Hostname: page-component-76fb5796d-zzh7m Total loading time: 0 Render date: 2024-04-29T08:42:33.368Z Has data issue: false hasContentIssue false

PHYSICAL, CHEMICAL, AND DISTRIBUTIONAL ASPECTS OF CANADIAN SPRINGS

Published online by Cambridge University Press:  31 May 2012

Robert O. van Everdingen*
Affiliation:
The Arctic Institute of North America, The University of Calgary, Calgary, Alberta, Canada T2N 1N4
Get access

Abstract

Springs, or points of natural, concentrated groundwater discharge, may be located in river or lake beds, or below mean sea level along the coast, but many are found some distance from surface-water bodies. Spring water commonly represents rain or snow-melt that has entered the ground at a higher elevation a number of years earlier.Measured springwater temperatures in Canada range from very cold (−2.9 °C) to hot (82.2 °C). Thermal spring waters, with temperatures above the local mean-annual air temperature, have undergone geothermal heating during deep subsurface circulation in areas of high topographic relief. Hot springs (>37 °C) are therefore found only in mountainous areas, in Alberta, British Columbia, Yukon, and the Northwest Territories. Spring locations are commonly controlled by major folding or faulting, or both, in the bedrock strata.Reported pH values in Canadian spring waters range from strongly acidic to alkaline (2.8 to >10.0). Low pH values (<4.0) are associated with high contents of dissolved Fe (up to 2600 mg·L−1) and other heavy metals (e.g. Zn up to 177 mg·L−1), resulting from the oxidation of metal sulfides. Measured redox potentials (Eh) range from −252 to +683 mV. Negative Eh values are found in spring waters that contain dissolved H2S and S2−, produced by bacterial reduction of dissolved sulfate.Total-dissolved-solids contents of Canadian spring waters are reported to range from as little as 32 to over 75 000 mg·L−1. Chemical composition also varies widely. Major anions include bicarbonate (up to 5960 mg·L−1), sulfate (up to 17 520 mg·L−1), and chloride (up to 44 300 mg·L−1). Major cations include calcium (up to 1823 mg·L−1), magnesium (up to 1190 mg·L−1), sodium (up to 27 100 mg·L−1, and potassium (up to 1568 mg·L−1). The chemical composition of each spring water reflects the mineral composition of the rock types with which the water has been in contact, as well as its subsurface residence time. In simplified terms, Ca–Mg/HCO3 waters come from carbonate rock (limestone, dolomite), Ca/SO4 waters from gypsum or anhydrite, and Na/Cl waters from salt beds.Springwater temperature and composition can both show gradual (seasonal) and sudden (incidental) variations. In springs that show seasonal variations, maximum temperature and mineralization occur near the end of winter; minimum values commonly occur during snowmelt. Sudden variations in temperature, mineralization, and discharge rate can occur during periods of heavy rain, if cold, non-mineralized rainwater enters spring conduits. Earthquakes may cause sudden changes in discharge rates and suspended-solids contents, without affecting water temperature or chemical composition.Information on Canadian spring locations, and on their physical and chemical character, is still spotty. As detailed knowledge about springs can be useful in both ecological and water-supply studies, an effort should be made to expand and refine the existing database.

Résumé

Les collections de phryganes provenant de 25 sources à travers le Canada révèlent quelques tendances générales et quelques différences de la faune des sources reliées à la région et à l'habitat. En général, le nombre d'espèces de phryganes dans les sources augmente avec la diversité d'habitat. Les limnocrènes et les rhéocrènes à faible courant et au substrat composé de petites particules maintiennent peu d'espèces de phryganes, mais peuvent avoir de grandes populations d'une seule espèce. Les espèces qui sont classées comme brouteurs, dechetiqueurs et prédateurs sont communes aux sources, mais les filtreurs sont rares.

Plusieurs genres mais peu d'espèces sont communs aux sources de l'est et de l'ouest du Canada. Quelques 35% des espèces observées dans cette étude sont de la famille Limnephilidae, mais le genre le plus fréquemment rencontré a été Parapsyche (Hydropsychidae), habituellement P. apicalis (Banks) dans l'est du Canada et P. elsis Milne dans l'ouest (C.B. et Alberta). D'autres genres communs autant à l'est qu'à l'ouest ont été Neophylax, Lepidostoma et Rhyacophila. Les genres communs uniquement retrouvés dans l'ouest ont été Anagapetus, Homophylax, Psychoglypha et Neothremma; tandis que Frenesia et Pseudostenophylax n'ont été retrouvé que dans l'est. Trois méthodes d'analyse — ordination par analyse de correspondance désorientée; ordination contrainte par analyse de correspondance canonique; et classification par analyse d'espèce indicateur duodirectionnelle — ont tous confirmé une différence géographique est/ouest en ce qui concerne les communautés de phryganes et ont démontré que l'elevation, l'étendue de l'origine de la nappe d'eau et la température estivale constituent des facteurs environnementaux qui influencent, sans pour autant être entièrement responsable, les distributions est/ouest des espèces. Les obstacles du passé et du présent à la migration semblent importants. La végétation riparienne, le courant, la taille des particules du substrat, la diversité du microhabitat et le pH, tous ont de fortes influences sur la composition des communautés des sources de l'est et de l'ouest. Les sources dans lesquelles les phryganes ont été principalement associées avec le traitement des détritus ont été dominées par Frenesia et par Lepidostoma dans l'est, mais par Homophylax dans l'ouest. Les râpeurs et les prédateurs n'ont été abondants qu'aux sources où la diversité en microhabitats, la vitesse du courant et le pH étaient relativement élevés.

