Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-v5vhk Total loading time: 0 Render date: 2024-06-25T03:59:01.022Z Has data issue: false hasContentIssue false

12 - Salt Marsh Ecosystems: Tidal Flow, Vegetation, and Carbon Dynamics

from Part IV - Coupling Fluvial and Aeolian Geomorphology, Hydrology/Hydraulics, and Ecosystems

Published online by Cambridge University Press:  27 October 2016

Simon M. Mudd
Affiliation:
University of Edinburgh
Sergio Fagherazzi
Affiliation:
Boston University
Edward A. Johnson
Affiliation:
University of Calgary
Yvonne E. Martin
Affiliation:
University of Calgary
Get access

Summary

Introduction

Salt marshes are coastal wetlands located in the intertidal zone that are populated by halophytic vegetation. Although they occupy a relatively small percentage of the Earth's surface, they have been the focus of intense ecological and geomorphological research for several decades due to their importance in filtering pollutants, buffering against coastal storms, serving as nurseries for commercial fisheries, and storing carbon (Barbier et al., 2011). These ecosystems were among the first to be studied in an ecogeomorphic context due to the clear feedbacks between plant populations and landscape formation (e.g., Redfield 1972).

Salt marsh ecosystems are found on all continents except for Antarctica in mid to high latitude locations (Figure 12.1), giving way to mangrove swamps in subtropical and tropical climates. Radiocarbon dating of basal peats suggests that salt marsh ecosystems became widespread sometime between 4000 and 6000 years ago (Allen, 2000). Prior to this, sea level rise associated with post-glacial melt water and ocean expansion was too rapid for marsh establishment. Rates of eustatic sea level rise have increased over the twentieth and into the twenty-first century (Church and White, 2011), and there are now growing concerns that accelerating rates of sea level rise, combined with a decrease of sediment availability due to river damming could threaten marsh ecosystems. The threat of marsh loss has focused research into the balance between plant growth, hydrodynamics, and sedimentation in salt marsh ecosystems.

In this chapter we will examine the close coupling between plant vitality, hydrodynamics on marsh surfaces, and sedimentation. These components of the ecogeomorphic system on salt marshes form a continuous loop: marsh plants respond to edaphic factors such as salinity and depth in the tidal frame, and these are controlled by the hydrodynamics of tidally induced floods and sediment deposition rates. Plants interact with flows through drag, and can affect sedimentation rates through trapping. We begin by examining the factors that control plant productivity on coastal marshes.

Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2016

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

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, 1155–1231, doi:10.1016/S0277-3791(99)00034–7.Google Scholar
Baptist, M. J., Babovic, V., Rodríguez Uthurburu, J. et al. (2007). On inducing equations for vegetation resistance. Journal of Hydraulic Research, 45, 435–50, doi:10.1080/00221686.2007.9521778.Google Scholar
Barbier, E. B., Hacker, S. D., Kennedy, C. et al. (2011). The value of estuarine and coastal ecosystem services. Ecological Monographs, 81, 169–93, doi:10.1890/10–1510.1.Google Scholar
Bayliss-Smith, T. P., Healey, R., Lailey, R., Spencer, T. and Stoddart, D. R. (1978). Tidal flows in salt marsh creeks. Estuarine Coastal Marine Science, 9, 235–55.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, 603–20, doi:10.1002/2013JF002967.Google Scholar
Bertness, M. D. and Ellison, A. M. (1987). Determinants of pattern in a New England salt marsh plant community. Ecological Monographs, 57, 129–47, doi:10.2307/1942621.Google Scholar
Bertness, M. D. and Silliman, B. R. (2008). Consumer control of salt marshes driven by human disturbance. Conservation Biology, 22, 618–23, doi:10.1111/j.1523–1739.2008.00962.x.Google Scholar
Blum, L. K. (1993). Spartina alterniflora root dynamics in a Virginia marsh. Marine Ecology Progress Series, 102, 169–78.Google Scholar
Blum, L. K. and Christian, R. R. (2004). Belowground production and decomposition along a tidal gradient in a Virginia salt marsh. In: Coastal and Estuarine Studies, ed. Fagherazzi, S., Marani, M. and Blum, L. K.. Washington, DC: American Geophysical Union, pp. 47–73.
