Hostname: page-component-76fb5796d-45l2p Total loading time: 0 Render date: 2024-04-25T17:30:02.771Z Has data issue: false hasContentIssue false

Energy expenditure in wild birds

Published online by Cambridge University Press:  28 February 2007

David M. Bryant
Affiliation:
Avian Ecology Unit, Department of Biological and Molecular Sciences, University of Stirling, Stirling FK9 4LA
Rights & Permissions [Opens in a new window]

Abstract

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Symposium on ‘Nutrition of wild and captive wild animals’ Plenary Lecture
Copyright
Copyright © The Nutrition Society 1997

References

REFERENCES

Aschoff, J. & Pohl, H. (1970). Der Ruheumsatz von Vogeln als Funktion der Tageszeit und der Korpergrosse. (The resting metabolism of birds as a function of time of day and body size). Journal für Ornithologie 111, 3847.CrossRefGoogle Scholar
Bennett, P. M. & Harvey, P. H. (1987). Active and resting metabolism in birds: allometry, phylogeny and ecology. Journal of Zoology, London 213, 327363.Google Scholar
Blaxter, K. (1989). Energy Metabolism in Animals and Man. Cambridge: Cambridge University Press.Google Scholar
Brody, S. (1945). Bioenergetics and Growth. New York: Reinhold.Google Scholar
Brown, J. H. & Maurer, B. A. (1986). Body size, ecological dominance, and Cope's rule. Nature 324, 248250.Google Scholar
Bryant, D. M. (1988). Energy expenditure and body mass changes as measures of reproductive costs in birds. Functional Ecology 2, 2334.Google Scholar
Bryant, D. M. (1989). Determination of respiration rates of free-living animals by the double-labelling technique. In Toward a More Exact Ecology vol. 30, pp. 85109 [Grubb, P. J. and Whittaker, J. B., editors]. Oxford: Blackwell Scientific Publications.Google Scholar
Bryant, D. M. (1991). Constraints on energy expenditure by birds. Acta XX Congressus Internationalis 4, 19892001.Google Scholar
Bryant, D. M., Hails, C. J. & Tatner, P. (1984). Reproductive energetics of two tropical bird species. Auk 101, 2537.Google Scholar
Bryant, D. M. & Tatner, P. (1988 a). The costs of brood provisioning: effects of brood size and food supply. Proceedings International Ornithological Congress XIX, 364379. Ottawa: University of Ottawa Press.Google Scholar
Bryant, D. M. & Tatner, P. (1988 b). Energetics of the annual cycle of Dippers Cinclus cinclus Ibis 130, 1738.Google Scholar
Bryant, D. M. & Tatner, P. (1991). Intraspecies variation in avian energy expenditure: correlates and constraints. Ibis 133, 236245.Google Scholar
Bryant, D. M. & Westerterp, K. R. (1980). The energy budget of the house martin Delichon urbica. Ardea 68, 91102.Google Scholar
Bryant, D. M. & Westerterp, K. R. (1983). Short-term variability in energy turnover by breeding house martins (Delichon urbica): a study using doubly labelled water. Journal of Animal Ecology 52, 524544.Google Scholar
Butler, P. J. (1981). Respiration during flight. In Advances in Physiological Sciences, vol. 10, pp. 155164 [Hutas, I. and Debreczeni, L. A., editors]. Oxford: Pergamon Press.Google Scholar
Calder, W. A. III (1984). Size, Function and Life History. Cambridge, MA: Harvard University Press.Google Scholar
Carlson, A., Moreno, J. & Alatalo, R. V. (1993). Winter metabolism of coniferous forest birds (Paridae) under arctic conditions: a study with doubly labeled water. Ornis Scandinavica 24, 161164.Google Scholar
Carpenter, F. L., Hixon, M. A., Russell, R. W., Paton, D. C. & Temeles, E. J. (1993). Interference asymmetries among age-sex classes of rufous hummingbirds during migratory stopovers. Behavioral Ecology and Sociobiology 33, 297304.Google Scholar
