Вы находитесь на странице: 1из 44

F I S H and F I S H E R I E S, 2003, 4, 147^190

Compensatory growth in fishes: a response to growth


depression

M Ali1,y, A Nicieza2 & R J Wootton1,

1
Institute of Biological Sciences, University of Wales Aberystwyth, Aberystwyth SY233DA,Wales, UK; 2Departamento de
Biolog|¤ a de Organismos y Sistemas, Universidad de Oviedo, E-33071 Oviedo, Spain


Abstract Correspondence:
Compensatory growth (CG) is a phase of accelerated growth when favourable condi- R J Wootton,
Institute of
tions are restored after a period of growth depression. CG reduces variance in size by
Biological Sciences,
causing growth trajectories to converge and is important to ¢sheries management, University of Wales
aquaculture and life history analysis because it can o¡set the e¡ects of growth arrests. Aberystwyth,
Compensatory growth has been demonstrated in both individually housed and Aberystwyth SY23
grouped ¢sh, typically after growth depression has been induced by complete or partial 3DA,Wales, UK
Tel.:
food deprivation. Partial, full and over-compensation have all been evoked in ¢sh,
þ44 1970 622346
although over-compensation has only been demonstrated when cycles of deprivation Fax:
and satiation feeding have been imposed. Individually housed ¢sh have shown that CG is þ44 1970 622350
partlya response to hyperphagiawhen rates of food consumptionare signi¢cantly higher E-mail:
thanthose in ¢shthat have notexperiencedgrowthdepression.The severityof thegrowth rjw@aber.ac.uk
y
depression increases the duration of the hyperphagic phase rather than maximum daily Present address:
Institute of Pure &
feeding rate. In many studies, growth e⁄ciencies were higher during CG. Changes in Applied Biology,
metabolic rate and swimming activity have not been demonstrated yet to playa role. Bahauddin
Periods of food deprivation induce changes in the storage reserves, particularly lipids, Zakariya
of ¢sh. Apart from the strong evidence for the restoration of somatic growth trajec- University, Multan,
Pakistan
tories, CG is a response to restore lipid levels. Although several neuro-peptides, includ-
ing neuropeptide-Y, are probably involved in the control of appetite, their role and the
Received15 July 2002
role of hormones, such as growth hormone (GH) and insulin-like growth factor (IGF),
Accepted 26 Feb 2003
in the hyperphagia associated with CG are still unclear.
The advantages of CG probably relate to size dependencies of mortality, fecundityand
diet that are characteristic of teleosts. These size dependencies favour a recovery from
the e¡ects of growth depression if environmental factors allow. High growth rates
may also impose costs, including adverse e¡ects on future development, growth, repro-
duction and swimming performance. Hyperphagia may lead to riskier behaviour in
the presence of predators. CG’s evolutionary consequences are largely unexplored. An
understanding of why animals grow at rates below their physiological capacity, an eva-
luation of the costs of rapid growth and the identi¢cation of the constraints on growth
trajectories represent major challenges for life-history theory.

Keywords growth e⁄ciency, growth regulation, hyperphagia, storage levels, trade-o¡s

Introduction 148

Characteristics of compensatory growth 149


Degree of compensation 149
Focus of compensation 150

# 2003 Blackwell Publishing Ltd 147


Compensatory growth in ¢shes M Ali et al.

Taxonomic distribution of compensatory response in teleosts 150

Methodological problems in studies on compensatory growth 155


Experimental studies 155
Field studies 157

Factors affecting compensatory growth in fish 158


E¡ect of feeding protocols including length and intensity of deprivation 158
In£uence of social factors 162
Ontogenetic changes 162
Seasonal variation 163
Sexual maturation and reproduction 164
Compensatory growth of body components 165

Proximate causes of compensatory growth 165


Hyperphagia 166
Responses to free access to food after a period of growth depression 166
E¡ects of food quality on hyperphagic response 168
Physiological basis of hyperphagia: consumption, absorption and evacuation rates 168
Physiological basis of compensatory growth: metabolic rate 169
Growth e⁄ciency 170

Proximate composition, metabolites and endocrines during food deprivation


and recovery 171
Lipids 171
Protein and RNA:DNA ratios 172
Carbohydrates 173
Water content 173
Endocrines 173

Modelling compensatory growth 174


Structural and storage model 174
Model based on speci¢c growth rate 175

Functional significance of compensatory growth 176


Size-dependent mortality 176
Size-dependent prey selection 176
Size-dependent ontogenetic changes 176
Size-dependent reproductive success 177