Boaga, J., D'Alpaos, A., Cassiani, G., Marani, M. and Putti, M. (2014). Plant-soil interactions in salt marsh environments: experimental evidence from electrical resistivity tomography in the Venice Lagoon. Geophysical Research Letters, 41, 2014GL060983, doi:10.1002/2014GL060983.Google Scholar
Bouma, T. J., Friedrichs, M., Klaassen, P. et al. (2009). Effects of shoot stiffness, shoot size and current velocity on scouring sediment from around seedlings and propagules. Marine Ecoogy Progress Series, 388, 293–7, doi:10.3354/meps08130.Google Scholar
Brain, M. J., Long, A. J., Petley, D. N., Horton, B. P. and Allison, R. J. (2011). Compression behaviour of minerogenic low energy intertidal sediments. Sedimentary Geology, 233, 28–41, doi:10.1016/j.sedgeo.2010.10.005.CrossRefGoogle Scholar
Buresh, R. J., DeLaune, R. D. and Patrick, W. H. (1980). Nitrogen and phosphorus distribution and utilization by Spartina alterniflora in a Louisiana gulf coast marsh. Estuaries, 3, 111–21, doi:10.2307/1351555.Google Scholar
Callaway, J. C. and Josselyn, M. N. (1992). The introduction and spread of smooth cordgrass (Spartina alterniflora) in South San Francisco Bay. Estuaries, 15, 218–26, doi:10.2307/1352695.Google Scholar
Callaway, J. C., DeLaune, R. D. and Patrick, W. H. P. Jr. (1997). Sediment accretion rates from four coastal wetlands along the Gulf of Mexico. Journal of Coastal Research, 13, 181–91.Google Scholar
Callaway, R. M., Jones, S., Ferren, W. R. Jr. and Parikh, A. (1990). Ecology of a mediterranean-climate estuarine wetland at Carpinteria, California: plant distributions and soil salinity in the upper marsh. Canadian Journal of Botany, 68, 1139–46, doi:10.1139/b90-144.Google Scholar
Carniello, L., Defina, A., Fagherazzi, S. and D'Alpaos, L. (2005). A combined wind wave–tidal model for the Venice lagoon, Italy. Journal of Geophysical Research-Earth Surface, 110, F04007, doi:10.1029/2004JF000232.Google Scholar
Chen, Z., Ortiz, A., Zong, L. and Nepf, H. (2012). The wake structure behind a porous obstruction and its implications for deposition near a finite patch of emergent vegetation. Water Resources Research, 48, W09517, doi:10.1029/2012WR012224.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, doi:10.1029/2002GB001917.Google Scholar
Church, J. A. and White, N. J. (2011). Sea-level rise from the late 19th to the early 21st century. Surveys in Geophysics, 32, 585–602, doi:10.1007/s10712-011–9119-1.Google Scholar
Colmer, T. D. and Flowers, T. J. (2008). Flooding tolerance in halophytes. New Phytologist, 179, 964–74, doi:10.1111/j.1469–8137.2008.02483.x.Google Scholar
Conley, D. J. (1997). Riverine contribution of biogenic silica to the oceanic silica budget. Limnology and Oceanography, 42, 774–7.Google Scholar
Conley, D., Schelske, C., Stoermer, E., (1993). Modification of the Biogeochemical Cycle of Silica with Eutrophication. Marine Ecology Progress Series, 101, 179–92, doi:10.3354/meps101179.Google Scholar
Coverdale, T. C., Axelman, E. E., Brisson, C. P. et al. (2013). New England salt marsh recovery: opportunistic colonization of an invasive species and its non-consumptive effects. PLoS ONE, 8, e73823, doi:10.1371/journal.pone.0073823.CrossRefGoogle Scholar
Dacey, J. W. H. and Howes, B. L. (1984). Water uptake by roots controls water table movement and sediment oxidation in short Spartina marsh. Science, 224, 487–9, doi:10.1126/science.224.4648.487.Google Scholar
