Effects on general hydro-ecology

From The Encyclopedia of Earth
Jump to: navigation, search

Effects of climate change on general hydro-ecology in the Arctic

February 9, 2010, 3:25 pm
July 11, 2012, 9:32 pm

This is Section 8.4.3 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

Streams and rivers, deltas, and estuaries (8.4.3.1)

A number of hydrologic shifts related to climate change will affect lakes (River and lake ice in the Arctic) and rivers (River and lake ice in the Arctic), including seasonal flow patterns, ice-cover thickness and duration, and the frequency and severity of extreme flood events. In the present climate, most streams and rivers originating within the Arctic have a nival regime in which snowmelt produces high flows and negligible flow occurs in winter. In areas of significant glaciers (Glaciers and ice sheets in the Arctic), such as on some Canadian and Russian islands, Greenland, and Svalbard, ice melt from glaciers can sustain flow during the summer, whereas many other streams produce summer flow only from periodic rainstorm events unless they are fed by upstream storage in lakes and ponds.

The subarctic contains a much broader range of hydrologic regimes, which vary from cold interior continental (comparable to those of the Arctic) to maritime regimes fed moisture directly from open seas even during winter. Overall, a warmer climate is very likely to lead to a shift toward a more pluvial runoff regime as a greater proportion of the annual precipitation falls as rain rather than snow; the magnitude of the peak of spring snowmelt declines; thawing permafrost (Permafrost in the Arctic) increases near-surface storage and reduces runoff peaks; and a more active groundwater system augments base flows.

Enhancement of winter flow will very probably lead to the development of a floating ice cover in some streams that currently freeze to the bed. This is very likely to be beneficial to the biological productivity of arctic streams and fish (Climate change effects on arctic freshwater fish populations) survival where winter freshwater habitat is limited to unfrozen pools[1]. For other arctic streams and rivers, warming is very likely to result in a shortened ice season and thinner ice cover (Section 6.7.3 (Effects of climate change on general hydro-ecology in the Arctic)). Since river ice is such a major controller of the ecology of northern streams and rivers, there are likely to be numerous significant impacts. Under conditions of overall annual temperature increases, a delay in the timing of freeze-up and an earlier breakup will very probably reduce the duration of river-ice cover. Data compiled over the last century or more indicate that changes in timing of these events are likely to be at a rate of approximately one day per 0.2°C increase in air temperature (see Sections 8.4.1 (Effects of climate change on general hydro-ecology in the Arctic) and 6.7.3 (Effects of climate change on general hydro-ecology in the Arctic))[2]. For freeze-up, higher water and air temperatures in the autumn combine to delay the time of first ice formation and eventual freeze-up. If there was also a reduction in the rate of autumn cooling, the interval between these two events would increase. Although all major ice types would continue to form, unless there were also significant changes in the flow regime, the frequency and magnitude of, for example, periods of major frazil ice growth will probably be reduced. This has implications for the types of ice that constitute the freeze-up cover and for the creation of unique under-ice habitats such as air cavities and those influenced by frazil concentrations[3].

Changes in the timing and duration of river ice formation will also alter the dissolved oxygen (DO) regimes of arctic lotic ecosystems. Following freeze-up and the elimination of direct water–atmosphere exchanges, DO concentrations steadily decline, sometimes to near-critical levels for river biota[4]. Reductions in ice-cover duration and a related increase in the number of open-water re-aeration zones are very likely to reduce the potential for this biologically damaging oxygen depletion. Such benefits will possibly be offset by the projected enhanced input of DOC and its subsequent oxidation[5], the rate and magnitude of which would also be increased as a result of the above-noted higher nutrient loading. Worst-case scenarios would develop on rivers where the flow is already comprised of poorly oxygenated groundwater, such as that supplied from extensive bogs and peatlands. Some rivers in the West Siberian Plain offer the best examples of this situation. Here, the River Irtysh drains large quantities of de-oxygenated water from vast peatlands into the River Ob, resulting in DO levels of only about 5% of saturation[6].

The greatest ice-related ecological impacts of climate change on arctic lotic systems are likely to result from changes in breakup timing and intensity. As well as favoring earlier breakup, higher spring air [[temperature]s] can affect breakup severity[7]. While thinner ice produced during a warmer winter would tend to promote a less severe breakup, earlier timing of the event could counteract this to some degree. Breakup severity also depends on the size of the spring flood wave. While greater and more rapid snowmelt runoff would favor an increase in breakup severity, the reverse is true for smaller snowpacks and more protracted melt. Hence, changes in breakup severity will vary regionally according to the variations in winter precipitation (Precipitation and evapotranspiration in the Arctic) and spring melt patterns.