Type
Research Article
Copyright
Copyright © Entomological Society of Canada 1991

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

Atchison, M.E. 1964. A study of springs and spring deposits in Flat River map area. B.Sc. thesis, University of British Columbia.Google Scholar
Baas Becking, L.G.M., Kaplan, I.R., and Moore, D.. 1960. Limits of the natural environment in terms of pH and oxidation-reduction potentials. J. Geol. 68(3): 243284.Google Scholar
Beschel, R.E. 1963. Sulphur springs at Gypsum Hill. McGill Univ. Axel Heiberg Island Res., Prelim. Rep. 1961–62. pp. 183187.Google Scholar
Brandon, L.V. 1965. Groundwater hydrology and water supply in the District of Mackenzie, Yukon Territory, and adjoining parts of British Columbia. Geol. Surv. Can., Pap. 64–39. 102 pp.Google Scholar
Bryan, J.E. 1973. The influence of pipeline development on fresh-water fishery resources of northern Yukon Territory. Env:-Soc. Program. N. Pipelines, Task Force on Northern Oil Development, Rep. 73–6. 63 pp.Google Scholar
Cairnes, C.E. 1927. Observations on Lillooet Valley B.C. Can. Mining J. 47(8): 140144.Google Scholar
Cecile, M.P., and Jones, L.D.. 1979. Note on radioactive barite sinter, Bonnet Plume map area, District of Mackenzie. Geol. Surv. Can., Pap. 79–1B: 416.Google Scholar
Clapp, C.H. 1914. Sharp Point hot spring, Vancouver Island, B.C. Geol. Surv. Can., Sum. Rep. 1913. pp. 8083.Google Scholar
Clark, I.D., Fritz, P., and Michel, F.A.. 1977. Isotope hydrogeology and geothermometry of the Mount Meager geothermal area. Can. J. Earth Sci. 19: 14541473.Google Scholar
Cole, L.H. 1915. The salt deposits of Canada and the salt industry. Can. Mines Branch, Rep. 325. 152 pp.Google Scholar
Crandall, J.T., and Sadlier-Brown, T.L.. 1976. Data on geothermal areas, Cordilleran Yukon, Northwest Territories and adjacent British Columbia. Rept. for Can. Dept. Energy, Mines and Resources, Contract 1SQ5-0136. 63 pp.Google Scholar
Duffell, S., and Souther, J.G.. 1964. Geology of Terrace map area, B.C. Geol. Surv. Can., Mem. 329. 117 pp.Google Scholar
Elworthy, R.T. 1918. Mineral springs of Canada, Part II — The chemical character of some Canadian mineral springs. Can. Mines Branch, Bull. 20. 17 pp.Google Scholar
Elworthy, R.T. 1926. Hot springs in Western Canada — Their radioactivity and chemical properties, pp. 1–33 in Investigations of Mineral Resources and the Mining Industry, 1925. Can. Mines Branch, Rep. 669.Google Scholar
Gabrielse, H., Blusson, S.L., and Roddick, J.A.. 1973. Geology of Flat River, Glacier Lake, and Wrigley Lake map areas, District of Mackenzie and Yukon Territory. Geol. Surv. Can., Mem. 366. 153 pp.Google Scholar
Gussow, W.C. 1953. Carboniferous stratigraphy and structural geology of New Brunswick, Canada. Am. Assoc. Petr. Geol. Bull. 37(7): 17131816.Google Scholar
Haites, T.B. 1959. Banff thermal springs, a fascinating problem. J. Alberta Soc. Petr. Geol. 7(2): 2332.Google Scholar
Kerr, F.A. 1948. Lower Stikine and western Iskut River areas, British Columbia. Geol. Surv. Can., Mem. 246. 94 pp.Google Scholar
Krouse, H.R., Cook, F.D., Sasaki, A., and Smejkal, V.. 1970. Microbiological isotope fractionation in springs of Western Canada. Proceedings of the International Conference on Mass Spectrometry, Kyoto, Japan. Univ. Tokyo Press, Tokyo, Japan, pp. 629639.Google Scholar
Krouse, H.R., and van Everdingen, R.O.. 1984. δ34 variations in vegetation and soil exposed to intense biogenic sulphide emissions near Paige Mountain N.W.T., Canada. Water, Air Soil Poll. 23: 6167.Google Scholar
Mazor, E., van Everdingen, R.O., and Krouse, H.R.. 1983. Noble-gas evidence for geothermal activity in a karstic terrain: Rocky Mountains, Canada. Geochim. Cosmochim. Ada 47: 11111115.Google Scholar
McAllister, D.E. 1969. Introduction of tropical fishes into a hotspring near Banff, Alberta. Can. Fid Nat. 83(1): 3135.Google Scholar
