Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-mp689 Total loading time: 0 Render date: 2024-04-16T08:22:58.711Z Has data issue: false hasContentIssue false

9 - Riparian management and the conservation of stream ecosystems and fishes

Published online by Cambridge University Press:  05 December 2015

Mark S. Wipfli
Affiliation:
University of Alaska Fairbanks
John S. Richardson
Affiliation:
University of British Columbia
Gerard P. Closs
Affiliation:
University of Otago, New Zealand
Martin Krkosek
Affiliation:
University of Toronto
Julian D. Olden
Affiliation:
University of Washington
Get access

Summary

ECOLOGICAL IMPORTANCE OF RIPARIAN HABITATS TO FRESHWATER ECOSYSTEMS

Riparian areas are the terrestrial environment adjacent to water that both influences and is influenced by the aquatic feature (Gregory et al., 1991; Naiman et al., 2010). Riparian areas along streams provide shade, sources of wood and organic matter, contribute to bank stability, filter sediments, take up excess nutrients from groundwater inputs, and other key processes that protect freshwaters (e.g. Naiman et al., 2010; Richardson & Danehy, 2007; Figure 9.1). Riparian areas also increase biodiversity through habitat complexity and close juxtaposition of aquatic and terrestrial environments (Quinn et al., 2004; Naiman et al., 2010). Alterations to riparian areas, despite their small area relative to the landscape, have disproportionate effects on habitats and fish communities (Naiman et al., 2010; Wipfli & Baxter, 2010). Key habitat losses and alterations are derived from modification of riparian areas by reducing instream habitat complexity (Bilby & Ward, 1989; Fausch & Northcote, 1992; Naiman et al., 2010), diminishing the productive basis of freshwater food webs (Belsky et al., 1999; Quinn et al., 2004), increasing nutrient, contaminant and sediment intrusion (Muscutt et al., 1993; Daniels & Gilliam, 1996; Nguyen et al., 1998; Waters, 1999).

Riparian and freshwater ecosystems are typically tightly coupled, especially in their natural states, and the linkages that couple them frequently exert strong influence on their associated invertebrate and fish fauna (e.g. Gregory et al., 1991; Naiman et al., 2010). Riparian habitats, and the condition of these habitats, further plays a key role in the ecology of these fresh waters, influencing critical processes such as water, nutrient and sediment delivery and dynamics; prey resources for fish and other consumers, and other organic materials exchanged between aquatic and terrestrial habitats (Nakano et al., 1999; Naiman et al., 2010); light and water temperature dynamics that in turn affect food web processes and fish metabolism and growth; aquatic physical habitat (wood); and terrestrial consumers that prey upon fishes (Bisson & Bilby, 1998; Naiman et al., 2010; Wipfli & Baxter, 2010). These processes in turn directly or indirectly influence fishes in freshwater systems (Wang et al., 2001; Pusey & Arthington, 2003; Allan, 2004; Richardson et al., 2010a).

Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2015

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

Agouridis, C. T., Workman, S. R., Warner, R. C. & Jennings, G. D. (2005). Livestock grazing management impacts on stream water quality: a review. Journal of the American Water Resources Association, 41, 591–606.CrossRefGoogle Scholar
Allan, J. D. (2004). Landscapes and riverscapes: the influence of land use on stream ecosystems. Annual Review of Ecology, Evolution, and Systematics, 35, 257–284.CrossRefGoogle Scholar
Allan, J. D., Wipfli, M. S., Caouette, J. P., Prussian, A. & Rodgers, J. (2003). Influence of streamside vegetation on inputs of terrestrial invertebrates to salmonid food webs. Canadian Journal of Fisheries and Aquatic Sciences, 60, 309–320.CrossRefGoogle Scholar
Anderson, N. H., Sedell, J. R., Roberts, L. M. & Triska, F. J. (1978). The role of aquatic invertebrates in processing of wood debris in coniferous forest streams. American Midland Naturalist, 100, 64–82.CrossRefGoogle Scholar
Antón, A., Elosegi, A., García-Arberas, L., Díez, J. & Rallo, A. (2011). Restoration of dead wood in Basque stream channels: effects on brown trout population. Ecology of Freshwater Fish, 20, 461–471.CrossRefGoogle Scholar
Armour, C. L., Duff, D. A. & Elmore, W. (1991). The effects of livestock grazing on riparian and stream ecosystems. Fisheries, 16, 7–11.Google Scholar
Baillie, B. R., Hicks, B. J., Den Heuvel, M. R., Kimberley, M. O. & Hogg, I. D. (2013). The effects of wood on stream habitat and native fish assemblages in New Zealand. Ecology of Freshwater Fish, 22, 553–566.CrossRefGoogle Scholar
Baker, N. E. & Baker, E. (2002). Important Bird Areas in Tanzania: A First Inventory. Dar es Salaam, Tanzania: Wildlife Conservation Society of Tanzania.Google Scholar
Baxter, C. V., Fausch, K. D. & Saunders, W. C. (2005). Tangled webs: reciprocal flows of invertebrate prey link streams and riparian zones. Freshwater Biology, 50, 201–220.CrossRefGoogle Scholar
Beechie, T. J., Sear, D. A., Olden, J. D., et al. (2010). Process-based principles for restoring river ecosystems. Bioscience, 60, 209–222.CrossRefGoogle Scholar