Costs of compensatory growth 177

Compensatory growth as a by-product of flexible growth and dynamic trade-offs 178

Future directions for research 179

Acknowledgements 181

References 181

growth depression may achieve the same size-at-age


Introduction
as conspeci¢cs experiencing environmental condi-
Many organisms exhibit faster growth during recov- tions that are more favourable. The response, which
ery from total or partial food deprivation than they tends to restore the original growth trajectory, is
do during periods of continuous food availability called compensatory, or catch-up, growth. The phe-
(Wilson and Osbourn 1960; Jobling1994). The conse- nomenon raises some important, but largely unan-
quence is that animals experiencing a period of swered, questions in the ¢elds of evolutionary and

148 # 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190


Compensatory growth in ¢shes M Ali et al.

applied ecology.Why do animals sometimes grow at a generally, the negative correlation between
rate below their physiological potential? What are growth rates in successive time periods (Ricker
the implications for the development of life-history 1975 and references therein).
theory? What are the implications for the manage- 2 To describe the response of increased individual
ment of exploited or endangered species? growth rates in a population after a reduction in
Compensatory growth has mostly been studied in population density has increased food supply
domesticated endotherms, although it has been (density-dependent; Ferreri and Taylor1996). Such
observed in a range of invertebrates and vertebrates an increase is predictable from the known rela-
(Wilson and Osbourn 1960; Sibly and Calow 1986). tionship between growth and feeding rate in ¢sh
Teleosts (referred to hereafter as ¢sh, unless other- (Weatherley and Gill 1987; Wootton 1998). Such
wise quali¢ed) are ectotherms with patterns of inde- ‘competitive release’ is often observed in predation
terminate growth, which allow the examination of experiments (Wootton1998).
the compensatory process at almost every stage of 3 To describe the growth resulting from an extended
their life cycles. Compensatory growth has been stu- period of feeding prior to some size-dependent
died, at least super¢cially, in more ¢sh species than life-cycle switch, such as smolti¢cation in salmo-
in any other vertebrate taxa. Research on ¢sh has nids (Nicieza and Bran‹a1993a,b).
also pioneered the analysis of the e¡ects of growth Only the ¢rst form of compensatory growth is the
depressors other than starvation, including unsea- primary subject of this review. It is still unclear what
sonably low temperatures, exposure to hypoxia or role compensatory growth plays in natural popula-
prophylactics, and reproductive e¡ort. There are also tions, or whether the phenomenon can be used in the
practical reasons for studies of growth compensation context of aquaculture to optimise ¢sh production.
in ¢sh. The group includes many species of economic
and social importance because of exploitation in
Characteristics of compensatory growth
commercial or recreational ¢sheries, or use in aqua-
culture. This review provides a framework on which Accelerated growth in response to previous food
further studies can build in a more integrated way restriction provides evidence that growth rates are
than has characterised studies on compensatory regulated. It suggests that animals can ‘evaluate’
growth so far. their achieved growth trajectory and adjust growth
Compensatory growth is identi¢ed by being signif- rates to bu¡er deviations from an ‘ideal’ trajectory
icantly faster than the growth rate of control animals (Hubbell 1971; but see Broekhuizen et al. 1994). The
that have not experienced growth depression, held expression of this growth regulation is dependent on
under comparable conditions. This accelerated several factors (Wilson and Osbourn 1960; Ryan
growth eventually declines to growth rates typical 1990). These include the nature, severity and dura-
of the control animals. Compensatory growth has tion of the under-nutrition, the stage of development
also been de¢ned as any growth that reduces the var- at the start of under-nutrition, the age at sexual
iance in the system (Atchley 1984) or as the negative maturity and the pattern of re-alimentation.
correlation between the growth of a trait in succes-
sive time intervals (Ricker 1969, 1975; Riska et al.
Degree of compensation
1984). Growth may be controlled by feedback
mechanisms, which adjust growth rates to achieve a In terms of a targeted growth trajectory (Monteiro
target trajectory (Tanner 1963; Monteiro and Fal- and Falconer1966; Riska et al.1984), the consequence
coner 1966; Hubbell 1971). An important conse- of a compensatory growth response is the attainment
quence of growth compensation is the convergence of a size status relative to the size achieved by an
of the growth trajectories of individuals, which organism that has not experienced any phase of
re£ects the ‘target-seeking’ consequence of the growth impairment. Implicitly, the size of the latter
catch-up growth. It is a process that tends to canalise organism is assumed to be optimum at every time, so
ontogenetic changes in size, and to bu¡er the e¡ects the ratio between the size of compensating and con-
of environmental variability (Wootton1998). trol animals when the compensatory response has
In the literature on ¢sh growth, the term ‘compen- abated provides a measure of the e¡ectiveness of the
satory growth’ has been used in several ways: compensation. In full compensation, the deprived
1 To describe the accelerated growth by an indivi- animals eventually achieve the same size at the same
dual after a period of growth depression, or more age as continuously fed contemporaries (Fig. 1). In

# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190 149


Compensatory growth in ¢shes M Ali et al.