Da Lio, C., D'Alpaos, A. and Marani, M. (2013). The secret gardener: vegetation and the emergence of biogeomorphic patterns in tidal environments. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 371, 20120367. doi:10.1098/rsta.2012.0367.Google Scholar
D'Alpaos, A., Mudd, S. M. and Carniello, L. (2011). Dynamic response of marshes to perturbations in suspended sediment concentrations and rates of relative sea level rise. Journal of Geophysical Research-Earth Surface, 116, F04020, doi:10.1029/2011JF002093.Google Scholar
D'Alpaos, A., Lanzoni, S., Marani, M., Fagherazzi, S. and Rinaldo, A. (2005). Tidal network ontogeny: channel initiation and early development. Journal of Geophysical Research-Earth Surface, 110, doi:10.1029/2004JF000182.Google Scholar
Darby, F. A. and Turner, R. E. (2008). Below- and aboveground Spartina alterniflora production in a Louisiana salt marsh. Estuaries and Coasts, 31, 223–31, doi:10.1007/s12237-007–9014-7.Google Scholar
Deegan, L. A., Johnson, D. S., Warren, R. S. et al. (2012). Coastal eutrophication as a driver of salt marsh loss. Nature, 490, 388–92, doi:10.1038/nature11533.Google Scholar
Dietrich, W. E. (1982). Settling velocity of natural particles. Water Resources Research, 18, 1615–26, doi:10.1029/WR018i006p01615.Google Scholar
Drake, D. C., Peterson, B. J., Galván, K. A. et al. (2009). Salt marsh ecosystem biogeochemical responses to nutrient enrichment: a paired 15N tracer study. Ecology, 90, 2535–46.Google Scholar
Erickson, J. E., Peresta, G., Montovan, K. J. and Drake, B. G. (2013). Direct and indirect effects of elevated atmospheric CO2 on net ecosystem production in a Chesapeake Bay tidal wetland. Global Change Biology, 19, 3368–78, doi:10.1111/gcb.12316.CrossRefGoogle Scholar
Fagherazzi, S. (2013). The ephemeral life of a salt marsh. Geology, 41(8), 943–44.Google Scholar
Fagherazzi, S. and Priestas, A. M. (2010) Sediments and water fluxes in a muddy coastline: interplay between waves and tidal channel hydrodynamics, Earth Surface Processes and Landforms, 35(3), 284–93.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-Earth Surface, 114(F3), doi:10.1029/2008JF001139.Google Scholar
Fagherazzi, S., Hannion, M. and D'Odorico, P. (2008). Geomorphic structure of tidal hydrodynamics in salt marsh creeks, Water Resources Research, 44, W02419, doi:10.1029/2007WR006289.CrossRefGoogle 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, 103(22), 8337–41.Google Scholar
Fagherazzi, S., Mariotti, G., Wiberg, P. and McGlathery, K. (2013a). Marsh collapse does not require sea level rise. Oceanography, 26(3), 70–7.Google Scholar
Fagherazzi, S., Palermo, C., Rulli, M. C., Carniello, L. and Defina, A. (2007). Wind waves in shallow microtidal basins and the dynamic equilibrium of tidal flats. Journal of Geophysical Research-Earth Surface, 112(F2), doi:10.1029/2006JF000572.Google Scholar
Fagherazzi, S., Bortoluzzi, A., Dietrich, W. E. et al. (1999). Tidal networks: 1. automatic network extraction and preliminary scaling features from digital terrain maps. Water Resources Research, 35, 3891–904, doi:10.1029/1999WR900236.Google Scholar
Fagherazzi, S., Wiberg, P. L., Temmerman, S. et al. (2013b). Fluxes of water, sediments, and biogeochemical compounds in salt marshes. Ecological Processes, 23, doi:10.1186/2192-1709-2–3.Google Scholar