Belsky, A., Matzke, A. & Uselman, S. (1999). Survey of livestock influences on stream and riparian ecosystems in the western United States. Journal of Soil and Water Conservation, 54, 419–431.Google Scholar
Benda, L. E. & Bigelow, P. (2014). On the patterns and processes of wood in northern California streams. Geomorphology, 209, 79–97.CrossRefGoogle Scholar
Benda, L. E. & Sias, J. C. (2003). A quantitative framework for evaluating the mass balance of in-stream organic debris. Forest Ecology and Management, 172, 1–16.CrossRefGoogle Scholar
Benda, L. E., Miller, D. J., Sias, J. C., et al. (2003). Wood recruitment processes and wood budgeting. In The Ecology and Management of Wood in World Rivers. Bethesda, MD: American Fisheries Society, pp. 49–73.Google Scholar
Bergfur, J., Demars, B. O. L., Stutter, M. I., Langan, S. J. & Friberg, N. (2012). The tarland catchment initiative and its effect on stream water quality and macroinvertebrate indices. Journal of Environmental Quality, 41, 314–321.CrossRefGoogle ScholarPubMed
Bilby, R. E. & Ward, J. W. (1989). Changes in characteristics and function of woody debris with increasing size of streams in western Washington. Transactions of the American Fisheries Society, 118, 368–378.2.3.CO;2>CrossRefGoogle Scholar
Bilby, R., Bisson, P. & Naiman, R. (1998). Function and distribution of large woody debris. In River Ecology and Management: Lessons from the Pacific Coastal Ecoregion. New York, NY: Springer-Verlag, pp. 373–398.Google Scholar
Bisson, P. A. & Bilby, R. E. (1998). Organic matter and trophic dynamics. In River Ecology and Management: Lessons from the Pacific Coastal Ecoregion. New York, NY: Springer, pp. 373–398.
Bisson, P. A., Bilby, R. E., Bryant, M. D., et al. (1987). Large woody debris in forested streams in the Pacific Northwest: past, present, and future. In Streamside Management: Forestry and Fishery Interactions. University of Washington, Seattle, WA: Institute of Forest Resources, pp. 143–190.Google Scholar
Bojsen, B. H. & Barriga, R. (2002). Effects of deforestation on fish community structure in Ecuadorian Amazon streams. Freshwater Biology, 47, 2246–2260.CrossRefGoogle Scholar
Bonnington, C., Weaver, D. & Fanning, E. (2008). The habitat preference of four kingfisher species along a branch of the Kilombero River, Southern Tanzania. African Journal of Ecology, 46, 424–427.CrossRefGoogle Scholar
Bugert, R. M., Bjornn, T. C. & Meehan, W. R. (1991). Summer habitat use by young salmonids and their responses to cover and predators in a small southeast Alaska stream. Transactions of the American Fisheries Society, 120, 474–485.Google Scholar
Canhoto, C., Calapez, R., Goncalves, A. L. & Moreira-Santos, M. (2013). Effects of eucalyptus leachates and oxygen on leaf-litter processing by fungi and stream invertebrates. Freshwater Science, 32, 411–424.CrossRefGoogle Scholar
Compton, J. E., Church, M. R., Larned, S. T. & Hogsett, W. E. (2003). Nitrogen export from forested watersheds in the Oregon Coast Range: the role of N2 fixing red alder. Ecosystems, 6, 773–785.CrossRefGoogle Scholar
Comte, L., Buisson, L., Daufresne, M. & Grenouillet, G. (2013). Climate-induced changes in the distribution of freshwater fish: observed and predicted trends. Freshwater Biology, 58, 625–639.CrossRefGoogle Scholar
Correll, D. L. (1998). The role of phosphorus in the eutrophication of receiving waters: a review. Journal of Environmental Quality, 27, 261–266.CrossRefGoogle Scholar
Culp, J. M., Scrimgeour, G. J. & Townsend, G. D. (1996). Simulated fine woody debris accumulations in a stream increase rainbow trout fry abundance. Transactions of the American Fisheries Society, 125, 472–479.2.3.CO;2>CrossRefGoogle Scholar
Czarnomski, N. M., Dreher, D. M., Snyder, K. U., Jones, J. A. & Swanson, F. J. (2008). Dynamics of wood in stream networks of the Western Cascades Range, Oregon. Canadian Journal of Forest Research, 38, 2236–2248.CrossRefGoogle Scholar
Daniels, R. & Gilliam, J. (1996). Sediment and chemical load reduction by grass and riparian filters. Soil Science Society of America Journal, 60, 246–251.CrossRefGoogle Scholar
Davies, J. N. & Boulton, A. J. (2009). Great house, poor food: effects of exotic leaf litter on shredder densities and caddisfly growth in 6 subtropical Australian streams. Journal of the North American Benthological Society, 28, 491–503.CrossRefGoogle Scholar
Dawson, K. (2002). Fish kill events and habitat losses of the Richmond River, NSW Australia: an overview. Journal of Coastal Research, 36, 216–221.CrossRefGoogle Scholar
Décamps, H., Fortune, M., Gazelle, F. & Pautou, G. (1988). Historical influence of man on the riparian dynamics of a fluvial landscape. Landscape Ecology, 1, 163–173.CrossRefGoogle Scholar
Di Giulio, R. T. & Hinton, D. E. (2008). The Toxicology of Fishes. Boca Raton, FL: Taylor and Francis Group.CrossRefGoogle Scholar
Diehl, S. (1988). Foraging efficiency of three freshwater fishes: effects of structural complexity and light. Oikos, 207–214.Google Scholar
Dolloff, C. A. & WarrenJr, M. L. (2003). Fish relationships with large wood in small streams. American Fisheries Society Symposium, 37, 179–193.Google Scholar