Figure 1 Idealised patterns of


growth compensation based on
Jobling (1994).

partial compensation, the deprived animals fail to primary body components. The e¡ects of nutritional
achieve the same size at the same age as nonres- conditions on body composition and physiological
tricted contemporaries, but do show relatively rapid state of teleosts are well understood (e.g. McMillan
growth rates, and may have better food conversion and Houlihan 1992; Jobling 1993; Mommsen 1998).
ratios during the re-alimentation period (Fig. 1). Over- The in£uence of undernutrition on the growth
compensation occurs when the animals that had dynamics of length and mass can di¡er substantially
experienced a restricted ration achieve a greater size (for instance, mass loss is often associated with no or
at the same age than nonrestricted animals (Fig. 1). negligible decrease in length). Therefore, the
The compensation is so strong that animals subject dynamics of growth recovery may also be di¡erent,
to variable food supplies exhibit a higher growth both in the degree and in the time of compensation
rate than those for whom food was continuously (e.g. Nicieza and Metcalfe 1997). In particular, there
available. This was observed in Lepomis hybrids by is evidence that recovery of lipid or energy levels
Hayward et al. (1997), but seems to be a rare outcome. can operate through a pathway separate from that
If the re-alimented ¢sh resume growing at a rate char- leading to compensation for structural growth (Bull
acteristic of the size reached at the end of the depriva- et al. 1996; Nicieza and Metcalfe 1997; Metcalfe et al.
tionperiod, no compensation has occurred (Fig. 1). 2002).

Focus of compensation Taxonomic distribution of compensatory


response in teleosts
In ¢sh, the term ‘compensatory growth’ is used
usually to describe increases in growth rates in The term ‘compensatory growth’ was ¢rst used in
whole body length or mass. However, the relative relation to mammals and has subsequently been
growth of body components can be altered by manip- demonstrated in a range of domesticated endotherms
ulation of environmental conditions, thus producing (Wilson and Osbourn 1960). Earlier studies on ¢sh
shape modi¢cation (Emerson 1986). The organism had reported negative correlations between growth
can experience a reduced growth of some body com- rates in successive periods (review in Zivkov 1982).
ponents without signi¢cant alteration in its whole Despite these early observations, there were few stu-
size. It remains unknown whether there can be com- dies on compensatory growth in ¢sh until the early
pensation for such uncoupling in the growth of di¡er- 1990s when the subject attracted attention, espe-
ent body parts with the ‘normal’ shape restored. cially in relation to aquaculture. The occurrence of
Compensatory responses can be examined in rela- compensatory growth has been reported for a
tion to the growth of di¡erent tissues, organs or restricted number of ¢sh taxa (Table 1). Even so, ¢sh

150 # 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190


# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190

Table 1 Taxonomic distribution of studies on growth compensation in teleosts and incidence of full compensation (FC), partial compensation (PC), over-compensation (OvC) and absence of
compensation (AC), where information is available.

Species (by family) Methods/experimental protocol Manipulated factor Effect Response variable Reference

Clupeidae
Clupea harengus (Herring) Group-housed, posthatch larvae Food supply FC Length, soluble protein Pedersen et al. (1990)
C. harengus Group-housed, posthatch larvae Food supply AC Length Pedersen (1993)
C. harengus Natural population, back-calculation None PC Length Moores and Winters (1982)

Salmonidae
Salmo salar (Atlantic salmon) Group-housed, mature vs. immature parr None FC Length Skilbrei (1990)
S. salar Group-housed Food supply PC Mass Thorpe et al. (1990)
S. salar Group-housed, juveniles Food supply FC Lipid content Metcalfe and Thorpe (1992)
S. salar Group-housed, preadults Food supply PC Mass Reimers et al. (1993)
S. salar Group-housed Temperature PC Mass Mortensen and Damsgård (1993)
S. salar Group-housed Salinity FC Mass Damsgård and Arnesen (1998)
S. salar Natural population, back-calculation None PC Length Nicieza and Braña (1993a,b)
S. salar Group-housed, juvenile None PC Length Nicieza et al. (1994a)
S. salar Group-housed Food supply, FC, PC Length, mass Nicieza and Metcalfe (1997)
temperature
S. salar Group-housed, juveniles Food supply FC Lipid content Bull and Metcalfe (1997)
S. salar Group-housed, juveniles Temperature FC Length Maclean and Metcalfe (2001)
S. salar Group-housed, juveniles Food supply PC, FC Length, lipid content Morgan and Metcalfe (2001)
S. salar Group-housed, postsmolts Food supply PC, FC Johansen et al. (2001)

Compensatory growth in ¢shes M Ali et al.


