Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-qxdb6 Total loading time: 0 Render date: 2024-04-26T23:47:24.140Z Has data issue: false hasContentIssue false

5 - Adaptations to hypoxia in fishes

Published online by Cambridge University Press:  05 June 2012

Göran E. Nilsson
Affiliation:
Universitetet i Oslo
Get access

Summary

Hypoxia in the aquatic environment

Both ocean and freshwater environments can challenge the inhabitants with large spatial and temporal variations in oxygen levels. As we pointed out in Chapter 1, oxygen has a low solubility and diffuses slowly in water. Further, the solubility of O2 in water falls with increases in temperature. At close to 0°C, air-saturated freshwater contains 10.2 ml O2 per liter, whereas at tropical temperatures (30°C) fresh water can only hold 5.9 ml O2 per liter when air saturated. These figures are even 20% or so lower in sea water, as salt reduces oxygen solubility (Table 1.1 in Chapter 1).

These physical factors make water breathing more challenging than air breathing, and particularly so when water oxygen levels are below air saturation. The oxygen that enters the water from the atmosphere, or is produced by photosynthesizing algae and phytoplankton, can be rapidly consumed by organisms and chemical oxidation reactions. There is no photosynthetic O2 production in the dark, and O2 diffusion is extremely slow in water (see Chapter 1), so oxygen movement to depth depends on convection, i.e. oxygen is carried to depth by water flow rather than diffusion. Surface waters generally have high oxygen content because of both photosynthesis and diffusion of oxygen from air. Aeration is increased by convection and mixing at the surface, a process that is strongly influenced by wind.

Type
Chapter
Information
Respiratory Physiology of Vertebrates
Life With and Without Oxygen
, pp. 131 - 173
Publisher: Cambridge University Press
Print publication year: 2010

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

Axelsson, M. and Fritsche, R. (1991). Effects of exercise, hypoxia and feeding on the gastrointestinal blood flow in the Atlantic cod Gadus morhua. J. Exp. Biol. 158, 181–98.Google ScholarPubMed
Bæuf, G. and Payan, P. (2001). How should salinity influence fish growth?Comp. Biochem. Physiol. 130C, 411–23.Google Scholar
Beamish, F. W. H. (1964). Respiration of fishes with special emphasis on standard oxygen consumption. III. Influence of oxygen. Can. J. Zool.. 42, 355–66.CrossRefGoogle Scholar
Bellwood, D. R. and Fisher, R. (2001). Relative swimming speeds in reef fish larvae. Mar. Ecol. Prog. Ser. 211, 299–303.CrossRefGoogle Scholar
Bernal, D., Dickson, K. A., Shadwick, R. E. and Graham, J. B. (2001). Analysis of the evolutionary convergence for high performance swimming in lamnid sharks and tunas. Comp. Biochem. Physiol. 129A, 695–726.CrossRefGoogle Scholar
Bernier, N. J., Craig, P. M. (2005) CRF-related peptides contribute to the stress response and the regulation of appetite in hypoxic rainbow trout. Am. J. Physiol. 289, R982–90.Google ScholarPubMed
Bernier, N., Harris, J., Lessard, J. and Randall, D. (1996). Adenosine receptor blockade and hypoxia-tolerance in rainbow trout and Pacific hagfish. I. Effects on anaerobic metabolism. J. Exp. Biol. 199, 485–95.Google ScholarPubMed
Bicego, K. C., Barros, R. C. and Branco, L. G. (2007). Physiology of temperature regulation: comparative aspects. Comp. Biochem. Physiol. 147A, 616–39.CrossRefGoogle Scholar
Booth, J. H. (1978). The distribution of blood flow in the gills of fish: application of a new technique to rainbow trout (Salmo gairdneri). J. Exp. Biol. 73, 119–30.Google Scholar
Booth, J. H. (1979a). The effect of oxygen supply, epinephrine and acetylcholine on the distribution of blood flow in trout gills. J. Exp. Biol. 83, 31–9.Google Scholar
Booth, J. H. (1979b). Circulation in trout gills: The relationship between branchial perfusion and the width of the lamellar blood space. Can. J. Zool. 57, 2185–93.CrossRefGoogle Scholar
Bosworth, C. A. 4th, Chou, C. W., Cole, R. B. and Rees, B. B. (2005). Protein expression patterns in zebrafish skeletal muscle: initial characterization and the effects of hypoxic exposure. Proteomics 5, 1362–71.CrossRefGoogle ScholarPubMed
Boutilier, R. G., Dobson, G., Hoeger, U. and Randall, D. J. (1987). Acute exposure to graded levels of hypoxia in rainbow trout (Salmo gairdneri): metabolic and respiratory adaptations. Respir. Physiol. 71, 69–82.CrossRefGoogle Scholar
Branson, B. A. and Hake, P. (1972). Observations of an accessory breathing mechanism in Piaractus nigripinnis (Cope). Zoologischer Anzeiger 189, 292–7.Google Scholar