Downing, A. S., van Nes, E. H., Mooij, W. M., & Scheffer, M. (2012). The resilience and resistance of an ecosystem to a collapse of diversity. PloS ONE, 7, e46135.CrossRefGoogle ScholarPubMed
Drake, D. C. & Naiman, R. J. (2007). Reconstruction of pacific salmon abundance from riparian tree-ring growth. Ecological Applications, 17, 1523–1542.CrossRefGoogle ScholarPubMed
Dunbar, M. J., Pedersen, M. L., Cadman, D., et al. (2010). River discharge and local-scale physical habitat influence macroinvertebrate LIFE scores. Freshwater Biology, 55, 226–242.CrossRefGoogle Scholar
Eaton, J. G. & Scheller, R. M. (1996). Effects of climate warming on fish thermal habitat in streams of the United States. Limnology and Oceanography, 41, 1109–1115.CrossRefGoogle Scholar
Edwards, E. D. & Huryn, A. D. (1996). Effect of riparian land use on contributions of terrestrial invertebrates to streams. Hydrobiologia, 337, 151–159.CrossRefGoogle Scholar
England, L. E. & Rosemond, A. D. (2004). Small reductions in forest cover weaken terrestrial-aquatic linkages in headwater streams. Freshwater Biology, 49, 721–734.CrossRefGoogle Scholar
Erős, T., Gustafsson, P., Greenberg, L. A. & Bergman, E. (2012). Forest–stream linkages: effects of terrestrial invertebrate input and light on diet and growth of brown trout (Salmo trutta) in a boreal forest stream. PloS ONE, 7, e36462.CrossRefGoogle Scholar
Fagan, W. F. (2002). Connectivity, fragmentation, and extinction risk in dendritic metapopulations. Ecology, 83, 3243–3249.CrossRefGoogle Scholar
Fagan, W. F., Unmack, P. J., Burgess, C., & Minckley, W. L. (2002). Rarity, fragmentation, and extinction risk in desert fishes. Ecology, 83, 3250–3256.CrossRefGoogle Scholar
Fausch, K. D. & Northcote, T. G. (1992). Large woody debris and salmonid habitat in a small coastal British Columbia stream. Canadian Journal of Fisheries and Aquatic Sciences, 49, 682–693.CrossRefGoogle Scholar
Feller, M. C. (2005). Forest harvesting and streamwater inorganic chemistry in western North America: a review. Journal of the American Water Resources Association, 41, 785–811.CrossRefGoogle Scholar
FEMAT (Forest Ecosystem Management Assessment Team). (1993). Forest Ecosystem Management: An Ecological, Economic, and Social Assessment. Portland, OR: US Forest Service, US Department of Commerce, National Oceanic and Atmospheric Administration, National Marine Fisheries Service, United States Bureau of Land Management, & US Fish and Wildlife Service.
Finlay, J. C., Hood, J. M., Limm, M. P., et al. (2011). Light-mediated thresholds in stream-water nutrient composition in a river network. Ecology, 92, 140–150.CrossRefGoogle Scholar
Fisher, S. G. & Likens, G. E. (1973). Energy flow in bear brook, New Hampshire: an integrative approach to stream ecosystem metabolism. Ecological Monographs, 43, 421–439.CrossRefGoogle Scholar
Flecker, A. S. & Taylor, B. W. (2004). Tropical fishes as biological bulldozers: density effects on resource heterogeneity and species diversity. Ecology, 85, 2267–2278.CrossRefGoogle Scholar
Fraser, N. & Metcalfe, N. (1997). The costs of becoming nocturnal: feeding efficiency in relation to light intensity in juvenile Atlantic salmon. Functional Ecology, 11, 385–391.CrossRefGoogle Scholar
Freeman, M. C., Pringle, C. M. & Jackson, C. R. (2007). Hydrologic connectivity and the contribution of stream headwaters to ecological integrity at regional scales. Journal of the American Water Resources Association, 43, 5–14.CrossRefGoogle Scholar
Fyles, J. W. & Fyles, I. H. (1993). Interaction of Douglas-fir with red alder and salal foliage litter during decomposition. Canadian Journal of Forest Research, 23, 358–361.CrossRefGoogle Scholar
Gomi, T., Sidle, R. C. & Richardson, J. S. (2002). Understanding processes and downstream linkages of headwater systems headwaters differ from downstream reaches by their close coupling to hillslope processes, more temporal and spatial variation, and their need for different means of protection from land use. Bioscience, 52, 905–916.Google Scholar
Greenwood, H., O'Dowd, D. J. & Lake, P. S. (2004). Willow (Salix rubens) invasion of the riparian zone in south-eastern Australia: reduced abundance and altered composition of terrestrial arthropods. Diversity and Distributions, 10, 485–492.CrossRefGoogle Scholar
Gregory, S. V., Swanson, F. J., McKee, W. A. & Cummins, K. W. (1991). An ecosystem perspective of riparian zones. Bioscience, 41, 540–551.CrossRefGoogle Scholar
Gurnell, A. M., Piegay, H., Swanson, F. J. & Gregory, S. V. (2002). Large wood and fluvial processes. Freshwater Biology, 47, 601–619.CrossRefGoogle Scholar
Hagen, E. M., Mctammany, M. E., Webster, J. R. & Benfield, E. F. (2010). Shifts in allochthonous input and autochthonous production in streams along an agricultural land-use gradient. Hydrobiologia, 655, 61–77.CrossRefGoogle Scholar
Hill, W. R., Mulholland, P. J. & Marzolf, E. R. (2001). Stream ecosystem responses to forest leaf emergence in spring. Ecology, 82, 2306–2319.CrossRefGoogle Scholar
Hocking, M. D. & Reynolds, J. D. (2011). Impacts of salmon on riparian plant diversity. Science, 331, 1609–1612.CrossRefGoogle ScholarPubMed