For regions that experience a more "thermal" or less dynamic ice breakup[8], the magnitude of the annual spring flood will very probably be reduced. For the many northern communities that historically located near river floodplains for ease of transportation access, reductions in spring ice-jam flooding would be a benefit. In contrast, however, reductions in the frequency and severity of ice-jam flooding would have a serious impact on river ecology since the physical disturbances associated with breakup scouring and flooding are very important to nutrient and organic matter dynamics, spring water chemistry, and the abundance and diversity of river biota[9]. Specifically, ice-induced flooding supplies the flux of sediment, nutrients, and water that is essential to the health of the riparian system; river deltas being particularly dependent on this process[10]. More generally, given that the magnitude and recurrence interval of water levels produced by ice jams often exceed those of open-water conditions, breakup is probably the main supplier of allochthonous organic material in cold-regions rivers[11]. In the same manner, breakup serves as an indirect driver of primary and secondary productivity through the supply of nutrients – a common limiting factor for productivity in cold-regions rivers. Even the mesoscale climate of delta ecosystems and spring plant growth depends on the timing and severity of breakup flooding[12].

River ice is also a key agent of geomorphological change and is responsible for the creation of numerous [erosion]]al and depositional features within river channels and on channel floodplains[13]. Since most geomorphological activity occurs during freeze-up and breakup, changes in the timing of these events are very unlikely to have any significant effect. If, however, climatic conditions alter the severity of such events, this is likely to affect particular geomorphological processes. Furthermore, breakup events affect the general processes of channel enlargement, scour of substrate habitat, and the removal and/or succession of riparian vegetation. All such major river-modifying processes would be altered by any climate-induced shift in breakup intensity.

In summary, if climate change alters the long-term nature of breakup dynamics, the structure and function of rivers and related delta ecosystems are very likely to be significantly altered with direct effects on in-channel and riparian biological productivity. If, for example, significant reductions in dynamic breakups and the related level of disturbance occur, this will reduce overall biological diversity and productivity, with the most pronounced effects on floodplain and delta aquatic systems.

Owing to the reduced ice-cover season and increased air [[temperature]s] during the open-water period, summer water temperatures will very probably rise. Combined with greater DOC and nutrient loadings, higher water temperatures are likely to lead to a general increase in total stream productivity, although it is unclear whether temperature will have a significant direct effect on the processing rate of additional particulate detrital material. Irons et al.[14], for example, found a comparable rate of litter processing by invertebrates in Michigan and Alaska and concluded that temperature was not a main factor. The effect of increased temperature on processing efficiency by "cold-climate" species of invertebrates, however, has not been evaluated. The effect of enhanced nutrient loading to arctic streams is more predictable. The current nutrient limitation of many arctic streams is such that even slight increases in available phosphorus, for example, will produce a significant increase in primary productivity[15]. Where productivity responses of stream biota are co-limited by phosphorus and nitrogen[16], increased loadings of both nutrients would be required to sustain high levels of enhanced productivity.

Table 8.1 summarizes the potential impacts of climate change on the dynamics of arctic estuaries[17]. The major factor affecting arctic estuarine systems given the degree of climate change projected by the ACIA-designated models will be the increase in freshwater discharge (Section 6.8.3 (Effects of climate change on general hydro-ecology in the Arctic)). In some arctic basins, such as the Chukchi Sea, there is presently very little freshwater runoff and consequently no estuarine zones. Increased river discharge could possibly create estuarine areas, providing new habitat opportunities for euryhaline species. In established estuarine systems, such as the Mackenzie River system and the Ob and Yenisey Rivers, increased freshwater input in summer[18] is likely to increase stratification, making these habitats more suitable for freshwater species and less suitable for marine species. There are likely to be shifts in species composition to more euryhaline and anadromous species. In addition, increased freshwater input is likely to deposit more organic material, changing estuarine biogeochemistry and perhaps increasing primary productivity, the positive effects of which will possibly be offset in part by increased resuspension of contaminated sediments in these systems.

Table 8.1. A synthesis of the potential effects of climate change on arctic estuarine systems from both the bottom:up and top-down ecological perspectives.[19]

bottom:up:
nutrients/production/biota etc.

Top-down:
humans/predators/biota etc.

  • More open water, more wind mixing, upwelling and greater nutrient availability for primary producers (+)
  • More open water, more light penetration especially seasonally hence more primary production (+); potential for increased UV radiation levels (-/?)
  • Decreased ice cover, decreased ice-associated algal production, and subsequent impacts on pelagic and benthic food webs (-)
  • Increased basin rainfall, increased export of carbon to nearshore (+)
  • Increased storms and open water, increased coastal erosion (-), increased sediment loads, nutrients and mixing (+), possibly increased productivity especially in late season (?) but offset by decreased light penetration (-)
  • Potential positive feedback to climate change processes (e.g., permafrost thawing, release of methane, and increased radiative forcing) (-)
  • Contaminant inputs, mobilization, or increased fluxes driven by temperature changes will increase availability and biomagnification of contaminants in food chains (-)
  • Shifting water masses and currents will affect biotic cues for habitat use and migrations of biota such as fish and marine mammals (?)
  • Redistribution of grazers will affect underlying trophic structure (-/?)
  • Climate-induced changes in freshwater, estuarine, and marine habitats seasonally used by anadromous fishes will affect distribution and suitability for use, with consequences for the prey communities and possibly fish availability for humans (-/?)
  • Physical absence or alteration of seasonality or characteristics of ice platforms will affect ice-associated biota (e.g., polar bears, seals, algae) (-), with cascading consequences for fish (+/-)
  • Increased open water will facilitate whale migrations (+) but increase predator risk to calves (-); shifts in whale populations may cascade through the trophic structure (e.g., shifted predation on fish by belugas; increased predation on plankton by bowhead whales) with unknown trophic consequences for anadromous and marine fish (?)