Feagin, R. A., Lozada-Bernard, S. M., Ravens, T. M. et al. (2009). Does vegetation prevent wave erosion of salt marsh edges? Proceedings of the National Academy of Sciences, 106(25), 10109–13, doi:10.1073/pnas.0901297106.Google Scholar
Fragoso, G. and Spencer, T. (2008). Physiographic control on the development of Spartina marshes. Science, 322, 1064, doi:10.1126/science.1159973.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, 53–65, doi:10.1016/j.geomorph.2013.04.026.Google Scholar
Gedan, K. B. and Bertness, M. D. (2010). How will warming affect the salt marsh foundation species Spartina patens and its ecological role? Oecologia, 164, 479–87, doi:10.1007/s00442-010–1661-x.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, 7–29, doi:10.1007/s10584-010-0003-7.Google Scholar
Gleason, M. L., Elmer, D. A., Pien, N. C. and Fisher, J. S. (1979). Effects of stem density upon sediment retention by salt marsh cord grass, Spartina alterniflora loisel . Estuaries, 2, 271–3, doi:10.2307/1351574.CrossRefGoogle Scholar
Hartmann, D.L., Klein Tank, A. M. G., Rusticucci, M. et al. (2013). Observations: atmosphere and surface. 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, ed. Stocker, T. F., Qin, D., Plattner, G.-K., et al. Cambridge: Cambridge University Press, pp. 159–254.
Hemminga, M., Huiskes, A., Steegstra, M. and van Soelen, J. (1996). Assessment of carbon allocation and biomass production in a natural stand of the salt marsh plant Spartina anglica using 13C. Marine Ecology Progress Series, 130, 169–78, doi:10.3354/meps130169.Google Scholar
Hoekstra, J. M., Molnar, J. L., Jennings, M. et al. (2010). The Atlas of Global Conservation: Changes, Challenges, and Opportunities to Make a Difference, ed. Molnar, J. L.. Berkeley, CA: University of California Press.
Houser, C. (2010). Relative importance of vessel-generated and wind waves to salt marsh erosion in a restricted fetch environment. Journal of Coastal Research, 230–40, doi:10.2112/08–1084.1.
Kiehn, W. M. and Morris, J. T. (2009). Relationships between Spartina alterniflora and Littoraria irroratain in a South Carolina salt marsh. Wetlands, 29, 818–25, doi:10.1672/08–178.1.Google Scholar
Kirby, C. J. and Gosselink, J. G. (1976). Primary production in a Louisiana gulf coast Spartina alterniflora marsh. Ecology, 57, 1052–9, doi:10.2307/1941070.Google Scholar
Kirwan, M. L. and Mudd, S. M. (2012). Response of salt-marsh carbon accumulation to climate change. Nature, 489, 550–3, doi:10.1038/nature11440.Google Scholar
Kirwan, M. L., Guntenspergen, G. R. and Langley, J. A. (2014). The temperature sensitivity of organic matter decay in tidal marshes. Biogeosciences Discussion, 11, 6019–37, doi:10.5194/bgd-11–6019-2014.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, 1982–9, doi:10.1111/j.1365–2486.2008.01834.x.Google Scholar
Kirwan, M. L., Murray, A. B. and Boyd, W. S. (2008). Temporary vegetation disturbance as an explanation for permanent loss of tidal wetlands. Geophysical Research Letters, 35, L05403, doi:10.1029/2007GL032681.Google Scholar
Kirwan, M. L., Guntenspergen, G. R., D'Alpaos, A. et al. (2010). Limits on the adaptability of coastal marshes to rising sea level. Geophysical Research Letters, 37(23), L23401, doi:10.1029/2010GL045489.Google Scholar
Krone, R. B. (1987). A method for simulating historic marsh elevations. In Coastal Sediments (1987), ed. Kraus, N. C.. New York: American Society of Civil Engineers, pp. 316–23.
