Effects of climate change on hydro-ecology of contributing basins in the Arctic

From The Encyclopedia of Earth
Jump to: navigation, search


February 9, 2010, 3:24 pm
May 7, 2012, 1:03 pm
This is Section 8.4.2 of the Arctic Climate Impact Assessment
Lead Authors: Frederick J.Wrona,Terry D. Prowse, James D. Reist; Contributing Authors: Richard Beamish, John J. Gibson, John Hobbie, Erik Jeppesen, Jackie King, Guenter Koeck, Atte Korhola, Lucie Lévesque, Robie Macdonald, Michael Power,Vladimir Skvortsov,Warwick Vincent; Consulting Authors: Robert Clark, Brian Dempson, David Lean, Hannu Lehtonen, Sofia Perin, Richard Pienitz, Milla Rautio, John Smol, Ross Tallman, Alexander Zhulidov

The regional patterns of projected changes in temperature and precipitation reviewed in the previous section are useful to understand some of the broad-scale effects that may occur. Specific effects, however, will be much more diverse and complex, even within regions of similar temperature and precipitation changes, because of intra-regional heterogeneity in freshwater systems and the surrounding landscapes that affect them. For example, elevational difference is one physical factor that will produce a complex altered pattern of snow (Snow cover in the Arctic) storage and runoff. Although warmer conditions are very likely to reduce the length of winter, snow accumulation could either decrease or increase, with the latter most likely to occur in higher-elevation zones where enhanced storm activity combined with orographic effects will probably increase winter snowfall. Increased accumulation is likely to be most pronounced at very high elevations above the elevated freezing level, where the summer season is likely to remain devoid of major melt events, thereby creating the conditions for the preservation of more semi-permanent snowpacks at high altitudes[1]. In contrast, temperature increases at lower elevations, especially in the more temperate maritime zones, are likely to increase rainfall and rain-on-snow runoff events. Snow patterns will be affected by a number of other factors, including vegetation, which is also projected to be altered by climate change (Section 7.5.3.2 (Effects of climate change on hydro-ecology of contributing basins in the Arctic)). For example, shifts from tundra vegetation to trees have led to greater snow interception and subsequent losses through sublimation[2], whereas shifts from tundra to shrubs have been shown to reduce snow losses[3], thereby affecting the magnitude of the snowpack available for spring melt.

An advance of the spring warming period means that snowmelt will occur during a period of lower insolation, which, other things being equal, will lead to a more protracted melt and less intense runoff. Traditional ecological knowledge indicates that through much of northern Canada, including the western Canadian Arctic and Nunavut, spring melt is already occurring earlier than in the past, and spring air temperatures are higher[4], although observations near eastern Hudson Bay indicate a delay in the initiation of spring melt[5]. See Chapter 3 (Effects of climate change on hydro-ecology of contributing basins in the Arctic) for local accounts of such changes in the Arctic. The effects of early and less intense spring melt will be most dramatic for catchments wholly contained within the northern latitudes, where snowmelt forms the major and sometimes only flow event of the year. Reductions in the spring peak will be accentuated where the loss of permafrost through associated warming increases the capacity to store runoff, although there will also be a compensating increase in summer base flow. Overall, the magnitude and frequency of high flows will decline while low flows will increase, thereby flattening the annual hydrograph. This impact is similar to that observed as a result of river regulation, and hence will tend to compound such effects.

Loss of permafrost or deepening of the active layer (seasonal melt depth; see Section 6.6.1 (Effects of climate change on hydro-ecology of contributing basins in the Arctic) for changes in permafrost) will also reduce the peak response to rainfall events in summer, increase infiltration, and promote groundwater flow. This is consistent with the analogue of northern basins where those with less permafrost but receiving comparable amounts of precipitation have a lowered and smaller range of discharge[6]. Changes in the rate of evapotranspiration (Precipitation and evapotranspiration in the Arctic) and its seasonal duration will also directly affect stream runoff from permafrost (Permafrost in the Arctic) basins. As suggested by the modeling results of Hinzman and Kane[7] for areas of Alaska, the greatest reduction in summer runoff is likely to occur in years experiencing light, uniformly spaced rainfall events whereas in years characterized by major rainfalls comprising most of the summer precipitation, total runoff volume is likely to be affected least.