Bras, M., Queenan, B. and Susin, S. A. (2005). Programmed cell death via mitochondria: different modes of dying. Biochemistry (Mosc.) 70, 231–9.Google ScholarPubMed
Braum, E. and Junk, W. J. (1982). Morphological adaptation of two Amazonian characoids (Pisces) for surviving in oxygen deficient waters, Int. Revue Res. Hydrobiol. 67, 869–86.Google Scholar
Brauner, C. J. and Randall, D. J. (1996). The interaction between oxygen and carbon dioxide movements in fishes. Comp. Biochem. Physiol. 113A 1, 83–90.Google Scholar
Brett, J. R. (1979) Environmental factors and growth. In Fish Physiology Vol. 8, Bioenergetics and Growth (eds. Hoar, W. S., Randall, D. J. and Brett, J. R.). San Diego: Academic Press, pp. 599–675.CrossRefGoogle Scholar
Brett, J. R. and Groves, T. D. D. (1979) Physiological energetics. In Fish Physiology Vol. 8, Bioenergetics and Growth, ed. Hoar, W. S., Randall, D. J. and Brett, J. R.. San Diego: Academic Press, pp. 280–352.Google Scholar
Brusilow, S. W. (2002). Hyperammoniemic encephalopathy. Medicine 81, 240–9.CrossRefGoogle Scholar
Butler, P. J. and Taylor, E. W. (1975). The effect of progressive hypoxia on respiration in the dogfish (Scyliorhinus canicula) at different seasonal temperatures. J. Exp. Biol. 63, 117–30.Google ScholarPubMed
Butler, P. J., Taylor, E. W., Capra, M. F. and Davison, W. (1978). The effect of hypoxia on the levels of circulating cathecholamines in the dogfish Scyliorhinus canicula. J. Comp. Physiol. B 127, 325–30.CrossRefGoogle Scholar
Burggren, W. W. (1982). ‘Air gulping’ improves blood oxygen transport during aquatic hypoxia in the goldfish Carassius auratus. Physiol. Zool. 55, 327–34.CrossRefGoogle Scholar
Burggren, W., McMahon, B. and Powers, D. (1991). Respiratory functions of blood. In Environmental and Metabolic Animal Physiology, ed. Prosser, C. L.. New York: Wiley-Liss, pp. 437–508.Google Scholar
Burleson, M. L., Wilhelm, D. R. and Smatresk, N. J. (2001). The influence of fish size on the avoidance of hypoxia and oxygen selection by largemouth bass. J. Fish Biol. 59, 1336–1349.Google Scholar
Chan, D. K. O. and Wong, T. M. (1977). Physiological adjustments to dilution of the external medium in the lip shark, Hemiscyllium plagiosum (Bennet). III. Oxygen consumption and metabolic rates. J. Exp. Zool. 200, 97–102.CrossRefGoogle Scholar
Chapman, L. J. and Hulen, K. G. (2001). Implications of hypoxia for the brain size and gill morphometry of mormyrid fishes. J. Zool. 254, 461–72.CrossRefGoogle Scholar
Chapman, L. J., Galis, F. and Shinn, J. (2000). Phenotypic plasticity and the possible role of genetic assimilation: hypoxia-induced trade-offs in the morphological traits of an African cichlid. Ecol. Lett. 3, 387–93.CrossRefGoogle Scholar
Chapman, L. J., Chapman, C. A., Nordlie, F. G. and Rosenberger, A. E. (2002). Physiological refugia: swamps, hypoxia tolerance and maintenance of fish diversity in the Lake Victoria region. Comp. Biochem. Physiol. 133A, 421–37.CrossRefGoogle Scholar
Chen, C. T., Liu, K. W. and Young, S. J. (1999). Preliminary report on Taiwan's whale shark fishery. In Elasmobranch Biodiversity, Conservation and Management, Proc. Int. Seminar and Workshop in Sabah, Malaysia, ed. Fowler, S. L., Reid, T. and Dipper, F. A.. IUCN, Gland, Switzerland, pp. 162–7.Google Scholar
Cruz-Neto, A. P. and Steffensen, J. F. (1997). The effects of acute hypoxia and hypercapnia on oxygen consumption of the freshwater European eel. J. Fish Biol. 50, 759–69.CrossRefGoogle Scholar
Davie, P. S. and Farrell, A. P. (1991). The coronary and luminal circulations of the myocardium of fishes. Can. J. Zool. 2, 158–164.Google Scholar
Davis, J. C. (1972) An infrared photographic technique useful for studying vascularization of fish gills. J. Fish. Res. Board Can. 29, 109–11.CrossRefGoogle Scholar
Boeck, G., Smet, H. and Blust, R. (1995). The effect of sublethal levels of copper on oxygen consumption and ammonia excretion in the common carp, Cyprinus carpio. Aquatic Toxicol. 32, 127–41.CrossRefGoogle Scholar
Farrell, A. P. (1985). A protective effect of adrenaline on the acidotic teleost heart. J. Exp. Biol. 116, 503–8.Google Scholar
Farrell, A. P. (2007). Tribute to P. L. Lutz: a message from the heart–why hypoxic bradycardia in fishes?J. Exp. Biol. 210, 1715–25.CrossRefGoogle Scholar
Farrell, A. P. and Jones, D. R. (1992). The heart. In Fish Physiology. Vol. 12A, ed. Hoar, W. S., Randall, D. J. and Farrell, A. P.. San Diego: Academic Press, pp. 1–88.Google Scholar