Casey, T. M. (1992). Energetics of locomotion. Advances in Comparative and Environmental Physiology, vol. 11, pp. 251275 [Alexander, M., editor]. Berlin: Springer Verlag.Google Scholar
Castro, G. & Myers, J. P. (1988). A statistical method to estimate the cost of flight in birds. Journal of Field Ornithology 59, 369380.Google Scholar
Cuthill, I. & Guilford, T. (1990). Perceived risk and obstacle avoidance in flying birds. Animal Behaviour 40, 188190.Google Scholar
Daan, S., Deerenberg, C. & Dijkstra, C. (1996). Increased daily work precipitates natural death in the kestrel. Journal of Animal Ecology 65, 539544.CrossRefGoogle Scholar
Dolnik, V. R. (1982). Time and energy budgets in free-living birds. Academy of Science USSR, Proceedings of the Zoological Institute 113, 137.Google Scholar
Drent, R. H. & Daan, S. (1980). The prudent parent: energetic adjustments in avian breeding. Ardea 68, 225252.Google Scholar
Evans, M. R. & Hatchwell, J. B. (1992). An experimental study of male adornment in the scarlet-tufted sunbird: II. The role of the elongated tail in mate choice and experimental evidence for a handicap. Behavioral Ecology and Sociobiology 29, 421427.Google Scholar
Feinsinger, P. & Chaplin, S. B. (1975). On the relationship between wing-disc loading and foraging strategy in hummingbirds. American Naturalist 109, 481497.Google Scholar
Furness, R. F. & Bryant, D. M. (1996). Effect of wind on field metabolic rates of breeding northern fulmars. Ecology 77, 11811188.Google Scholar
Gauthier, M. & Thomas, D. W. (1993). Nest site selection and cost of nest building by cliff swallows (Hirundo pyrhonotta). Canadian Journal of Zoology 71, 11201123.Google Scholar
Gill, F. (1990). Ornithology. New York: Freeman.Google Scholar
Goldstein, D. L. & Nagy, K. A. (1985). Resource utilization by desert quail: time, energy, food and water. Ecology 66, 378387.CrossRefGoogle Scholar
Greenwalt, C. H. (1962). Dimensional relationships for flying animals. Smithsonion Miscellaneous Collections 144, 146.Google Scholar
Hails, C. J. (1979). A comparison of flight energetics in hirundines and other birds. Comparative Biochemistry and Physiology 63A, 581585.CrossRefGoogle Scholar
Hails, C. J. (1983). The metabolic rate of tropical birds. Condor 85, 6165.CrossRefGoogle Scholar
Hammond, K. A. & Diamond, J. (1997). Maximal sustained energy budgets in humans and animals. Nature 386, 457462.Google Scholar
Harvey, P. H. & Pagel, M. S. (1991). The Comparative Method in Evolutionary Biology. Oxford: Oxford University Press.Google Scholar
Houston, A. I., McNamara, J. M.& Hutchinson, J. M. C. (1993). General results concerning the trade-off between gaining energy and avoiding predation. Philosophical Transactions of the Royal Society, London 341B, 375398.Google Scholar
Johnstone, I. G. (1994). Space use by passerine birds: a study of territory economics in robins Erithacus rubecula and dippers Cinclus cinclus. PhD thesis, University of Stirling.Google Scholar
Karasov, W. H., Brittingham, M. C. & Temple, S. A. (1992). Daily energy and expenditure by black-capped chickadees (Parus atricapillus) in winter. Auk 109, 393395.Google Scholar
Kendeigh, S. C., Dolnik, V. R. & Gavrilov, V. M. (1977). Avian energetics. Granivorous Birds in Ecosystems. International Biological Programme, vol. 12, pp. 127204 [Pinowski, J. K. and Kendeigh, S. C., editors]. Cambridge: Cambridge University Press.Google Scholar
King, J. R. (1974). Seasonal allocation of time and energy resources in birds. Avian Energetics, vol. 15, pp. 485 [Paynter, R. A., editor]. Cambridge, MA: Nuttall Ornithological Club.Google Scholar