Cascading consequences from a human perspective as generally: positive (+); negative (-); neutral (0); or unknown (?).

A secondary impact of increased freshwater discharge that is of serious concern, particularly for Siberian rivers that traverse large industrialized watersheds, is the potential for increased contaminant input. The Ob and Yenisey Rivers, for example, have high levels of organochlorine contamination compared to the Lena River[20], which is considered relatively pristine[21]. Larsen et al.[22] noted that arctic fishes have a life strategy that involves intensive feeding in spring and summer, allowing for the buildup of lipid stores and coping with food shortages in winter. The high body-lipid content of arctic fishes may make them more vulnerable to lipid-soluble pollutants such as polycyclic aromatic hydrocarbons (PAHs) or polychlorinated biphenyls (PCBs). In addition, reduced sea-ice coverage that leads to increased marine traffic is likely to have cascading negative consequences (e.g., pollution, risk of oil spills) for estuarine systems.

Arctic deltas provide overwintering habitat for many species that tolerate brackish waters. These areas are maintained as suitable habitat by a combination of continuous under-ice freshwater flow and the formation of the nearshore ice barrier in the stamukhi zone (area of grounded, nearshore ice pressure ridges). As [[temperature]s] rise, the seasonal ice zone of estuaries is likely to expand and the ice-free season lengthen[23]. Disruption of either the flow regime or the ice barrier could possibly have profound effects on the availability of suitable overwintering habitat for desired fish species. Given that such habitat is probably limited and hence limits population abundance, the consequences for local fisheries will probably be significant. In addition, in early winter, subsistence and commercial fisheries target fish that overwinter in deltas. Thinning ice is likely to limit access to these fisheries.

Similar to freshwater systems, ecological control of marine systems can be viewed from bottom:up (i.e., nutrients–production–biota linkages) and/or top-down (i.e., human activities–predators–keystone biota) perspectives[24]. The special role of ice as both a habitat and a major physical force shaping the estuarine and nearshore arctic environment suggests that climate change will work in both modes to affect these systems[25]. One example is the loss of the largest epishelf lake (fresh and brackish water body contained behind the ice shelf) in the Northern Hemisphere with the deterioration and break up of the Ward Hunt ice shelf[26]. The loss of this nearshore water body has affected a unique community of marine and freshwater planktonic species, as well as communities of cold-tolerant microscopic algae and animals that inhabited the upper ice shelf.

Lakes, ponds, and wetlands (8.4.3.2)

Lentic systems north of the Arctic Circle contain numerous small to medium lakes and a multitude of small ponds and wetland systems. Relatively deep lakes are primarily contained within alpine or foothill regions such as those of the Putorana Plateau in the lower basin of the Yenisey River. One very large and deep lake, Great Bear Lake (Northwest Territories, Canada), is found partly within the Arctic Circle. Variations in its water budget primarily depend on flows from its contributing catchment, comprised largely of interior plains lowlands and exposed bedrock north of 60° N. Its southern counterpart, Great Slave Lake, provides a strong hydrologic contrast to this system. Although also part of the main stem Mackenzie River basin and wholly located north of 60° N, its water budget is primarily determined by inflow that originates from Mackenzie River headwater rivers located much further to the south. Moreover, its seasonality in water levels reflects the effects of flow regulation and climatic variability in one of its major tributaries, the Peace River, located about 2,000 kilometers (km) upstream in the Rocky Mountain headwaters of western Canada[27]. As such, the Mackenzie River system offers the best example of a northern lentic system that is unlikely to be significantly affected by changes in hydrologic processes operating within the north (e.g., direct lake evaporation and precipitation) but will be dependent principally on changes in water-balance processes operating well outside the Arctic.

The other major arctic landscape type that contains large, although primarily shallower, lakes is the coastal plains region found around the circumpolar north. As mentioned previously, these shallow systems depend on snowmelt as their primary source of water, with rainfall gains often negated by evapotranspiration during the summer. Evaporation from these shallow water bodies is very likely to increase as the ice-free season lengthens. Hence, the water budget of most lake, pond, and wetland systems is likely to depend more heavily on the supply of spring meltwater to produce a positive annual water balance, and these systems are more likely to dry out during the summer. Another possible outcome of climate change is a shift in vegetation from non-transpiring lichens and mosses to vascular plants as [[temperature]s] rise and the growing season extends[28], potentially exacerbating water losses. However, factors such as increasing cloud cover and summer precipitation will possibly mitigate these effects.