Farrell, A. P. and Stecyk, J. A. W. (2007). The heart as a working model to explore themes and strategies for anoxic survival in ectothermic vertebrates. Comp. Biochem. Physiol. 147A, 300–12.CrossRefGoogle Scholar
Farrell, A. P., MacLeod, K. and Chancey, B. (1986). Intrinsic mechanical properties of the perfused rainbow trout heart and the effects of catecholamines and extracellular calcium under control and acidotic conditions. J. Exp. Biol. 125, 319–45.Google ScholarPubMed
Farrell, A. P., Sobin, S. S., Randall, D. J. and Crosby, S. (1980). Intralamellar blood flow patterns in fish gills. Am. J. Physiol. 239, R428–36.Google ScholarPubMed
Farrell, A. P., Hinch, S. G., Cooke, S. J., et al. (2008). Pacific Salmon in hot water: applying aerobic scope models and biotelemetry to predict the success of spawning migrations. Physiol. Biochem. Zool. 81, 697–708.CrossRefGoogle ScholarPubMed
Felipo, V. and Butterworth, R. F. (2002). Neurobiology of ammonia. Prog. Neurobiol. 67, 259–79.CrossRefGoogle ScholarPubMed
Fernandes, M. N. and Rantin, F. T. (1989). Respiratory responses of Oreochromis niloticus (Pisces, Cichlidae) to environmental hypoxia under different thermal conditions. J. Fish Biol. 35, 509–19.CrossRefGoogle Scholar
Fernandes, M. N., Rantin, F. T., Kalinin, A. L. and Moron, S. E. (1994). Comparative study of gill dimensions of 3 Erythrinid species in relation to their respiratory function. Can. J. Zool. 72, 160–5.CrossRefGoogle Scholar
Foss, A., Evensen, T. H. and Oiestad, V. (2002). Effects of hypoxia and hyperoxia on growth and food conversion efficiency in the spotted wolfish Anarhichas minor (Olafsen). Aquaculture Res. 33, 437–44.CrossRefGoogle Scholar
Frederich, M. and Pörtner, H. O. (2000). Oxygen limitation of thermal tolerance defined by cardiac and ventilatory performance in spider crab, Maja squinado. Am. J. Physiol. 279, R1531–8.Google ScholarPubMed
Frimodt, C. (1995). Multilingual Illustrated Guide to the World's Commercial Coldwater Fish. Oxford: Fishing News Books.Google Scholar
Fry, F. E. J (1971) The effect of environmental factors on the physiology of fish. In Fish Physiology Vol. 6, ed. Hoar, W. S. and Randall, D. J.. New York: Academic Press, pp. 1–98.Google Scholar
Fry, F. E. J. and Hart, J. S. (1948). The relation of temperature to oxygen consumption in the goldfish. Biol. Bull. 94, 66–77.CrossRefGoogle ScholarPubMed
Gallaugher, P. and Farrell, A. P. (1998). Hematocrit and blood oxygen-carrying capacity. In Fish Physiology Vol. 17, Fish Respiration, ed. Perry, S. F. and Tufts, B. L.. New York: Academic Press, pp. 185–227.Google Scholar
Gesser, H. and Jorgensen, E. (1982). pHi, contractility and Ca-balance under hypercapnic acidosis in the myocardium of different vertebrate species. J. Exp. Biol. 96, 405–12.Google ScholarPubMed
Gilmore, K. M. (1998) Causes and consequences of acid-base disequilibria. In Fish Physiology Vol. 17, Fish Respiration, ed. Perry, S. F. and Tufts, B. L.. New York: Academic Press, pp. 321–348.Google Scholar
Gonzalez, R. J. and McDonald, D. G. (1992). The relationship between oxygen consumption and ion loss in a freshwater fish. J. Exp. Biol. 163, 317–32.Google Scholar
Gooding, R. M., Neill, W. H. and Dizon, A. E. (1981). Respiration rates and low-oxygen tolerance limits in skipjack tuna, Katsuwonus pelamis. Fisheries Bull. 79, 31–48.Google Scholar
Gracey, A. Y. (2007). Interpreting physiological responses to environmental change through gene expression profiling. J. Exp. Biol. 210, 1584–92.CrossRefGoogle ScholarPubMed
Gracey, A. Y., Troll, J. V. and Somero, G. N. (2001). Hypoxia-induced gene expression profiling in the euryoxic fish Gillichthys mirabilis. Proc. Natl. Acad. Sci. USA 98, 1993–8.CrossRefGoogle ScholarPubMed
Gray, I. E. (1954). Comparative study of the gill area of marine fishes. Biol. Bull. 107, 219–55.CrossRefGoogle Scholar
Hand, S. C. and Hardewig, I. (1996). Downregulation of cellular metabolism during environmental stress: mechanisms and implications. Ann. Rev. Physiol. 58, 539–63.CrossRefGoogle ScholarPubMed
Hochachka, P. W., Buck, L. T., Doll, C. J. and Land, S. C. (1996). Unifying theory of hypoxia tolerance: molecular metabolic defense and rescue mechanisms for surviving oxygen lack. Proc. Natl. Acad. Sci. USA 93, 9493–8.CrossRefGoogle ScholarPubMed
Holeton, G. F. and Randall, D. J. (1967a). The effect of hypoxia upon the partial pressure of gases in blood and water afferent and efferent to the gills of rainbow trout. J. Exp. Biol. 46, 317–27.Google ScholarPubMed
Holeton, G. F., and Randall, D. J. (1967b). Changes in blood pressure in the rainbow trout during hypoxia. J. Exp. Biol. 46, 297–305.Google ScholarPubMed