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, 106(15), doi:10.1073/pnas.0807695106.Google Scholar
Leonard, L. A. and Croft, A. L. (2006). The effect of standing biomass on flow velocity and turbulence in Spartina alterniflora canopies. Estuarine, Coastal and Shelf Science, 69, 325–36. doi:10.1016/j.ecss.2006.05.004.Google Scholar
Leonard, L. A. and Luther, M. E. (1995). Flow hydrodynamics in tidal marsh canopies. Limnology and Oceanography, 40, 1474–84, doi:10.4319/lo.1995.40.8.1474.Google Scholar
López, F. and García, M. (2001). Mean flow and turbulence structure of open-channel flow through non-emergent vegetation. Journal of Hydraulic Engineering, 127, 392–402, doi:10.1061/(ASCE)0733–9429(2001)127:5(392).Google Scholar
Marani, M., Da Lio, C. and D'Alpaos, A. (2013). Vegetation engineers marsh morphology through multiple competing stable states. Proceedings of the National Academy of Sciences, 110(9), 3259–63.Google Scholar
Marani, M., Lanzoni, S., Silvestri, S. and Rinaldo, A. (2004).Tidal landforms, patterns of halophytic vegetation and the fate of the lagoon of Venice. Journal of Marine Systems, 51, 191–210, doi:10.1016/j.jmarsys.2004.05.012.Google Scholar
Marani, M., D'Alpaos, A., Lanzoni, S., Carniello, L. and Rinaldo, A. (2007). Biologically controlled multiple equilibria of tidal landforms and the fate of the Venice lagoon. Geophysical Research Letters, 34(11), L11402, doi:10.1029/2007GL030178.Google Scholar
Marani, M., D'Alpaos, A., Lanzoni, S., Carniello, L. and Rinaldo, A. (2010). The importance of being coupled: stable states and catastrophic shifts in tidal biomorphodynamics. Journal of Geophysical Research-Earth Surface, 115(F4), F04004, doi:10.1029/2009JF001600.CrossRefGoogle Scholar
Marani, M., Belluco, E., D'Alpaos, A. et al. (2003). On the drainage density of tidal networks. Water Resources Research, 39, 1040, doi:10.1029/2001WR001051.Google Scholar
Mariotti, G. and Fagherazzi, S. (2010). A numerical model for the coupled long-term evolution of salt marshes and tidal flats. Journal of Geophysical Research-Earth Surface, 115(F1), F01004, doi: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, 110(14), 5353–6.Google Scholar
McGlathery, K. J., Reidenbach, M. A., D'Odorico, P. et al. (2013). Nonlinear dynamics and alternative stable states in shallow coastal systems. Oceanography, 26(3), 220–31.Google Scholar
McKee, K. L. and Patrick, W. H. (1988). The relationship of smooth cordgrass (Spartina alterniflora) to tidal datums: a review. Estuaries, 11, 143–51, doi:10.2307/1351966.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, 65–73, doi:10.1111/j.1466–882X.2004.00075.x.Google Scholar
Mcleod, E., Chmura, G. L., Bouillon, S. et al. (2011). A blueprint for blue carbon: toward an improved understanding of the role of vegetated coastal habitats in sequestering CO2 . Frontiers in Ecology and the Environment, 9, 552–60, doi:10.1890/110004.Google Scholar
Mendelssohn, I. A. and Kuhn, N. L. (2003). Sediment subsidy: effects on soil-plant responses in a rapidly submerging coastal salt marsh. Ecological Engineering, 21, 115–28, doi:10.1016/j.ecoleng.2003.09.006.Google Scholar
Mendelssohn, I. A. and Morris, J. T. (2000). Eco-physiological controls on the productivity of Spartina alterniflora loisel . In Concepts and Controversies in Tidal Marsh Ecology, ed. Weinstein, M. P. and Kreeger, D. A.. Dordrecht, Netherlands: Kluwer Academic, pp. 59–80.