Hofer, N. & Richardson, J. S. (2007). Comparisons of the colonisation by invertebrates of three species of wood, alder leaves, and plastic ‘leaves’ in a temperate stream. International Review of Hydrobiology, 92, 647–655.CrossRefGoogle Scholar
Hynes, H. B. N. (1975). The stream and its valley. Verhandlungen des Internationalen Verein Limnologie, 19, 1–15.Google Scholar
Johnson, L. B., Richards, C., Host, G. E. & Arthur, J. W. (1997). Landscape influences on water chemistry in midwestern stream ecosystems. Freshwater Biology, 37, 193–208.CrossRefGoogle Scholar
Johnson, L. B., Breneman, D. H. & Richards, C. (2003). Macroinvertebrate community structure and function associated with large wood in low gradient streams. River Research and Applications, 19, 199–218.CrossRefGoogle Scholar
Jones, J. I., Murphy, J. F., Collins, A. L., et al. (2012). The impact of fine sediment on macroinvertebrates. River Research and Applications, 28, 1055–1071.Google Scholar
Jönsson, M., Hylander, S., Ranåker, L., Nilsson, P. A. & Brönmark, C. (2011). Foraging success of juvenile pike Esox lucius depends on visual conditions and prey pigmentation. Journal of Fish Biology, 79, 290–297.CrossRefGoogle ScholarPubMed
Kail, J., Hering, D., Muhar, S., Gerhard, M. & Preis, S. (2007). The use of large wood in stream restoration: experiences from 50 projects in Germany and Austria. Journal of Applied Ecology, 44, 1145–1155.CrossRefGoogle Scholar
Kano, Y., Miyazaki, Y., Tomiyama, Y., et al. (2013). Linking mesohabitat selection and ecological traits of a fish assemblage in a small tropical stream (Tinggi River, Pahang Basin) of the Malay Peninsula. Zoological Science, 30, 178–184.CrossRefGoogle Scholar
Katano, O., Aonuma, Y., Nakamura, T. & Yamamoto, S. (2003). Indirect contramensalism through trophic cascades between two omnivorous fishes. Ecology, 84, 1311–1323.CrossRefGoogle Scholar
Kawaguchi, Y. & Nakano, S. (2001). Contribution of terrestrial invertebrates to the annual resource budget for salmonids in forest and grassland reaches of a headwater stream. Freshwater Biology, 46, 303–316.CrossRefGoogle Scholar
Kawaguchi, Y., Taniguchi, Y. & Nakano, S. (2003). Terrestrial invertebrate inputs determine the local abundance of stream fishes in a forested stream. Ecology, 84, 701–708.CrossRefGoogle Scholar
Kiffney, P. M., Richardson, J. S. & Bull, J. P. (2003). Responses of periphyton and insects to experimental manipulation of riparian buffer width along forest streams. Journal of Applied Ecology, 40, 1060–1076.CrossRefGoogle Scholar
Kiffney, P. M., Richardson, J. S. & Bull, J. P. (2004). Establishing light as a causal mechanism structuring stream communities in response to experimental manipulation of riparian buffer width. Journal of the North American Benthological Society, 23, 542–555.2.0.CO;2>CrossRefGoogle Scholar
Kirby, J. S., Holmes, J. S. & Sellers, R. M. (1996). Cormorants Phalacrocorax carbo as fish predators: an appraisal of their conservation and management in Great Britain. Biological Conservation, 75, 191–199.CrossRefGoogle Scholar
Kominoski, J. S., Shah, J. J. F., Canhoto, C., et al. (2013). Forecasting functional implications of global changes in riparian plant communities. Frontiers in Ecology and the Environment, 11, 423–432.CrossRefGoogle Scholar
Langford, T. E. L., Langford, J. & Hawkins, S. J. (2012). Conflicting effects of woody debris on stream fish populations: implications for management. Freshwater Biology, 57, 1096–1111.CrossRefGoogle Scholar
Lapointe, N. W. R., Cooke, S. J., Imhof, J. G., et al. (2014). Principles for ensuring healthy and productive freshwater ecosystems that support sustainable fisheries. Environmental Reviews, 22, 110–134.CrossRefGoogle Scholar
Leach, J. A., Moore, R. D., Hinch, S. G. & Gomi, T. (2012).Estimation of forest harvesting-induced stream temperature changes and bioenergetic consequences for cutthroat trout in a coastal stream in British Columbia, Canada. Aquatic Sciences, 74, 427–441.CrossRefGoogle Scholar
Lecerf, A., Patfield, D., Boiche, A., et al. (2007). Stream ecosystems respond to riparian invasion by Japanese knotweed (Fallopia japonica). Canadian Journal of Fisheries and Aquatic Sciences, 64, 1273–1283.CrossRefGoogle Scholar
Ledger, M. E. & Hildrew, A. G. (2001). Growth of an acid-tolerant stonefly on epilithic biofilms from streams of contrasting pH. Freshwater Biology, 46, 1457–1470.CrossRefGoogle Scholar
Lee, P., Smyth, C. & Boutin, S. (2004). Quantitative review of riparian buffer width guidelines from Canada and the United States. Journal of Environmental Management, 70, 165–180.CrossRefGoogle ScholarPubMed
Leopold, L. B. & O'Brien Marchand, M. (1968). On the quantitative inventory of the riverscape. Water Resources Research, 4, 709–717.CrossRefGoogle Scholar
Leroy, C. J. & Marks, J. C. (2006). Litter quality, stream characteristics and litter diversity influence decomposition rates and macroinvertebrates. Freshwater Biology, 51, 605–617.CrossRefGoogle Scholar
Lodge, D. M., Williams, S., MacIsaac, H. J., et al. (2006). Biological invasions: recommendations for US policy and management. Ecological Applications, 16, 2035–2054.CrossRefGoogle Scholar