Hughes, G. M. and Umezawa, S.-I. (1968). On respiration in the dragonet Callionymus lyra L. J. Exp. Biol. 49, 565–82.Google ScholarPubMed
Ip, Y. K., Peh, B. K., Tam, W. L.Wong, W. P. and Chew, S. F. (2005). Effects of intra-peritoneal injection with NH4Cl, urea or NH4Cl + urea on nitrogen excretion and metabolism in the African lungfish Protopterus dolloi, J. Exp. Zool. 303A, 272–82.CrossRefGoogle Scholar
Jensen, F. B., Fago, A. and Weber, R. E. (1998). Hemoglobin structure and functions. In Fish Physiology, Vol. 17, Fish Respiration, ed. Perry, S. F. and Tufts, B. L.. San Diego: Academic Press, pp. 1–40.Google Scholar
Jibb, L. A. and Richards, J. G. (2008) AMP-activated protein kinase activity during metabolic rate depression in the hypoxic goldfish, Carassius auratus. J. Exp. Biol. 211, 3111–22.CrossRefGoogle ScholarPubMed
Johansson, D., Nilsson, G. E. and Törnblom, E. (1995). Effects of anoxia on energy metabolism in crucian carp brain slices studied with microcalorimetry. J. Exp. Biol. 198, 853–9.Google ScholarPubMed
Ju, Z., Wells, M. C., Heater, S. J. and Walter, R. B. (2007). Multiple tissue gene expression analyses in Japanese medaka (Oryzias latipes) exposed to hypoxia. Comp. Biochem. Physiol. 145C, 134–44.Google Scholar
Korshunov, S. S., Korkina, O. V., Ruuge, E.Kskulachev, V. P. and Starkov, A. A. (1998). Fatty acids as natural uncouplers preventing generation of O2− and H2O2 by mitochondria in the resting state. FEBS Lett. 435, 215–18.CrossRefGoogle ScholarPubMed
Kottelat, M., Britz, R., Hui, T. H. and Witte, K.-E. (2006). Paedocypris, a new genus of Southeast Asian cyprinid fish with a remarkable sexual dimorphism, comprises the world's smallest vertebrate. Proc. R. Soc. Lond. B, Biol. Sci. 273, 895–9.CrossRefGoogle ScholarPubMed
Kramer, D. L. and McClure, M. (1982). Aquatic surface respiration, a widespread adaptation to hypoxia in tropical fishes. Env. Biol. Fish. 7, 47–55.CrossRefGoogle Scholar
Krauss, S., Zhang, C. Y. and Lowell, B. B. (2005). The mitochondrial uncoupling-protein homologues. Nature Rev. Mol. Cell Biol. 6, 248–61.CrossRefGoogle ScholarPubMed
Kudo, H., Kato, A. and Hirose, S. (2007). SourceFluorescence visualization of branchial collagen columns embraced by pillar cells. J. Histochem. Cytochem. 55, 57–62.CrossRefGoogle ScholarPubMed
Kutty, M. N. (1968). Respiratory quotients in goldfish and rainbow trout. J. Fish. Res. Bd. Can. 25, 1689–728.CrossRefGoogle Scholar
Lai, J. C, Kakuta, I., Mok, H. O., Rummer, J. L. and Randall, D. (2006). Effects of moderate and substantial hypoxia on erythropoietin levels in rainbow trout kidney and spleen. J. Exp. Biol. 209, 2734–8.CrossRefGoogle ScholarPubMed
Lam, K. K. Y. (1999). Hydrography, nutrients and phytoplankton, with special reference to a hypoxic event, at an experimental artificial reef at Hoi Ha Wan, Hong Kong. Asian Mar. Biol. 16, 35–64.Google Scholar
Landry, C. A., Steele, S. L., Manning, S. and Cheek, A. O. (2007). Long term hypoxia suppresses reproductive capacity in the estuarine fish, Fundulus grandis. Comp. Biochem. Physiol. 148A, 317–23.CrossRefGoogle Scholar
Lannig, G., Bock, C., Sartoris, F. J. and Pörtner, H. O. (2004). Oxygen limitation of thermal tolerance in cod, Gadus morhua L., studied by magnetic resonance imaging and on-line venous oxygen monitoring. Am. J. Physiol. 287, R902–10.Google ScholarPubMed
Lu, G., Mak, Y. T., Wai, S. M., et al. (2005). Hypoxia-induced differential apoptosis in the central nervous system of the sturgeon (Acipenser shrenckii). Microscopy Res. Tech. 68, 258–63.CrossRefGoogle Scholar
Matey, V., Richards, J. G., Wang, Y. X., et al. (2008). The effect of hypoxia on gill morphology and ionoregulatory status in the Lake Qinghai scaleless carp, Gymnocypris przewalskii. J. Exp. Biol. 211, 1063–74.CrossRefGoogle ScholarPubMed
McKenzie, D. J., Wong, S, Randall, D. J., Egginton, S, Taylor, E. W. andd Farrell, A. P. (2004). The effects of sustained exercise and hypoxia upon oxygen tensions in the red muscle of rainbow trout. J. Exp. Biol. 207, 3629–37.CrossRefGoogle ScholarPubMed
Muusze, B., Marcon, J., Thillart, G. and Almeida-Val, V. (1998). Hypoxia tolerance of Amazon fish, respirometry and energy metabolism of the cichlid Astronotus ocellatus. Comp. Biochem. Physiol. 120A, 151–6.CrossRefGoogle Scholar
Nikinmaa, M. and Salama, A. (1998). Oxygen transport in fish. In Fish Physiology, Vol. 17 Fish Respiration, ed. Perry, S. F. and Tufts, B. L.. San Diego: Academic Press, pp. 141–84.Google Scholar
Nilsson, S. (1986). Control of gill blood flow. In Fish Physiology: Recent Advances, ed. Nilsson, S. and Holmgren, S.. London: Croom Helm, pp. 87–101.CrossRefGoogle Scholar