Kleiber, M. (1975). The Fire of Life: An Introduction to Animal Energetics. New York: R. E. Krieger.Google Scholar
Lifson, N.& McClintock, R. (1966). Theory of use of the turnover rates of body water for measuring energy and material balance. Journal of Theoretical Biology 12 4674.Google Scholar
Lima, S. L. (1986). Predation risk and unpredictable feeding conditions: determinants of body mass in wintering birds. Ecology 67, 377385.Google Scholar
McNamara, J. M. & Houston, A. I. (1986). The common currency for behavioral decisions. American Naturalist 127. 358378.Google Scholar
McNamara, J. M. & Houston, A. I. (1987). Starvation and predation as factors limiting population size. Ecology 68, 15151519.CrossRefGoogle Scholar
Masman, D. & Klaassen, M. (1987). Energy expenditure during free-flight in trained and free-living Eurasian kestrels. Auk 104, 603616.Google Scholar
Mock, P. J. (1991). Daily allocation of time and energy of western bluebirds feeding nestlings. Condor 93, 598611.Google Scholar
Moller, A. P. (1991). Influence of wing and tail morphology on the duration of song flight in skylarks. Behavioral Ecology and Sociobiology 28, 309314.CrossRefGoogle Scholar
Moreno, J. & Carlson, A. (1989). Clutch size and the costs of incubation in the pied flycatcher Ficedula hypoleuca. Ornis Scandinavica 20, 123128.Google Scholar
Moreno, J., Carlson, A. & Alatalo, R. (1988). Winter energetics of coniferous forest tits Paridae in the north: the implications of body size. Functional Ecology 2, 163170.Google Scholar
Moreno, J., Gustafsson, L., Carlson, A. & Part, T. (1991). The cost of incubation in relation to clutch size in the collared flycatcher. Ibis 133, 186192.Google Scholar
Moreno, J. & Sanz, J. J. (1994). The relationship between the energy expenditure during incubation and clutch size in the pied flycatcher Ficedula hypoleuca. Journal of Avian Biology 25, 125130.Google Scholar
Nagy, K. A. (1980). CO2 production in animals: analysis of potential errors in the doubly-labeled water method. American Journal of Physiology 238, R466R473.Google Scholar
Nagy, K. A. (1987). Field metabolic rate and food requirement scaling in mammals and birds. Ecological Monographs 57, 111128.Google Scholar
Nagy, K. A. (1994). Field bioenergetics of mammals: what determines field metabolic rates. Australian Journal of Zoology 42, 4354.Google Scholar
Norberg, U. M. (1990). Vertebrate Flight: Mechanics, Physiology, Morphology, Ecology and Evolution. Berlin: Springer Verlag.Google Scholar
Pedley, T. J. (1977). Scale Effects in Animal Locomotion. London: Academic Press.Google Scholar
Pennycuick, C. J. (1989). Bird Flight Performance. Oxford: Oxford University Press.Google Scholar
Peters, R. H. (1983). The Ecological Implications of Body Size. New York: Cambridge University Press.CrossRefGoogle Scholar
Piersma, T. & Morrison, R. I. G. (1994). Energy expenditure and water turnover of incubating ruddy turnstones: high costs under high arctic climatic conditions. Auk 111, 366376.Google Scholar
Powers, D. R. & Conley, T. M. (1994). Field metabolic rate and food consumption of two sympatric hummingbird species in southeastern Arizona. Condor 96, 141150.Google Scholar
Powers, D. R. & Nagy, K. A. (1988). Field metabolic rate and food consumption by free-living Anna's hummingbirds (Calypte anna). Physiological Zoology 61, 500506.Google Scholar
Rayner, J. M. V. (1993). On aerodynamics and the energetics of vertebrate flapping flight. Fluid Dynamics in Biology, vol. 141, pp. 351400 [Cheer, A. Y. and van Dam, C. P., editors]. Providence, RI: American Mathematical Society.Google Scholar
