulates marsh elevation gain, counterbalancing sea-level rise. Proceedings of the National Academy of Sciences of the USA, 106(15):61826186.Google Scholar
Langley, J. A., and Megonigal, J. P. 2010. Ecosystem response to elevated CO2 levels limited by nitrogen-induced plant species shift. Nature, 466:9699.Google Scholar
Langley, J. A., Mozdzer, T. J., Shepard, K. A., Hagerty, S. B., and Megonigal, J. P. 2013. Tidal marsh plant responses to elevated CO2, nitrogen fertilization, and sea level rise. Global Change Biology, 19: 14951503.Google Scholar
Lee, S. -Y., Hamlet, A. F., and Grossman, E. E. 2016. Impacts of climate change on regulated streamflow, hydrologic extremes, hydropower production, and sediment discharge in the Skagit River Basin. Northwest Science, 90(1):2343.Google Scholar
Lenssen, G. M., Lamers, J., Stroetenga, M., and Rozema, J. 1993. Interactive effects of atmospheric CO2 enrichment, salinity and flooding on growth of C3 (Elymus athericus) and C4 (Spartina anglica) salt marsh species. Vegetatio, 104/105:379388.Google Scholar
Leuzinger, S., Luo, Y., Beier, C., Dieleman, W., Vicca, S., and Korner, C. 2011. Do global change experiments overestimate impacts on terrestrial ecosystems? Trends in Ecology & Evolution, 26:236241.Google Scholar
Linthurst, R. A., and Seneca, E. D. 1981. Aeration, nitrogen and salinity as determinants of Spartina alterniflora Loisel. Growth response. Estuaries, 4(1):5363.Google Scholar
Lissner, J., Mendelssohn, I. A., Lorenzen, B., Brix, H., Mckee, K. L., and Miao, S. 2003. Interactive effects of redox intensity and phosphate availability on growth and nutrient relations of Cladium jamaicense (Cyperaceae), American Journal of Botany 90: 736748.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
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
McKee, K., Rogers, K., and Saintilan, N. 2012. Response of salt marsh and mangrove wetlands to changes in atmospheric CO2, climate, and sea level. Global Change Ecology and Wetlands, 1:6396.Google Scholar
Megonigal, J. P., Hines, M. E., and Visscher, P. T. 2004. Anaerobic metabolism: linkages to trace gases and aerobic processes. pp. 317424. In: Biogeochemistry, Schlesinger, W.H. (ed.), Elsevier-Pergamon, Oxford, UK.Google Scholar
Mendelssohn, I. A., and McKee, K. L. 1992. Indicators of environmental stress in wetland plants. pp. 603624. In: Ecological Indicators, McKenzie, D. H., Hyatt, D. E., and MacDonald, V. J. (eds.), Elsevier Applied Science, New York, NY, USA.Google Scholar
Mendelssohn, I. A., McKee, K. L., Hester, M. W., Lin, Q., McGinnis, T., and Willis, J. 2006. Brown marsh task II.1: integrative approach to understanding the causes of salt marsh dieback – determination of salt marsh species tolerance limits to potential environmental stressors. Report submitted to the Louisiana Department of Natural Resources, Baton Rouge, LA.Google Scholar
Mendelssohn, I. A., and Morris, J. T. 2000. Ecophysiological controls on the growth of Spartina alterniflora. pp. 5980. In: Concepts and Controversies in Tidal Marsh Ecology. Weinstein, N. P.,, and Kreeger, D. A. (eds.). Kluwer Academic Publishers, New York.Google Scholar
Mendelssohn, I. A., Sorrell, B. K., Brix, H., Schierup, H. H., Lorenzen, B., and Maltby, E. 1999. Controls on soil cellulose decomposition along a salinity gradient in a Phragmites australis wetland in Denmark. Aquatic Botany, 64:381398.Google Scholar
Miller, W. D., Neubauer, S. C., and Anderson, I. C. 2001. Effects of sea level induced disturbance on high salt marsh metabolism. Estuaries, 24:357367.Google Scholar
Montagna, P. A., and Ruber, E. 1980. Decomposition of Spartina alterniflora in different seasons and habitats of a northern Massachusetts salt marsh, and a comparison with other Atlantic regions. Estuaries, 3:6164.Google Scholar
Morris, J. T. 1988. Pathways and controls of the carbon cycle in salt marshes. pp. 497510. In: The Ecology and Management of Wetlands, Volume 1, Ecology of Wetlands, Hook, D. D., (ed.), Croom Helm Ltd., Beckenham, UK.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
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(10):28692877.Google Scholar
Morrissey, E. M., Gillespie, J. L., Morina, J. C., and Franklin, R. B. 2014. Salinity affects microbial activity and soil organic matter content in tidal wetlands. Global Change Biology, 20:13511362.Google Scholar
Mote, P. W., and Salathé, E. P. Jr 2010. Future climate in the Pacific Northwest. Climatic Change, 102:2950.CrossRefGoogle Scholar
Munns, R. 2002. Comparative physiology of salt and water stress. Plant, Cell and Environment, 25:239250.Google Scholar
Naidoo, G., McKee, K. L., and Mendelssohn, I. A. 1992. Anatomical and metabolic responses to waterlogging and salinity in Spartina alterniflora and S. patens (Poaceae). American Journal of Botany, 79(7):765770.Google Scholar
Nyman, J. A., and DeLaune, R. D. 1991. CO2 emission and soil Eh responses to different hydrological regimes in fresh, brackish and saline marsh soils. Limnology and Oceanography, 36:14061414.CrossRefGoogle Scholar
Nyman, J. A., DeLaune, R. D., Roberts, H. H.,and Patrick, W. H. 1993. Relationship between vegetation and soil formation in a rapidly submerging coastal marsh. Marine Ecology Progress Series, 96:269279.Google Scholar
Odum, W. E. 1988. Comparative ecology of tidal freshwater and salt marshes. Annual Review of Ecology and Systematics, 19:147176.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
Pennings, S. C., Grant, M. -B., and Bertness, M. D. 2005. Plant zonation in low-latitude salt marshes: disentangling the roles of flooding, salinity and competition. Journal of Ecology, 93:159167.Google Scholar
Poljakoff-Mayber, A. 1988. Ecological-physiological studies on the responses of higher plants to salinity and drought. Science Review of Arid Zone Research, 6:163183.Google Scholar
Polley, W. H., Johnson, H. B.,and Derner, J. D. 2003. Increasing CO2 from subambient to superambient concentrations alters species composition and increases above-ground biomass in a C3/C4 grassland. New Phytologist, 160:319327.Google Scholar
Pozo, J., and Colino, R. 1992. Decomposition processes of Spartina maritime in a salt marsh of the Basque Country. Hydrobiologia, 231:165175.Google Scholar
Rahmstorf, A. 2007. A semi-empirical approach to projecting future sea-level rise. Science, 315:368370.Google Scholar
Rasse, D. P., Peresta, G., and Drake, B. G. 2005. Seventeen years of elevated CO2 exposure in a Chesapeake Bay wetland: sustained but contrasting responses of plant growth and CO2 uptake. Global Change Biology, 11:369377.Google Scholar
Ratliff, K. M., Braswell, A. E., and Marani, M. 2015. Spatial response of coastal marshes to increased atmospheric CO2. Proceedings of the National Academy of Sciences of the USA, 112(51):1558015584.Google Scholar
Reed, D. J. 1995. The response of coastal marshes to sea-level rise: survival or submergence? Earth Surface Processes, 20:3948.Google Scholar
Reich, P. B. 2009. Elevated CO2 reduces losses of plant diversity caused by nitrogen deposition. Science, 326:13991402.Google Scholar
Rozema, J., Dorel, F., Janissen, R., Lenssen, G. M., Broekman, R. A., Ar, W. J., and Drake, B. G. 1991. Effect of elevated CO2 on growth, photosynthesis and water relations of salt marsh grass species. Aquatic Botany, 39:4555.Google Scholar
Saintilan, N., Wilson, N., Rogers, K., Rajkaran, A., and Krauss, W. K. 2014. Mangrove expansion and salt marsh decline at mangrove poleward limits. Global Change Biology, 20:147157.Google Scholar
Schile, L. M., Callaway, J. C., Suding, K. N., and Kelly, N. M. 2017. Can community structure track sea-level rise? Stress and competitive controls in tidal wetlands. Ecology and Evolution, 7:12761285.Google Scholar
Sharpe, J. P., and Baldwin, H. A. 2012. Tidal marsh plant community response to sea-level rise: a mesocosm study. Aquatic Botany, 101:3440.Google Scholar
Shaver, G. R., Canadell, J., Chapin, F. S., Gurevitch, J., Harte, J., Henry, G., Ineson, P., et al. 2000. Global warming and terrestrial ecosystems: a conceptual framework for analysis. Bioscience, 50:871882.Google Scholar
Shea, M. L. 1977. Photosynthesis and photorespiration in relation to phenotypic forms of Spartina alterniflora. PhD thesis, Yale University, New Haven, Connecticut.Google Scholar
Short, F. T., Kosten, S., Morgan, P. A., Malone, S., and Moore, G. E. 2016. Impacts of climate change on submerged and emergent wetland plants. Aquatic Botany, 135:317.Google Scholar
Smith, S. M. 2009. Multi-decadal changes in salt marshes of Cape Cod, MA: photographic analyses of vegetation loss, species shifts and geomorphic change. Northeastern Naturalist, 16:183208.Google Scholar
Stagg, C. L., Schoolmaster, D. R. Jr., Piazza, S. C., Snedden, G., Steyer, G. D., Fischenich, C. J., and McComas, R. W. 2017. A landscape-scale assessment of above- and belowground primary production in coastal wetlands: implications for climte change-induced community shifts. Estuaries and Coasts, 40:856879.Google Scholar
Stevenson, J. C., Kearney, M. S., and Pendleton, E. C. 1985. Sedimentation and erosion in a Chesapeake Bay brackish marsh system. Marine Geology, 67:213235.Google Scholar
Stralberg, D., Brennan, M., Callaway, J. C., Wood, J. K., Schile, L. M., Jongsomjit, D., Kelly, M., Parker, V. T., and Crooks, S. 2011. Evaluating tidal marsh sustainability in the face of sea-level rise: a hybrid modeling approach applied to San Francisco Bay. PLoS ONE, 6(11): e27388.Google Scholar
Sundareshwar, P. V., Morris, J. T., Koepfler, E. K., and Fornwalt, B. 2003. Phosphorus limitation of coastal ecosystem processes. Science, 299:563565.Google Scholar
Syvitski, J. P. M., Vörösmarty, C. J., Kettner, A. J., and Green, P. 2005. Impact of humans on the flux of terrestrial sediment to the global coastal ocean. Science, 308:376380.Google Scholar
Teal, J. M., and Howes, B. L. 1996. Interannual variability of a salt-marsh ecosystem. Limnology and Oceanography, 41:802809.Google Scholar
Tobias, V. D., and Nyman, J. A. 2017. Leaf tissue indicators of flooding stress in the above- and belowground biomass of Spartina patens. Journal of Coastal Research, 33(2):309320.Google Scholar
Turner, R. E., Swenson, E. M., and Milan, C. S. 2000. Organic and inorganic contributions to vertical accretion in salt marsh sediments. pp. 583595. In: Concepts and Controversies in Tidal Marsh Ecology, Weinstein, M.P., and Kreeger, D.A. (eds.), Springer, Dordrecht, The Netherlands,.Google Scholar
Turner, R. E., Swenson, E. M., Milan, C. S., Lee, J. M. and Oswald, T. A. 2004. Below-ground biomass in healthy and impaired salt marshes. Ecological Research, 19:2935.Google Scholar
Turner, R. E. 2011. Beneath the salt marsh canopy: loss of soil strength with increasing nutrient loads. Estuaries and Coasts, 34:10841093.Google Scholar
Urban, O. 2003. Physiological impacts of elevated CO2 concentration ranging from molecular to whole plant responses. Photosynthetica, 41:920.Google Scholar
Valiela, I., Teal, J. M., Allen, S. D., Van Etten, R., Goehringer, D., and Volkmann, S. 1985. Decomposition in salt marsh ecosystems: the phases and major factors affecting disappearance of above-ground organic matter. Journal of Experimental Marine Biology and Ecology, 89:2954.Google Scholar
van Dobben, H. F., and Slim, P. A. 2011. Past and future plant diversity of a coastal wetland driven by soil subsidence and climate change. Climatic Change, 110: 597618.Google Scholar
Visser, J. M., Duke-Sylvester, S. M., Carter, J., and Broussard, W. P. III 2013. A computer model to forecast wetland vegetation changes resulting from restoration and protection in coastal Louisiana. Journal of Coastal Research, 67(4):5159.Google Scholar
Wasson, K., Endris, R. C., Perry, D. C., Woolfolk, A., Beheshti, K., Rodriguez, M., et al. 2017. Eutrophication decreases salt marsh resilience through proliferation of algal mats. Biological Conservation 212:111.Google Scholar
Watson, E. B., Wigand, C., Davey, E. W., Andrews, H. M., Bishop, J., and Raposa, K. B. 2017. Wetland loss patterns and inundation-productivity relationships prognosticate widespread salt marsh loss for southern New England. Estuaries and Coasts, 40:662681.Google Scholar
Wells, J. T., and Coleman, J. M. 1987. Wetland loss and the subdelta life cycle. Estuarine, Coastal and Shelf Science, 25:111125.Google Scholar
White, K. P., Langley, J. A., Cahoon, D. R., and Megonigal, J. P. 2012. C3 and C4 biomass allocation responses to elevated CO2 and nitrogen: contrasting resource capture strategies. Estuaries and Coasts, 35:10281035.Google Scholar
Williams, K., Pinzon, Z. S., Stumpf, R. P., and Raabe, E. A. 1999. Sea-level rise and coastal forests on the Gulf of Mexico. Open-file report 99-441, United States Geological Survey, St. Petersburg, FL.Google Scholar
Willis, J. M., and Hester, M. W. 2004. Interactive effects of salinity, flooding, and soil type on Panicum hemitomon. Wetlands, 24(1):4350.Google Scholar
Wolf, A. A., Drake, B. G., Erickson, J. E., and Megonigal, J. P. 2007. An oxygen-mediated positive feedback between elevated carbon dioxide and soil organic matter decomposition in a simulated anaerobic wetland. Global Change Biology, 13:20362044.Google Scholar