Loss of permafrost (Permafrost in the Arctic) increases the potential for many northern shallow lotic systems to dry out from a warmer temperature regime. Ponds are likely to become coupled with the groundwater system and drain if losses due to downward percolation and evaporation are greater than resupply by spring snowmelt and summer precipitation. Patchy arctic wetlands are particularly sensitive to permafrost degradation that can link surficial waters to the supra-permafrost groundwater system. Those along the southern limit of permafrost, where increases in temperature are most likely to eliminate the relatively warm permafrost, are at the highest risk of drainage[29]. Traditional ecological knowledge from Nunavut and eastern arctic Canada indicates that recently there has been enhanced drying of lakes and rivers, as well as swamps and bogs, enough to impair access to traditional hunting grounds and, in some instances, fish migration[30] (see Section 3.4.5 (Effects of climate change on general hydro-ecology in the Arctic) for detailed discussion of a related case study).

Warming of surface permafrost, however, will very probably enhance the formation of thermokarst wetlands, ponds, and drainage networks, particularly in areas characterized by concentrations of massive ground ice. Thawing of such ice concentrations, however, is very likely to lead to dramatic increases in terrain slumping and subsequent sediment transport and deposition in rivers, lakes, deltas, and nearshore marine environments. This is likely to produce distinct changes in channel geomorphology in systems where sediment transport capacity is limited, and will probably have a significant impact on the aquatic ecology of the receiving water bodies. Catastrophic drainage of permafrost-based lakes that are now in a state of thermal instability, such as those found along the western arctic coast of Canada, is also very likely[31]. Losses of thermokarst lakes within lowlying deltaic areas are also likely to result from rising sea levels. Marine inundation resulting from continually rising sea level commonly drains lakes in the outer portion of the Mackenzie Delta (northern Richards Island[32]). Moreover, Mackay J.[33] estimated that one lake per year has drained in the Tuktoyaktuk coastlands of northern Canada over the last few thousand years. Future, more pronounced rises in sea level are likely to accelerate this process.

Changes in the water balance of northern wetlands are especially important because most wetlands in permafrost regions are peatlands, which can be sources or sinks of carbon and CH4 depending on the depth of the water table (see also Section 8.4.4.4 (Effects of climate change on general hydro-ecology in the Arctic)). An analysis by Rouse et al.[34] of subarctic sedge fens in a doubled-CO2 climate suggested that increases in temperature (4°C) would reduce water storage in northern peatlands even with a small and persistent increase in precipitation. While acknowledging that storage changes depend on variability in soil moisture and peat properties, projected declines in the water table were 10 to 20 centimeters (cm) over the summer.

As the ice cover of northern lakes and ponds becomes thinner, forms later, and breaks up earlier ([[Section 6.7.3 (Effects of climate change on general hydro-ecology in the Arctic)]2]), concomitant limnological changes are very likely. Thinner ice covers with less snow cover will increase the under-ice receipt of solar radiation, thereby increasing under-ice algal production and oxygen[35] and reducing the potential for winter anoxia and fish kills. Lower water levels, which reduce under-ice water volumes and increase the likelihood of winterkill, could possibly counteract this effect. Similarly, greater winter precipitation on a thinner ice cover is very likely to promote the formation of more highly reflective snow and white-ice layers. Such layers would reduce radiation penetration well into the spring because they also tend to delay breakup compared to covers comprised of only black ice. Notably, the ACIA-designated models project that incident radiation will decline. Reductions are likely to be relatively small (i.e., 10–12 W/m2 in May-June between 1981–2000 and 2071–2090, Section 4.4.4 (Effects of climate change on general hydro-ecology in the Arctic)), however, compared to the major reductions that are likely to result from greater reflective loss from enhanced white-ice formation.

A longer ice-free season will also increase the length of the stratified season and generally increase the depth of mixing (Box 8.4), although the magnitude and duration of the effects will depend on factors such as basin depth and area. This is likely to lower oxygen concentrations in the hypolimnion and increase stress on cold-water organisms[36]. Furthermore, such an enhancement of mixing processes and reduction in ice cover will probably increase the potential for many northern lakes and ponds to become contaminant sinks (Section 8.7 (Effects of climate change on general hydro-ecology in the Arctic)).

Box 8.4. Lake-ice duration and water column stratification: Lake Saanajärvi, Finnish Lapland

Lake Saanajärvi (maximum depth 24 m, area 0.7 km2; 69º N, 20º E) is the key Finnish site in the European research projects Mountain Lake Research and European Mountain Lake Ecosystems: Regionalisation, Diagnostics and Socio-economic Evaluation. Lake Saanajärvi has been intensively monitored since 1996.The data presented here have been published in several papers, including those by Korhola et al.[37], Rautio et al. [38], Sorvari and Korhola [39], and Sorvari et al. [40].The mean annual temperature of the area is -2.6 ºC, and annual precipitation is approximately 400 mm.The catchment area is mostly covered by bare rocks and alpine vegetation. Lake Saanajärvi is a dimictic, ultra-oligotrophic, clear-water lake.The lake is ice-free for nearly four months of the year, with highly oxygenated waters, and is strongly stratified for two months after spring overturn. Phytoplankton biomass and densities are low[41], consisting predominantly of chrysophytes and diatoms. Bacterial biomass is low as well, and zooplankton are not very abundant. Freshwater shrimp (Gammarus lacustris) are common and form an important food source for fish, which include Arctic char and brown trout (Salmo trutta lacustris).