Reyer, H-U. A. & Westerterp, K. R. (1985). Parental energy expenditure: a proximate cause of helper recruitment in the pied kingfisher Ceryle rudis. Behavioral Ecology and Sociobiology 17, 363369.Google Scholar
Ricklefs, R. E. (1991). Structures and transformations of life histories. Functional Ecology 5, 174183.CrossRefGoogle Scholar
Ricklefs, R. E. & Starck, J. M. (1996). Application of phylogenetically independent contrasts: a mixed progress report. Oikos 77, 167172.Google Scholar
Saether, B. -E. (1994). Reproductive strategies in relation to prey size in altricial birds: homage to Elton, Charles. American Naturalist 144, 285299.CrossRefGoogle Scholar
Saether, B. E., Andersen, R. & Pedersen, H. C. (1993). Regulation of parental effort in a long-lived seabird an experimental manipulation of the cost of reproduction in the antarctic petrel Thalassoica antarctica. Behavioral Ecology and Sociobiology 33, 147150.Google Scholar
Sibley, C. G. & Ahlquist, J. E. (1985). The phylogeny and classification of the passerine birds, based on comparisons of the genetic material, DNA. In Proceedings of the XVIII International Ornithological Congress, Moscow (1982), pp. 83121 [Ilyichev, V. N. and Gavrilov, V. M., editors]. Moscow: Navka Publications.Google Scholar
Speakman, J. R. & Racey, P. A. (1988). The doubly labelled water technique for measurement of CO2 production: principles and problems. Science Progress Series 72, 227237.Google Scholar
Steams, S. C. (1992). The Evolution of Life Histories. Oxford: Oxford University Press.Google Scholar
Storer, R. W. (1971). Classification of birds. In Avian Biology, vol. 1, pp. 149188 [Famer, D. S., King, J. R. and Parkes, K. C., editors]. New York: Academic Press.Google Scholar
Tatner, P. (1990). Energetic demands during brood rearing in the Wheatear Oenanthe oenanthe. Ibis 132, 423435.Google Scholar
Tatner, P. & Bryant, D. M. (1986). Flight cost of a small passerine measured using doubly-labelled water: implications for energetics studies. Auk 103, 169180.Google Scholar
Tatner, P. & Bryant, D. M. (1989). Doubly-labelled water technique for measuring energy expenditure. Techniques in Comparative Respiratory Physiology: An Experimental Approach, vol. 37, pp. 77112 [Bridges, C. R. and Butler, P. J., editors]. Cambridge: Cambridge University Press.Google Scholar
Tatner, P. & Bryant, D. M. (1993). Interspecific variation in daily energy expenditure during avian incubation. Journal of Zoology, London 231, 215232.Google Scholar
Thomas, A. L. R. (1993). On the aerodynamics of birds tails. Philosophical Transactions of the Royal Society, London 340B, 361380.Google Scholar
Thomas, D. W., Brigham, R. M. & LaPierre, H. (1996). Field metabolic rates and body mass changes in common poorwills (Phalaenoptilus nuttallii: Caprimulgidae). Ecoscience 3, 7074.Google Scholar
Trevelyan, R., Harvey, P. H. & Pagel, M. D. (1990). Metabolic rates and life histories in birds. Functional Ecology 4, 135141.Google Scholar
Tucker, V. A. (1973). Bird metabolism during flight: evaluation of a theory. Journal of Experimental Biology 58, 689709.CrossRefGoogle Scholar
Utter, J. M. (1971). Daily energy expenditure of free-living purple martins (Progne subis) and mockingbirds (Minus polyglottos) with a comparison of two northern populations of mockingbirds. PhD Thesis, Rutgers University, USA.Google Scholar
Walsberg, G. E. (1983). Avian ecological energetics. In Avian Biology, vol. 7, pp. 161220 [Farner, D. S. and King, J. R., editors]. New York: Academic Press.Google Scholar
Weathers, W. W. (1979). Climate adaptation in avian standard metabolic rate. Oecologia 42, 8189.Google Scholar