Moffett, K. B., Gorelick, S. M., McLaren, R. G. and Sudicky, E. A. (2012). Salt marsh ecohydrological zonation due to heterogeneous vegetation–groundwater–surface water interactions. Water Resources Research, 48, W02516, doi:10.1029/2011WR010874.Google Scholar
Morris, J. T. and Haskin, B. (1990). A 5-yr record of aerial primary production and stand characteristics of Spartina alterniflora . Ecology, 71, 2209–17, doi:10.2307/1938633.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, 2869–77.Google Scholar
Morris, J., Sundberg, K. and Hopkinson, C. (2013). Salt marsh primary production and its responses to relative sea level and nutrients in estuaries at Plum Island, Massachusetts, and North Inlet, South Carolina, USA. Oceanography, 26, 78–84, doi:10.5670/oceanog.2013.48.Google Scholar
Mudd, S. M., D'Alpaos, A. and Morris, J. T. (2010). How does vegetation affect sedimentation on tidal marshes? Investigating particle capture and hydrodynamic controls on biologically mediated sedimentation. Journal of Geophysical Research-Earth Surface, 115, F03029, doi:10.1029/2009JF001566.Google Scholar
Mudd, S. M., Howell, S. M. and Morris, J. T. (2009). Impact of dynamic feedbacks between sedimentation, sea-level rise, and biomass production on near-surface marsh stratigraphy and carbon accumulation. Estuarine, Coastal and Shelf Science, 82, 377–89, doi:10.1016/j.ecss.2009.01.028.Google Scholar
Mudd, S. M., Fagherazzi, S., Morris, J. T. and Furbish, D. J. (2004). Flow, sedimentation, and biomass production on a vegetated salt marsh in South Carolina: toward a predictive model of marsh morphologic and ecologic evolution. In Coastal and Estuarine Studies, ed. Fagherazzi, S., Marani, M. and Blum, L. K.. Washington, DC: American Geophysical Union, pp. 165–88.
Nepf, H. M. (1999). Drag, turbulence, and diffusion in flow through emergent vegetation. Water Resources Research, 35, 479–89, doi:10.1029/1998WR900069.Google Scholar
Nepf, H. M. (2012). Flow and transport in regions with aquatic vegetation. Annual Review of Fluid Mechanics, 44, 123–42, doi:10.1146/annurev-fluid-120710-101048.Google Scholar
Neumeier, U. and Amos, C. L. (2006). The influence of vegetation on turbulence and flow velocities in European salt-marshes. Sedimentology, 53, 259–77, doi:10.1111/j.1365–3091.2006.00772.x.Google Scholar
Nikora, N. and Nikora, V. (2007). A viscous drag concept for flow resistance in vegetated channels. Proceedings of the Congress–International Association for Hydraulic Research, 32(1), 158.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, 370–80, doi:10.1016/j.ecss.2006.05.041.Google Scholar
Palmer, M. R., Nepf, H. M., Petterson, T. J. R. and Ackerman, J. D. (2004). Observations of particle capture on a cylindrical collector: implications for particle accumulation and removal in aquatic systems. Limnology and Oceanography, 49, 76–85, doi:10.2307/3597612.Google Scholar
Paramor, O. A. L. and Hughes, R. G. (2004).The effects of bioturbation and herbivory by the polychaete Nereis diversicolor on loss of saltmarsh in south-east England. Journal of Applied Ecology, 41, 449–63, doi:10.1111/j.0021–8901.2004.00916.x.Google Scholar
Pennings, S. C. and Callaway, R. M. (1992). Salt marsh plant zonation: the relative importance of competition and physical factors. Ecology, 73, 681–90, doi:10.2307/1940774.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, 159–67, doi:10.1111/j.1365–2745.2004.00959.x.Google Scholar