Lovell, S. T. & Sullivan, W. C. (2006). Environmental benefits of conservation buffers in the united states: evidence, promise, and open questions. Agriculture, Ecosystems & Environment, 112, 249–260.CrossRefGoogle Scholar
Lu, Y. H., Bauer, J. E., Canuel, E. A., et al. (2014). Effects of land use on sources and ages of inorganic and organic carbon in temperate headwater streams. Biogeochemistry, 119, 275–292.CrossRefGoogle Scholar
Maser, C. & Sedell, J. R. (1994). From the Forest to the Sea: The Ecology of Wood in Streams, Rivers, Estuaries, and Oceans. Delray Beach, FL: St. Lucie Press.Google Scholar
Mason, C. & MacDonald, S. (1982). The input of terrestrial invertebrates from tree canopies to a stream. Freshwater Biology, 12, 305–311.CrossRefGoogle Scholar
Mathewson, D., Hocking, M. & Reimchen, T. (2003). Nitrogen uptake in riparian plant communities across a sharp ecological boundary of salmon density. BMC Ecology, 3, 4.CrossRefGoogle ScholarPubMed
May, C. L. & Gresswell, R. E. (2003). Processes and rates of sediment and wood accumulation in headwater streams of the Oregon Coast Range, USA. Earth Surface Processes and Landforms, 28, 409–424.CrossRefGoogle Scholar
McGlynn, B. L. & McDonnell, J. J. (2003). Quantifying the relative contributions of riparian and hillslope zones to catchment runoff. Water Resources Research, 39, 1310.CrossRefGoogle Scholar
McKay, S. K., Schramski, J. R., Conyngham, J. N. & Fischenich, J. C. (2013). Assessing upstream fish passage connectivity with network analysis. Ecological Applications, 23, 1396–1409.CrossRefGoogle ScholarPubMed
Meador, M. R. & Goldstein, R. M. (2003). Assessing water quality at large geographic scales: relations among land use, water physicochemistry, riparian condition, and fish community structure. Environmental Management, 31, 504–517.CrossRefGoogle ScholarPubMed
Millar, R. G. (2000). Influence of bank vegetation on alluvial channel patterns. Water Resources Research, 36, 1109–1118.CrossRefGoogle Scholar
Miller, J., Chanasyk, D., Curtis, T., Entz, T. & Willms, W. (2010). Influence of streambank fencing with a cattle crossing on riparian health and water quality of the lower Little Bow River in southern Alberta, Canada. Agricultural Water Management, 97, 247–258.CrossRefGoogle Scholar
Monadjem, A. (1996). Habitat associations of birds along the Sabie River, South Africa. African Journal of Ecology, 34, 75–78.CrossRefGoogle Scholar
Montgomery, D. R., Collins, B. D., Buffington, J. M. & Abbe, T. B. (2003). Geomorphic effects of wood in rivers. American Fisheries Society Symposium, 37, 21–47.Google Scholar
Moore, R. D. & Wondzell, S. M. (2005). Physical hydrology and the effects of forest harvesting in the Pacific Northwest: a review. Journal of the American Water Resources Association, 41, 763–784.CrossRefGoogle Scholar
Moore, R. D., Spittlehouse, D. L. & Story, A. (2005). Riparian microclimate and stream temperature response to forest harvesting: a review. Journal of the American Water Resources Association, 41, 813–834.CrossRefGoogle Scholar
Morgan, R. P. C. (2009). Soil Erosion and Conservation. Malden, MA: Wiley-Blackwell.Google Scholar
Motomori, K., Mitsuhashi, H. & Nakano, S. (2001). Influence of leaf litter quality on the colonization and consumption of stream invertebrate shredders. Ecological Research, 16, 173–182.CrossRefGoogle Scholar
Muscutt, A., Harris, G., Bailey, S. & Davies, D. (1993). Buffer zones to improve water quality: a review of their potential use in UK agriculture. Agriculture, Ecosystems & Environment, 45, 59–77.CrossRefGoogle Scholar
Naiman, R. J., Decamps, H. & McClain, M. E. (2010). Riparia: Ecology, Conservation, and Management of Streamside Communities. Burlington, MA: Academic Press.Google Scholar
Nakano, S., Miyasaka, H. & Kuhara, N. (1999). Terrestrial–aquatic linkages: riparian arthropod inputs alter trophic cascades in a stream food web. Ecology, 80, 2435–2441.Google Scholar
Neal, C., Ormerod, S. J., Langan, S. J., Nisbet, T. R. & Roberts, J. (2004). Sustainability of UK forestry: contemporary issues for the protection of freshwaters, a conclusion. Hydrology and Earth System Sciences, 8, 589–595.CrossRefGoogle Scholar
Nguyen, M., Sheath, G., Smith, C. & Cooper, A. (1998). Impact of cattle treading on hill land: 2. Soil physical properties and contaminant runoff. New Zealand Journal of Agricultural Research, 41, 279–290.CrossRefGoogle Scholar
Opperman, J. J., Galloway, G. E., Fargione, J., et al. (2009). Sustainable floodplains through large-scale reconnection to rivers. Science, 326, 1487–1488.CrossRefGoogle ScholarPubMed
Parmesan, C. & Yohe, G. (2003). A globally coherent fingerprint of climate change impacts across natural systems. Nature, 421, 37–42.CrossRefGoogle ScholarPubMed
Parris, K. M. & Lindenmayer, D. B. (2004). Evidence that creation of a Pinus radiata plantation in south-eastern Australia has reduced habitat for frogs. Acta Oecologica, 25, 93–101.CrossRefGoogle Scholar
Pejchar, L. & Mooney, H. A. (2009). Invasive species, ecosystem services and human well-being. Trends in Ecology & Evolution, 24, 497–504.CrossRefGoogle ScholarPubMed