Changes in the water balance will vary by regional climate and surface conditions, but particular areas and features are believed to be especially sensitive to such alterations. Such is the case for the unglaciated lowlands of many arctic islands where special ecological niches, such as found at Polar Bear Pass on Bathhurst Island or Truelove Lowland on Devon Island, are produced by unique hydro-climatic regimes and are largely dependent on ponded water produced by spring snowmelt. On a broad scale, arctic islands and coastal areas are likely to experience significant changes in local microclimates that will probably affect water balance components, especially evaporation rates. Here, longer open-water seasons in the adjacent marine environments are likely to enhance the formation of fog and low clouds and reduce associated solar radiation. Increased water vapor and lower energy flux would thereby offset any potential increase in evaporation resulting from higher air [[temperature]s][8]).

Large regional differences in water balance will also occur because of differences in plant communities (see also Section 7.4.1 (Effects of climate change on hydro-ecology of contributing basins in the Arctic)). For example, surface drying of open tundra is restricted when non-transpiring mosses and lichens overlie the tundra. Over the longer term, a longer growing season combined with a northward expansion of more shrubs and trees will very probably increase evapotranspiration. Quite a different situation is very likely to exist over the multitude of wetlands that occupy so much of the northern terrain. Although evaporation is inhibited after initial surface drying on those wetlands covered by sphagnum moss or lichen, evapotranspiration continues throughout the summer in wetlands occupied by vascular plants over porous peat soils, and only slows as the water table declines. Higher summer temperatures have the ability to dry such wetlands to greater depths, but their overall storage conditions will depend on changes in other water balance components, particularly snowmelt and rainfall inputs.

As the active layer deepens and more unfrozen flow pathways develop in the permafrost (Permafrost in the Arctic), an enhancement of geochemical weathering and nutrient release is very likely (e.g., phosphorus[9]; see also Section 6.6.1.3 (Effects of climate change on hydro-ecology of contributing basins in the Arctic)). Ultimately, this is very likely to affect productivity in arctic freshwater systems such as Toolik Lake, Alaska (Box 8.3). In the short term, the chemical composition of surface runoff and groundwater flows is very likely to change. In addition, suspended sediment loads will very probably increase as a result of thermokarst erosion, particularly in ice-rich locations. Suspended sediment and nutrient loading of northern freshwater systems will probably also increase as land subsidence, slumping, and landslides increase with permafrost degradation, as traditional ecological knowledge has documented in the western Canadian Arctic where the depth of the active layer has increased[10]. Thermokarst erosion is very likely to continue until at least the large near-surface ice deposits are depleted and new surface flow patterns stabilize. Such fluvial-morphological adjustment is likely to be very lengthy, of the order of hundreds of years, considering the time that has been estimated for some northern rivers to reach a new equilibrium after experiencing a major shift in their suspended-sediment regimes[11]. A major reason for such a protracted period is the time it takes for new vegetation to colonize and stabilize the channel landforms. The stabilization that will occur in the Arctic under climate change is further complicated by the projected change in vegetation regimes, particularly the northward advance of shrubs and trees (see [[Section 7.5.3.2 (Effects of climate change on hydro-ecology of contributing basins in the Arctic)]2]). Such vegetation shifts will cause further changes in stream water chemistry by altering dissolved organic carbon (DOC) concentrations. Current data indicate that DOC is negatively correlated with latitude[12] and decreases with distance from treeline[13] and along gradients from boreal forest (Arctic boreal forest environments) to tundra[14]. Hence, as vegetation shifts from mosses and lichens to grasses and woody species, runoff is very likely to contain increasing concentrations of DOC and particulate detrital material. Verification of enhanced DOC supply associated with northward treeline advance is provided by various paleolimnological and paleoclimatic studies[15]. Although such increases will be long-term, given the slow rates of major vegetation shifts (see also [[Section 7.5.3.2 (Effects of climate change on hydro-ecology of contributing basins in the Arctic)]3]), earlier increases in DOC and DIC are very likely to result from the earlier thermal and mechanical erosion of the permafrost landscape (see also [[Section 6.6.1.3 (Effects of climate change on hydro-ecology of contributing basins in the Arctic)]2]). Zepp et al.[16] and Häder et al.[17] provided comprehensive reviews of the projected interactive effects of changes in |ultra-violet (UV) radiation levels and climate on DOC and dissolved inorganic carbon (DIC) and related aquatic biogeochemical cycles.