Oncorhynchus mykiss (rainbow trout) Group-housed Food supply PC, FC, OvC Body components Weatherley and Gill (1981)
O. mykiss Group-housed Food supply FC Mass Dobson and Holmes (1984)
O. mykiss Group-housed Food supply FC Mass Smith (1987)
O. mykiss Group-housed, cycles of Food supply FC Mass Kindschi (1988)
deprivation and refeeding
O. mykiss Group-housed Food supply PC Mass, body components Quinton and Blake (1990)
O. mykiss Group-housed Food supply FC,PC Mass Teskeredzic et al. (1995)
O. mykiss Group-housed Food supply Mass Jobling and Koskela (1996)
O. mykiss Group-housed H2O2 prophylactic FC Mass Speare and Arsenault (1997)
treatment

Oncorhynchus kisutch Group-housed Food supply FC Mass Damsgård and Dill (1998)
(Coho salmon)
Oncorhynchus nerka Group-housed Food supply PC, FC Mass Bilton and Robins (1973)
151

(Sockye salmon)
152

Compensatory growth in ¢shes M Ali et al.


Table 1 continued

Species (by family) Methods/experimental protocol Manipulated factor Effect Response variable Reference

Salvelinus alpinus (Arctic charr) Individually housed Food supply PC Mass Miglavs and Jobling (1989a,b)
S. alpinus Group-housed, cycles of deprivation and refeeding Food supply PC Mass Jobling et al. (1993)
S. alpinus Group-housed Temperature PC Mass Mortensen and Damsgård (1993)

First feeding Coregonid larva Group-housed Food supply Mass Dabrowski et al. (1986)
Characiformes
Piractus brachypomus Group-housed Food supply FC Mass Koppe et al. (1993)
Cyprinidae
Cyprinus carpio (carp), Leuciscus Natural populations, back-calculation PC, FC Length Zivkov (1982)
cephalus, Alburnus alburnus,
Rutilus rutilus (roach) Carassius
auratus (goldfish)
Cyprinus carpio, Leuciscus cephalus, Natural populations, back-calculation Length Zivkov (1996)
Rutilus rutilus, Abramis brama
Cyprinus carpio Group-housed: Food supply NC Mass Schwarz et al. (1985)
protein or energy deprived
# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190

Tor putitora Group-housed Tandon and Johal (1983a)


Labeo rohita, Group-housed Tandon and Johal (1983b)
Cirrhina mrigala, Catla catla
Cyprinus carpio Group-housed Johal and Tandon (1986)
Cyprinus carpio Group-housed Food supply Bastrop et al. (1991).
Scardinus erythropthalmus (rudd), Group-housed Food supply Mass Wieser et al. (1992)
Leuciscus cephalus (chub)
Chalcalburnus chalcoides mento
(Danubian bleak)
Phoxinus phoxinus Individually housed Food supply NC, FC Mass Russell and Wootton (1992)
(European minnow)

Phoxinus phoxinus Individually housed Food supply FC Mass Zhu et al. (2001)
Rutilus rutilus Group-housed Food supply Mass, enzyme Mendez and Wieser (1993)
concentrations
Carassius auratus gibelio Group-housed Food supply FC Mass Qian et al. (2000)
(Gibel carp)
C. auratus gibelio Individually housed Food supply FC Mass Xie et al. (2001)
Abramis brama Group-housed Raikova-Petrova and Zivkov (1998)
# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190

Ictaluridae
Ictalurus puntatus, Channel catfish Group-housed Food supply FC Mass Kim and Lovell (1995)
I. punctatus Group-housed, cycles of deprivation Food supply FC Mass Gaylord and Gatlin (2001)
I. punctatus Group-housed, cycles of deprivation Food supply NC Mass Gaylord et al. (2001)
Heterobranchus longifilis Group-housed, single deprivation and cyclic Food supply PC, FC Mass Luquet et al. (1995)
(African catfish) deprivation

Cyprinodontidae
Cypronodon nevadensis Group-housed Gerking and Raush (1979)