Redfield, A. C. (1972). Development of a New England salt marsh. Ecological Monographs, 42, 201–37, doi:10.2307/1942263.CrossRefGoogle Scholar
Rietkerk, M. and van de Koppel, J. (2008). Regular pattern formation in real ecosystems. Trends in Ecology and Evolution, 23, 169–75, doi:10.1016/j.tree.2007.10.013.Google Scholar
Rinaldo, A., Fagherazzi, S., Lanzoni, S., Marani, M. and Dietrich, W. E. (1999). Tidal networks: 2. watershed delineation and comparative network morphology. Water Resources Research, 35, 3905–17, doi:10.1029/1999WR900237.Google Scholar
Schwarz, C., Ye, Q.H., van der Wal, D. et al. (2014). Impacts of salt marsh plants on tidal channel initiation and inheritance: salt marsh plants channel development. Journal of Geophysical Research-Earth Surface, 119, 385–400, doi:10.1002/2013JF002900.CrossRefGoogle Scholar
Shepard, C. C., Crain, C. M. and Beck, M. W. (2011). The protective role of coastal marshes: a systematic review and meta-analysis. PLoS ONE, 6, e27374, doi:10.1371/journal.pone.0027374.Google Scholar
Shi, Z., Hamilton, L. J. and Wolanski, E. (2000). Near-bed currents and suspended sediment transport in saltmarsh canopies. Journal of Coastal Research, 16, 909–14.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, 1803–6, doi:10.1126/science.1118229.Google Scholar
Silvestri, S., Defina, A. and Marani, M. (2005). Tidal regime, salinity and salt marsh plant zonation. Estuarine, Coastal and Shelf Science, 62, 119–30, doi:10.1016/j.ecss.2004.08.010.Google Scholar
Struyf, E. and Conley, D. J. (2012). Emerging understanding of the ecosystem silica filter. Biogeochemistry, 107, 9–18.Google Scholar
Struyf, E., Dausse, A., Van Damme, S. et al. (2006). Tidal marshes and biogenic silica recycling at the land sea interface. Limnology and Oceanography, 51, 838–46, doi:10.4319/lo.2006.51.2.0838.Google Scholar
Tanino, Y. and Nepf, H. (2008). Laboratory investigation of mean drag in a random array of rigid, emergent cylinders. Journal of Hydraulic Engineering, 134, 34–41, doi:10.1061/(ASCE)0733–9429(2008)134:1(34).Google Scholar
Temmerman, S., De Vries, M. B. and Bouma, T. J. (2012). Coastal marsh die-off and reduced attenuation of coastal floods: a model analysis. Global and Planetary Change, 92–93, 267–74, doi:10.1016/j.gloplacha.2012.06.001.Google Scholar
Temmerman, S., Bouma, T.J., Govers, G. and Lauwaet, D. (2005a). Flow paths of water and sediment in a tidal marsh: relations with marsh developmental stage and tidal inundation height. Estuaries, 28, 338–52, doi:10.1007/BF02693917.Google Scholar
Temmerman, S., Bouma, T. J., Govers, G. et al. (2005b). Impact of vegetation on flow routing and sedimentation patterns: three-dimensional modeling for a tidal marsh. Journal of Geophysical Research-Earth Surface, 110, F04019, doi:10.1029/2005JF000301.Google Scholar
Temmerman, S., Bouma, T. J., van de Koppel, J. et al. (2007). Vegetation causes channel erosion in a tidal landscape. Geology, 35, 631–4, doi:10.1130/G23502A.1.Google Scholar
Temmerman, S., Meire, P., Bouma, T. J. et al. (2013). Ecosystem-based coastal defence in the face of global change. Nature, 504, 79–83, doi:10.1038/nature12859.Google Scholar
Tonelli, M., Fagherazzi, S. and Petti, M. (2010). Modeling wave impact on salt marsh boundaries. Journal of Geophysical Research-Oceans, 115, C09028, doi:10.1029/2009JC006026.Google Scholar