Box 8.3. Ecological transitions in Toolik Lake, Alaska, in the face of changing climate and catchment characteristics

Toolik Lake (maximum depth 25 m, area 1.5 km2) lies in the foothills north of the Brooks Range, Alaska, at 68º N, 149º W. The river study site is the headwaters of the Kuparuk River. Details of the research project and related publications are available on the Arctic Long-Term Ecological Research site.The mean annual temperature of the area is -9 ºC, and annual precipitation is approximately 300 to 400 mm. Permafrost is 200 m thick with an active layer up to 46 cm deep. Acidic tussock tundra covers the hillslopes. Sedges dominate a small area of wetlands in the study site, while the dry uplands have a cover of lichens and heaths. Lakes and streams are ultra-oligotrophic, and are ice-free from July to September with strong summer stratification and oxygen saturation. Stream flow is nival, and carries DOC-enriched spring runoff from peaty catchments to Toolik Lake. Primary producers in Toolik Lake consist of 136 species of phytoplankton, dominated by chrysophytes with dinoflagellates and cryptophytes, as well as diatoms. Annual primary productivity averages 12 g C/m2 and is co-limited by nitrogen and phosphorus. Zooplankton are sparse. Fish species are lake trout, Arctic grayling, round whitefish (Prosopium cylindraceum), burbot, and slimy sculpin (Cottus cognatus), which feed on benthic chironomid larvae and snails, the latter controlling epilithic algae in the lake. Dissolved organic carbon drives microbial productivity (5–8 g C/m2/yr).

Figure for Box 8.3. Bacterial productivity and chlorophyll over the spring and summer of 2002 in Toolik Lake (averaged over depths of 3 to 12 m; adapted from Crump et al., 2003).

The average air temperature of northern Alaska has increased by nearly 2 ºC over the past 30 years. Warming of Alaskan waters will possibly have a detrimental effect on adult grayling, which grow best during cool and wet summers and which may actually lose
weight during warm and dry summers[18]. Approximately 20 km from Toolik[19], permafrost temperatures at 20 m depth increased from -5.5 to -4.5 ºC between 1991 and 2000. This warming of frozen soils probably accounts for recent increases in stream- and lake-water alkalinity.

Climate has been shown to have a significant control on the vegetation of the site, which in turn has affected aquatic resources for productivity. Runoff from thawing soils within the catchment of Toolik Lake has affected lake productivity in a number of ways. Dissolved organic carbon from excretion, leaching, and decomposition of plants in the catchment, along with associated humic materials, has been found to reduce photosynthesis in the lake and absorb 99% of the UV-B radiation in the upper 20 cm[20]. In spring, meltwater carries terrestrially derived DOC and abundant nutrients. Upon reaching the lake, meltwater flows cause a two-week high in bacterial productivity (~50% of the annual total) beneath the lake-ice cover (see figure).This peak in production, which takes place at 2 ºC, illustrates that bacteria are carbon- and energy-limited, not temperature-limited, and as such will be indirectly affected by climate change. Phytoplankton biomass and primary production peak soon after the ice leaves the lake, and as solar radiation peaks.The lake stratifies so rapidly that no spring turnover occurs, causing oxygen-depleted bottom waters to persist over the summer.This effect is very likely to be amplified with higher temperatures, and will probably reduce the habitat available to fish species such as lake trout.