Gadidae
Gadus morhua (Atlantic cod) Natural population Beacham (1981)
G. morhua Group-housed, cycles of deprivation and Food supply NC, FC Mass Jobling et al. (1994)
single deprivation

Gasterosteidae
Gasterosteus aculeatus Full-compensation Food supply FC Mass Zhu et al. (2001)
(three-spined stickleback)

Cichlidae
Oreochromis mossambicus Group-housed Temperature Gonad size Chmilevskii (1994)
O. mossambicus Group-housed Chmilevskii (1998)
O. mossambicus Group-housed Chmilevskii (1996)
O. mossambicus Group-housed Food supply Christensen and McLean (1998)
O. niloticus Group-held, juveniles Density FC Mass Vera-Cruz and Mair (1994)
O. niloticus Group-housed Density FC Mass Basiao et al. (1996)
O. mossambicus  O. niloticus hybrids Group-housed in sea water Food supply PC, FC Mass Wang et al. (2002)

Sparidae

Compensatory growth in ¢shes M Ali et al.


Pagrus pagrus Group-held, demand feeders FC Mass Rueda et al. (1998)

Carcharhinidae
Trichiurus japonicus Natural population Length Haweet et al. (1996)

Centropomidae
Lates calcarifer Aranyakananda et al. (1996)

Anarhicadidae
Anarhicas minor Group-housed Oxygen concentration Mass Foss and Imsland (2002)

Percichthyidae
Golden perch, Macquaria ambigua Group-housed Food supply Lipid, protein Collins and Anderson 1995
153
154

Compensatory growth in ¢shes M Ali et al.


Table 1 continued

Species (by family) Methods/experimental protocol Manipulated factor Effect Response variable Reference

Percidae
Perca flavescens Yellow perch Individually housed; cycles of deprivation Food supply PC, FC Mass Hayward and Wang (2001)
and refeeding
Stizostedion lucioperca (pike-perch) Group-housed Zivkov and Raikova-Petrova (1991)

Centrarchidae
Lepomis cyanellus  L. macrochirus Individually housed; cycles of deprivation Food supply OvC, FC, PC Mass Hayward et al. (1997) Whitledge
(hybrid sunfish) and refeeding et al. (1998)
L. cyanellus  L. macrochirus Group-housed; cycles of deprivation and feeding Food supply PC Mass Hayward et al. (2000)

Pomatomidae
Pomatomus saltatrix Group-housed PC Buckel et al. (1998)

Pleuronectidae
Pleuronectes americanus Group-housed Length Chambers et al. (1988)
(Winter flounder)
P. americanus Group-housed, pre- and postmetamorphic Length Bertram et al. (1993).
# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190

Hippoglossus hipoglossus Group-housed Food supply Björnsson et al. (1992)


Pleuronectus asper Group-housed Food supply PC Mass Paul et al. (1995)
(Alaska yellowfin sole)
Compensatory growth in ¢shes M Ali et al.