Perna, C. N., Cappo, M., Pusey, B. J., Burrows, D. W. & Pearson, R. G. (2012). Removal of aquatic weeds greatly enhances fish community richness and diversity: an example from the Burdekin River Floodplain, tropical Australia. River Research and Applications, 28, 1093–1104.CrossRefGoogle Scholar
Peterson, D. P., Rieman, B. E., Young, M. K. & Brammer, J. A. (2010). Modeling predicts that redd trampling by cattle may contribute to population declines of native trout. Ecological Applications, 20, 954–966.CrossRefGoogle ScholarPubMed
Piccolo, J. J. & Wipfli, M. S. (2002). Does red alder (Alnus rubra) in upland riparian forests elevate macroinvertebrate and detritus export from headwater streams to downstream habitats in southeastern Alaska?Canadian Journal of Fisheries and Aquatic Sciences, 59, 503–513.CrossRefGoogle Scholar
Pollock, M. M., Pess, G. R. & Beechie, T. J. (2004). The importance of beaver ponds to coho salmon production in the Stillaguamish River Basin, Washington, USA. North American Journal of Fisheries Management, 24, 749–760.CrossRefGoogle Scholar
Poole, G. C. & Berman, C. H. (2001). An ecological perspective on in-stream temperature: natural heat dynamics and mechanisms of human-caused thermal degradation. Environmental Management, 27, 787–802.CrossRefGoogle ScholarPubMed
Power, M. E. (2003). Life cycles, limiting factors, and behavioral ecology of four loricariid catfishes in a Panamanian stream. Catfishes, 2, 581–600.Google Scholar
Pringle, C. M. (2003). What is hydrologic connectivity and why is it ecologically important?Hydrological Processes, 17, 2685–2689.CrossRefGoogle Scholar
Pringle, C. M. & Hamazaki, T. (1997). Effects of fishes on algal response to storms in a tropical stream. Ecology, 78, 2432–2442.Google Scholar
Pusey, B. J. & Arthington, A. H. (2003). Importance of the riparian zone to the conservation and management of freshwater fish: a review. Marine and Freshwater Research, 54, 1–16.CrossRefGoogle Scholar
Quinn, J. M., Boothroyd, I. K. & Smith, B. J. (2004). Riparian buffers mitigate effects of pine plantation logging on New Zealand streams: 2. Invertebrate communities. Forest Ecology and Management, 191, 129–146.CrossRefGoogle Scholar
Rader, R. B., Belish, T., Young, M. K. & Rothlisberger, J. (2007). The scotopic visual sensitivity of four species of trout: a comparative study. Western North American Naturalist, 67, 524–537.CrossRefGoogle Scholar
Ranganath, S. C., Hession, W. C. & Wynn, T. M. (2009). Livestock exclusion influences on riparian vegetation, channel morphology, and benthic macroinvertebrate assemblages. Journal of Soil and Water Conservation, 64, 33–42.CrossRefGoogle Scholar
Reid, M. A., Ogden, R. & Thoms, M. C. (2011). The influence of flood frequency, geomorphic setting and grazing on plant communities and plant biomass on a large dryland floodplain. Journal of Arid Environments, 75, 815–826.CrossRefGoogle Scholar
Reimchen, T., Mathewson, D., Hocking, M., Moran, J. & Harris, D. (2003). Isotopic evidence for enrichment of salmon-derived nutrients in vegetation, soil, and insects in riparian zones in coastal British Columbia. American Fisheries Society Symposium, 34, 59–70.Google Scholar
Richardson, J. S. & Danehy, R. J. (2007). A synthesis of the ecology of headwater streams and their riparian zones in temperate forests. Forest Science, 53, 131–147.Google Scholar
Richardson, J. S., Shaughnessy, C. R. & Harrison, P. G. (2004). Litter breakdown and invertebrate association with three types of leaves in a temperate rainforest stream. Archiv für Hydrobiologie, 159, 309–325.CrossRefGoogle Scholar
Richardson, J. S., Bilby, R. E. & Bondar, C. A. (2005). Organic matter dynamics in small streams of the Pacific Northwest. Journal of the American Water Resources Association, 41, 921–934.CrossRefGoogle Scholar
Richardson, J. S., Taylor, E., Schluter, D., Pearson, M. & Hatfield, T. (2010a). Do riparian zones qualify as critical habitat for endangered freshwater fishes?Canadian Journal of Fisheries and Aquatic Sciences, 67, 1197–1204.CrossRefGoogle Scholar
Richardson, J. S., Zhang, Y. X. & Marczak, L. B. (2010b). Resource subsidies across the land–freshwater interface and responses in recipient communities. River Research and Applications, 26, 55–66.CrossRefGoogle Scholar
Richardson, J. S., Naiman, R. J. & Bisson, P. A. (2012). How did fixed-width buffers become standard practice for protecting freshwaters and their riparian areas from forest harvest practices?Freshwater Science, 31, 232–238.CrossRefGoogle Scholar
Rier, S. T. & Stevenson, R. J. (2002). Effects of light, dissolved organic carbon, and inorganic nutrients on the relationship between algae and heterotrophic bacteria in stream periphyton. Hydrobiologia, 489, 179–184.CrossRefGoogle Scholar
Roni, P. & Beechie, T. (2013). Introduction to restoration: key steps for designing effective programs and projects. In Stream and Watershed Restoration: A Guide to Restoring Riverine Processes and Habitats. Chichester: Wiley-Blackwell, pp. 1–10.Google Scholar
Roon, D. A. (2011). Ecological effects of invasive European birdcherry on salmonid food webs in Alaska streams. Masters of Science Thesis. Fairbanks, AK: University of Alaska Fairbanks.