Wolters, M., Garbutt, A., and Bakker, J. P. 2005. Salt-marsh restoration: evaluating the success of de-embankments in north-west Europe. Biological Conservation, 123:249268.Google Scholar
Woodrow, I. E., and Berry, J. A. 1988. Enzymatic regulation of photosynthetic CO2 fixation in C3 plants. Annual Review of Plant Physiology and Plant Molecular Biology, 39:533594.CrossRefGoogle Scholar
Woodward, F. I., Thompson, G. B., and McKee, I. F. 1991. The effects of elevated concentrations of carbon dioxide on individual plants, populations, communities, and ecosystems. Annals of Botany, 67:2338.Google Scholar
Wu, W., Huang, H., Biber, P., and Bethel, M. 2017. Litter decomposition of Spartina alterniflora and Juncus roemerianus: implications of climate change in salt marshes. Journal of Coastal Research, 33(2):372384.Google Scholar
Yeo, A. 1999. Predicting the interaction between the effects of salinity and climate change on crop plants. Scientia-Horticulturae-Amsterdam, 78:159174.Google Scholar
Zedler, J. B. 1983. Freshwater impacts in normally hypersaline marshes. Estuaries, 6:346355.Google Scholar
Zedler, J. B., Williams, P., and Boland, J. 1986. Catastrophic events reveal the dynamic nature of salt-marsh vegetation in southern California. Estuaries, 9(1):7580.Google Scholar

References

Able, K. W., Hagan, S. M., and Brown, S. A. 2003. Mechanisms of marsh habitat alteration due to Phragmites: Response of young-of-the-year mummichog (Fundulus heteroclitus) to treatment for Phragmites removal. Estuaries 26:484494.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
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
Amsberry, L., Baker, M. A., Ewanchuk, P. J., and Bertness, M. D. 2000. Clonal integration and the expansion of Phragmites australis. Ecological Applications, 10: 11101118.Google Scholar
Anderson, C. M., and Treshow, M. 1980. A review of environmental and genetic factors that affect height in Spartina alterniflora Loisel. (Salt marsh cord grass). Estuaries, 3:168176.Google Scholar
Anderson, M. G. 1995. Interaction between Lythrum salicaria and native organisms: a critical review. Environmental Management, 19: 225231.Google Scholar
Ayres, D. R., Smith, D. L., Zaremba, K., Klohr, S., and Strong, D. R. 2004. Spread of exotic cordgrasses and hybrids (Spartina sp.) in the tidal marshes of San Francisco Bay, California, USA. Biological Invasions, 6:221231.Google Scholar
Ayres, D. R., Strong, D. R., and Baye, P. 2003. Spartina foliosa (Poaceae)–a common species on the road to rarity. Madrono, 50:209213.Google Scholar
Balouskus, R. G., and Targett, T. E. 2012. Egg deposition by Atlantic silverside, Menidia menidia: substrate utilization and comparison of natural and altered shoreline type. Estuaries and Coasts, 35:11001109.Google Scholar
Bart, D., Burdick, D., Chambers, R. and Hartman, J. M. 2006. Human facilitation of Phragmites australis invasions in tidal marshes: a review and synthesis. Wetlands Ecology and Management, 14:5365.Google Scholar
Belknap, D. F. and Wilson, K. R. 2014. Invasive green crab impacts on salt marshes in Maine-sudden increase in erosion potential. Abstract. Northeast Section, Geological Society of America. 24 March 2014, Lancaster, PA.Google Scholar
Benoit, L. K., and Askins, R. A. 1999. Impact of the spread of Phragmites on the distribution of birds in Connecticut tidal marshes. Wetlands, 19: 194208.Google Scholar
Bertness, M. D. 1984. Habitat and community modification by an introduced herbivorous snail. Ecology, 65: 370381.Google Scholar
Bertness, M. D., Brisson, C. P., Bevil, M. C., and Crotty, S. M. 2014. Herbivory drives the spread of salt marsh die-off. PLoS ONE, 9(3): e92916. doi:10.1371/journal.pone.0092916Google Scholar
Blank, R., and Young, J. 1997. Influence of invasion of perennial pepperweed on soil properties. pp. 1113 In Management of Perennial Pepperweed (Tall Whitetop). Special Report 972. Corvallis, OR: U.S. Department of Agriculture, Agricultural Research Service; Oregon State University, Agricultural Experiment Station.Google Scholar
Blossey, B., Skinner, L. C. and Taylor, J. 2001. Impact and management of purple loosestrife (Lythrum salicaria) in North America. Biodiversity and Conservation, 10:17871807.Google Scholar
Boyer, K. E. and Burdick, A. P. 2010. Control of Lepidium latifolium (perennial pepperweed) and recovery of native plants in tidal marshes of the San Francisco Estuary. Wetlands Ecology and Management, 18(6): 731743.Google Scholar
Buchsbaum, R. N., Catena, J., Hutchins, E., and Pirri, M. J. 2006. Changes in salt marsh vegetation, Phragmites australis, and nekton in response to increased tidal flushing in a New England salt marsh. Wetlands, 26:544557.Google Scholar
Burdick, D. M., and Konisky, R. A. 2003. Determinants for expansion of Phragmites australis, common reed, in natural and impacted coastal marshes. Estuaries, 26: 407416.Google Scholar
Burdick, D. M., and Roman, C. T. 2012. Salt marsh responses to tidal restriction and restoration. A summary of experiences. pp. 373382 In: Roman, C.T. and Burdick, D.M. (eds.) Tidal Marsh Restoration: A Synthesis of Science and Practice. Island Press. Washington.Google Scholar
Burdick, D., Peter, C., and Moore, G. E. 2013. Phase II of Tidal Marsh Restoration at Steedman Woods Reserve at York, Maine. Final report to Museums of Old York; accessed from UNH Scholars Repository: https://scholars.unh.edu.Google Scholar
Casazza, M. L., Overton, C. T., Bui, T. -V. D., Hull, J. M., Albertson, J. D., Bloom, V. K., Bobzien, S., et al. 2016. Endangered species management and ecosystem restoration: finding the common ground. Ecology and Society, 21(1):19.Google Scholar
Chambers, R. M., Meyerson, L. A., and Dibble, K. L. 2012. Ecology of Phragmites australis and responses to tidal restoration. pp. 8196. In: Roman, C.T. and Burdick, D.M. (eds.) Tidal Marsh Restoration. Island Press. Washington, DC.Google Scholar
Chambers, R. M., Meyerson, L. A., and Saltonstall, K. 1999. Expansion of Phragmites australis into tidal wetlands of North America. Aquatic Botany, 64:261273.Google Scholar
Chambers, R. M., Osgood, D. T., Bart, D. J., and Montalto, F. 2003. Phragmites australis invasion and expansion in tidal wetlands: interactions among salinity, sulfide, and hydrology. Estuaries, 26:398406.Google Scholar
Chapman, J. W., Carlton, J. T., Bellinger, M. R, and Blakeslee, A. M. H. 2007. Premature refutation of a human-mediated marine species introduction: the case history of the marine snail Littorina littorea in the Northwestern Atlantic. Biological Invasions, 9:9951008.Google Scholar
Chen, H., Qualls, R. G., and Miller, M. C. 2002. Adaptive responses of Lepidium latifolium to soil flooding: biomass allocation, adventitious rooting, aerenchyma formation and ethylene production. Environmental and Experimental Botany, 48:119128.Google Scholar
Connolly, B. A. and Hale, I. L. 2016. Lepidium latifolium (Brassicaceae): invasive perennial pepperweed observed in Rhode Island. Rhodora, 118(974):229231.CrossRefGoogle Scholar
Coverdale, T. C., Altieri, A. H., and Bertness, M. D. 2012. Belowground herbivory increases vulnerability of New England salt marshes to die-off. Ecology, 93:20852094.Google Scholar
Daehler, C. C., and Strong, D. R. 1996. Status, prediction and prevention of introduced cordgrass Spartina spp. invasions in Pacific estuaries, USA. Biological Conservation, 78:5158.Google Scholar
Daehler, C., and Strong, D. 1997. Hybridization between introduced smooth cordgrass (Spartina alterniflora; Poaceae) and native California cordgrass (S. foliosa) in San Francisco Bay, California, USA. American Journal of Botany, 84:607611.Google Scholar
Davidson, T. M., and de Rivera, C. E. 2010. Accelerated erosion of saltmarshes infested by the non-native burrowing crustacean Sphaeroma quoianum. Marine Ecology Progress Series, 419:129136.Google Scholar
Denoth, M., and Myers, J. H. 2007. Competition between Lythrum salicaria and a rare species: combining evidence from experiments and long-term monitoring. Plant Ecology, 191:153161.Google Scholar
Dibble, K. L., andMeyerson, L. A. 2012. Tidal flushing restores the physiological condition of fish residing in degraded salt marshes. PLoS ONE 7(9): e46161. doi:10.1371/journal.pone.0046161Google Scholar
Dibble, K. L., and Meyerson, L. A. 2013. The effects of plant invasion and ecosystem restoration on energy flow through salt marsh food webs. Estuaries and Coasts, 35Google Scholar
Dibble, K. L., Pooler, P. S, and Meyerson, L. A. 2013. Impacts of plant invasions can be reversed through restoration: a regional meta-analysis of faunal communities. Biological Invasions, 15:17251737.Google Scholar
Elmer, W. H., LaMondia, J. A., Useman, S., Mendelssohn, I. A., Schneider, R. W., Jimenez-Gasco, M. M., Marra, R. E., and Caruso, F. L. 2013. Sudden vegetation dieback in Atlantic and Gulf Coast salt marshes. Plant Disease, 97: 436445.Google Scholar
Eiswerth, M., Singletary, L., Zimmerman, J., Johnson, W. 2005. Dynamic benefit–cost analysis for controlling perennial pepperweed (Lepidium latifolium): A Case StudyWeed Technology 19: 237243.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: RG1002.Google Scholar
Farnsworth, E. J., and Ellis, D. R. 2001. Is purple loosestrife (Lythrum salicaria) an invasive threat to freshwater wetlands? Conflicting evidence from several ecological metrics. Wetlands, 21:199209.Google Scholar
Feng, J., Huang, Q., Qi, F., Guo, J., and Lin, G. 2015. Utilization of exotic Spartina alterniflora by fish community in the mangrove ecosystem of Zhangjiang Estuary: evidence from stable isotope analyses. Biological Invasions, 7: 21132121.Google Scholar
Flanagan, R. J., Mitchell, R. J., and Karron, J. D. 2010. Increased relative abundance of an invasive competitor for pollination, Lythrum salicaria, reduces seed number in Mimulus ringens. Oecologia, 164: 445454.Google Scholar
Ford, M. A., and Grace, J. B. 1998. Effects of vertebrate herbivores on soil processes, plant biomass, litter accumulation and soil elevation changes in a coastal marsh. Journal of Ecology, 86: 974982.Google Scholar
Grosholz, E. 2010. Avoidance by grazers facilitates spread of an invasive hybrid plant. Ecology Letters, 13:145153.Google Scholar
Guntenspergen, G. R., Keough, J. R., and Weinstein, M. P. 2003. Phragmites technical forum and workshop: synthesis of scientific knowledge and management needs. Estuaries, 26:18.Google Scholar
Hauber, D. P., Saltonstall, K., White, D. A., and Hood, C. S. 2011. Genetic variation in the common reed, Phragmites australis, in the Mississippi River Delta marshes: Evidence for multiple introductions. Estuaries and Coasts, 34:851862.Google Scholar
Hazelton, E. L. G., Mozdzer, T. J., Burdick, D. M., Kettenring, K. M., and Whigham, D. F. 2014. Phragmites australis management in the United States: 40 years of methods and outcomes. AoB Plants 6: plu001.Google Scholar
Hershner, C., and Havens, K. J. 2008. Managing invasive aquatic plants in a changing system: strategic consideration of ecosystem services. Conservation Biology, 22: 544550.Google Scholar
Hierro, J. L., and Callaway, R. M. 2003. Allelopathy and exotic plant invasion. Plant and Soil, 256:2939.Google Scholar
Hindell, J. S., and Warry, F. Y. 2010. Nutritional support of estuary perch (Macquaria colonorum) in a temperate Australian inlet: Evaluating the relative importance of invasive Spartina. Estuarine, Coastal and Shelf Science, 90:159167.Google Scholar
Holdredge, C., Bertness, M. D., and Altieri, A. H. 2009. Role of crab herbivory in die-off of New England salt marshes. Conservation Biology, 23: 672679.Google Scholar
Hughes, Z. J., FitzGerald, D. M., Wilson, C. A., Pennings, S. C., Wieski, K., and Mahadevan, A.. 2009. Rapid headward erosion of marsh creeks in response to relative sea level rise. Geophysical Research Letters 36: 5.Google Scholar
Jodoin, Y., Lavoie, C. L., Villeneuve, P., Theriault, M., Beaulieu, J., and Belzile, F. 2008. Highways as corridors and habitats for the invasive common reed Phragmites australis in Quebec, Canada. Journal of Applied Ecology, 45:459466.Google Scholar
Jefferies, R. L., Jano, A. P., and Abraham, K. F. 2006. A biotic agent promotes large-scale catastrophic change in the coastal marshes of Hudson Bay. Journal of Ecology, 94:234242.Google Scholar
Kahn, H. 1973. Paris Notebook, doldrums for French research. New Scientist 60:797–98.Google 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
Kerr, D. W., Hogle, I. B., Ort, B. S., and Thornton, W. J. 2016. A review of 15 years of Spartina management in the San Francisco Estuary. Biological Invasions, 18:22472266.Google Scholar
Kiehn, W. M., and Morris, J. T. 2009. Relationships between Spartina alterniflora and Littoraria irrorata in a South Carolina salt marsh. Wetlands, 29:818825.Google Scholar
Kiviat, E. 2013. Ecosystem services of Phragmites in North America with emphasis on habitat functions. AoB PLANTS 5: plt008. https://doi.org/10.1093/aobpla/plt008Google Scholar
Knight, I. A., Wilson, B. E., Gill, M., Aviles, L., Cronin, J. T., Nyman, J. A., Schneider, S. A., and Diaz, R. 2018. Invasion of Nipponaclerda biwakoensis (Hemiptera: Aclerdidae) and Phragmites australis die-back in southern Louisiana, USA. Biological Invasions, 20:27392744.Google Scholar