600px-ACIA Box 8.3.png Figure for Box 8.4. Comparison of diatom assemblage changes with regional and arctic-wide temperature anomalies, showing (a) principal components analysis (PCA) primary axis scores derived from the correlation matrices of the diatom percentage counts from the five study sites; (b) spring (March–May) temperature anomalies for northwestern Finnish Lapland, smoothed using a 10-year low-pass filter; (c) trend in mean annual air temperature in northwestern Finnish Lapland, smoothed using a 10-year low-pass filter; and (d) standardized proxy arctic-wide summer-weighted annual temperature, plotted as departure from the mean (panels a–c from Sorvari et al., 2000; panel d from Overpeck).

Changes in water temperature and stratification of Lake Saanajärvi have been associated with climate changes in Finnish Lapland over the past 200 years[42]. Mean annual air temperatures in Finnish Lapland, as in much of the Arctic, rose 1 to 2 ºC following the Little Ice Age. During this period of warming, diatom communities changed from benthic–periphytic to pelagic, Cladocera increased in abundance, and chrysophytes became less numerous.These changes have been shown to be associated with increased rates of organic matter accumulation and increased concentrations of algal pigments during the climatic warming[43] (see figure). After a period of cooling from the 1950s to the 1970s, air temperatures in the Arctic continued to rise. More recently, interannual variability in temperatures has been shown to account for changes in the thermal gradient and mixing of Lake Saanajärvi surface waters. For example, Lake Saanajärvi normally stratifies in early July, two weeks after ice breakup, and retains a distinct, steep thermocline at a depth of 10 to 12 m throughout the summer. In 2001, this summer stratification was broken after a period of slight cooling in early August, after which the lake was only weakly stratified. In 2002, on the other hand, spring and summer temperatures were extremely warm. Spring ice breakup was early and waters warmed quickly, resulting in a very sharp thermocline that was stable during the entire summer stratification period.

Future temperature increases are therefore very likely to affect the thermal structure of lakes in Finnish Lapland and throughout the Arctic, which is likely to have dramatic consequences for lake biota. Rising mean annual temperatures are very likely to influence the duration of summer stratification and the stability and depth of the thermocline in Finnish lakes. As such, many of the presently isothermal lakes are likely to become dimictic as temperatures increase. In addition, the prolonged thermal stratification that is likely to accompany rising temperatures could possibly lead to lower oxygen concentrations and increased phosphorus concentrations in the hypolimnion, benefiting nutrient-limited primary production. As spring temperatures rise and the ice-free period extends, not only is thermal stratification likely to stabilize, but production in many high latitude lakes could possibly peak twicerather than once during the open-water season[44]. On a broader scale, changes in lake stratification and water mixing will probably affect species composition[45].

With a longer and warmer ice-free season, total primary production is likely to increase in all arctic lakes and ponds, and especially in the oligotrophic high-arctic ponds that are currently frozen for a majority of the year[46]. Similar to the situation for arctic lotic systems, an enhanced supply of nutrients and organic matter from the more biologically productive contributing basins is likely to boost primary productivity[47]. Again, however, there are likely to be offsetting effects because of reductions in light availability resulting from enhanced turbidity due to higher inputs of DOC and suspended sediment. Hecky and Guildford[48] noted that analogous factors caused a switch from nutrient limitation, which is a common control of primary production in northern lakes, to light limitation.

Chapter 8: Freshwater Ecosystems and Fisheries
8.1. Introduction (Effects of climate change on general hydro-ecology 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