Nilsson, G. E. (1991). The adenosine receptor blocker aminophylline increases anoxic ethanol production in crucian carp. Am. J. Physiol. 261, R1057–60.Google Scholar
Nilsson, G. E. (1992). Evidence for a role of GABA in metabolic depression during anoxia in crucian carp (Carassius carassius L.). J. Exp. Biol. 164, 243–59.Google Scholar
Nilsson, G. E. (1996). Brain and body oxygen requirements of Gnathonemus petersii, a fish with an exceptionally large brain. J. Exp. Biol. 199, 603–7.Google ScholarPubMed
Nilsson, G. E. (2007). Gill remodeling in fish – a new fashion or an ancient secret?J. Exp. Biol. 210, 2403–9.CrossRefGoogle ScholarPubMed
Nilsson, G. E. and Östlund-Nilsson, S. (2008). Does size matter for hypoxia tolerance in fish?Biol. Rev. 83, 173–89.CrossRefGoogle Scholar
Nilsson, G. E. and Renshaw, G. M. C. (2004). Hypoxic survival strategies in two fishes: extreme anoxia tolerance in the North European crucian carp and natural hypoxic preconditioning in a coral-reef shark. J. Exp. Biol. 20, 3131–9.CrossRefGoogle Scholar
Nilsson, G. E., Hobbs, J.-P. A. and Östlund-Nilsson, S. (2007a). A tribute to P. L. Lutz: Respiratory ecophysiology of coral-reef teleosts. J. Exp. Biol. 210, 1673–86.CrossRefGoogle Scholar
Nilsson, G. E., Hylland, P. and Löfman, C. O. (1994). Anoxia and adenosine induce increased cerebral blood flow in crucian carp. Am. J. Physiol. 267, R590–5.Google ScholarPubMed
Nilsson, G. E., Rosén, P. and Johansson, D. (1993). Anoxic depression of spontaneous locomotor activity in crucian carp quantified by a computerized imaging technique. J. Exp. Biol. 180, 153–63.Google Scholar
Nilsson, G. E., Crawley, N., Lunde, I. G. and Munday, P. L. (2009). Elevated temperature reduces the respiratory scope of coral reef fishes. Global Change Biol. 15, 1405–12.CrossRef
Nilsson, G. E., Hobbs, J.-P. A., Munday, P. L. and Östlund-Nilsson, S. (2004). Coward or braveheart: extreme habitat fidelity through hypoxia tolerance in a coral-dwelling goby. J. Exp. Biol. 27, 33–9.CrossRefGoogle Scholar
Nilsson, G. E., Östlund-Nilsson, S., Penfold, R. and Grutter, A. S. (2007b). From record performance to hypoxia tolerance – respiratory transition in damselfish larvae settling on a coral reef. Proc. R. Soc. Lond. B, Biol. Sci. 274, 79–85.CrossRefGoogle ScholarPubMed
Olson, K. R. (2008). Hydrogen sulfide and oxygen sensing: implications in cardiorespiratory control. J. Exp. Biol. 211, 2727–34.CrossRefGoogle ScholarPubMed
Olson, K. R., Dombkowski, R. A., Russell, M. J., et al. (2006). Hydrogen sulfide as an oxygen sensor/transducer in vertebrate hypoxic vasoconstriction and hypoxic vasodilation. J. Exp. Biol. 209, 4011–23.CrossRefGoogle ScholarPubMed
Ong, K. J., Stevens, E. D. and Wright, P. A. (2007). Gill morphology of the mangrove killifish (Kryptolebias marmoratus) is plastic and changes in response to terrestrial air exposure. J. Exp. Biol. 210, 1109–15.CrossRefGoogle ScholarPubMed
Östlund-Nilsson, S and Nilsson, G. E. (2004). Breathing with a mouth full of eggs: respiratory consequences of mouthbrooding in cardinalfishes. Proc. R. Soc. Lond. B, Biol. Sci. 271, 1015–22.CrossRefGoogle Scholar
Pärt, P.Tuurala, H., Nikinmaa, M. and Kiessling, A. (1984). Evidence for a non-respiratory intralamellar shunt in perfused rainbow trout gills. Comp. Biochem. Physiol. 79A, 29–34.CrossRefGoogle Scholar
Pelster, B. and Randall, D. J. (1998) The physiology of the Root effect. In Fish Physiology, Vol. 17 Fish Respiration, ed. Perry, S. F. and Tufts, B. L.San Diego: Academic Press., pp. 321–48.Google Scholar
Perry, S. F and Gilmour, K. M. (1996). Consequences of catecholamine release on ventilation and blood oxygen transport during hypoxia and hypercapnia in an elasmobranch Squalus acanthias and a teleost Oncorhynchus mykiss. J. Exp. Biol. 199, 2105–18.Google Scholar
Pettersson, K. (1983). Adrenergic control of oxygen transfer in perfused gills of the cod, Gadus morhua. J. Exp. Biol. 102, 327–335.Google Scholar
Pichavant, K., Person-Le-Ruyet, J., Bayon, N., et al. (2000) Effects of hypoxia on growth and metabolism of juvenile turbot. Aquaculture 188, 103–44.CrossRefGoogle Scholar
Poon, W. L., Hung, C. Y., Nakano, K. and Randall, D. J. (2007). An in vivo study of common carp (Cyprinus carpio L.) liver during prolonged hypoxia. Comp. Biochem. Physiol. D 2, 295–302.Google Scholar
Pörtner, H. O. and Knust, R. (2007). Climate change affects marine fishes through the oxygen limitation of thermal tolerance. Science 315, 95–7.CrossRefGoogle ScholarPubMed