Konisky, R. A. and Burdick, D. M. 2004. Effects of stressors on invasive and halophytic plants. of New England salt marshes: a framework for predicting response to tidal restoration. Wetlands, 24: 434447.Google Scholar
Lelong, B., Lavoie, C., Jodoin, Y., and Belzile, F. 2007. Expansion pathways of the exotic common reed (Phragmites australis): a historical and genetic analysis. Diversity and Distributions, 13:430437.Google Scholar
Levin, L. A., Neira, C., and Grosholz, E. D. 2006. Invasive cordgrass modifies wetland trophic function. Ecology, 87:419432.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
Li, B., Liao, C., Zhang, X., Chen, H., Wang, Q., Chen, Z., Gan, X. et al. 2009. Spartina alterniflora invasions in the Yangtze River Estuary, China: An overview of current status and ecosystem effects. Ecological Engineering, 35:511520.Google Scholar
Lyeik, K. A. 1989. Lepidium latifolium L., a sea-shore species in Norway. Blyttia, 47:109113.Google Scholar
Marchant, C. J. 1967. Evolution in Spartina (Gramineae): I. The history and morphology of the genus in Britain. Botanical Journal, 60:124.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
Meyerson, L. A., Saltonstall, K., and Chambers, R. M. 2009. Phragmites australis in coastal marshes of North America: A historical and ecological perspective. pp. 5782. In: Human Impacts on Salt Marshes: A Global Perspective, ed. Silliman, B.R., Bertness, M. D., Grosholz, E. D. (eds.) University of California Press, Berkeley.Google Scholar
Miller, D. L., Smeins, F. E., Webb, J. W., and Yager, L. 2005. Mid-Texas, USA coastal marsh vegetation pattern and dynamics as influenced by environmental stress and snow goose herbivory. Wetlands, 25:648658.Google Scholar
Minchinton, T. E. 2002. Precipitation during El Niño correlates with increasing spread of Phragmites australis in New England, USA, coastal marshes. Marine Ecology Progress Series, 242: 305309.Google Scholar
Minchinton, T. E., and Bertness, M. D. 2003. Disturbance mediated competition and the spread of Phragmites australis in a coastal marsh. Ecological Applications 13: 14001416.Google Scholar
Minchinton, T. E., Simpson, J. C., and Bertness, M. D. 2006. Mechanisms of exclusion of native coastal marsh plants by an invasive grass. Journal of Ecology, 94: 342354.Google Scholar
Moore, G. E., Burdick, D. M., Peter, C. R. and Keirstead, D. R. 2011. Mapping soil pore water salinity of tidal marsh habitats using electromagnetic induction in Great Bay Estuary, USA. Wetlands, 31:309318.Google Scholar
Moore, G. E., Burdick, D. M., Peter, C. R. and Keirstead, D. R. 2012. Belowground biomass of Phragmites australis in coastal marshes. Northeast Naturalist, 19:611626.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
Morse, N. B., Pellissier, P. A., Cianciola, E. N., Bereton, R. L., Sullivan, M. M., Shonka, N. K., Wheeler, T. B., and McDowell, W. H. 2014. Novel ecosystems in the Anthopocene: a revision of the novel ecosystem concept for pragmatic applications. Ecology and Society, 19:12.Google Scholar
Neira, C., Levin, L A., Grosholz, E. D. and Mendoza, G. 2007. Influence of invasive Spartina growth stages on associated macrofaunal communities. Biological Invasions 9:975993.Google Scholar
Nyman, J. A., Walters, R. J., Delaune, R. D. and Patrick, W. H. Jr. 2006. Marsh vertical accretion via vegetative growth. Estuarine, Coastal and Shelf Science, 69:370380.Google Scholar
Orson, R. A. 1999. A paleoecological assessment of Phragmites australis in New England tidal marshes: changes in plant community structure during the last few millennia. Biological Invasions, 1:149158.Google Scholar
Orth, J. F., Gammon, M., Abdul-Basir, F., Stevenson, R. D., Tsirelson, D., Ebersole, J., Speak, S. and Kesseli, R. 2006. Natural history, distribution, and management of Lepidium latifolium (Brassicaceae) in New England. Rhodora, 108: 103118.Google Scholar
Overton, C. T., Casazza, M. L., Takekawa, J. Y., Strong, D. R., and Holyoak, M. 2014. Tidal and seasonal effects on survival rates of the endangered California clapper rail: Does invasive Spartina facilitate greater survival in a dynamic environment? Biological Invasions, 16:18971914.Google Scholar
Patten, K. 2002. Smooth cordgrass (Spartina alterniflora) control with Imazapyr. Weed Technology, 16:826–32.Google Scholar
Peter, C. R., and Burdick, D. M. 2010. Can plant competition and diversity reduce the growth and survival of exotic Phragmites australis invading a tidal marsh? Estuaries and Coasts, 33: 12261236.Google Scholar
Portnoy, J. W., and Valiela, I. 1997. Short-term effects of salinity reduction and drainage on salt-marsh biogeochemical cycling and Spartina (cordgrass) production. Estuaries, 20:569578.Google Scholar
Raichel, D. L., Able, K. W., and Hartman, J. M. 2003. The influence of Phragmites (common reed) on the distribution, abundance, and potential prey of a resident marsh fish in the Hackensack Meadowlands, New Jersey. Estuaries, 26: 511521.Google Scholar
Raposa, K. B., McKinney, R. A., Wigand, C., Hollister, J. W., Lovall, C., Szura, K., Gurak, J. A. Jr., et al. 2018. Top-down and bottom-up controls on southern New England salt marsh crab populations. PeerJ, 6, e4876. doi.org/10.7717/peerj.4876Google Scholar
Renz, M. J., and Blank, R. R. 2004. Influence of perennial pepperweed (Lepidium latifolium) biology and plant-soil relationships on management and restoration. Weed Technology, 18:13591363.Google Scholar
Renz, M. J., DiTomaso, , and J. M. 1998. The effectiveness of mowing and herbicides to control perennial pepperweed in rangeland and roadside habitats. Proceedings of the California Weed Science Society, 50:178.Google Scholar
Renz, M. J., and DiTomaso, J. M. 1999. Biology and control of perennial pepperweed. Proceedings of the California Weed Science Society, 51: 1316.Google Scholar
Renz, M. J., and DiTomaso, J. M. 2006. Early season mowing improves the effectiveness of chlorsulfuron and glyphosate for control of perennial pepperweed (Lepidium latifolium). Weed Technology, 20:3236.Google Scholar
Reynolds, L. K., and Boyer, K. E. 2010. Perennial pepperweed (Lepidium latifolium): Properties of invaded tidal marshes. Invasive Plant Science and Management, 3(2): 130138.Google Scholar
Robbins, W. W., Bellue, M. K., and Ball, W. S. 1951. Weeds of California. Sacramento, CA: California Department of Agriculture.Google Scholar
Rohmer, T., Kerr, D., and Hogle, I. 2014. San Francisco Estuary Invasive Spartina Project 2013 ISP monitoring and treatment report. Prepared for the California State Coastal Conservancy, Oakland, California, USA. www.spartinaGoogle Scholar
Rooth, J. E., and Stevenson, J. C. 2000. Sediment deposition patterns in Phragmites australis communities: Implications for coastal areas threatened by rising sea-level. Wetlands Ecology and Management, 8:173183.Google Scholar
Rudrappa, T., Bonsall, J., Gallagher, J. L., Seliskar, D. M., and Bais, H. P. 2007. Root-secreted allelochemical in the noxious weed Phragmites australis deploys a reactive oxygen species response and microtubule assembly disruption to execute rhizotoxicity. Journal of Chemical Ecology, 33: 18981918.Google Scholar
Saltonstall, K. 2002. Cryptic invasion by a non-native genotype of the common reed, Phragmites australis, into North America. Proceedings of the National Academy of Sciences of the USA, 99:24452449Google Scholar
Saltonstall, K., Peterson, P. M., and Soreng, R. 2004. Recognition of Phragmites australis subsp. americanus (Poaceae: Arundinaceae) in North America: Evidence from morphological and genetic analyses. Sida, 21:683692.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 US 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
Simenstad, C. A., and Thom, R. M. 1995. Spartina alterniflora (smooth cordgrass) as an invasive halophyte in Pacific Northwest Estuaries. Hortus Northwest, 6:912; 38–40.Google Scholar
Smith, J. A. M. 2013. The role of Phragmites australis in mediating inland salt marsh migration in a mid-Atlantic estuary. PLoS ONE 8(5): e65091. doi.org/10.1371/journal.pone.0065091Google Scholar
Smith, S. M. 2009. Multi-decadal changes in salt marshes of Cape Cod, MA: photographic analyses of vegetation loss, species shifts, and geomorphic change. Northeastern Naturalist, 16:183208.Google Scholar
Smith, S. M., Roman, C. T., James-Pirri, M. J., Chapman, , Portnoy, K. J., and Gwilliam, E. 2009. Responses of plant communities to incremental hydrologic restoration of a tide-restricted salt marsh in southern New England (Massachusetts, U.S.A.). Restoration Ecology, 17:606618.Google Scholar
Strong, D. R., and Ayres, D. R. 2013. Ecological and evolutionary misadventures of Spartina. Annual Review of Ecology, Evolution, and Systematics, 44:389410.Google Scholar
Talley, T. S., Crooks, J. A., and Levin, L. A. 2001. Habitat utilization and alteration by the invasive burrowing isopod, Sphaeroma quoyanum, in California salt marshes. Marine Biology, 138:561573.Google Scholar
Tang, M., and Kristensen, E. 2010. Associations between macrobenthos and invasive cordgrass, Spartina anglica, in the Danish Wadden Sea. Helgoland Marine Research, 64:321329.Google Scholar
Tavernia, B. G., and Reed, J. M. 2012. The impact of exotic purple loosestrife (Lythrum salicaria) on wetland bird abundances. The American Midland Naturalist, 168: 352363.Google Scholar
Thornton, W. 2018. How do transplant source, restoration site, and herbivory influence Pacific cordgrass restoration? Master’s thesis, San Francisco State University.Google Scholar
Trocki, C. L., and Paton, P. W. C. 2006. Assessing habitat selection by foraging egrets in salt marshes at multiple spatial scales. Wetlands, 26:307312.Google Scholar
Tyrrell, M., Dionne, M. and Edgerly, J. 2008. Physical factors mediate effects of grazing by a non-indigenous snail species on saltmarsh cordgrass (Spartina alterniflora) in New England marshes. ICES Journal of Marine Science, 65:746752.Google Scholar
Uddin, N., Robinson, R. W., Buultjen, A., Al Haruna, A. U., and Shampa, S. H. 2017. Role of allelopathy of Phragmites australis in its invasion processes, Journal of Experimental Marine Biology and Ecology, 486: 237244.Google Scholar
Wan, S., Qin, P., Liu, J., and Zhou, H. 2009. The positive and negative effects of exotic Spartina alterniflora in China. Ecological Engineering, 35:444452.Google Scholar
Weber, W. A. 1989. Additions to the flora of Colorado. Phytologia, 67:429437.Google Scholar
Wilson, C. A., Hughes, Z. J., and FitzGerald, D. M. 2012. The effects of crab bioturbation on Mid-Atlantic saltmarsh tidal creek extension: geotechnical and geochemical changes. Estuarine, Coastal and Shelf Science, 106: 3344.Google Scholar
Windham, L., and Lathrop, V. 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
Windham, L., and Meyerson, L. A.. 2003. Effects of common reed (Phragmites australis) expansions on nitrogen dynamics of tidal marshes of the northeastern U.S. Estuaries, 26: 452464.Google Scholar
Young, J. A., Palmquist, D. E., and Wotring, S. O. 1997. The invasive nature of Lepidium latifolium: A review, pp. 5968. In: Brock, J. A., Wade, M. Pysek, P., Green, D. (eds.). Plant Invasions: Studies from North America and Europe. Backhuys Publishers, Leiden, The Netherlands.Google Scholar
Young, J. A., Palmquist, D. E., and Blank, R. R. 1998. The ecology and control of perennial pepperweed. Weed Technology, 12(2): 402405.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

Allen, J. R. L. 1989. Evolution of salt-marsh cliffs in muddy and sandy systems: a qualitative comparison of British west-coast estuaries. Earth Surface Processes and Landforms 14(1): 8592.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(12):11551231.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(6): 14021410.Google Scholar
Amos, C. L., Bergamasco, A., Umgiesser, G., Cappucci, S., Cloutier, D., DeNat, L., Flindt, 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
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(2): 169193.Google Scholar
Barbier, E. B., Georgiou, I. Y., Enchelmeyer, I. Y., and Reed, D. J. 2013. The value of wetlands in protecting Southeast Louisiana from hurricane storm surges. PLoS ONE 8: e58715. https://doi.org/10.1371/journal.pone.0058715.Google Scholar
Beland, M., Biggs, T. W., Roberts, D. A., Peterson, S. H., Kokaly, R. F., and Piazza, S. 2017. Oiling accelerates loss of salt marshes, southeastern Louisiana, PLoS ONE 12(8): e0181197.Google Scholar
Bell, F. W. 1997. The economic valuation of saltwater marsh supporting marine recreational fishing in the southeastern United States. Ecological Economics 21: 243254.Google Scholar
Bendoni, M. 2015. Salt marsh edge erosion due to wind-induced waves. PhD Thesis. University of Florence-TU Braunschweig.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., Solari, L., 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
Bilkovic, D., Mitchell, M., Davis, J., Andrews, E., King, A., Mason, P., Herman, J., Tahvildari, N., and Davis, J. 2017. Review of boat wake wave impacts on shoreline erosion and potential solutions for the Chesapeake Bay. STAC Publication Number 17-002, Edgewater, MD.Google Scholar