See Also

Citation

Committee, I. (2012). Effects of climate change on general hydro-ecology in the Arctic. Retrieved from http://editors.eol.org/eoearth/wiki/Effects_on_general_hydro-ecology
  1. Craig, P.C., 1989. An introduction to anadromous fishes in the Alaskan Arctic. In: D.W. Norton (ed.). Research Advances on Anadromous Fish in Arctic Alaska and Canada. Biological Papers of the University of Alaska, 24:27–54.–Prowse, T.D., 2001a. River-ice ecology. I: hydrology, geomorphic, and water-quality aspects. Journal of Cold Regions Engineering, 15:1–16.–Prowse, T.D., 2001b. River-ice ecology. II: biological aspects. Journal of Cold Regions Engineering, 15:17–33.
  2. Magnuson, J.J., D.M. Robertson, B.J. Benson, R.H. Wynne, D.M. Livingstone, T. Arai, R.A. Assel, R.G. Barry, V. Card, E. Kuusisto, N.G. Granin, T.D. Prowse, K.M. Stewart and V.S. Vuglinski, 2000. Historical trends in lake and river ice cover in the Northern Hemisphere. Science, 289:1743–1746.
  3. Brown, R.S., S.S. Stanislawski and W.C. Mackay, 1994. Effects of frazil ice on fish. In: T.D. Prowse (ed.). Proceedings of the Workshop on Environmental Aspects of River Ice. National Hydrology Research Institute, Saskatoon, Symposium No. 12, pp. 261–278.–Cunjak, R.A., T.D. Prowse and D.L. Parrish, 1998. Atlantic salmon (Salmo salar) in winter: the season of parr discontent? Canadian Journal of Fisheries and Aquatic Sciences, 55(S1):161–180.–Prowse, T.D., 2001b. River-ice ecology. II: biological aspects. Journal of Cold Regions Engineering, 15:17–33.
  4. Chambers, P.A., G.J. Scrimgeour and A. Pietroniro, 1997. Winter oxygen conditions in ice-covered rivers: the impact of pulp mill and municipal effluents. Canadian Journal of Fisheries and Aquatic Sciences, 54:2796–2806.–Power, G., R. Cunjak, J. Flannagan and C. Katopodis, 1993. Biological effects of river ice. In: T.D. Prowse and N.C. Gridley (eds.). Environmental Aspects of River Ice. National Hydrology Research Institute, Saskatoon, Science Report No. 5, pp. 97–119.–Prowse, T.D., 2001a. River-ice ecology. I: hydrology, geomorphic, and water-quality aspects. Journal of Cold Regions Engineering, 15:1–16.–Prowse, T.D., 2001b. River-ice ecology. II: biological aspects. Journal of Cold Regions Engineering, 15:17–33.
  5. Schreier, H., W. Erlebach and L. Albright, 1980.Variations in water quality during winter in two Yukon rivers with emphasis on dissolved oxygen concentration. Water Research, 14:1345–1351.–Whitfield, P.H. and B. McNaughton, 1986. Dissolved-oxygen depressions under ice cover in two Yukon rivers. Water Resources Research, 22:1675–1679.
  6. Harper, P.P., 1981. Ecology of streams at high latitudes. In: M.A. Lock and D.D.Williams (eds.). Perspectives in Running Water Ecology, pp. 313–337. Plenum Press.–Hynes, H.B.N., 1970. The Ecology of Running Waters. Liverpool University Press, 555pp.
  7. Prowse, T.D. and S. Beltaos, 2002. Climatic control of river-ice hydrology: a review. Hydrological Processes, 16:805–822.
  8. Gray, D.M. and T.D. Prowse, 1993. Snow and floating ice. In: D.R. Maidment (ed.). Handbook of Hydrology, pp. 7.1–7.58. McGraw-Hill.
  9. Cunjak, R.A., T.D. Prowse and D.L. Parrish, 1998. Atlantic salmon (Salmo salar) in winter: the season of parr discontent? Canadian Journal of Fisheries and Aquatic Sciences, 55(S1):161–180.–Prowse, T.D. and J.M. Culp, 2003. Ice break-up: a neglected factor in river ecology. Canadian Journal of Civil Engineering, 30:145–155.–Scrimgeour, G.J., T.D. Prowse, J.M. Culp and P.A. Chambers, 1994. Ecological effects of river ice break-up: a review and perspective. Freshwater Biology, 32:261–275.
  10. Lesack, L., R.E. Hecky and P. Marsh, 1991.The influence of frequency and duration of flooding on the nutrient chemistry of the Mackenzie Delta lakes. In: P. Marsh and C.S.L. Ommanney (eds.). Mackenzie Delta: Environmental Interactions and Implications for Development, National Hydrology Research Institute, Saskatoon, Symposium No. 4, pp. 19–36.–Marsh, P. and M. Hey, 1989. The flooding hydrology of Mackenzie Delta lakes near Inuvik, N.W.T., Canada. Arctic, 42:41–49.–Prowse, T.D. and F.M. Conly, 2001. Multiple-hydrologic stressors of a northern delta ecosystem. Journal of Aquatic Ecosystem Stress and Recovery, 8:17–26.
  11. Prowse, T.D. and J.M. Culp, 2003. Ice break-up: a neglected factor in river ecology. Canadian Journal of Civil Engineering, 30:145–155.–Scrimgeour, G.J., T.D. Prowse, J.M. Culp and P.A. Chambers, 1994. Ecological effects of river ice break-up: a review and perspective. Freshwater Biology, 32:261–275.