Weathers, W. W., Koenig, W. D. & Stanback, M. T. (1990). Breeding energetics and thermal ecology of the acorn woodpecker in central coastal California. Condor 92, 341359.Google Scholar
Weathers, W. W. & Nagy, K. A. (1980). Simultaneous doubly labelled water (3H H 180) and time budget estimates of daily energy expenditure in Phinopepla nitens. Auk 97, 861867.Google Scholar
Weathers, W. W. & Paton, D. C. (1997). Summer field metabolic rate and water intake rate in Superb Fairy-wrens and a White-throated Treecreeper. emu (In the Press).Google Scholar
Weathers, W. W., Paton, D. C. & Seymour, R. S. (1996). Field metabolic rate and water flux of nectarivorous honeyeaters. Australian Journal of Zoology 44, 445460.Google Scholar
Weathers, W. W. & Siegel, R. B. (1995). Body size establishes the scaling of avian postnatal metabolic rate: an interspecific analysis using phylogenetically independent contrasts. Ibis 137, 532542.Google Scholar
Weathers, W. W. & Stiles, F. G. (1989). Energetics and water balance in free-living tropical hummingbirds. Condor 91, 324331.CrossRefGoogle Scholar
Weathers, W. W. & Sullivan, K. A. (1989). Juvenile foraging proficiency, parental effort, and avian reproductive success. Ecological Monographs 59, 223246.Google Scholar
Weathers, W. W. & Sullivan, K. A. (1991). Foraging efficiency of parent juncos and their young. Condor 93, 346353.Google Scholar
Weathers, W. W. & Sullivan, K. A. (1993). Seasonal patterns of time and energy allocation by birds. Physiological Zoology 66, 511536.Google Scholar
Webster, M. D. & Weathers, W. W. (1990). Heat produced as a by-product of foraging activity contributes to thermoregulation by verdins, Auriparus flaviceps. Physiological Zoology 63, 777794.Google Scholar
West, G. B., Brown, J. H. & Enquist, B. J. (1997). A general model for the origin of allometric scaling laws in biology. Science 276, 122126.Google Scholar
Westerterp, K. R. & Bryant, D. M. (1984). Energetics of free-existence in swallows and martins (Hirundinidae) during breeding: a comparative study using doubly labelled water. Oecologia 62, 376381.Google Scholar
Westerterp, K. & Drent, R. H. (1985). Energetic costs and energy saving mechanisms in parental care of free-living passerine birds as determined by D28 method. Proceedings of the XVII International Ornithological Congress, Moscow (1982), pp. 392398 [Ilyichev, V. N. and Gavrilov, V. M., editors]. Moscow: Navka Publications.Google Scholar
Williams, J. B. (1987). Field metabolism and food consumption of savannah sparrows during the breeding season. Auk 104, 277289.Google Scholar
Williams, J. B. (1988). Field metabolism of tree swallows during the breeding season. Auk 105, 706714.Google Scholar
Williams, J. B. (1993). Energetics of incubation in free-living orange-breasted sunbirds in South Africa. Condor 95, 115126.Google Scholar
Williams, J. B., Bradshaw, D. & Schmidt, L. (1995). Field metabolism and water requirements of Spinifex pigeons (Geophaps plumifera) in Western Australia. Australian Journal of Zoology 43, 116.Google Scholar
Williams, J. B. & Du Plessis, M. A. (1996). Field metabolism and water flux of Sociable Weavers Philetairus socius in the Kalahari Desert. Ibis 138, 168171.Google Scholar
Williams, J. B., Withers, P. C., Bradshaw, S. D. & Nagy, K. A. (1991). Metabolism and water flux of captive and free-living Australian parrots. Australian Journal of Zoology 39, 131142.Google Scholar
Wilson, R. P. & Culik, B. M. (1993). Activity-specific metabolic rates from doubly labeled water studies: are activity costs underestimated? Ecology 74, 12851287.Google Scholar