Boesch, D. F., and Turner, R. E., 1984. Dependence of fishery species on salt marshes: The role of food and refugeEstuaries 7: 460468.Google Scholar
Booij, N., Ris, R. C., and Holthuijsen, L. H. 1999. A third-generation wave model for coastal regions: 1. Model description and validation. Journal of Geophysical Research: Oceans 104 (C4): 76497666.Google Scholar
Breugem, W. A., and Holthuijsen, L., 2007. Generalized shallow water wave growth from Lake George. Journal of Waterway Port Coastal and Ocean Engineering 133: 2337.Google Scholar
Bromberg, K. D., and Bertness, M. D. 2005. Reconstructing New England salt marsh losses using historical maps. Estuaries 28: 823832.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 (1): 7388.Google Scholar
Chauhan, P. P. S. 2009. Autocyclic erosion in tidal marshes. Geomorphology 110(3): 4557.Google Scholar
Chen, Y., Collins, M. B., and Thompson, C. E. L. 2011. Creek enlargement in a low-energy degrading saltmarsh in southern England. Earth Surface Processes and Landforms 36: 767778.Google Scholar
Chen, Y., Thompson, C. E. L., and Collins, M. B. 2012. Saltmarsh creek bank stability: Biostabilisation and consolidation with depth. Continental Shelf Research 35: 6474.Google Scholar
Chmura, G. L., Anisfeld, S. C., Cahoon, D. R., and Lynch, J. C. 2003. Global carbon sequestration in tidal saline wetland soils. Global Biogeochemical Cycles 17:1111.Google Scholar
Cola, S., Sanavia, L., Simonini, P., and Schrefler, B. A. 2008. Coupled thermohydromechanical analysis of Venice lagoon salt marshes. Water Resources Research 44(5): W00C05. https://doi.org/10.1029/2007WR006570Google Scholar
Coops, H., Van der Brink, F.W.B., and Van der Velde, G.. 1996. Growth and morphological responses of four helophyte species in an experimental water-depth gradient. Aquatic Botany 54: 1124.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
Costanza, R., Pérez-Maqueo, O., Martinez, M. L., Sutton, P., Anderson, S. J., and Mulder, K. 2008. The value of coastal wetlands for hurricane protection. Ambio 37: 241248.Google Scholar
Coverdale, C. T., Bertness, M. D., and Altieri, A. H. 2013. Regional ontogeny of New England salt marsh die-off. Conservation Biology 27(5): 10411048.Google Scholar
Davidson, T. M., and de Rivera, C. E. 2010. Accelerated erosion of saltmarshes infested by a non-native burrowing crustacean Sphaeroma quoianum. Marine Ecology Progress Series 419: 129136.Google Scholar
D’Alpaos, A., Lanzoni, S., Marani, M., and Rinaldo, A. 2007. Landscape evolution in tidal embayments: Modeling the interplay of erosion, sedimentation, and vegetation dynamics. Journal of Geophysical Research: Earth Surface 112 (F1): https://doi.org/10.1029/2006JF000537.Google Scholar
Day, J. W., Scarton, F., Rismondo, A., and Are, D. 1998. Rapid deterioration of a salt marsh in Venice Lagoon, Italy. Journal of Coastal Research 14: 583590.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
Emerton, L., and Kekulandala, L. 2003. Assessment of the economic value of Muthurajawela wetland. Occasional Papers of IUCN Sri Lanka, IUCN-World Conservation Union, Sri Lanka Country Office, Colombo (Sri Lanka), Volume 4.Google Scholar
Fagherazzi, S. and Wiberg, P. L. 2009. Importance of wind conditions, fetch, and water levels on wave-generated shear stresses in shallow intertidal basins. Journal of Geophysical Research, Vol. 114, F03022: DOI: 10.1029/2008JF001139.Google Scholar
Fagherazzi, S., Kirwan, M. L., Mudd, S. M., Guntenspergen, G. R., Temmerman, S., D’Alpaos, A., Koppel, J., et al. 2012. Numerical models of salt marsh evolution: Ecological, geomorphic, and climatic factors. Reviews of Geophysics 50 (1): https://doi.org/10.1029/2011RG000359.Google Scholar
Feagin, R. A., Irish, J. L., Möller, I., Williams, A. M., Colón-Rivera, R. J., and Mousavi, M. E. 2009. Does vegetation prevent wave erosion of salt marsh edges? Proceedings of the National Academy of Sciences of the USA, 106(25): 1010910113.Google Scholar
FitzGerald, D., Hughes, Z., and Rosen, P. 2011. Boat wake impact and their role in shore erosion processes, Boston Harbor Islands National Recreation Area, Natural Resource Report NPS/NERO/NRR-2011/403, National Park Service, Fort Collins, Colorado.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
Fredsoe, J., and Deigaard, R. 1993. Mechanics of Coastal Sediment Transport. Advanced Series on Ocean Engingeering, vol. 3, World Science, Singapore.Google Scholar
French, G. T. 1990. Historical shoreline changes in response to environmental conditions in west Delaware Bay. MA thesis. University of Maryland College Park.Google Scholar
Gabet, E. J. 1998. Lateral migration and bank erosion in a saltmarsh tidal channel in San Francisco Bay, California. Estuaries 21 (4): 745753.Google Scholar
Gazetas, G. 1991. Foundation vibrations. Ed. by Fang, H. I. Foundation Engineering Handbook. . Kluwer Academic Publishers, Massachusetts, USA, pp. 553593.Google Scholar
Gedan, K. B., Silliman, B. R., and Bertness, M. D. 2009. Centuries of humandriven change in salt marsh ecosystems. Annual Review of Marine Science 1: 117141.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 (1): 729.Google Scholar
Gosselink, J. G. and Pope, R. M. 1974. The Value of The Tidal Marsh. (LSUSG-74-03, Center for Wetland Resources, Baton Rouge: Louisiana StateUniversity.Google Scholar
Gray, D. H., and Leiser, A. T.. 1982. Biotechnical slope protection and erosion control. Van Nostrand Reinhold Company Inc, New York.Google Scholar
Houser, C. 2010. Relative importance of vessel-generated and wind waves to salt marsh erosion in a restricted fetch environment. Journal of Coastal Research: 230–240.Google Scholar
Howes, N. C., FitzGerald, D. M., Hughes, Z. J., Georgiou, I. Y., Kulp, M. A., Miner, M. D., Smith, J. M., and Barras, J. A. 2010. Hurricane-induced failure of low salinity wetlands. Proceedings of the National Academy of Sciences of the USA 107 (32): 1401414019.Google Scholar
Huges, Z. J., FitzGerald, D. M., Howes, N. C., and Rosen, P. S. 2007. The impact of natural waves and ferry wakes on bluff erosion and beach morphology, Boston Harbor, USA. Journal of Coastal Research, SI 50 (Proceedings of the 9th International Coastal Symposium), 497–501. Gold Coast, Australia, ISSN 0749.0208.Google Scholar
Hughes, Z. J., FitzGerald, D. M., Wilson, C. A., Pennings, S. C., Wieski, K., and Mahadevan, A. 2009. Rapid headward erosion of marsh creeks in response to relative sea level rise. Geophysical Research Letters 36, L03602, doi:10.1029/2008GL036000.Google Scholar
Jadhav, R. S., and Chen, Q. 2013. Probability distribution of wave heights attenuated by salt marsh vegetation during tropical cyclone. Coastal Engineering 82: 4755.Google Scholar
Jadhav, R. S., Chen, Q., and Smith, J. M. 2013. Spectral distribution of wave energy dissipation by salt marsh vegetation. Coastal Engineering 77: 99107.Google Scholar
Karimpour, A., Chen, Q., and Twilley, R. R. 2016. A field study of how wind waves and currents may contribute to the deterioration of saltmarsh fringe. Estuaries and Coasts 39: 935950.Google Scholar
Kennish, M. J. 2001. Coastal salt marsh system in the U.S.: A review of anthropogenic impacts. Journal of Coastal Research 17(3): 731748.Google Scholar
Keulegan, G. H. and Carpenter, L. H. 1958. Forces on cylinders and plates in an oscillating fluid. Journal of Research of the National Bureau of Standards 60 (5): 423440.Google Scholar
King, S. E., and Lester, J. N. 1995. The value of salt marsh as a sea defence. Marine Pollution Bulletin 30 (3):180189.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(23): https://doi.org/10.1029/2010GL045489Google Scholar
Kirwan, M. L., and Murray, A. B. 2007. A coupled geomorphic and ecological model of tidal marsh evolution, Proceedings of the National Academy of Science of the USA, 104, 61186122.Google Scholar
Kirwan, M. L., and Murray, A. B. 2008. Tidal marshes as disequilibrium landscapes? Lags between morphology and Holocene sea level change. Geophysical Research Letters, 35, L24401, doi:10.1029/%202008GL036050.Google Scholar
Kirwan, M. and Temmerman, S. 2009. Coastal marsh response to historical and future sea-level acceleration. Quaternary Science Reviews 28(17): 18011808.Google Scholar
Kobayashi, N., Raichle, A. W. and Asano, T. 1993. Wave attenuation by vegetation. Journal of Waterway Port Coastal and Ocean Engineering 119(1): 3048.Google Scholar
Knutson, T. R., McBride, J. L., Chan, J., Emanuel, K., Holland, G., Landsea, C., Held, I., Kossin, J. P., Srivastava, A. K., and Sugi, M. 2010. Tropical cyclones and climate change, Nature Geoscience 3: 157163.Google Scholar
Le Hir, P., Monbet, Y., and Orvain, F., 2007. Sediment erodability in sediment transport modeling: Can we account for biota effects? Continental Shelf Research 27: 11161142.Google Scholar
Leonardi, N. and Fagherazzi, S. 2014. How waves shape salt marshes. Geology, 42 (10): 887890.Google Scholar
Leonardi, N. and Fagherazzi, S. 2015. Effect of local variability in erosional resistance on large-scale morphodynamic response of salt marshes to wind waves and extreme events. Geophysical Research Letters 42: 58725879, doi:10.1002/2015GL064730.Google Scholar
Leonardi, N., Ganju, N. K. and Fagherazzi, S. 2015. 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 (1): 564568.Google Scholar
Leonardi, N., Defne, Z., Ganju, N. K., and Fagherazzi, S. 2016. Salt marsh erosion rates and boundary features in a shallow Bay. Journal of Geophysical Research Earth Surface, 121: 18611875.Google Scholar
Madsen, P. A., and Sørensen, O. R. 1992. A new form of the Boussinesq equations with improved linear dispersion characteristics. Part 2: A slowly varying bathymetry. Coastal Engineering, 18, 183204.Google Scholar
Malkin, A. Y. and Isayev, A. Y. 2006. Rheological Concepts, Methods and Applications. ChemTech, Toronto.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 (21): https://doi.org/10.1029/2011GL048995.Google Scholar
Mariotti, G. and Fagherazzi, S. 2010. A numerical model for the coupled longterm evolution of salt marshes and tidal flats. Journal of Geophysical Research: Earth Surface 115 (F1): https://doi.org/10.1029/2009JF001326.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 (14): 53535356.Google Scholar
Mariotti, G. and Fagherazzi, S. 2013. Wind waves on a mudflat: The influence of fetch and depth on bed shear stresses. Continental Shelf Research, 60, S99S110.Google Scholar
Mariotti, G., Fagherazzi, S., Wiberg, P. L., McGlathery, K. J., Carinello, L., and Defina, A. 2010. Influence of storm surges and sea level on shallow tidal basin erosive proecesses. Journal of Geophysical Research, 115, C11012. https://doi.org/10.1029/2009JC005892.Google Scholar
Maurmeyer, E. M. 1978. Geomorphology and evolution of transgressive estuarine washover barrier along the western shore of Delaware Bay. PhD thesis. University of Delaware, Newark.Google Scholar
Maynord, S. 2001. Boat waves on Johnson Lake and Kenai River, Alaska. Technical Report U.S. Army Corps of Engineers. (No. ERDC/CHL-TR-01-31.Google Scholar
McLoughlin, S. M., Wiberg, P. L., Safak, I., and McGlathery, K. J. 2014. Rates and forcing of marsh edge erosion in a shallow coastal bay. Estuaries and Coasts, 38 : 620638.Google Scholar
Mendez, F. J., and Losada, I. J. 2004. An empirical model to estimate the propagation of random breaking and nonbreaking waves over vegetation fields. Coastal Engineering, 51 (2): 103118.Google Scholar
Minkoff, D. R., Escapa, M., Ferramola, F. E., Maraschin, S. D., Pierini, J. O., Perillo, G. M. E., and Delrieux, C. 2006. Effects of crabe-halophytic plant interactions on creek growth in a S. W. Atlantic salt marsh: A Cellular Automata model. Estuarine Coastal Shelf Science, 69, 403413.Google Scholar
Mirtskhoulava, T. E. 1991. Scouring by flowing water of cohesive and non cohesive beds. Journal of Hydraulic Research, 29 (3):341354.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 (10):727731.Google Scholar
Möller, I. and Spencer, T. 2002. Wave dissipation over macro-tidal salt marshes: Effects of marsh edge typology and vegetation change. Journal of Coastal Research, 36(1):506521.Google Scholar
Morris, P. H., Graham, J., and Williams, D. J. 1992. Cracking in drying soils. Canadian Geotechnical Journal, 29 (2): 263277.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 (10): 28692877.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
Ojea, E., Martin-Ortega, J., and Chiabai, A. 2012. Defining and classifying ecosystem services for economic valuation: the case of forest water services. Environmental Science and Policy, 19–20: 115.Google Scholar
Ozeren, Y., Wren, D. G., and Wu, W. 2013. Experimental investigation of wave attenuation through model and live vegetation. Journal of Waterway, Port, Coastal, and Ocean Engineering 140 (5): https://doi.org/10.1061/(ASCE)WW.1943-5460.0000251Google Scholar
Phillips, J. D. 1985. Aspat Bay analysis of the shoreline erosion, Delaware Bay, New Jersey. PhD thesis. Rutgers University, New Brunswick.Google Scholar
Price, F. 2006. Quantification, Analysis, and Management of Intracoastal Waterway Channel Margin Erosion in the Guana Tolomato Matanzas National Estuarine Research Reserve, Florida. National Estuarine Research Reserve Technical Report Series 2006:1.Google Scholar