Pörtner, H. O., Mark, F. C. and Bock, C. (2004). Oxygen limited thermal tolerance in fish? – Answers obtained by nuclear magnetic resonance techniques. Respir. Physiol. Neurobiol. 141, 243–60.CrossRefGoogle ScholarPubMed
Prosser, C. L., Barr, L. M., Pinc, R. D. and Lauer, C. Y. (1957). Acclimation of goldfish to low concentrations of oxygen. Physiol. Zool. 30, 137–41.CrossRefGoogle Scholar
Randall, D. J. (1982). The control of respiration and circulation in fish during exercise and hypoxia. J. Exp. Biol. 100, 275–88.Google Scholar
Randall, D. J. and Daxboeck, C. (1984). Oxygen and carbon dioxide transfer across fish gills. In Fish Physiology, Vol. XA, ed. Hoar, W. S. and Randall, D. J.. New York: Academic Press, pp. pp. 263–314.Google Scholar
Randall, D. J. and Ip, Y. K. (2006). Ammonia as a respiratory gas in water and air-breathing fishes. Respir. Physiol. Neurobiol. 154, 216–25.CrossRefGoogle ScholarPubMed
Randall, D. J. and Smith, J. C. (1967). The regulation of cardiac activity in fish in a hypoxic environment. Physiol. Zool. 40, 104–13.CrossRefGoogle Scholar
Randall, D. J., Holeton, G. F. and Stevens, E. Don. (1967). The exchange of oxygen and carbon dioxide across the gills of rainbow trout. J. Exp. Biol. 46, 339–48.Google ScholarPubMed
Rausch, R. N., Crawshaw, L. I. and Wallace, H. L. (2000). Effects of hypoxia, anoxia, and endogenous ethanol on thermoregulation in goldfish, Carassius auratus. Am. J. Physiol. 278, R545–55.Google ScholarPubMed
Renshaw, G. M. C, Kerrisk, C. B. and Nilsson, G. E. (2002). The role of adenosine in the anoxic survival of the epaulette shark, Hemiscyllium ocellatum. Comp. Biochem. Physiol. 131B, 133–41.CrossRefGoogle Scholar
Richards, J. G., Sardella, B. A. and Schulte, P. M. (2008). Regulation of pyruvate dehydrogenase in the common killifish, Fundulus heteroclitus, during hypoxia exposure. Am. J. Physiol. 295, R979–90.Google ScholarPubMed
Robb, T. and Abrahams, M. V. (2003). Variation in tolerance to hypoxia in a predator and prey species: an ecological advantage of being small?J. Fish Biol. 62, 1067–81.CrossRefGoogle Scholar
Routley, M. H., Nilsson, G. E. and Renshaw, G. M. C. (2002). Exposure to hypoxia primes the respiratory and metabolic responses of the epaulette shark to progressive hypoxia. Comp. Biochem. Physiol. A 131, 313–21.CrossRefGoogle ScholarPubMed
Saint-Paul, U. (1988). Diurnal routine O2 consumption at different O2 concentration by Colossoma macropomum and Colossoma brachypomum (Teleostei, Serrasalmidae). Comp. Biochem. Physiol. 89A, 675–82.CrossRefGoogle Scholar
Saunders, R. L. (1962). The irrigation of the gills in fishes. II. Efficiency of oxygen uptake in relation to respiratory flow activity and concentrations of oxygen and carbon dioxide. Can. J. Zool. 40, 817–62.CrossRefGoogle Scholar
Schaack, S. and Chapman, L. J. (2003). Interdemic variation in the African cyprinid Barbus neumayeri: correlations among hypoxia, morphology, and feeding performance. Can. J. Zool. 81, 430–40.CrossRefGoogle Scholar
Schmidt-Nielsen, K. (1984). Scaling: Why is Animal Size so Important?Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Schurmann, H. and Steffensen, J. F. (1997). Effects of temperature, hypoxia and activity on the metabolism of juvenile Atlantic cod. J. Fish Biol. 50, 1166–80.Google Scholar
Secor, D. H. and Gunderson, T. E. (1998). Effects of hypoxia and temperature on survival, growth, and respiration of juvenile Atlantic sturgeon, Acipenser oxyrinchus. Fish. Bull. 96, 603–13.Google Scholar
Shang, E. H. H and Wu, R. S. S. (2004) Aquatic hypoxia is a teratogen and affects fish embryonic developmentEnviron. Sci. Tech. 38, 4763.CrossRefGoogle ScholarPubMed
Shang, E. H. H., Yu, R. M. K. and Wu, R. S. S. (2006). Hypoxia affects sex differentiation and development, leading to a male-dominated population in zebrafish (Danio rerio). Environ. Sci. Tech. 40, 3118–22.Google Scholar
Shepard, M. P. (1955). Resistance and tolerance of young speckled trout (Salvelinus fontinalis) to oxygen lack, with special reference to low oxygen acclimation. J. Fish Res. Bd. Can. 12, 387–433.CrossRefGoogle Scholar
Smith, R. W., Cash, P., Ellefsen, S. and Nilsson, G. E. (2009). Proteomic changes in the crucian carp brain during exposure to anoxia. Proteomics, 9, 2217–29.CrossRefGoogle ScholarPubMed
Smith, R. W., Houlihan, D. F., Nilsson, G. E. and Brechin, J. G. (1996). Tissue specific changes in protein synthesis rates in vivo during anoxia in crucian carp. Am. J. Physiol. 271, R897–904.Google ScholarPubMed
Sloman, K. A., Wood, C. M., Scott, G. R., et al. (2006). Tribute to R. G. Boutilier: the effect of size on the physiological and behavioural responses of oscar, Astronotus ocellatus, to hypoxia. J. Exp. Biol. 209, 1197–205.CrossRefGoogle Scholar