studies constitute the best source of information on the degrees of freedom used for testing the experi-
compensatory growth in ectotherms. mental e¡ect (Ruohonen et al. 2001).
A disproportionate number of studies of compen- The use of grouped ¢sh also introduces two more
satory growth have focused on six salmonid species. factors that are usually ignored in the analysis and
The number of studies describing compensatory interpretation: density and numerical abundance.
growth in cyprinids is lower, but there are data on Any conclusions from a study will strictly be valid
about 13 species. To our knowledge, there are only a only for the density of ¢sh used unless density is
few studies on compensatory growth for some taxo- deliberately manipulated as an experimental factor.
nomic groups whose biology is well researched (e.g. If density is manipulated, care needs to be taken to
Poeciliids, Cichlids). The lack of information is even avoid changing numerical abundance simulta-
more striking for strictly marine species. Moreover, neously. Equally, an experimental manipulation of
there is evidence of interspeci¢c variation in patterns numerical abundance per group as a factor must
of growth compensation, so species having similar avoid changing density. In some studies of compensa-
geographical ranges, similar social behaviour, and tory growth, the numerical size of groups in tanks
similar diets can have distinct compensatory capa- has been reduced during an experiment to provide
cities and di¡erent mechanisms of compensation intermediate samples. This manipulation introduces
(e.g. Sogard and Olla 2002).This accentuates the need another confounding factor, changing both density
for a wider taxonomic coverage in the study of com- and numerical abundance simultaneously. The use
pensatory growth. A lack of comparative studies of individually housed ¢sh, if they do not show patho-
using the same protocols means that it is unclear logical behaviours, avoids all these problems.
whether these reports are describing the same phe- Growth and consumption rates have an allometric
nomenon or whether several di¡erent causal path- relationship to body mass in ¢sh (Jobling 1994;
ways can lead to the observation of unusually high Wootton 1998). Size-speci¢c rates tend to decline
growth rates. with size. Fish that have experienced growth depres-
sion will be smaller than control ¢sh, and so may
exhibit higher speci¢c growth and consumption
Methodological problems in studies on
rates simply because of allometry. Consequently,
compensatory growth
comparisons for these variables between control and
experimental ¢sh should be made for ¢sh of approxi-
Experimental studies
mately the same size, either by experimental manipu-
An evaluation of the experimental evidence for com- lation or by covariance adjustment.
pensatory growth is hampered by the prevalence of A third problem, particularly with earlier studies,
poor experimental design and inappropriate statisti- is the failure to use appropriate statistical analyses.
cal analyses, which obscure the valid identi¢cation A common error was the use of serial t-tests to make
of a compensatory response. A common £aw is the multiple comparisons where analysis of variance
misidenti¢cation of the appropriate level of replica- with preplanned contrasts was required (Underwood
tion in relation to the detection of experimental 1997). In some studies, a series of analyses of variance
e¡ects, leading to the problem of pseudoreplication have been used, where repeated-measures anova
(see Underwood 1997 for an excellent discussion). should have been used to analyse changes in growth
The use of grouped ¢sh in experiments can be justi- and consumption over time.
¢ed on two grounds. First, some species characteristi- These problems mean that some highly cited stu-
cally live in shoals and, if kept in isolation, show dies may not have demonstrated real growth com-
pathological behaviour with reduced appetite and pensation, but simply the plasticity of ¢sh growth in
poor growth (e.g. Stirling 1976). Second, many relation to environmental variables, including food
experiments on compensatory growth have been in availability, ¢sh density and abundance, social inter-
relation to aquaculture, a context in which ¢sh are actions and even unique tank e¡ects. Table 2 pro-
maintained at high densities. However, if grouped vides a critique of methodologies of some studies
¢sh are used, it has to be recognised that the unit of considered in further detail below. Despite these di⁄-
replication for testing an experimental treatment is culties, which are not unique to studies of compensa-
the tank, not the individual ¢sh in the tank, even if tory growth, the body of experimental evidence
individuals are identi¢able. The misuse of individual suggests that compensatory growth in ¢sh is a real
¢sh as the unit of replication leads to an in£ation of phenomenon.

# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190 155


156

Compensatory growth in ¢shes M Ali et al.


Table 2 Survey of experimental designs and analyses in a selection of studies on compensatory growth in ¢sh.

Species studied Experimental design Statistical analysis Reference

Oncorhynchus nerka No replication of treatments. Group-housed, but Invalid use of serial t-tests to compare treatments Bilton and Robins (1973)
individuals within tanks treated as valid replicates.
Densities changed during treatment.
O. mykiss No replication of treatments. Group-housed, but ANOVA and multiple range tests. No adjustment Weatherley and Gill (1981)
individuals within tanks treated as valid replicates. for size differences.
Densities changed during experiment.
O. mykiss Inadequate replication with unbalanced design. Invalid use of t-test. No adjustment for size effects. Dobson and Holmes (1984)
Group-housed, but individuals within tanks
treated as valid replicates.
Salvelinus alpinus Individually housed fish as valid replicates. Invalid use of serial t-tests. No adjustment for Miglavs and Jobling(1989a,b)
size differences
O. mykiss No replication of treatments. Group-housed, but Invalid use of Kruskal–Wallis and t-tests. Quinton and Blake (1990)
individuals within tanks treated as valid replicates. No adjustment for size effects.
Densities changed during experiment.
Salmo salar No replication of treatments. Group-housed, but Invalid use of serial t-tests. Metcalfe and Thorpe (1992)
# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190

individuals within tanks treated as valid


replicates. Indirect estimate of rate of
food consumption
Phoxinus phoxinus Individually housed fish as valid replicates. Size adjustment by covariance, preplanned Russell and Wootton (1992)
comparisons, but invalid use of serial rather
than repeated measures ANOVA.
Chalcalburnis chalcoides, Leuciscus No replication of treatments. Group-housed, Invalid use of ANOVA Wieser et al. (1992)
cephlaus, Scardinius erythrothalamus but individuals within tanks treated as valid replicates.
S. salar No replication of treatments. Reimers et al. (1993)
S. salar and Salvelinus alpinus No replication of treatments. Group-housed, but Invalid use of ANOVA with multiple comparisons. Mortensen and Damsgard (1993)
individuals within tanks treated as valid replicates
S. alpinus Group-housed, with three tanks per treatment, tank Valid use of Kruskal–Wallis, but invalid use of Jobling et al. (1993)
means used as valid replicates. Densities serial Mann–Whitney U-tests.
changed during experiment.
Ictalurus punctatus No replication of treatments. Group-housed, Invalid use of ANOVA and multiple range tests Kim and Lovell (1995)
but individuals within tanks treated as valid replicates.
Pleuronectes platessa No replication of treatments. Group-housed, but Invalid use of serial t-tests Paul et al. (1995)
individuals within tanks treated as valid replicates.
Compensatory growth in ¢shes M Ali et al.