Rosenberg, D. M. & Resh, V. H. (1993). Freshwater Biomonitoring and Benthic Macroinvertebrates. New York, NY: Chapman & Hall.Google Scholar
Sakamaki, T. & Richardson, J. S. (2013). Nonlinear variation of stream–forest linkage along a stream-size gradient: an assessment using biogeochemical proxies of in-stream fine particulate organic matter. Journal of Applied Ecology, 50, 1019–1027.CrossRefGoogle Scholar
Sarr, D. A. (2002). Riparian livestock exclosure research in the western United States: a critique and some recommendations. Environmental Management, 30, 516–526.CrossRefGoogle ScholarPubMed
Sato, T., Watanabe, K., Kanaiwa, M., et al. (2011). Nematomorph parasites drive energy flow through a riparian ecosystem. Ecology, 92, 201–207.CrossRefGoogle ScholarPubMed
Sayer, E. J. (2006). Using experimental manipulation to assess the roles of leaf litter in the functioning of forest ecosystems. Biological Reviews, 81, 1–31.Google ScholarPubMed
Schindler, D. E., Hilborn, R., Chasco, B., et al. (2010). Population diversity and the portfolio effect in an exploited species. Nature, 465, 609–612.CrossRefGoogle Scholar
Schlaepfer, M. A., Sax, D. F. & Olden, J. D. (2011). The potential conservation value of non-native species. Conservation Biology, 25, 428–437.CrossRefGoogle ScholarPubMed
Sepúlveda, M. A., Bartheld, J. L., Monsalve, R., Gómez, V. & Medina-Vogel, G. (2007). Habitat use and spatial behaviour of the endangered southern river otter (Lontra provocax) in riparian habitats of Chile: conservation implications. Biological Conservation, 140, 329–338.CrossRefGoogle Scholar
Spänhoff, B. & Meyer, E. I. (2004). Breakdown rates of wood in streams. Journal of the North American Benthological Society, 23, 189–197.2.0.CO;2>CrossRefGoogle Scholar
Stevens, C. E., Paszkowski, C. A. & Foote, A. L. (2007). Beaver (Castor canadensis) as a surrogate species for conserving anuran amphibians on boreal streams in Alberta, Canada. Biological Conservation, 134, 1–13.CrossRefGoogle Scholar
Strayer, D. L. (2010). Alien species in fresh waters: ecological effects, interactions with other stressors, and prospects for the future. Freshwater Biology, 55, 152–174.CrossRefGoogle Scholar
Ström, L., Jansson, R., Nilsson, C., Johansson, M. E. & Xiong, S. (2011). Hydrologic effects on riparian vegetation in a boreal river: an experiment testing climate change predictions. Global Change Biology, 17, 254–267.CrossRefGoogle Scholar
Suttle, K. B., Power, M. E., Levine, J. M. & McNeely, C. (2004). How fine sediment in riverbeds impairs growth and survival of juvenile salmonids. Ecological Applications, 14, 969–974.CrossRefGoogle Scholar
Swanson, F. J., Gregory, S., Sedell, J. & Campbell, A. (1982). Land–water interactions: the riparian zone. In Analysis of Coniferous Forest Ecosystems in the Western United States. Stroudsburg, PA: Hutchinson Ross Publishing Co., pp. 267–291.Google Scholar
Tabacchi, E., Lambs, L., Guilloy, H., et al. (2000). Impacts of riparian vegetation on hydrological processes. Hydrological Processes, 14, 2959–2976.3.0.CO;2-B>CrossRefGoogle Scholar
Thoms, M. C. (2003). Floodplain–river ecosystems: lateral connections and the implications of human interference. Geomorphology, 56, 335–349.CrossRefGoogle Scholar
Tockner, K. & Stanford, J. A. (2002). Riverine flood plains: present state and future trends. Environmental Conservation, 29, 308–330.CrossRefGoogle Scholar
Vannote, R. L., Minshall, G. W., Cummins, K. W., Sedell, J. R. & Cushing, C. E. (1980). The river continuum concept. Canadian Journal of Fisheries and Aquatic Sciences, 37, 130–137.CrossRefGoogle Scholar
Wagenhoff, A., Townsend, C. R., Phillips, N. & Matthaei, C. D. (2011). Subsidy-stress and multiple-stressor effects along gradients of deposited fine sediment and dissolved nutrients in a regional set of streams and rivers. Freshwater Biology, 56, 1916–1936.CrossRefGoogle Scholar
Wallace, J. B. & Benke, A. C. (1984). Quantification of wood habitat in sub-tropical coastal-plain streams. Canadian Journal of Fisheries and Aquatic Sciences, 41, 1643–1652.CrossRefGoogle Scholar