Priestas, A. M. and Fagherazzi, S. 2011. Morphology and hydrodynamics of wave-cut gullies. Geomorphology 131 (1): 113.Google Scholar
Riffe, K. C., Henderson, S. M., and Mullarney, J. C. 2011. Wave dissipation by flexible vegetation. Geophysical Research Letters 38 (18): https://doi.org/10.1029/2011GL048773.Google Scholar
Schwimmer, R. A. 2001. Rates and processes of marsh shoreline erosion in Rehoboth Bay, Delaware, USA. Journal of Coastal Research 17 (3): 672683.Google Scholar
Schwimmer, R. A. and Pizzuto, J. E. 2000. A model for the evolution of marsh shorelines. Journal of Sedimentary Research 70 (5): 10261035.Google Scholar
Selby, M. J. 1993. Hillslope Materials and Processes. Oxford University Press.Google Scholar
Silliman, B. R., Van de Koppel, J., McCoy, M. W., Diller, J., Kasozi, G. N., Earl, K., Adams, P. N., and Zimmerman, A. R. 2012. Degradation and resilience in Louisiana salt marshes after the BP–Deepwater Horizon oil spill. Proceedings of the National Academy of Sciences of the USA, 109 (28): 1123411239.Google Scholar
Silinski, A., Heuner, M., Schoelynck, J., Puijalon, S., Schroder, U., Fuchs, E., Troch, P., Bouma, T. J., Meire, P., and Temmerman, S. 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(3): e0118687.Google Scholar
Smith, S. M. 2009. Multi-decadal changes in salt marshes of Cape Cod, MA: Photographic analyses of vegetation loss, species shifts, and geomorphic change. Northeastern Naturalist 16(2): 183208.Google Scholar
Smith, S. M. 2015. Vegetation change in salt marshes of Cape Cod national seashore (Massachusetts, USA) between 1984 and 2013. Wetlands, 35(1): 127136.Google Scholar
Smith, J. A. M. 2013. The role of Phragmites australis in mediating inland salt marsh migration in a mid-Atlantic estuary. PLoS ONE 8(5): e65091.doi:10.1371/journal.pone.0065091.Google Scholar
Sorensen, R. M. 1973. Water waves produced by ships. Journal of the Waterways, Harbors and Coastal Engineering Division, 99(2), 245256.Google Scholar
Spencer, T., Moller, I., Rupprecht, F., Bouma, T. J., van Wesenbeeck, B. K., Kudella, M., Paul, M., et al. 2015. Salt marsh surface survives true-to-scale simulated storm surges. Earth Surface Processes and Landforms 41(4): 543552.Google Scholar
Suzuki, T. 2011. Wave dissipation over vegetation fields. PhD thesis. Delft University of Technology, Netherlands.Google Scholar
Swisher, M. 1982. The rates and causes of coastal erosion around a transgressive coastal lagoon, Rehoboth Bay, Delaware. MA thesis. University of Delaware, Newark.Google Scholar
Temmerman, S., Meire, P., Bouma, T., Herman, P. M. J., Ysebaert, T., and  De Vriend, H. J.. 2013. Ecosystem-based coastal defence in the face of global changeNature 5047983Google Scholar
Thorne, C. R. and Tovey, N. K. 1981. Stability of composite river banks. Earth Surface Processes and Landforms 6 (5): 469484.Google Scholar
Tonelli, M., Fagherazzi, S., and Petti, M. 2010. Modeling wave impact on salt marsh boundaries. Journal of Geophysical Research: Oceans 115 (C9). https://doi.org/10.1029/2009JC006026.Google Scholar
Trosclair, K. J. 2013. Wave transformation at a saltmarsh edge and resulting edge erosion: observation and modeling. PhD thesis, University of New Orleans Theses and Dissertations, Paper 1777.Google Scholar
Tuan, Q. T. and Oumeraci, H. 2012. Numerical modelling of wave overtopping-induced erosion of grassed inner sea-dike slopes. Natural Hazards 63 (2): 417447.Google Scholar
Valiela, I. and Teal, J. M. 1979. The nitrogen budget of a salt marsh ecosystem. Nature 280 (5724): 652656.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 ecosytems. The American Naturalist 165 (1): E112.Google Scholar
Van der Meer, J. W., Verheij, H. J., Lindenberg, J., Van Hoven, A., and Hoffmans, G. J. C. M. 2007. Wave overtopping and strenght of inner slopes of dikes. Tech. rep. 05i028. in Dutch. WL|Delft Hydraulics, Geodelft.Google Scholar
Van Der Wal, Z.De Graaf, G., and  Lasthuizen, K. 2008. What’s valued most similarities and differences between the organizational values of the public and private sector? Public Administration 86465482.Google Scholar
Van Eerdt, M. M. 1985. Salt marsh cliff stability in the Oosterschelde. Earth Surface Processes and Landforms 10 (2): 95106.Google Scholar
Van Eerdt, M. M. 1985. The influence of vegetation on erosion and accretion in salt marshes of the Oosterschelde, The Netherlands. Vegetation 62 (1–3): 367373.Google Scholar
Watson, E. B., Oczkowski, A. J., Wigand, C., Hanson, A. R., Dawey, E. W., Crosby, S. C., Johnson, R. L., and Andrews, H. M. 2014. Nutrient enrichment and precipitation changes do not enhance resiliency of salt marshs to sea level rise in the Northeastern U.S. Climatic Change 125: 501509.Google Scholar
Webster, P. J., Holland, G. J., Curry, J. A., and Chang, H. -R 2005. Changes in tropical cyclone number, duration and intensity in a warming environment, Science, 309(5742): 18441846.Google Scholar
Weston, N. B. 2014. Declining sediments and rising seas: an unfortunate convergence for tidal wetlands. Estuaries and Coasts 37: 123.Google Scholar
Wilson, C. A. and Allison, M. A. 2008. An equilibrium profile model for retreating marsh shorelines in southeast Louisiana. Estuarine, Coastal and Shelf Science 80 (4): 483494.Google Scholar
Winterwerp, J. C. and Van Kesteren, W. G. M. 2004. Introduction to the physics of cohesive sediment dynamics in the marine environment. Ed. by Van Loon, T.. Vol. 56. Developments in Sedimentology. Elsevier, Amsterdam, the Netherlands.Google Scholar
Winterwerp, J. C., Kesteren, W. G. M., Prooijen, B. C., and Jacobs, W. 2012. A conceptual framework for shear flow–induced erosion of soft cohesive sediment beds. Journal of Geophysical Research: Oceans 117 (C10): https://doi.org/10.1029/2012JC008072.Google Scholar
Young, I. R. and Verhagen, L. A. 1996. The growth of fetch limited waves in water of finite depth. Part 1. Total energy and peak frequency. Coastal Engineering 29 (1):4778.Google Scholar

References

Abraham, K. F., Jefferies, R. L., and Alisauskas, R. T. 2005. The dynamics of landscape change and snow geese in mid‐continent North America. Global Change Biology, 11(6): 841855.Google Scholar
Abson, D. J., von Wehrden, H., Baumgärtner, S., Fischer, J., Hanspach, J., Härdtle, W., Heinrichs, H. et al. 2014. Ecosystem services as a boundary object for sustainability. Ecological Economics, 103: 2937.Google Scholar
Arriaga Cabrera, L., Vázquez Domínguez, E., González Cano, J., Jiménez Rosenberg, R., Muñoz López, E., and Aguilar Sierra, V. (coordinators). 1998. Regiones marinas prioritarias de México. Comisión Nacional para el Conocimiento y uso de la Biodiversidad. México.Google Scholar
Arkema, K. K., and Samhouri, J. F. 2012. Linking ecosystem health and services to inform marine ecosystem-based management. American Fisheries Society Symposium, 79: 925.Google Scholar
Barbier, E. B. 2012. Progress and challenges in valuing coastal and marine ecosystem services. Review of Environmental Economics and Policy, 6: 119.Google Scholar
Bernatchez, P., and Fraser, C. 2012. Evolution of coastal defence structures and consequences for beach width trends, Québec, Canada. Journal of Coastal Research, 28(6): 15501566.Google Scholar
Borchert, S. M., Osland, M. J., Enwright, N. M., and Griffith, K. T. 2018. Coastal wetland adaptation to sea level rise: Quantifying potential for landward migration and coastal squeeze. Journal of Applied Ecology, 55: 28762887.Google Scholar
Brinson, M. M., Christian, R. R., and Blum, L. K. 1995. Multiple states in the sea-level induced transition from terrestrial forest to estuary. Estuaries, 18(4): 648659.Google Scholar
Brown de Colstoun, E. C., Huang, C., Wang, P., Tilton, J. C., Tan, B., Phillips, J., Niemczura, S., Ling, P.-Y., and Wolfe, R. E. 2017. Global Man-made Impervious Surface (GMIS) Dataset from Landsat. Palisades, NY: NASA Socioeconomic Data and Applications Center (SEDAC). https://doi.org/10.7927/H4P55KKF. Accessed 01 01 2017.Google Scholar
Burrough, P. A., and McDonnel, R. A. 1999. Principles of Geographical Information Systems. Oxford University Press.Google Scholar
CBD. 2015. Convention on Biological Diversity. TARGET 11 - Technical Rationale extended (provided in document COP/10/INF/12/Rev.1). www.cbd.int/sp/targets/rationale/target-11/.Google Scholar
Commission for the Environment, 2020. North American Environmental Atlas. www.cec.org/tools-and-resources/north-american-environmental-atlas/map-filesGoogle Scholar
CEC. 2015a. Commission for Environmental Cooperation. North American Protected Areas Network. www2.cec.org/nampan/.Google Scholar
CEC. 2015b. Commission for Environmental Cooperation. North America's Blue Carbon: Assessing the role of coastal habitats in the Continent's carbon budget. www.cec.org/bluecarbon.Google Scholar
CGIAR. 2015. Consultative Group for International Agricultural Research-Consortium of Spatial Information. Digital Elevation Model. www.cgiar-csi.org/data.Google Scholar
Chmura, G. L., Burdick, D., and Moore, G. 2012. Recovering salt marsh ecosystem services through tidal restoration, In: Tidal Marsh Restoration, ed. Roman, C.T. and Burdick, D.M., 233251. Washington, DC: Island Press.Google Scholar
Church, J. A., Clark, P. U., Cazenave, A., Gregory, J. M., Jevrejeva, S., Levermann, A., M. A.Merrifield, A., et al. 2013. Sea level change. In: Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Stocker, T.F., Qin, D., Plattner, G.-K., Tignor, M., Allen, S.K., Boschung, J., Nauels, A., Xia, Y., Bex, V. and Midgley, P.M. (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA.Google Scholar
Cooper, J. K., Li, J., and Montagnes, D. J. 2012. Intermediate fragmentation per se provides stable predator‐prey metapopulation dynamics. Ecology Letters, 15: 856863.Google Scholar
Costanza, R., dArge, R., deGroot, 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
Crain, C. M., Kroeker, K., and Halpern, B. S. 2008. Interactive and cumulative effects of multiple human stressors in marine systems. Ecology Letters, 11: 13041315.Google Scholar
Crooks, S. 2004. The effect of sea‐level rise on coastal geomorphology. Ibis, 146: 1820.Google Scholar
Doody, P. J. 2004. “Coastal Squeeze”: An historical perspective. Journal of Coastal Conservation, 10: 129138.Google Scholar
Elvidge, C. D., Tuttle, B. T., Sutton, P. C., Baugh, K. E., Howard, A. T., Milesi, C., Bhaduri, B. and Nemani, R. 2007. Global distribution and density of constructed impervious surfaces. Sensors, 7: 19621979.Google Scholar
Enwright, N. M., Griffith, K. T., and Osland, M. J. 2016. Barriers to and opportunities for landward migration of coastal wetlands with sea‐level rise. Frontiers in Ecology and the Environment, 14(6): 307316.Google Scholar
ESRI. 2015. Environmental Systems Research Institute. How slope works. http://resources.arcgis.com/en/help/main/10.2/index.html#//009z000000vz000000.Google Scholar
Erwin, K. L. 2009. Wetlands and global climate change: The role of wetland restoration in a changing world. Wetlands Ecology and Management, 17: 7184.Google Scholar
Ewing, L. C. 2015. Resilience from coastal protection. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 373(2053): 20140383.Google Scholar
Feagin, R. A., Martínez, 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
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
FitzGerald, D. M., and Hughes, Z. 2019. Marsh processes and their response to climate change and sea-level rise. Annual Review of Earth and Planetary Sciences, 47: 481517.Google Scholar
Game, E. T., Grantham, H. S., Hobday, A. J., Pressey, R. L., Lombard, A. T., Beckley, L. E., Gjerde, K., Bustamante, R., Possingham, H. P., and Richardson, A. J. 2009. Pelagic protected areas: The missing dimension in ocean conservation. Trends in Ecology and Evolution, 24: 360369.Google Scholar
Guerry, A. D., Polasky, S., Lubchenco, J., Chaplin-Kramer, R., Daily, G. C., Griffin, R., and Ruckelshaus, M., et al. 2015. Natural capital and ecosystem services informing decisions: From promise to practice. Proceedings of the National Academy of Sciences 112:73487355Google Scholar
Hitch, A. T., Purcell, K. M., Martin, S. B., Klerks, P. L., and Leberg, P. L. 2011. Interactions of salinity, marsh fragmentation and submerged aquatic vegetation on resident nekton assemblages of coastal marsh ponds. Estuaries and Coasts, 34: 653662.Google Scholar
Hobday, A. J., Hartog, J. R., Timmiss, T. and Fielding, J. 2010. Dynamic spatial zoning to manage southern bluefin tuna (Thunnus maccoyii) capture in a multi-species longline fishery. Fisheries Oceanography, 19: 243253.Google Scholar
Hughes, Z. J., FitzGerald, D. M., Wilson, C. A., Pennings, S. C., Więski, K., and Mahadevan, A. 2009. Rapid headward erosion of marsh creeks in response to relative sea level rise. Geophysical Research Letters, 36(3): https://doi.org/10.1029/2008GL036000.Google Scholar
Hussein, A. H. 2009. Modeling of sea-level rise and deforestation in submerging coastal ultisols of Chesapeake Bay. Soil Science Society of America Journal 73(1): 185196.Google Scholar
Hyrenbach, K., Keiper, C., Allen, S., Ainley, D., and Anderson, D. 2006. Use of marine sanctuaries by far‐ranging predators: Commuting flights to the California Current System by breeding Hawaiian albatrosses. Fisheries Oceanography, 15: 95103.Google Scholar