Field studies
Nicieza and Metcalfe (1997)

Saether and Jobling (1999)


There are at least three possible approaches to the
study of compensatory growth under natural condi-

Metcalfe et al. (2002)


Boujard et al. (2000)
tions. Observational studies compare the responses

Zhu et al. (2001)

Zhu et al. (2003)


of animals of equivalent reproductive and develop-
mental status, whose growth history was a¡ected
di¡erentially by some identi¢ed or unknown envir-
onmental factor. Experimental ¢eld studies address
compensatory responses that result from the manip-
ulation of the environment either to induce growth

apply Bonferroni adjustment for multiple comparisons.


depression in some individuals or to relax an existing
Failure to apply Bonferroni adjustment for multiple
for size, preplanned, but nonorthogonal contrasts.

size. Pre-planned orthogonal contrasts. Failure to


factor constraining growth rates such as a high risk
Repeated measures with ANCOVA adjustment for
Repeated measures, with ANCOVA adjustment
Repeated measures ANOVA, but contrasts not

of predation. A third approach is to analyse growth


rates under natural conditions after exposure of the
preplanned. Failure to apply Bonferroni

ANOVA and size-adjustment by ancova

experimental animals to controlled conditions in the


adjustment for multiple comparisons
Repeated measures, with ANCOVA

laboratory.
Invalid use of ANOVA and t-tests.

The demonstration of compensatory growth in the


wild requires estimates of the size of individuals or
groups of individuals at di¡erent times. Recapture
adjustment for size.

rates of individually marked ¢sh must be su⁄ciently


high to generate adequate sample sizes, which
comparisons

requires tagging large numbers of ¢sh. The ¢eld site


used must not impose conditions such as low tem-
peratures or food limitation that preclude the detec-
tion of compensatory responses. An additional
di⁄culty in manipulative ¢eld studies is to obtain
refeeding phase, with tank means as valid replicates

individuals of equivalent developmental status, but


No replication of treatments. Group-housed, with all
treatments combined during re-alimentation phase,

Group-housed, with two tanks per treatment, tank

Group-housed with 3 tanks per treatment in used

individuals within tanks treated as valid replicates

di¡ering in their recent growth history. Indirect


No replication of treatments. Group-housed, but

methods, including the back-calculation of size at


Individually housed fish as valid replicates.
but individuals treated as valid replicates

known ages, can avoid the costs and di⁄culties asso-


Individually housed as valid replicates

ciated with tracking marked individuals (Nicieza


and Bran‹a 1993a,b), but have limitations. They are
means used as valid replicates

only applicable to some species, do not allow the pre-


cise de¢nition of the time intervals, and provide no
information about the environmental factors in£u-
encing growth dynamics. These di⁄culties asso-
ciated with ¢eld studies raise doubts about the
relevance of compensatory responses in natural
populations.
Most of the empirical evidence for compensatory
growth comes from manipulative experiments, in
Studies are presented in historical order.

which a reduction of food intake is imposed either


directly or indirectly. To assess the role that compen-
satory growth can play as an important process of
growth regulation in natural populations, a ¢rst step
Gasterosteus aculeatus,
Scophthalmus maximus

Gasterosteus aculeatus

is to de¢ne the constraints in the habitat and ecology


of the population that might evoke a compensatory
response. The spatio-temporal variation in the envir-
P. phoxinus
Salmo salar

onment must be su⁄cient for a variable fraction of


O. mykiss

S. salar

the population to experience a potentially reversible


reduction in growth rates. If the entire population

# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190 157


Compensatory growth in ¢shes M Ali et al.

experiences an unpredictable reduction in growth


rates, the demonstration of compensatory growth
will be di⁄cult because of the problem of identifying
an undisturbed growth trajectory. In this situation,
an alternative method of detecting a process of com-
pensation would be to use an intercohort analysis of
growth patterns over a time series for which data on
relevant environmental variables are available.
Another factor that indirectly induces growth
depression is density. In natural conditions, the com-
pensatory adjustment would require the existence of
some behavioural mechanism (e.g. condition-depen-
dent dispersal).