McLean, F.H., and Kindle, E.D.. 1950. Geology of northeastern British Columbia. Geol. Surv. Can., Mem. 259. 236 pp.Google Scholar
Michel, F.A. 1977. Hydrogeologic studies of springs in the Central Mackenzie Valley, North-west Territories, Canada. M.Sc. thesis, University of Waterloo, Waterloo, Ont.185 pp.Google Scholar
Paquet, R. 1964. Mineral waters of Quebec. Can. Mining Metallur. Bull. 57(629): 939944.Google Scholar
Rutulis, M. 1983. Springs in Southern Manitoba. Wat. Resour. Branch, Manitoba Dept. Nat. Resour., Prelim. Rep. 23 pp.Google Scholar
Satterly, J., and Elworthy, R.T.. 1917. Mineral springs of Canada, Part I — The radioactivity of some Canadian mineral springs. Can. Mines Brch, Bull. 16. 60 pp.Google Scholar
Smejkal, V., Cook, F.D., and Krouse, H.R.. 1971. Studies of sulfur and carbon isotope fractionation with micro-organisms isolated from springs of Western Canada. Geochim. Cosmochim. Acta 35: 787800.Google Scholar
Souther, J.G., and Halstead, E.C.. 1969. Mineral and thermal waters of Canada. Proceedings of Symposium II, 23rd International Geology Congress. Academia, Prague, pp. 225256.Google Scholar
TERA Environmental Consultants. 1984. Natural history study of mineral and thermal springs in Canada. Report for Parks Canada, Vancouver, B.C.37 pp. + tables.Google Scholar
Tyrrell, J.B. 1892. Report on northwestern Manitoba with portions of the adjacent districts of Assiniboia and Saskatchewan. Geol. Surv. Can., A. Rep. 5: 1E225E.Google Scholar
van Everdingen, R.O. 1969. The Ink Pots — A group of karst springs in the Rocky Mountains near Banff. Can. J. Earth Sci. 6: 545554.Google Scholar
van Everdingen, R.O. 1970 a. Seasonal variations, Sulphur Mountain Hot Springs, Banff, Alberta, Inl. Wat. Branch, Tech. Bull. 33. 11 pp.Google Scholar
van Everdingen, R.O. 1970 b. The Paint Pots, Kootenay National Park, British Columbia — Acid spring water with extreme heavy-metal content. Can. J. Earth Sci. 7(3): 831852.Google Scholar
van Everdingen, R.O. 1971. Surface-water composition in Southern Manitoba reflecting discharge of saline subsurface water and subsurface solution of evaporites. Geol. Assoc. Can., Spec. Pap. 9. pp. 343352.Google Scholar
van Everdingen, R.O. 1972. Thermal and mineral springs in the Southern Rocky Mountains of Canada. Environ. Can. Rep. EN36-415/1972. 151 pp.Google Scholar
van Everdingen, R.O. 1974. Groundwater in permafrost regions of Canada, pp. 8393in Permafrost Hydrology. Proceedings of the Workshop Seminar, Can. Nat. Comm. Int. Hydrol. Decade.Google Scholar
van Everdingen, R.O. 1984. Dirty-water events at Rocky Mountain hot springs and their correlation with other short-lived phenomena. Can. J. Earth Sci. 21(9): 9971007.Google Scholar
van Everdingen, R.O. 1987. The importance of permafrost in the hydrological regime. Chapter 9, pp. 243276in Healey, M.C., and Wallace, R.R. (Eds.), Canadian Aquatic Resources. Can. Bull. Fish. Aquat. Sci. 215.Google Scholar
van Everdingen, R.O., Shakur, M.A., and Krouse, H.R.. 1982. Isotope geochemistry of dissolved, precipitated, airborne, and fallout sulfur species associated with springs near Paige Mountain, Norman Range, N.W.T. Can. J. Earth Sci. 19(7): 13951407.Google Scholar
van Everdingen, R.O., Shakur, M.A., and Krouse, H.R.. 1985 a. Role of corrosion by H2SO4 fallout in cave development in a travertine deposit — Evidence from sulfur and oxygen isotopes. Chem. Geol. 49: 205211.Google Scholar
van Everdingen, R.O., Shakur, M.A., and Michel, F.A.. 1985 b. Oxygen- and sulfur-isotope geochemistry of acidic groundwater discharge in British Columbia, Yukon, and District of Mackenzie, Canada. Can. J. Earth Sci. 22(11): 16891695.Google Scholar
Warren, P.S. 1927. Banff area, Alberta. Geol. Surv. Can., Mem. 153. 94 pp.Google Scholar
Williams, D.D., and Hogg, I.D.. 1988. Ecology and production of invertebrates in a Canadian coldwater spring – springbrook system. Holarctic Ecol. 11: 4154.Google Scholar