Söderström, V., Renshaw, G. M. C. and Nilsson, G. E. (1999). Brain blood flow and blood pressure during hypoxia in the epaulette shark (Hemiscyllium ocellatum), a hypoxia tolerant elasmobranch. J. Exp. Biol. 202, 829–35.Google Scholar
Soivio, A. and Tuurala, H. (1981). Structural and circulatory responses to hypoxia in the secondary lamellae of Salmo gairdneri gills at two temperatures. J. Comp. Physiol. B 145, 37–43.CrossRefGoogle Scholar
Sollid, J., Weber, R. E. and Nilsson, G. E. (2005). Temperature alters the respiratory surface area of crucian carp Carassius carassius and goldfish Carassius auratus. J. Exp. Biol. 208, 1109–16.CrossRefGoogle ScholarPubMed
Sollid, J., Angelis, P.Gundersen, K. and Nilsson, G. E. (2003). Hypoxia induces adaptive and reversible gross-morphological changes in crucian carp gills. J. Exp. Biol. 206, 3667–73.CrossRefGoogle ScholarPubMed
Sørensen, C., Øverli, Ø, Summers, C. H. and Nilsson, G. E. (2007). Social regulation of neurogenesis in teleosts. Brain Behav. Evol. 70, 239–46.CrossRefGoogle ScholarPubMed
Stecyk, J. A. W. and Farrell, A. P. (2006). Regulation of the cardiorespiratory system of common carp (Cyprinus carpio) during severe hypoxia at three seasonal acclimation temperatures. Physiol. Biochem. Zool. 79, 614–27.CrossRefGoogle ScholarPubMed
Stecyk, J. A. W. and Farrell, A. P. (2007). Effects of extracellular changes on spontaneous heart rate of normoxia- and anoxia-acclimated turtles (Trachemys scripta). J. Exp. Biol. 210, 421–31.CrossRefGoogle Scholar
Stecyk, J. A. W., Stensløkken, K.-O., Farrell, A. P. and Nilsson, G. E. (2004). Maintained cardiac pumping in anoxic crucian carp. Science 306, 77.CrossRefGoogle ScholarPubMed
Stensløkken, K.-O., Sundin, L. and Nilsson, G. E. (1999). Cardiovascular and gill microcirculatory effects of endothelin-1 in Atlantic cod: evidence for pillar cell contraction. J. Exp. Biol. 202, 1151–7.Google ScholarPubMed
Stensløkken, K.-O., Sundin, L. and Nilsson, G. E. (2006). Endothelin receptors in teleost fish: cardiovascular effects and branchial distribution. Am. J. Physiol. 290, R852–60.Google ScholarPubMed
Stensløkken, K.-O., Sundin, L., Renshaw, G. M. C. and Nilsson, G. E. (2004). Adenosinergic and cholinergic control mechanisms during hypoxia in the epaulette shark (Hemiscyllium ocellatum), with emphasis on branchial circulation. J. Exp. Biol. 207, 4451–61.CrossRefGoogle Scholar
Stensløkken, K.-O., Ellefsen, S., Stecyk, J. A. W., Dahl, M. B., Nilsson, G. E. and Vaage, J. (2008). Differential regulation of AMP-activated kinase and AKT kinase in response to oxygen availability in crucian carp (Carassius carassius). Am. J. Physiol. 295, R1403–14.Google Scholar
Storey, K. B. and Storey, J. M. (2004). Metabolic rate depression in animals: transcriptional and translational controls. Biol. Rev. 79, 207–33.CrossRefGoogle ScholarPubMed
Sultana, N., Nag, K., Kato, A. and Hirose, S. (2007). Pillar cell and erythrocyte localization of fugu ETA receptor and its implication. Biochem. Biophys. Res. Comm. 355, 149–55.CrossRefGoogle Scholar
Sundin, L. and Nilsson, S. (1992). Arterio-venous branchial blood flow in the Atlantic cod Gadus morhua. J. Exp. Biol. 165, 73–84.Google Scholar
Sundin, L. and Nilsson, G. E. (1997). Neurochemical mechanisms behind gill microcirculatory responses to hypoxia in trout: in-vivo microscopy study. Am. J. Physiol. 272, R576–85.Google ScholarPubMed
Sundin, L. and Nilsson, G. E. (1998a). Acute defence mechanisms against haemorrhage from mechanical gill injury in rainbow trout. Am. J. Physiol. 275, R460–5.Google Scholar
Sundin, L. and Nilsson, G. E. (1998b). Endothelin redistributes blood flow through the lamellae of rainbow trout gills: evidence for pillar cell contraction. J. Comp. Physiol. B 168, 619–23.CrossRefGoogle Scholar
Taylor, E. W. and Barrett, D. J. (1985). Evidence of a respiratory role for the hypoxic bradycardia in the dogfish Scyliorhinus canicula L. Comp. Biochem. Physiol. 80A, 99–102.CrossRefGoogle Scholar
Taylor, J. C. and Miller, J. M. (2001). Physiological performance of juvenile southern flounder, Paralichthys lethostigma, in chronic and episodic hypoxia. J. Exp. Mar. Biol. Ecol. 258, 195–214.CrossRefGoogle ScholarPubMed
Tervonen, V., Ruskoaho, H., Lecklin, T., Ilves, M. and Vuolteenaho, O. (2002). Salmon cardiac natriuretic peptide is a volume-regulating hormone. Am. J. Physiol. 283, E353–61.Google ScholarPubMed
Thomas, P., Rahman, S., Kummer, J. and Khan, I., (2005). Neuroendocrine Changes Associated with Reproductive Dysfunction in Atlantic Croaker after Exposure to Hypoxia. Baltimore: Society of Environmental Toxicology and Chemistry, p. 59.Google Scholar