  12. Gill, D., 1974. Significance of spring breakup to the bioclimate of the Mackenzie River Delta. In: J.C. Reed and J.E. Sater (eds.). The Coast and Shelf of the Beaufort Sea: Proceedings of a Symposium on Beaufort Sea Coast and Shelf Research, pp. 543–544. Arctic Institute of North America.–Hirst, S.M., 1984. Effects of spring breakup on microscale air temperatures in the Mackenzie River Delta. Arctic, 37:263–269.–Prowse, T.D., 2001a. River-ice ecology. I: hydrology, geomorphic, and water-quality aspects. Journal of Cold Regions Engineering, 15:1–16.
  13. Prowse, T.D., 2001a. River-ice ecology. I: hydrology, geomorphic, and water-quality aspects. Journal of Cold Regions Engineering, 15:1–16.–Prowse, T.D. and N.C. Gridley (eds.), 1993. Environmental Aspects of River Ice. National Hydrology Research Institute, Saskatoon, Science Report No. 5, 155pp.
  14. Irons, J.G., M.W. Oswood, R.J. Stout and C.M. Pringle, 1994. Latitudinal patterns in leaf litter breakdown: is temperature really important? Freshwater Biology, 32:401–411.
  15. Flanagan, K., E. McCauley, F. Wrona and T.D. Prowse, 2003. Climate change: the potential for latitudinal effects on algal biomass in aquatic ecosystems. Canadian Journal of Fisheries and Aquatic Sciences, 60:635–639.
  16. Peterson, B.J., L. Deegan, J. Helfrich, J.E. Hobbie, M. Hullar, B. Moller, T.E. Ford, A. Hershey, A. Hiltner, G. Kipphut, M.A. Lock, D.M. Fiebig, V. McKinley, M.C. Miller, J.R.Vestal, J. Robie, R. Ventullo and G.Volk, 1993. Biological responses of a tundra river to fertilization. Ecology, 74:653–672.
  17. Carmack, E.C. and R.W. Macdonald, 2002. Oceanography of the Canadian Shelf of the Beaufort Sea: a setting for marine life. Arctic, 55(S1):29–45.
  18. Peterson, B.J., R.M. Holmes, J.W. McClelland, C.J.Vorosmarty, R.B. Lammers, A.I. Shiklomanov, I.A. Shiklomanov and S. Rahmstorf, 2002. Increasing river discharge to the Arctic Ocean. Science, 298:2171–2173.
  19. Carmack, E.C. and R.W. Macdonald, 2002. Oceanography of the Canadian Shelf of the Beaufort Sea: a setting for marine life. Arctic, 55(S1):29–45.
  20. Zhulidov, A.V., J.V. Headley, D.F. Pavlov, R.D. Robarts, L.G. Korotova, V.V. Fadeev, O.V. Zhulidova, Y. Volovik and V. Khlobystov, 1998. Distribution of organochlorine insecticides in rivers of the Russian Federation. Journal of Environmental Quality, 27:1356–1366.
  21. Guieu, C., W.W. Huang, J.-M. Martin and Y.Y. Yong, 1996. Outflow of trace metals into the Laptev Sea by the Lena River. Marine Chemistry, 53:255–267.
  22. Larsen, L.H., A. Evenset and B. Sirenko, 1995. Linkages and impact hypotheses concerning valued ecosystem components (VECs) invertebrates, fish, the coastal zone and large river estuaries and deltas. International Northern Sea Route Programme, working paper 12, 39pp. +app.
  23. Carmack, E.C. and R.W. Macdonald, 2002. Oceanography of the Canadian Shelf of the Beaufort Sea: a setting for marine life. Arctic, 55(S1):29–45.
  24. Parsons, T.R., 1992.The removal of marine predators by fisheries and the impact of trophic structure. Marine Pollution Bulletin, 25:51–53.
  25. Carmack, E.C. and R.W. Macdonald, 2002. Oceanography of the Canadian Shelf of the Beaufort Sea: a setting for marine life. Arctic, 55(S1):29–45.
  26. Mueller, D.R., W.F. Vincent and M.O. Jeffries, 2003. Break-up of the largest Arctic ice shelf and associated loss of an epishelf lake. Geophysical Research Letters, 30: doi:10.1029/2003GL017931.
  27. Gibson, J.J., T.D. Prowse and D.L. Peters, in press. Hydroclimatic controls on water balance and water level variability in Great Slave Lake. Hydrological Processes.–Peters, D.L. and T.D. Prowse, 2001. Regulation effects on the lower Peace River, Canada. Hydrological Processes, 15:3181–3194.
  28. 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.
  29. Woo, M.-K., A.G. Lewkowicz and W.R. Rouse, 1992. Response of the Canadian permafrost environment to climatic change. Physical Geography, 134:287–317.
  30. 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.
  31. Mackay, J.R., 1992. Lake stability in an ice-rich permafrost environment. Examples from the Western Arctic Coast. In: R.D. Roberts and M.L. Bothwell (eds.). Aquatic Ecosystems in Semi-Arid Regions. Implications for Resource Management. National Hydrology Research Institute, Saskatoon, Symposium Series 7, pp. 1–26.–Marsh, P. and N. Neumann, 2001. Processes controlling the rapid drainage of two ice-rich permafrost-dammed lakes in NW Canada. Hydrological Processes, 15:3433–3446.–Marsh, P. and N. Neumann, 2003. Climate and hydrology of a permafrost dammed lake in NW Canada. In: M. Phillips, S.M. Springman and L.U. Arenson (eds.). Permafrost: Proceedings of the 8th International Conference on Permafrost, Zurich, Switzerland, 21–25 July 2003, Vol. 2, pp. 729:734. International Permafrost Association, Longyearbyen.