Jarvis, A., Reuter, H. I., Nelson, A., and Guevara, E. 2008. Hole-filled SRTM for the globe version 4. CGIAR-CSI SRTM 90m database: http://srtm.csi.cgiar.org.Google Scholar
Jolicoeur, S., and O’Carroll, S. 2007. Sandy barriers, climate change and long-term planning of strategic coastal infrastructures, Îles-de-la-Madeleine, Gulf of St. Lawrence (Québec, Canada). Landscape and Urban Planning, 81(4): 287298.Google Scholar
Jones, S., Bosch, A. C., and Strange, E. 2009. Vulnerable species: the effects of sea-level rise on coastal habitats. In Coastal Sensitivity to Sea-Level Rise: A Focus on the Mid-Atlantic Region. A report by the U.S. Climate Change Science Program and the Subcommittee on Global Change Research., ed. Titus, J. G., Anderson, K. E., Cahoon, D. R., Gesch, D. B., Gill, S. K., Gutierrez, B. T., Thieler, E. R. and Williams, S. J., pp. 73–84: US Environmental Protection Agency, Washington, DC.Google Scholar
JPL. 2015. Jet Propulsion Laboratory. Shuttle Radar Topography Mission. www2.jpl.nasa.gov/srtm/statistics.html: National Aeronautics and Space Administration.Google Scholar
Kelley, J. T., Gehrels, W. R., and Belknap, D. F. 1995. Late Holocene relative sea-level rise and the geological development of tidal marshes at Wells, Maine, USA. Journal of Coastal Research, 11: 136153.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
Lambeck, K., and Bard, E. 2000. Sea-level change along the French Mediterranean coast for the past 30 000 years. Earth and Planetary Science Letters, 175(3–4): 203222.Google Scholar
Lewison, R., Hobday, A. J., Maxwell, S., Hazen, E., Hartog, J. R., Dunn, D. C., Briscoe, D., et al. 2015. Dynamic ocean management: Identifying the critical ingredients of dynamic approaches to ocean resource management. BioScience, 65: 486498.Google Scholar
Martínez, M. L., Mendoza-González, G., Silva-Casarín, R., and Mendoza-Baldwin, E. 2014. Land use changes and sea level rise may induce a “coastal squeeze” on the coasts of Veracruz, Mexico. Global Environmental Change, 29: 180188.Google Scholar
McGranahan, G., Balk, D., and Anderson, B. 2007. The rising tide: Assessing the risks of climate change and human settlements in low elevation coastal zones. Environment and Urbanization, 19: 1737.Google Scholar
McLeod, E., Chmura, G. L., Bouillon, S., Salm, R., Björk, M., Duarte, C. M., and Silliman, B.R. 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(10): 552560.Google Scholar
MEA. 2005. Millennium Ecosystem Assessment. Ecosystems and Human Well-being: Synthesis: Island Press Washington, DC.Google Scholar
NOAA. 2010. National Oceanic and Atmospheric Administration-National Center for Environmental Information. Global Distribution and Density of Constructed Impervious Surfaces. http://ngdc.noaa.gov/eog/dmsp/download_global_isa.html: NOAA.Google Scholar
Nordlie, F. G. 2003. Fish communities of estuarine salt marshes of eastern North America, and comparisons with temperate estuaries of other continents. Reviews in Fish Biology and Fisheries, 13(3): 281325.Google Scholar
Pendleton, E. A., Thieler, E. R., and Williams, S. J. 2010. Importance of coastal change variables in determining vulnerability to sea-and lake-level change. Journal of Coastal Research, 26: 176183.Google Scholar
Pontee, N. 2013. Defining coastal squeeze: A discussion. Ocean and Coastal Management, 84: 204207Google Scholar
Raabe, E. A., and Stumpf, R. P. 2016. Expansion of tidal marsh in response to sea-level rise: Gulf Coast of Florida, USA. Estuaries and Coasts, 39(1): 145157.Google Scholar
Redfield, A. C. 1972. Development of a New England salt marsh. Ecological Monographs, 42(2): 201237.Google Scholar
Salgado, K., and Martínez, M. L. 2017. Is ecosystem-based coastal defense a realistic alternative? Exploring the evidence. Journal of Coastal Conservation, 21(6): 837848.Google Scholar
Schleupner, C. 2008. Evaluation of coastal squeeze and its consequences for the Caribbean island Martinique. Ocean and Coastal Management, 51: 383390.Google Scholar
Sutton-Grier, A. E., Wowk, K., and Bamford, H. 2015. Future of our coasts: The potential for natural and hybrid infrastructure to enhance the resilience of our coastal communities, economies and ecosystems. Environmental Science and Policy, 51: 137148.Google Scholar
Sweet, W. V., Kopp, R. E., Weaver, C. P., Obeysekera, J., Horton, R. M., Thieler, E. R., and Zervas, C.. 2017. Global and regional sea level rise scenarios for the United States. NOAA Technical Report NOS CO-OPS 083. NOAA/NOS Center for Operational Oceanographic Products and ServicesGoogle Scholar
Thorne, K., MacDonald, G., Guntenspergen, G., Ambrose, R., Buffington, K., Dugger, B., Freeman, C., et al. 2018. US Pacific coastal wetland resilience and vulnerability to sea-level rise. Science Advances, 4(2): 3270.Google Scholar
Titus, J. G., and Neumann, J. E.. 2009. Implications for decisions. In coastal sensitivity to sea-level rise: A focus on the mid-Atlantic region. A report by the U.S. Climate Change Science Program and the Subcommittee on Global Change Research., ed. Titus, J. G., Anderson, K. E., Cahoon, D. R., Gesch, D. B., Gill, S. K., Gutierrez, B. T., Thieler, E. R. and Williams, S. J., pp. 141–156: US Environmental Protection Agency, Washington DC, USA.Google Scholar
Tomaselli, V., Tenerelli, P., and Sciandrello, S. 2012. Mapping and quantifying habitat fragmentation in small coastal areas: A case study of three protected wetlands in Apulia (Italy). Environmental Monitoring and Assessment, 184: 693713.Google Scholar
Torio, D. D., and Chmura, G. L. 2013. Assessing coastal squeeze of tidal wetlands. Journal of Coastal Research, 29: 233243.Google Scholar
Torio, D. D., and Chmura, G. L. 2015. Impacts of sea level rise on marsh as fish habitat. Estuaries and Coasts, 38: 12881303.Google Scholar
UNFCCC. 2015. United Nation Framework Convention on Climate Change. NAMAs, Nationally Appropriate Mitigation Actions. http://unfccc.int/focus/mitigation/items/7172.php.Google Scholar
White, E. P., and Brown, J. H. 2005. The template: patterns and processes of spatial variation. In Ecosystem Function in Heterogeneous Landscapes, ed. Lovett, G. M., Jones, C. G., Turner, M. G., and Weathers, K. C., pp. 3147: Springer, New York, USA.Google Scholar
White, E. and Kaplan, D. 2017. Restore or retreat? Saltwater intrusion and water management in coastal wetlands. Ecosystem Health and Sustainability, 3(1): e01258.Google Scholar
Zacheis, A., Hupp, J. W., and Ruess, R. W. 2001. Effects of migratory geese on plant communities of an Alaskan salt marsh. Journal of Ecology, 89(1): 5771.Google Scholar

References

Adamowicz, S. C. and O’Brien, K. M.. 2012. Drakes Island tidal restoration, science, community and compromise. In Roman, C. T. and Burdick, D. M. (eds.) Tidal Marsh Restoration: A Synthesis of Science and Practice. pp. 315332. Washington, DC: Island Press.Google Scholar
Allison, M. A., and Meselhe, E. A. 2010. The use of large water and sediment diversions in the lower Mississippi River (Louisiana) for coastal restoration. Journal of Hydrology, 387: 346e360.Google Scholar
Allison, M. A., Demas, C. R., Ebersole, B. A., Kleiss, B. A., Little, C. D., Meselhe, E. A., Powell, N. J., Pratt, T. C., and Vosburg, B. M. 2012. A water and sediment budget for the lower Mississippi–Atchafalaya River in flood years 2008–2010: Implications for sediment discharge to the oceans and coastal restoration in Louisiana. Journal of Hydrology, 432: 8497.Google Scholar
Barnes, S., Bond, C., Burger, N., Anania, K., Strong, A., Weilant, S., and Virgets, S. 2015. Economic evaluation of coastal land loss in Louisiana. Lousiana State University and the Rand Corporation. Published online. http://coastal.la.gov/economic-evaluation-of-land-loss-in-louisiana/Google Scholar
Batker, D., de la Torre, I., Costanza, R., Day, J., Swedeen, P., Boumans, R., and Bagstad, K.. 2014. The threats to the values of ecosystem goods and services of the Mississippi Delta. In Day, J., Kemp, P., Freeman, A., and Muth, D. (eds.) Perspectives on the Restoration of the Mississippi Delta. pp. 155174. Springer, New York.Google Scholar
Baumann, R., Day, J. W., and Miller, C. 1984. Mississippi deltatic wetland survival: Sedimentation versus coastal submergence. Science, 224: 10931095.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: 488449.Google Scholar
Boesch, D. F., Josselyn, M. N., Mehta, A. J., Morris, J. T., Nuttle, W. K., Simenstad, C. A., and Swift, D. J. Scientific assessment of coastal wetland loss, restoration and management in Louisiana. J. Coast. Res. 1994, i–v, 1103.Google Scholar
Bozek, C. and Burdick, D. M. 2005. Impacts of seawalls on saltmarsh plant communities in the Great Bay Estuary, New Hampshire USA. Wetlands Ecology and Management, 13: 553568.Google Scholar
Bromberg, K. D., and Bertness, M. D. 2005. Reconstructing New England salt marsh losses using historical maps. Estuaries, 28: 823832.Google Scholar
Burdick, D. M., Dionne, M., Boumans, R. M., and Short, F. T. 1997. Ecological responses to tidal restorations of two northern New England salt marshes. Wetlands Ecology and Management, 4: 129144.Google Scholar
Caffey, R. H. and Schexnayder, M. 2002. Fisheries implications of freshwater diversions., in: An Interpretive Topic Series on Louisiana Coastal Wetland Restoration, Coastal Wetland Planning, Preservation, and Restoration Act (eds.), National Sea Grant Library No. LSU-G-02-003.Google Scholar
Cahoon, D. R., Reed, D. J., Day, J. W. Jr., Steyer, G. D., Boumans, R. M., Lynch, J. C., McNally, D., and Latif, N. 1995. The influence of Hurricane Andrew on Sediment Distribution in Louisiana Coastal Marshes. Journal of Coastal Research Special Issue, 21: 280294.Google Scholar
Callaway, J. C., DeLaune, R. D. and Patrick, W. H. Jr. 1997. Sediment accretion rates from four coastal wetlands along the Gulf of Mexico. Journal of Coastal Research, 13: 181191.Google Scholar
Chabreck, R. H, and Palmisano, A. W. 1973. The effects of Hurricane Camille on the marshes of the Mississippi River Delta. Ecology, 54: 11181123.Google Scholar
Clewell, A. F., and Aronson, J. 2013. Ecological Restoration: Principles, Values, and Structures of an Emerging Profession. 2nd ed. Island Press, Washington, DC, USA.Google Scholar
Coastal Protection and Restoration Authority of Louisiana (CPRA). Integrated Ecosystem Restoration and Hurricane Protection: Louisiana’s Comprehensive Master Plan for a Sustainable Coast. Baton Rouge, LA, USA, 2007.Google Scholar
CPRA (Coastal Protection and Restoration Authority of Louisiana). Louisiana’s Coastal Master Plan for a Sustainable Coast. Baton Rouge, LA, USA, 2012.Google Scholar
CPRA (Coastal Protection and Restoration Authority of Louisiana. Louisiana’s Coastal Master Plan for a Sustainable Coast; Coastal Protection and Restoration Authority of Louisiana: Baton Rouge, LA, USA, 2017.Google Scholar
Comín, F. A., Romero, J. A., Hernández, O., and Menéndez, M. 2001. Restoration of wetlands from abandoned rice fields for nutrient removal, and biological community and landscape diversity. Restoration Ecology, 9 (2): 201208.Google Scholar
Condrey, R., Hoffman, P., and Evers, D. 2014. The last naturally active delta complexes of the Mississippi River (LNDM): Discovery and implications. In Day, J., Kemp, P., Freeman, A., and Muth, D. (eds.) Perspectives on the Restoration of the Mississippi Delta. Estuaries of the World. pp. 3350. Springer, New York.Google Scholar
Cornu, C. E., and Sadro, S. 2002. Physical and functional responses to experimental marsh surface elevation manipulation in Coos Bay’s South Slough. Restoration Ecology, 10: 474486.Google Scholar
Couvillion, B. R., Barras, J. A., Steyer, G. D., Sleavin, W. , Fischer, M., Beck, H., Trahan, N., Griffin, B., Heckman, D. 2017. Land Area Change in Coastal Louisiana from 1932 to 2010, U.S. Geological Survey: Reston, VA, USA.Google Scholar
Coverdale, T. C., Altieri, A. H., and Bertness, M. D. 2012. Belowground herbivory increases vulnerability of New England salt marshes to die-off. Ecology, 93: 20852094.Google Scholar
Craig, N. J., Turner, R. E., and Day, J. W. 1979. Land loss in coastal Louisiana (U.S.A.). Environmental Management, 3: 133144.Google Scholar
Craft, C. 2015. Creating and Restoring Wetlands: From Theory to Practice. 1st edn. Elsevier, Inc. The Netherlands.Google Scholar
Culbertson, J. B., Valiela, I., Pickart, M., Peacock, E. E., and Reddym, C. M. 2008. Long-term consequences of residual petroleum on salt marsh grass. Journal of Applied Ecology, 45: 12841292.Google Scholar
Davis, D. W. 1993. Crevasses on the Lower Course of the Mississippi River. Coastal Zone’93, vol. 1, pp. 360378, July 19–23, New Orleans, Louisiana.Google Scholar
Davis, D. W. 2000. Historical perspective on crevasses, levees, and the Mississippi River. In Colten, C. E. (ed.) Transforming New Orleans and its Environs. pp. 84106. Pittsburgh: University of Pittsburgh Press.Google Scholar
Davis, J. L., Currin, C. A., O’Brien, C. Raffenburg, , and Davis, C. A. 2015. Living shorelines: Coastal resilience with a blue carbon benefit. PLoS ONE 10:e0142595. doi:10.1371/journal.pone.0142595Google Scholar