Future increases in average air temperature and precipitation are very likely to further affect freshwater systems at Toolik. Lakes will very probably experience early breakup and higher water temperatures. Stream waters are very likely to warm as well, and runoff is very likely to increase, although evapotranspiration could possibly offset increased precipitation. As waters warm, primary production in lakes and rivers at the site is very likely to increase, although most species of aquatic plants and animals are unlikely to change over the 21st century. Lake and river productivity are also very likely to increase in response to changes in the catchment, in particular, temperature increases in permafrost soils and increased weathering and release of nutrients. Increased precipitation will also affect nutrient supply to freshwater systems, and is likely to result in increased decomposition of organic matter in soils[21], formation of inorganic nitrogen compounds, and increased loss of nitrogen from land to water.The shift in terrestrial vegetation to predominantly shrubs is very likely to cause greater loading of DOC and humic materials in streams and lakes, and a reduction in UV-B radiation penetration. However, increases in organic matter are likely to have detrimental effects on the stream population of Arctic grayling at this site, resulting in their disappearance in response to high oxygen depletion. Lake trout, on the other hand, are likely to survive but their habitat will probably be slightly reduced by the combination of reduced deep-water oxygen and warmer surface waters.

Changes in freshwater catchments with climate change will affect not only loadings of nutrients, sediments, DOC, and DIC to freshwater systems but also the transport and transformation of contaminants. Contaminant transport from surrounding catchments to freshwaters is likely to increase as permafrost degrades and perennial snow melts[22]. The contaminants released from these frozen stores, and those originating from long-range transport and deposition in contributing basins, can then be stored in sediments or metabolized and biomagnified through the food web. Section 8.7 (Effects of climate change on hydro-ecology of contributing basins in the Arctic) discusses this topic in more detail.

Chapter 8: Freshwater Ecosystems and Fisheries
8.1. Introduction (Effects of climate change on hydro-ecology of contributing basins in the Arctic)
8.2. Freshwater ecosystems in the Arctic
8.3. Historical changes in freshwater ecosystems
8.4. Climate change effects
8.4.1. Broad-scale effects on freshwater systems
8.4.2. Effects on hydro-ecology of contributing basins
8.4.3. Effects on general hydro-ecology
8.4.4. Changes in aquatic biota and ecosystem structure and function
8.5. Climate change effects on arctic fish, fisheries, and aquatic wildlife
8.5.1. Information required to project responses of arctic fish
8.5.2. Approaches to projecting climate change effects on arctic fish populations
8.5.3. Climate change effects on arctic freshwater fish populations
8.5.4. Effects of climate change on arctic anadromous fish
8.5.5. Impacts on arctic freshwater and anadromous fisheries
8.5.6. Impacts on aquatic birds and mammals
8.6. Ultraviolet radiation effects on freshwater ecosystems
8.7. Global change and contaminants
8.8. Key findings, science gaps, and recommendations