  32. Dallimore, A., C.J. Schröder-Adams and S.R. Dallimore, 2000. Holocene environmental history of thermokarst lakes on Richards Island, Northwest Territories, Canada: Theocamoebians as paleolimnological indicators. Journal of Paleolimnology, 23:261–283.
  33. Mackay, J.R., 1992. Lake stability in an ice-rich permafrost environment. Examples from the Western Arctic Coast. In: R.D. Roberts and M.L. Bothwell (eds.). Aquatic Ecosystems in Semi-Arid Regions. Implications for Resource Management. National Hydrology Research Institute, Saskatoon, Symposium Series 7, pp. 1–26.
  34. 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.
  35. Prowse, T.D. and R.L. Stephenson, 1986.The relationship between winter lake cover, radiation receipts and the oxygen deficit in temperate lakes. Atmosphere-Ocean, 24:386–403.
  36. 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.
  37. KKorhola, A., S. Sorvari, M. Rautio, P.G. Appleby, J.A. Dearing,Y. Hu, N. Rose, A. Lami and N.G. Cameron, 2002a.A multi-proxy analysis of climate impacts on the recent development of subarctic Lake Sannajärvi in Finnish Lapland. Journal of Paleolimnology, 28(1):59–77.
  38. Rautio, M., S. Sorvari and A. Korhola, 2000. Diatom and crustacean zooplankton communities, their seasonal variability, and representation in the sediments of subarctic Lake Saanajärvi. Journal of Limnology, 59(Suppl.1):81–96.
  39. Sorvari, S. and A. Korhola, 1998. Recent diatom assemblage changes in subarctic Lake Saanajärvi, NW Finnish Lapland, and their paleoenvironmental implications. Journal of Paleolimnology, 20:205–215.
  40. Sorvari, S., M. Rautio and A. Korhola, 2000. Seasonal dynamics of subarctic Lake Saanajärvi in Finnish Lapland. In:W.D.Williams (ed.). Internationale Vereinigung für Theoretische und Angewandte Limnologie: Verhandlungen, 27th Congress, Dublin, 1998, 27:507–512.-- Sorvari, S., A. Korhola and R.Thompson, 2002. Lake diatom response to recent Arctic warming in Finnish Lapland. Global Change Biology, 8:153–163.
  41. Forsström, L., 2000. Seasonal variability of phytoplankton in Lake Saanajärvi. M.Sc.Thesis, University of Helsinki, 53pp. (In Finnish- Forsström, L., S. Sorvari, A. Korhola and M. Rautio, in press. Seasonality of phytoplankton in subarctic Lake Saanajärvi in NW Finnish Lapland. Polar Biology.;Rautio et al., Op. cit
  42. Alexandersson, H. and B. Eriksson, 1989. Climate fluctuations in Sweden 1860–1987. Reports of Meteorology and Climatology, 58:1–54. Swedish Meteorological and Hydrological Institute, Stockholm.;-- Sorvari et al., 2002, Op. cit.;-- Tuomenvirta, H. and R. Heino, 1996. Climatic changes in Finland – recent findings. Geophysica, 32:61–75.
  43. Korhola, A., H. Olander and T. Blom, 2000a. Cladoceran and chironomid assemblages as quantitative indicators of water depth in subarctic Fennoscandian lakes. Journal of Paleolimnology, 24:43–54.;-- Sorvari and Korhola, 1998, Op. cit.;-- Sorvari et al., 2002, Op. cit.;
  44. e.g.,Catalan, J., M.Ventura, A. Brancelj, I. Granados, H.Thies, U. Nickus, A. Korhola, A.F. Lotter, A. Barbieri, E. Stuchlik, L. Lien, P. Bitusik, T. Buchaca, L. Camarero, G.H. Goudsmit, J. Kopacek, G. Lemcke, D.M. Livingstone, B. Müller, M. Rautio, M. Sisko, S. Sorvari, F. Sporka, O. Strunecky and M.Toro, 2000. Seasonal ecosystem variability in remote mountain lakes: implications for detecting climatic signals in sediment records. Journal of Paleolimnology, 28(1):25–46.;-- Hinder, B., M. Gabathuler, B. Steiner, K. Hanselmann and H.R. Preisig, 1999. Seasonal dynamics and phytoplankton diversity in high mountain lakes (Jöri lakes, Swiss Alps). Journal of Limnology, 58:152–161.;-- Lepistö, L., 1999. Phytoplankton assemblages reflecting the ecological status of lakes in Finland. Monographs of the Boreal EnvironmentResearch, 16. 97pp.;-- Lotter, A.F. and C. Bigler, 2000. Do diatoms in the Swiss Alps reflect the length of ice-cover? Aquatic Sciences, 62:125–141.;Medina-Sánchez et al., 1999;-- Rautio et. al., 2000, Op. cit.
  45. e.g., diatoms; Agbeti, M.D., J.C. Kingston, J.P. Smol and C.Watters, 1997 Comparison of phytoplankton succession in two lakes of different mixing regimes. Archiv für Hydrobiologie, 140:37–69.
  46. Douglas, M.S.V. and J.P. Smol, 1999. Freshwater diatoms as indicators of environmental change in the High Arctic. In: E.F. Stoermer and J.P. Smol (eds.).The Diatoms: Applications for the Environmental and Earth Sciences, pp. 227–244. Cambridge University Press.
  47. 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.
  48. Hecky, R.E. and S.J. Guildford, 1984. Primary productivity of Southern Indian Lake before, during and after impoundment and Churchill River Diversion. Canadian Journal of Fisheries and Aquatic Sciences, 41:591–604.