Thompson, E. B. (1998) Special topic: apoptosis, Ann. Rev. Physiol. 60, 525–32.CrossRefGoogle ScholarPubMed
Tsui, T. K. N., Randall, D. J., Hanson, L., Farrell, A. P., Chew, S. F. and Ip, Y. K. (2004) Dogmas and controversies in the handling of nitrogenous wastes: ammonia tolerance in the oriental weatherloach, Misgurnus anguillicaudatus. J. Exp. Biol. 207, 1977–83.CrossRefGoogle ScholarPubMed
Tufts, B. L. and Perry, S. F. (1998) Carbon dioxide transport and excretion. In Fish Physiology, Vol. 17 Fish Respiration, ed. Perry, S. F. and Tufts, B. L.. New York: Academic Press, pp. 229–81.Google Scholar
Tufts, B. L. and Randall, D. J. (1989) The functional significance of adrenergic pH regulation in fish erythrocytes. Can. J. Zool. 67, 235–8.CrossRefGoogle Scholar
Tuurala, H., Egginton, S. and Soivio, A. (1998). Cold exposure increases branchial water-blood barrier thickness in the eel. J. Fish Biol. 53, 451–5.Google Scholar
Ultsch, G. R., Jackson, D. C. and Moalli, R. (1981). Metabolic oxygen conformity among lower vertebrates–the toadfish revisited. J. Comp. Physiol. 142, 439–43.CrossRefGoogle Scholar
Val, A. L. (2000). Organic phosphates in the red blood cells of fish. Comp. Biochem. Physiol. 125A, 417–35.CrossRefGoogle Scholar
Val, A. L. and Almeida-Val, V. M. F. (1995). Fishes of the Amazon and their Environment: Physiological and Biochemical Features. Heidelberg: Springer Verlag.CrossRefGoogle Scholar
Linden, A.Verhoye, M. and Nilsson, G. E. (2001). Does anoxia induce cell swelling in carp brains? Dymanic in vivo MRI measurements in crucian carp and common carp. J. Neurophysiol. 85, 125–33.CrossRefGoogle Scholar
Meer, D. L., Thillart, G. E., Witte, F., et al. (2005). Gene expression profiling of the long-term adaptive response to hypoxia in the gills of adult zebrafish. Am. J. Physiol. 289, R1512–19.Google ScholarPubMed
Ginneken, V., Nieveen, M., VanEersel, R., Thillart, G. and Addink, A. (1996). Neurotransmitter levels and energy status in brain of fish species with and without the survival strategy of metabolic depression. Comp. Biochem. Physiol. 114A, 189–96.CrossRefGoogle Scholar
Waversveld, J., Addink, A. D. F. and Thillart, G. (1989). Simultaneous direct and indirect calorimetry on normoxic and anoxic goldfish. J. Exp. Biol. 142, 325–35.Google Scholar
Verheyen, E., Blust, R. and Decleir, W. (1994). Metabolic rate, hypoxia tolerance and aquatic surface respiration of some lacustrine and riverine African cichlid fishes. Comp. Biochem. Physiol. A 107, 403–11.CrossRefGoogle Scholar
Wahlqvist, I. and Nilsson, S. (1980). Adrenergic control of the cardio-vascular system of the Atlantic cod, Gadus morhua, during ‘stress’. J. Comp. Physiol. 137, 145–50.CrossRefGoogle Scholar
Walsh, P. J., Veauvy, C. M., McDonald, M. D., Pamenter, M. E., Buck, L. T. and Wilkie, M. P. (2007). Piscine insights into comparisons of anoxia tolerance, ammonia toxicity, stroke and hepatic encephalopathy. Comp. Biochem. Physiol. 147A, 332–43.CrossRefGoogle Scholar
Wang, S-H., Yuen, S. F., Randall, D. J., et al. (2008) Hypoxia inhibits fish spawning via LH-dependent final oocyte maturation. Comp. Biochem. Physiol. 148C, 363–9.Google Scholar
Watson, W. and WalkerJr, J. J. (2004). The world's smallest vertebrate, Schindleria brevipinguis, a new paedomorphic species in the family Schindleriidae (Perciformes: Gobioidei). Records of the Australian Museum 56, 139–42.CrossRefGoogle Scholar
Winemiller, K. O. (1989). Development of dermal lip protuberances for aquatic surface respiration in South American characid fishes. Copeia, 1989, 382–90.CrossRefGoogle Scholar
Wood, C. M. (2001) Toxic responses of the gill. In Target Organ Toxicity in Marine and Freshwater Teleosts, Vol. 1 – Organs, ed. Schlenk, D. and Benson, W. H.. London: Taylor & Francis, pp. 1–89.Google Scholar
Wu, R. S. S., Zhou, B. S., Randall, D. J., Woo, N. Y. S. and Lam, P. K. S. (2003). Aquatic hypoxia is an endocrine disruptor and impairs fish reproduction. Envir. Sci. Tech. 37, 1137–41.CrossRefGoogle ScholarPubMed
Yoshikawa, H., Ishida, Y., Kawata, K., Kawai, F. and Kanamori, M. (1995). Electroencephalograms and cerebral blood-flow in carp, Cyprinus carpio, subjected to acute hypoxia. J. Fish Biol. 46, 114–22.CrossRefGoogle Scholar
Zhou, B. S., Wu, R. S. S., Randall, D. J. and Lam, P. K.S. (2001) Bioenergetics and RNA/DNA ratios in the common carp (Cyprinus carpio) under hypoxiaJ. Comp. Physiol. B 171, 49–57.CrossRefGoogle ScholarPubMed

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
×