References


Citation

Committee, I. (2012). Effects of climate change on hydro-ecology of contributing basins in the Arctic. Retrieved from http://editors.eol.org/eoearth/wiki/Effects_of_climate_change_on_hydro-ecology_of_contributing_basins_in_the_Arctic
  1. Woo, M.-K., 1996. Hydrology of northern North America under global warming. In: J.A.A. Jones (ed.). Regional Hydrological Responses to Climate Change, pp. 73–86. Kluwer Academic Publishers.
  2. Pomeroy, J.W., D.M. Gray and P.G. Landine, 1993. The Prairie Blowing Snow Model: characteristics, validation, operation. Journal of Hydrology, 144:165–192.
  3. Liston, G.E., J.P. McFadden, M. Sturm and R.A. Pielke Sr., 2002. Modelled changes in arctic tundra snow, energy, and moisture fluxes due to increased shrubs. Global Change Biology, 8:17–32.
  4. Krupnik, I. and D. Jolly (eds.), 2002. The Earth is Faster Now: Indigenous Observations of Arctic Environmental Change. Arctic Research Consortium of the United States, Fairbanks, Alaska, 384pp.
  5. McDonald, M.A., L. Arragutainaq and Z. Novalinga, 1997. Voices from the Bay: Traditional Ecological Knowledge of Inuit and Cree in the Hudson Bay Bioregion. Canadian Arctic Resources Committee and Environmental Committee of the Municipality of Sanikiluaq, Ottawa, Canada.
  6. Rouse,W.R., M.S.V. Douglas, R.E. Hecky, A.E. Hershey, G.W. Kling, L. Lesack, P. Marsh, M. McDonald, B.J. Nicholson, N.T. Roulet and J.P. Smol, 1997. Effects of climate change on the freshwaters of Arctic and subarctic North America. Hydrological Processes, 11:873–902.
  7. Hinzman, L.D. and D.L. Kane, 1992. Potential response of an Arctic watershed during a period of global warming. Journal of Geophysical Research, 97(D3):2811–2820.
  8. Rouse,W.R., M.S.V. Douglas, R.E. Hecky, A.E. Hershey, G.W. Kling, L. Lesack, P. Marsh, M. McDonald, B.J. Nicholson, N.T. Roulet and J.P. Smol, 1997. Effects of climate change on the freshwaters of Arctic and subarctic North America. Hydrological Processes, 11:873–902.
  9. Hobbie, J.E., B.J. Peterson, N. Bettez, L. Deegan, W.J. O’Brien, G.W. Kling, G.W. Kipphut, W.B. Bowden and A.E. Hershey, 1999. Impact of global change on the biogeochemistry and ecosystems of an arctic freshwater system. Polar Research, 18:207–214.
  10. Krupnik, I. and D. Jolly (eds.), 2002. The Earth is Faster Now: Indigenous Observations of Arctic Environmental Change. Arctic Research Consortium of the United States, Fairbanks, Alaska, 384pp.
  11. Church, M., 1995. Geomorphic response to river flow regulation: case studies and time-scales. Regulated Rivers: Research and Management, 11:3–22.
  12. Fallu, M.-A. and R. Pienitz, 1999. Diatomées lacustres de Jamésie-Hudsonie (Québec) et modèle de reconstitution des concentrations de carbone organique dissous. Ecoscience, 6:603–620.–Rühland, K.M. and J.P. Smol, 1998. Limnological characteristics of 70 lakes spanning arctic treeline from Coronation Gulf to Great Slave Lake in the central Northwest Territories, Canada. International Review of Hydrobiology, 83:183–203.
  13. Korhola, A., J. Weckström and T. Blom, 2002b. Relationships between lake and land-cover features along latitudinal vegetation ecotones in arctic Fennoscandia. Archiv für Hydrobiologie, Supplementbände (Monograph Studies), 139(2):203–235.–Pienitz, R. and J.P. Smol, 1994.The ecology and physicochemical characteristics of lakes in the subarctic and arctic regions of the Yukon Territory, Fennoscandia (Finland, Norway), the Northwest Territories and Northern Quebec. In: P.B. Hamilton (ed.). Proceedings of the Fourth Arctic-Antarctic Diatom Symposium, pp. 31–43. Canadian Technical Report of Fisheries and Aquatic Sciences No. 1957.