Torres, R. and Styles, R. (2007). Effects of topographic structure on salt marsh currents. Journal of Geophysical Research-Earth Surface, 112, F02023, doi:10.1029/2006JF000508.Google Scholar
Turner, R. E., Howes, B. L., Teal, J. M. et al. (2009). Salt marshes and eutrophication: an unsustainable outcome. Limnology and Oceanography, 54, 1634–42, doi:10.4319/lo.2009.54.5.1634.Google Scholar
Ursino, N., Silvestri, S. and Marani, M. (2004). Subsurface flow and vegetation patterns in tidal environments. Water Resources Research, 40(5), doi:10.1029/2003WR002702.Google Scholar
Valiela, I. and Cole, M. L. (2002). Comparative evidence that salt marshes and mangroves may protect seagrass meadows from land-derived nitrogen loads. Ecosystems, 5, 92–102.Google Scholar
Valiela, I., Teal, J. M. and Persson, N. Y. (1976). Production and dynamics of experimentally enriched salt marsh vegetation: belowground biomass. Limnology and Oceanography, 21, 245–52, doi:10.4319/lo.1976.21.2.0245.Google Scholar
Vandenbruwaene, W., Bouma, T. J., Meire, P. and Temmerman, S. (2013). Bio-geomorphic effects on tidal channel evolution: impact of vegetation establishment and tidal prism change. Earth Surface Processes and Landforms, 38, 122–32, doi:10.1002/esp.3265.Google Scholar
Vandenbruwaene, W., Temmerman, S., Bouma, T. J. et al. (2011). Flow interaction with dynamic vegetation patches: Implications for biogeomorphic evolution of a tidal landscape. Journal of Geophysical Research-Earth Surface, 116, F01008, doi:10.1029/2010JF001788.Google Scholar
van Wesenbeeck, B. K., van de Koppel, J., Herman, P. M. J. and Bouma, T. J. (2008). Does scale-dependent feedback explain spatial complexity in salt-marsh ecosystems? Oikos, 117, 152–9, doi:10.1111/j.2007.0030–1299.16245.x.Google Scholar
Voss, C. M., Christian, R. R. and Morris, J. T. (2013). Marsh macrophyte responses to inundation anticipate impacts of sea-level rise and indicate ongoing drowning of North Carolina marshes. Marine Biology, 160, 181–94, doi:10.1007/s00227-012–2076-5.Google Scholar
Vieillard, A. M., Fulweiler, R. W., Hughes, Z. J. and Carey, J. C. (2011). The ebb and flood of silica: quantifying dissolved and biogenic silica fluxes from a temperate saltmarsh. Estuarine, Coastal and Shelf Science, 95, 415–23, doi:10.1016/j.ecss.2011.10.012.Google Scholar
Wamsley, T. V., Cialone, M. A., Smith, J. M., Atkinson, J. H. and Rosati, J. D. (2010). The potential of wetlands in reducing storm surge. Ocean Engineering, 37, 59–68, doi:10.1016/j.oceaneng.2009.07.018.Google Scholar
White, B. L. and Nepf, H. M. (2008). A vortex-based model of velocity and shear stress in a partially vegetated shallow channel. Water Resources Research, 44, W01412, doi:10.1029/2006WR005651.Google Scholar
Wiegert, R. G., Chalmers, A. G. and Randerson, P. F. (1983). Productivity gradients in salt marshes: the response of Spartina alterniflora to experimentally manipulated soil water movement. Oikos, 41, 1–6, doi:10.2307/3544339.Google Scholar
Wigand, C., Brennan, P., Stolt, M., Holt, M. and Ryba, S. (2009). Soil respiration rates in coastal marshes subject to increasing watershed nitrogen loads in southern New England, USA. Wetlands, 29, 952–63, doi:10.1672/08–147.1.Google Scholar
Yang, S. L. (1998). The role of Scirpus marsh in attenuation of hydrodynamics and retention of fine sediment in the Yangtze Estuary. Estuarine, Coastal and Shelf Science, 47, 227–33, doi:10.1006/ecss.1998.0348.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
×