Day, J., Agboola, J., Chen, Z., D’Elia, C., Forbes, D., Giosan, L., Kemp, P., et al. 2016c. Approaches to defining deltaic sustainability in the 21st century. Estuarine, Coastal and Shelf Science, 183: 275291.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
Day, J. W., Britsch, L., Hawes, S., Shaffer, G., Reed, D., and Cahoon, D. 2000. Pattern and process of land loss in the Mississippi Delta: A spatial and temporal analysis of wetland habitat change. Estuaries and Coasts, 23:425438.Google Scholar
Day, J., Cable, J., Lane, R., and Kemp, G. 2016b. Sediment deposition at the Caernarvon crevasse during the great Mississippi flood of 1927: Implications for coastal restoration. Water, 8 (38): doi:10.3390/w8020038.Google Scholar
Day, J., Colten, C., and Kemp, G. P. 2019. Mississippi Delta restoration and protection: shifting baselines, diminishing resilience, and growing non-sustainability. In: Wolanski, E. . Day, J, Elliott, M., and Ramnachandran, R. (eds.) Coasts and Estuaries: The Future. pp. 167186. Elsevier, New York.Google Scholar
Day, J. W. and Erdman, J.. 2018. Sustainable pathways for Mississippi delta restoration – Pathways to a Sustainable Future. Springer, Cham, Switzerland.Google Scholar
Day, J., Kemp, G. P., Freeman, A. M., and Muth, D. (Eds.). 2014. Perspectives on the Restoration of the Mississippi Delta – The Once and Future Delta. Springer, New York.Google Scholar
Day, J. W., Lane, R. R., D’Elia, C. F., Wiegman, A. R. H., Rutherford, J. S., Shaffer, G. P., Brantley, C. G., and Kemp, G. P. 2016a. Large infrequently operated river diversions for Mississippi delta restoration. Estuarine, Coastal and Shelf Science, 183: 292303.Google Scholar
Deegan, L. A., Johnson, D. S., Warren, R. S., Peterson, B. J., Fleeger, J. W., Fagherazzi, S. and Wollheim, W. 2012. Coastal eutrophication as a driver of salt marsh loss. Nature 490: 388392.Google Scholar
DeLaune, R. D., Kongchum, M., White, J. R., and Jugsujinda, A. 2013. Freshwater diversions as an ecosystem management tool for maintaining soil organic matter accretion in coastal marshes. Catena, 107: 139e144.Google Scholar
Edwards, K. R., and Proffitt, C. E. 2003. Comparison of wetland structural characteristics between created and natural salt marshes in southwest Louisiana, USA. Wetlands, 23(2): 344356.Google Scholar
Edwards, K. R., and Mills, K. P. 2005. Aboveground and belowground productivity of Spartina alterniflora (Smooth Cordgrass) in natural and created Louisiana salt marshes. Estuaries, 28: 252265.Google Scholar
Elsey-Quirk, T., and Adamowicz, S. 2016. Influence of physical manipulations on short-term salt marsh morphodynamics: Examples from the north and mid-Atlantic coast, USA. Estuaries and Coasts, 39:423439.Google Scholar
Elsey‐Quirk, T., Middleton, B. A. and Proffitt, C. E. 2009. Seed dispersal and seedling emergence in a created and a natural salt marsh on the Gulf of Mexico coast in Southwest Louisiana, USA. Restoration Ecology, 17(3): 422432.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
Ford, M. A., and Grace, J. B. 1998. Effects of vertebrate herbivores on soil processes, plant biomass, litter accumulation and soil elevation changes in a coastal marsh. Journal of Ecology, 86: 974982.Google Scholar
Forès, E. 1992. Desecación de la laguna de la Encanyissada: un procedimiento para disminuir los niveles de eutrofia. Butlletí del Parc Nat Delta de l’Ebre, 7: 2631.Google Scholar
Forès, E., Espanya, A., and Morales, F. 2002. Regeneración de la laguna costera de La Encanyissada (Delta del Ebro). Una experiencia de biomanipulación. Escosistemas 2002/2.Google Scholar
Frame, G. W., Mellander, M. K., Adamo, D. A.. 2006. Big Egg Marsh experimental restoration in Jamaica Bay, New York. In: Harmon, D. (ed.) People, Places, and Parks: Proceedings of the 2005 George Wright Society Conference on Parks, Protected Areas, and Cultural Sites. pp. 123130. Hancock, Michigan. The George Wright Society.Google Scholar
Friess, D. A., Krauss, K. W., Horstman, E. M., Balke, T., Bouma, T. J., Galli, D. and Webb, E. L. 2012. Are all intertidal wetlands naturally created equal? Bottlenecks, thresholds and knowledge gaps to mangrove and saltmarsh ecosystems. Biological Reviews, 87: 346366.Google Scholar
Gagliano, S. M., Meyer-Arendt, K. J., and Wicker, K. M. 1981. Land loss in the Mississippi River deltaic plain. Gulf Coast Association of Geological Societies Transactions, 31:295300Google Scholar
Genua-Olmedo, A., Alcaraz, C., Caiola, N., and Ibáñez, C. 2016. Sea level rise impacts on rice production: The Ebro Delta as an example. Science of The Total Environment, 571: 12001210.Google Scholar
Giosan, L., Syvitski, J., Constantinescu, S., and Day, J. 2014. Protect the world’s deltas. Nature, 516: 3133.Google Scholar
Gittman, R. K., Popowich, A. M., Bruno, J. F., and Peterson, C. H. 2014. Marshes with and without sills protect estuarine shorelines from erosion better than bulkheads during a category 1 hurricane. Ocean & Coastal Management, 102: 94102.Google Scholar
Gittman, R. K., Peterson, C. H., Currin, C. A., Fodrie, F. J., Piehler, M. F. and Bruno, J. F. 2015. Living shorelines can enhance the nursery role of threatened estuarine habitats. Ecological Applications, 26: 249263.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
Guntenspergen, G. R., Cahoon, D. R., Grace, J., Stayer, G. D., Fournet, S., Townson, M. A., and Foote, A. L. 1995. Disturbance and recovery of Louisiana coastal marsh landscape from the impacts of Hurricane Andrew. Journal of Coastal Research Special Issue, 21: 324339.Google Scholar
Hackney, C. T., Brady, S., Stemmy, L., Boris, M., Dennis, C., Hancock, T., O’Bryon, M., Tilton, C. and Barbee, E. 1996. Does intertidal vegetation indicate specific soil and hydrologic conditions. Wetlands, 16: 8994.Google Scholar
Hatton, R. S., DeLaune, R. D., and Patrick, W. H. Jr 1983. Sedimentation, accretion, and subsidence in marshes of Barataria Basin, Louisiana. Limnology and Oceanography, 28: 494502.Google Scholar
Herrera-Silveira, J., Lara-Domíinguez, A., Day, J., Yáñez-Arancibia, A., Ojeda, S. M., Hernández, C. T., and Kemp, G. P. 2019. Ecosystem functioning and sustainable management in coastal systems with high reshwater input in the Southern Gulf of Mexico and Yucatan Peninsula. In: Wolanski, E., Day, J., Elliott, M., and Ramnachandran, R. (eds.) Coasts and Estuaries: The Future. pp. 377397. Elsevier, New York.Google Scholar
Hazelton, E. L. G., Mozdzer, T. J., Burdick, D. M., Kettenring, K. M., and Whigham, D. F. 2014. Phragmites australis management in the United States: 40 years of methods and outcomes. AoB Plants, 6: plu001.Google Scholar
Hedges, P., Kriwoken, L. K. and Patten, K. 2003. A review of Spartina management in Washington State, US. Journal of Aquatic Plant Management, 41: 8290.Google Scholar
Hinkle, R., and Mitsch, W. J. 2005. Salt marsh vegetation recovery at salt hay farm wetland restoration sites on Delaware Bay. Ecological Engineering, 25: 240251.Google Scholar
Ibáñez, C., Canicio, A., Curcó, A., Day, J. W., and Prat, N. 1996b. Evaluation of vertical accretion and subsidence rates. MEDDELT Final Report, Ebre Delta Plain Working Group. University of Barcelona, Barcelona, Spain.Google Scholar
Ibáñez, C., Canicio, A., and Day, J. W. 1997. Morphologic development, relative sea level rise and sustainable management of water and sediment in the Ebre Delta, Spain. Journal of Coastal Conservation 3:191202.Google Scholar
Ibañez, C., James, P. Day, J. W., Day, J. N., and Prat, N. 2010. Vertical accretion and relative sea level rise in the Ebro delta wetlands (Catalonia, Spain). Wetlands, 30: 979988.Google Scholar
Ibáñez, C., and Prat, N. 2003. The environmental impact of the Spanish Hydrological Plan on the lower Ebro river and delta. Water Resources Development, 19(3): 485500.Google Scholar
Ibáñez, C., Prat, N., and Canicio, A. 1996a. Changes in the hydrology and sediment transport produced by large dams on the lower Ebro river and its estuary. Regulated Rivers, 12(1): 5162.Google Scholar
Jefferies, R. L., Jano, A. P. and Abraham, K. F. 2006. A biotic agent promotes large-scale catastrophic change in the coastal marshes of Hudson Bay. Journal of Ecology, 94: 234242.Google Scholar
Jones, S. F., Stagg, C. A. Krauss, K. W. and Hester, M. W. 2016. Tidal saline wetland regeneration of sentinel vegetation types in the northern Gulf of Mexico: An overview. Estuarine, Coastal and Shelf Science, 174: A1A10.Google Scholar
Keddy, P. A., Campbell, D., McFalls, T., Shaffer, G. P., Moreau, R., Dranguet, C., and Heleniak, R. 2007. The wetlands of Lakes Pontchartrain and Maurepas: Past, present and future. Environmental Reviews, 15: 4377.Google Scholar
Kemp, G. P., Day, J. W. and Freeman, A. M. 2014. Restoring the sustainability of the Mississippi River Delta. Ecological Engineering, 65: 131146Google Scholar
Kemp, P., Day, J., Rogers, D., Giosan, L., and Peyronnin, N. 2016. Enhancing mud supply to the Mississippi River delta: Dam bypassing and coastal restoration. Estuarine, Coastal and Shelf Science, 183: 304313.Google Scholar
Keown, M. P., Dardeau, E. A. Jr., and Causey, E. M. 1986. Historic trends in the sediment flow regime of the Mississippi River. Water Resources Research, 22: 15551564.Google Scholar
Kesel, R. H. 1989. The role of the Mississippi River in wetland loss in southeastern Louisiana, U.S.A. Environmental Geology, 13:183193.Google Scholar
Kesel, R. H., Yodis, E. G., and McCraw, D. J. 1992. An approximation of the sediment budget of the lower Mississippi River prior to major human modification. Earth Surface Processes and Landforms, 17: 711722.Google Scholar
Kim, W., Mohrig, D., Twilley, R., Paola, C., and Parker, G. 2009. Is it feasible to build new land in the Mississippi River Delta? Eos, Transactions, American Geophysical Union, 90: 373374Google Scholar
Kirwan, M. L., and Blum, L. K. 2011. Enhanced decomposition offsets enhancedIbáñez C, Prat N (2003) The environmental impact of the Spanish Hydrological Plan on the lower Ebro river and delta. Water Resources Development, 19(3): 485500.Google Scholar
Ko, J., and Day, J. 2004. A review of ecological impacts of oil and gas development on coastal ecosystems in the Mississippi delta. Ocean and Coastal Management, 47: 671691.Google Scholar
Konisky, R. A., Burdick, D. M., Dionne, , and Neckles, M. H. A. 2006. A regional assessment of salt marsh restoration and monitoring in the Gulf of Maine. Restoration Ecology, 14: 516525.Google Scholar
La Peyre, M. K., Gossman, B., and Piazza, B. P. 2009. Short- and long-term response of deteriorating brackish marshes and open-water ponds to sediment enhancement by thin-layer dredge disposal. Estuaries and Coasts, 32: 390402.Google Scholar
LCWCRTF (Louisiana Coastal Wetlands Conservation and Restoration Task Force). 2006. The 2006 Evaluation Report to the U.S. Congress on the Effectiveness of Coastal Wetlands Planning, Protection and Restoration Act Projects. U.S. Army Corps of Engineers, New Orleans.Google Scholar
LCWCRTF (Louisiana Coastal Wetlands Conservation and Restoration Task Force). 2012. The 2012 Evaluation Report to the U.S. Congress on the Effectiveness of Coastal Wetlands Planning, Protection and Restoration Act Projects. U.S. Army Corps of Engineers, New Orleans.Google Scholar
Lopez, J., Henkel, T., Moshogianis, A., Baker, A., Boyd, E., Hillmann, E., Connor, P., and Baker, D. B., 2014. Examination of deltaic processes of Mississippi River outlets Caernarvon Delta and Bohemia Spillway in southeast Louisiana. Gulf Coast Association of Geological Societies Transactions, 64: 707e708.Google Scholar
MacBroom, J. G., and Schiff, R. 2012. Predicting the hydrologic response of salt marshes to tidal restoration. In Roman, C.T. and Burdick, D.M. (eds.) Tidal Marsh Restoration: A Synthesis of Science and Practice. pp. 315332. Washington, DC: Island Press.Google Scholar
Mitsch, W. J. and Gosselink, J. G. 2015. Wetlands, 5th ed. John Wiley & Sons, Inc., Hoboken, NJ.Google Scholar
Morgan, P. A., Burdick, D. M., and Short, F. T. 2009. The functions and values of fringing salt marshes in northern New England, USA. Estuaries and Coasts, 32: 483495.Google Scholar
Nagarkar, M., and Raulund-Rasmussen, K. 2016. An appraisal of adaptive management planning and implementation in ecological restoration: Case studies from the San Francisco Bay Delta, USA. Ecology and Society, 21(2): 43.Google Scholar
Naquin, J. D., Lui, K. B, McCloskey, T. A., and Bianchette, T. A., 2014. Storm deposition induced by hurricanes in a rapidly subsiding coastal zone. Journal of Coastal Research Special Issue, 70: 308313.Google Scholar
NAS (National Academies of Sciences, Engineering, and Medicine). 2017. Effective Monitoring to Evaluate Ecological Restoration in the Gulf of Mexico. Washington, DC: The National Academies Press.Google Scholar
Needelman, B. A., Crooks, S., Shumway, C. A. J., Titus, G. Takacs, R. and Hawkes, J. E. 2012. Restore-Adapt-Mitigate: Responding to Climate Change through Coastal Habitat Restoration. Washington, DC: Restore America’s Estuaries.Google Scholar
Nicholls, R. J., and Cazenave, A. 2010. Sea-level rise and its impact on coastal zones. Science, 328: 15171520.Google Scholar