  14. Vincent, W.F. and J.E. Hobbie, 2000. Ecology of Arctic lakes and rivers. In: M. Nuttall and T.V. Callaghan (eds.). The Arctic: Environment, People, Policy, pp. 197–231. Harwood Academic Press.
  15. Korhola, A. and J. Weckström, 2005. Paleolimnological studies in arctic Fennoscandia and the Kola Peninsula (Russia). In: R. Pienitz, M.S.V. Douglas and J.P. Smol (eds.). Long-Term Environmental Change in Arctic and Antarctic Lakes, pp. 381–418. Springer.–Seppä, H. and J. Weckström, 1999. Holocene vegetational and limnological changes in the Fennoscandian tree-line area as documented by pollen and diatom records from Lake Tsuolbmajavri, Finland. Ecoscience, 6:621–635.–Solovieva, N. and V.J. Jones, 2002. A multiproxy record of Holocene environmental changes in the central Kola Peninsula, northwest Russia. Journal of Quaternary Science, 17:303–318.
  16. Zepp, R.G., T.V. Callaghan and D.J. Erickson III, 2003. Interactive effects of ozone depletion and climate change on biogeochemical cycles. In: J.F. Bornman, K. Solomon, and J.C. van der Leun (eds.). Environmental Effects of Ozone Depletion and its Interactions with Climate Change: 2002 Assessment. Photochemical and Photobiological Sciences, 2:51–61.
  17. Häder, D.-P., H.D. Kumar, R.C. Smith and R.C. Worrest, 2003. Aquatic ecosystems: effects of solar ultraviolet radiation and interactions with other climatic change factors. In: J.F. Bornman, K. Solomon, and J.C. van der Leun (eds.). Environmental Effects of Ozone Depletion and its Interactions with Climate Change: 2002 Assessment. Photochemical and Photobiological Sciences, 2:39–50.
  18. Deegan, L.A., H.E. Golden, C.J. Harvey and B.J. Peterson, 1999. Influence of environmental variability on the growth of age-0 and adult Arctic grayling.Transactions of the American Fisheries Society, 128:1163–1175.
  19. Hobbie, J.E., G. Shaver, J. Laundre, K. Slavik, L.A. Deegan, J. O’Brien, S. Oberbauer and S. MacIntyre, 2003. Climate forcing at the Arctic LTER site. In: D. Greenland, D.G. Goodin and R.C. Smith (eds.). Climate Variability and Ecosystem Response in Long-Term Ecological Research (LTER) Sites, pp. 74–91. Oxford University Press.
  20. Morris, D.P., H.E. Zagarese, C.E.Williamson, E.G. Balseiro, B.R. Hargreaves, B. Modenutti, R. Moeller and C. Queimalinos, 1995.The attenuation of solar UV radiation in lakes and the role of dissolved organic carbon. Limnology and Oceanography, 40:1381–1391.
  21. Clein, J.S., B.L. Kwiatkowski, A.D. McGuire, J.E. Hobbie, E.B. Rastetter, J.M. Melillo and D.W. Kicklighter, 2000. Modelling carbon responses of tundra ecosystems to historical and projected climate: a comparison of a plot- and a global-scale ecosystem model to identify process-based uncertainties. Global Change Biology, 6(S1):127–140.
  22. Blais, J.M., D.W. Schindler, D.C.G. Muir, L.E. Kimpe, D.B. Donald and B. Rosenberg, 1998. Accumulation of persistent organochlorine compounds in mountains of western Canada. Nature, 395:585–588.–Blais, J.M., D.W. Schindler, D.C.G. Muir, M. Sharp, D.B. Donald, M. Lafreniere, E. Braekevelt and W.M.J. Strachan, 2001. Melting glaciers: a major source of persistent organochlorines to subalpine Bow Lake in Banff National Park, Canada. Ambio, 30:410–415.–McNamara, J.P., D.L. Kane and L.D. Hinzman, 1999. An analysis of an arctic channel network using a digital elevation model. Geomorphology, 29:339–353.