Wallace, J. B., Eggert, S. L., Meyer, J. L. & Webster, J. R. (1997). Multiple trophic levels of a forest stream linked to terrestrial litter inputs. Science, 277, 102–104.CrossRefGoogle Scholar
Wang, L., Lyons, J., Kanehl, P. & Bannerman, R. (2001). Impacts of urbanization on stream habitat and fish across multiple spatial scales. Environmental Management, 28, 255–266.CrossRefGoogle ScholarPubMed
Waters, T. F. (1999). Long-term trout production dynamics in Valley Creek, Minnesota. Transactions of the American Fisheries Society, 128, 1151–1162.2.0.CO;2>CrossRefGoogle Scholar
Webster, J. R., Benfield, E. F., Ehrman, T. P., et al. (1999). What happens to allochthonous material that falls into streams? A synthesis of new and published information from Coweeta. Freshwater Biology, 41, 687–705.CrossRefGoogle Scholar
Webster, J. R., Morkeski, K., Wojculewski, C. A., et al. (2012). Effects of hemlock mortality on streams in the southern Appalachian Mountains. American Midland Naturalist, 168, 112–131.CrossRefGoogle Scholar
Wellington, C. G., Mayer, C. M., Bossenbroek, J. M. & Stroh, N. A. (2010). Effects of turbidity and prey density on the foraging success of age 0 year yellow perch Perca flavescens. Journal of Fish Biology, 76, 1729–1741.CrossRefGoogle Scholar
Wetzel, R. G. & Likens, G. E. (2000). Inorganic nutrients: nitrogen, phosphorus, and other nutrients. In Limnological Analyses. New York, NY: Springer Science, pp. 85–112.CrossRefGoogle Scholar
Whiteway, S. L., Biron, P. M., Zimmermann, A., Venter, O. & Grant, J. W. A. (2010). Do in-stream restoration structures enhance salmonid abundance? A meta-analysis. Canadian Journal of Fisheries and Aquatic Sciences, 67, 831–841.CrossRefGoogle Scholar
Wilzbach, M. A., Cummins, K. W. & Hall, J. D. (1986). Influence of habitat manipulations on interactions between cutthroat trout and invertebrate drift. Ecology, 67, 898–911.CrossRefGoogle Scholar
Winterbourn, M. (1990). Interactions among nutrients, algae and invertebrates in a New Zealand mountain stream. Freshwater Biology, 23, 463–474.CrossRefGoogle Scholar
Wipfli, M. S. (1997). Terrestrial invertebrates as salmonid prey and nitrogen sources in streams: contrasting old-growth and young-growth riparian forests in southeastern Alaska, USA. Canadian Journal of Fisheries and Aquatic Sciences, 54, 1259–1269.CrossRefGoogle Scholar
Wipfli, M. S. & Baxter, C. V. (2010). Linking ecosystems, food webs, and fish production: subsidies in salmonid watersheds. Fisheries, 35, 373–387.CrossRefGoogle Scholar
Wipfli, M. S. & Gregovich, D. P. (2002). Export of invertebrates and detritus from fishless headwater streams in southeastern Alaska: implications for downstream salmonid production. Freshwater Biology, 47, 957–969.CrossRefGoogle Scholar
Wipfli, M. S. & Musslewhite, J. (2004). Density of red alder (Alnus rubra) in headwaters influences invertebrate and detritus subsidies to downstream fish habitats in Alaska. Hydrobiologia, 520, 153–163.CrossRefGoogle Scholar
Wipfli, M. S., Richardson, J. S. & Naiman, R. J. (2007). Ecological linkages between headwaters and downstream ecosystems: transport of organic matter, invertebrates, and wood down headwater channels. Journal of the American Water Resources Association, 43, 72–85.CrossRefGoogle Scholar
Wright, J. P. & Flecker, A. S. (2004). Deforesting the riverscape: the effects of wood on fish diversity in a Venezuelan piedmont stream. Biological Conservation, 120, 439–447.CrossRefGoogle Scholar
Wright, J. P., Jones, C. G. & Flecker, A. S. (2002). An ecosystem engineer, the beaver, increases species richness at the landscape scale. Oecologia, 132, 96–101.CrossRefGoogle ScholarPubMed
Young, R. A. & Wiersma, J. (1973). The role of rainfall impact in soil detachment and transport. Water Resources Research, 9, 1629–1636.CrossRefGoogle Scholar
Zeigler, M. P., Brinkman, S. F., Caldwell, C. A., et al. (2013). Upper thermal tolerances of Rio Grande cutthroat trout under constant and fluctuating temperatures. Transactions of the American Fisheries Society, 142, 1395–1405.CrossRefGoogle 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
×