Nicholls, R. J., Hutton, C. W., Lazar, , Allan, A., Adger, W., Adams, H., Wolff, J., Rahman, M. and Salehin, M. 2016. Integrated assessment of social and environmental sustainability dynamics in the Ganges-Brahmaputra-Meghna delta, Bangadesh. Estuarine, Coastal and Shelf Science, 183: 370381.Google Scholar
Nittrouer, J. A., Best, J. L., Brantley, C., Cash, R. W., Czapiga, M., Kumar, P., and Parker, G., 2012. Mitigating land loss in coastal Louisiana by controlled diversion of Mississippi River sand. Nature Geoscience, 5: 534537.Google Scholar
NOAA. 2015. Guidance for Considering the Use of Living Shorelines. Living Shorelines Workgroup. www.habitatblueprint.noaa.gov/living-shorelines/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
Paola, C., Twilley, R., Edmonds, D., Kim, W., Mohrig, D., and Parker, G. 2011. Natural processes in delta restoration: Application to the Mississippi delta. Annual Reviews of Marine Science, 3: 6791.Google Scholar
Peterson, S. B., Teal, J. M., and Mitsch, W. J. eds. 2005. Delaware Bay Salt Marsh Restoration. Ecological Engineering, Special issue25: 199314.Google Scholar
Pethick, J. 2002. Estuarine and tidal wetland restoration in the United Kingdom: policy versus practice. Restoration Ecology, 10: 431437.Google Scholar
Peyronnin, N. S., Green, M., Richards, C. P., Owens, A., Reed, D., Chamberlain, J., Groves, D. G., Rhinehart, K., and Belhadjali, K. 2013. Louisiana’s 2012 coastal master plan: Overview of a science-based and publicly informed decision-making process. Journal of Coastal Research, 67, 115.Google Scholar
Peyronnin, N. S., Caffey, R. H., Cowan, J. H. Jr., Justic, D., Kolker, A. S., Laska, S. B., McCorquodale, A., et al. 2017. Optimizing sediment diversion operations: Working group recommendations for integrating complex ecological and social landscape interactions. Water, 9: 368. DOI: 10.3390/w9060368Google Scholar
Pont, D., Day, J. and Ibáñez, C. 2017. The impact of two large floods (1993–1994) on sediment deposition in the Rhone delta: Implications for sustainable management. Science of the Total Environment, 609: 251262.Google Scholar
Powers, S. P., and Boyer, K. E. 2013. Marine restoration ecology. In Bertness, M. D., Bruno, J., Silliman, B., and Stachowicz, J. et al. (eds.). Marine Community Ecology and Conservation. pp. 495516. Sunderland, MA: Sinauer Associates.Google Scholar
Prado, P., Alcaraz, C., Jornet, L., Caiola, N. and Ibáñez, C. 2017. Effects of enhanced hydrological connectivity on Mediterranean salt marsh fish assemblages with emphasis on the endangered Spanish toothcarp (Aphanius iberus). PeerJ 5:e3009. DOI 10.7717/peerj.3009.Google Scholar
Proffitt, C. E., Chiasson, R. L., Owens, A. B., Edwards, K. R., and Travis, S. E. 2005. Spartina alterniflora genotype influences facilitation and suppression of high marsh species colonizing an early successional salt marsh. Journal of Ecology, 93(2): 404416.Google Scholar
Proffitt, C. E., and Young, J. 1999. Salt marsh plant colonization, growth and dominance on large mudflats created using dredged sediments. In Rozas, L. R et al. (eds.) Recent Research in Coastal Louisiana: Natural System Function and Response to Human Influence. pp. 218228. Louisiana Sea Grant College Program, Baton Rouge, Louisiana.Google Scholar
Reiner, E. L. 2012. Restoration of tidally restricted salt marshes at Rumney Marsh, Massachusetts. In Roman, C. T. and Burdick, D. M. (eds.) Tidal Marsh Restoration a Synthesis of Science and Practice. pp. 355370. Washington, DC: Island Press.Google Scholar
Ray, G. L. 2007. Thin layer disposal of dredged material on marshes: A review of the technical and scientific literature. ERDC/EL Technical Notes Collection (ERDC/EL TN-07-1), Vicksburg, MS: U.S. Army Engineer Research and Development Center. 2007.Google Scholar
Restore America’s Estuaries. 1999. Principles of Estuarine Habitat Restoration. Arlington Virginia. www.edc.uri.edu/restoration/html/resource/rae-erf.pdfGoogle Scholar
Roberts, H. H. 1997. Dynamic changes of the holocene Mississippi River delta plain: The delta cycle. Journal of Coastal Research, 13(3): 605627.Google Scholar
Rogers, B. D., Herke, W. H., and Knudsen, E. E. 1992. Effects of three different water-control structures on the movement of coastal fishes and macrocrustaceans. Wetlands, 12: 106120.Google Scholar
Roman, C. T. and Burdick, D. M. 2012. Tidal Marsh Restoration: A Synthesis of Science and Management. Island Press, Washington, DC.Google Scholar
Rozsa, R. 1995. Tidal restoration in Connecticut. In Dreyer, G. D. and Niering, W. A. (eds.) Tidal Marshes of Long Island Sound: Ecology, History and Restoration. Bulletin no. 34. pp. 5165. The Connecticut College Arboretum, New London, Connecticut.Google Scholar
Rutherford, J., Day, J., D’Elia, C., Wiegman, A., Willson, C., Caffey, R., Shaffer, G., Lane, R., and Batker, D. 2018. Evaluating trade-offs of a large, infrequent sediment diversion for restoration of a forested wetland in the Mississippi delta. Estuarine, Coastal and Shelf Science, 203: 8089.Google Scholar
Saltonstall, K. 2002. Cryptic invasion by a non-native genotype of the common reed, Phragmites australis, into North America. Proceedings of the National Academy of Sciences of the USA 99: 24452449.Google Scholar
Sasser, C., Holm, G., Evers, E., and Shaffer, G. 2018. The nutria in Louisiana: A current and historical perspective. In Day, J. and Erdman, J. (eds.) Mississippi Delta Restoration – Pathways to a Sustainable Future. pp. 3960. Springer, Glam, Switzerland.Google Scholar
Saucier, R. R. 1963. Recent Geomorphic History of the Pontchartrain basin. Coastal Studies Series No. 9. Louisiana State University Press, Baton Rouge, LA.Google Scholar
Simenstad, C. A., Reed, D. and Ford, M. 2006. When is restoration not? Incorporating landscape-scale processes to restore self-sustaining ecosystems in coastal wetland restoration. Ecological Engineering, 26: 2739.Google Scholar
Simenstad, C. A., and Thom, R. M. 1996. Functional equivalency trajectories of the restored Gog-Le-Hi-Te estuarine wetland. Ecological Applications, 6: 3856.Google Scholar
Slocum, M. G., Mendelssohn, I. A., and Kuhn, N. L., 2005. Effects of sediment slurry enrichment on salt marsh rehabilitation: plant and soil responses over seven years. Estuaries, 28: 519528.Google Scholar
Smith, S. M., Roman, C. T., James-Pirri, M., Chapman, K., Portnoy, , and Gwilliam, J. E. 2009. Responses of plant communities to incremental hydrologic restoration of a tide-restricted salt marsh in southern New England (Massachusetts, USA). Restoration Ecology, 17: 606618.Google Scholar
Society for Ecological Restoration International Science & Policy Working Group. 2004. The SER International Primer on Ecological Restoration. Tucson Az: Society for Ecological Restoration International.Google Scholar
Staver, L. W. 2015. Ecosystem dynamics in tidal marshes constructed with fine grained, nutrient rich dredged material. Dissertation, University of Maryland, College Park, MD.Google Scholar
Streever, 2000. Spartina alterniflora marshes on dredge material: a critical review of the ongoing debate over success. Wetlands Ecology and Management, 8: 295316.Google Scholar
Swenson, E. M., and Turner, R. E. 1987. Spoil banks: effects on a coastal marsh water-level regime. Estuarine, Coastal and Shelf Science, 24:599609.Google Scholar
Syvitski, J., Kettner, A., Overeem, I., Hutton, E., Hannon, M. Brakenridge, G., Day, J., et al. 2009. Sinking deltas due to human activities. Nature Geoscience, 2: 681686.Google Scholar
Teal, J., and Teal, M. 1969. Life and Death of the Salt Marsh. Ballantine Books, New York.Google Scholar
Teal, J. M, and Weinstein, M. P. 2002. Ecological engineering, design, and construction considerations for marsh restorations in Delaware Bay, USA. Ecological Engineering, 18: 607618.Google Scholar
Tockner, K., Malard, F., and Ward, J. V. 2000. An extension of the flood pulse concept. Hydrological Processes, 14: 28612883.Google Scholar
Turner, R. E. and Streever, B. 2002. Approaches to Coastal Wetland Restoration: Northern Gulf of Mexico. Kugler Publications, The Netherlands.Google Scholar
Twilley, R. R., Couvillion, B. R., Hossain, I., Kaiser, C., Owens, A. B., and Steyer, G. D. 2014. Coastal Louisiana Ecosystem Assessment and Restoration Program: The role of ecosystem forecasting in evaluating restoration planning in the Mississippi River Deltaic Plain. American Fisheries Society Symposium. v. 64, p. 29–46.Google Scholar
Twilley, R. R., Couvillion, B. R., Hossain, I., Kaiser, C., Owens, A. B., Steyer, G. D., and Visser, J. M. 2018 Coastal Louisiana ecosystem assessment and restoration program: The role of ecosystem forecasting in evaluating restoration planning in the Mississippi river deltaic plain. American Fisheries Society Symposium, 64, 2946.Google Scholar
Twilley, R. R., and Rivera-Monroy, V. 2009. Sediment and nutrient tradeoffs in restoring Mississippi River Delta: restoration vs. eutrophication. Journal of Contemporary Water Research & Education, 141: 3944.Google Scholar
Tyrrell, M., Dionne, M. and Edgerly, J. 2008. Physical factors mediate effects of grazing by a non-indigenous snail species on saltmarsh cordgrass (Spartina alterniflora) in New England marshes. ICES Journal of Marine Science, 65: 746752.Google Scholar
USGCRP. 2017. Climate Science Special Report: Fourth National Climate Assessment, Volume I [Wuebbles, D.J., Fahey, D.W., Hibbard, K.A., Dokken, D.J., Stewart, B.C., and Maycock, T.K. (eds.)]. U.S. Global Change Research Program, Washington, DC, USAGoogle Scholar
Valoppi, L., 2018, Phase 1 studies summary of major findings of the South Bay Salt Pond Restoration Project, South San Francisco Bay, California: U.S. Geological Survey Open-File Report 2018–1039, 58 p., plus appendixes, https://doi.org/10.3133/ofr20181039.Google Scholar
Wiegman, A., Day, J., D’Elia, C., Rutherford, J., Morris, J., Roy, E., Lane, R., Dismukes, D., and Synder, B. 2017. Modeling impact of sea-level rise, oil price, and management strategy on the costs of sustaining Mississippi delta marshes with hydraulic dredging. STOTEN.Google Scholar
Wilber, P. 1993. Managing Dredged Material Via Thin-Layer Disposal in Coastal Marshes. USACE Technical Note EEDP-01-32.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
Wolters, M., Garbutt, A. and Bakker, J. P. 2005. Salt-marsh restoration: evaluating the success of de-embankments in north-west Europe. Biological Conservation, 123: 249268.Google Scholar
Wolters, M., Garbutt, A., Bekker, R. M., Bakker, J. P, and Carey, P. D. 2008. Restoration of saltmarsh vegetation in relation to site suitability, species pool and dispersal traits. Journal of Applied Ecology, 45: 904912.Google Scholar

References

Armitage, A. R., Highfield, W. E., Brody, S. D. and Louchouarn, P. 2015The contribution of mangrove expansion to salt marsh loss on the Texas Gulf coastPLoS ONE 10(5): e0125404. doi:10.1371/journal.pone.0125404.Google Scholar
CPRA (Coastal Protection and Restoration Authority of Louisiana). 2017. Louisiana’s Comprehensive Master Plan for a Sustainable Coast. Coastal Protection and Restoration Authority of Louisiana. Baton Rouge, LA.Google Scholar
Day, J., Agboola, J., Chen, Z., D'Elia, C., Forbes, D., Giosan, L., Kemp, P. et al. 2016. Approaches to defining deltaic sustainability in the 21st century. Estuarine, Coastal and Shelf Science. 183: 275291.Google Scholar
Day, J., Colten, C. and Kemp, G. P. 2019. Mississippi Delta restoration and protection: Shifting baselines, diminishing resilience, and growing non-sustainability. In: Wolanski, E., Day, J., Elliott, M. and Ramnachandran, R. (eds). Coasts and Estuaries: The Future. Elsevier, Amsterdam, pp. 173–192.Google Scholar
De Groot, R., Brander, L., Van Der Ploeg, S., Costanza, R., Bernard, F., Braat, L., Christie, M., Crossman, N., Ghermandi, A., and Hein, L. 2012. Global estimates of the value of ecosystems and their services in monetary units. Ecosystem Services., 1: 5061.Google Scholar
Donnelly, J. P., and Bertness, M. D. 2001. Rapid shoreward encroachment of salt marsh cordgrass in response to accelerated sea-level rise. Proceedings of the National Academy of Sciences of the USA, 98:1421814223.Google Scholar
He, Q., Bertness, M. D., Bruno, J. 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
Mcowen, C., Weatherdon, L., Bochove, J., Sullivan, E., Blyth, S., Zockler, C., Stanwell-Smith, D., Kingston, N. and Martin, C. 2017A global map of saltmarshesBiodiversity Data Journal5 (5): e11764.  doi:10.3897/bdj.5.e11764Google Scholar
Mehvar, S., Filatova, T., Dastgheib, A., de Ruyter van Steveninck, E., and Ranasinghe, R. 2018, Quantifying economic value of coastal ecosystem services: A review. Journal of Marine Science and Engineering, 6: 5; doi:10.3390/jmse6010005Google 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. Journal of Ecology, 83(10):28692877.Google Scholar
Weston, N.B. 2014. Declining sediments and rising seas: an unfortunate convergence for tidal wetlands Estuaries and Coasts, 37:123Google Scholar
Wu, W., Huang, H., Biber, P., and Bethel, M. 2017. Litter decomposition of Spartina alterniflora and Juncus roemerianus: implications of climate change in salt marshes. Journal of Coastal Research, 33(2): 372384.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
×