Factors affecting compensatory


growth in fish
Most studies in ¢sh have provoked compensatory
growth bya period of total or partial food deprivation.
The consequences of changing food qualityon growth
are also susceptible to compensation (Bjo«rnsson et al.
1992). It has also been induced after a period of
unseasonably low temperatures, which reduced the
rate of food consumption in juvenile Atlantic salmon,
Salmo salar (Salmonidae; Mortensen and DamsgQrd
1993; Nicieza and Metcalfe 1997; Maclean and
Metcalfe 2001) or Atlantic cod, Gadus morhua (Gadi-
dae; Purchase and Brown 2001) and during treat-
ment of diseased rainbow trout, Oncorhynchus
mykiss (Salmonidae) with hydrogen peroxide (Speare
and Arsenault 1997). Atlantic salmon smolts showed
growth compensation after a short period of growth
depression associated with a move from fresh to sea
water (DamsgQrd and Arnesen 1998). Spotted wolf-
¢sh, Anarhichas minor (Anarhichadidae), displayed
compensatory growth after being returned to nor-
mal oxygen conditions following 75 days in hypoxic Figure 2 Full growth compensation demonstrated by
conditions (Foss and Imsland 2002). In other studies, individually housed, juvenile gibel carp, Carassius auratus
the origin of the reduced growth and the subsequent gibelio, after periods of total food deprivation. , controls
compensation has not been clear (Zivkov 1982, 1996; (fed to satiation daily); &, deprivation for1 week; ~,
deprivation for 2 weeks prior to re-alimentation. Means for
Zivkov and Raikova-Petrova 1991; Nicieza and Bran‹a
9^10 ¢sh. (a) Growth in mass; (b) mean daily speci¢c growth
1993a,b; Nicieza et al.1994a).
rate, Gw (% day1).

E¡ect of feeding protocols including length


single period of deprivation, or periods of deprivation
and intensity of deprivation
and refeeding have alternated in cycles. Not surpris-
Experimental studies on compensatory growth ingly, this diversity of protocols and in some studies
using food deprivation to induce growth depression poor experimental design and analysis have led to a
can be classi¢ed along several dimensions. The diversity of reported results.
experiments have used either individually or group- Clear evidence of the pattern of compensatory
housed ¢sh as the unit of replication. Deprivation growth, with either full or partial compensation
has either been total or partial. There has been a (Fig. 2), has emerged from experiments with

158 # 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190


Compensatory growth in ¢shes M Ali et al.

individually housed ¢sh exposed to a single period of group-housed Atlantic salmon juveniles in two ways.
deprivation. Full compensation following 1 or One group had a reduced food availability (a ration
2 weeks of total deprivation was achieved by two that was expected to give about 10% of maximum
size-classes of juvenile European minnow, Phoxinus growth). The second group experienced a period of
phoxinus (Cyprinidae; Russell and Wootton 1992; unseasonably low temperatures, which reduced the
Zhu et al. 2001), juvenile gibel carp, Carassius auratus rate of food consumption of the fry. At the re-imposi-
gibelio (Cyprinidae; Xie et al. 2001), and was achieved tion of ad libtum rations and restoration of control
or approached by the three-spined stickleback, Gas- temperatures, the low-temperature ¢sh were smaller
terosteus aculeatus (Gasterosteidae; Zhu et al. 2001). than the food-restricted juveniles. The subsequent
In these experiments, the deprived ¢sh reached the pattern of compensatory growth di¡ered between
mass of control ¢sh after 2^4 weeks of refeeding. the groups. The restricted food group showed full
When two groups of three-spined sticklebacks were compensation within 31 weeks. The low-tempera-
reduced by di¡erent deprivation protocols to a mean ture group commenced a phase of compensatory
mass of 80% of control ¢sh fed ad libitum, they growth a week later than the food-restricted group
showed full compensation in mass and lipid concen- and showed only partial compensationafter31 weeks.
tration (Zhu et al. 2003). The trajectories of recovery It is not clear whether the di¡erences were an e¡ect
were similar for both groups. In the minnows, 4 days of the di¡erent protocols used to achieve a phase of
of deprivation were not su⁄cient to evoke a detect- slow growth or of the intensity of the growth retarda-
able compensatory response (Russell and Wootton tion. There is growing evidence that periods of food
1992), and after a week of deprivation, the response shortage or low temperatures can provoke di¡erent
was weak but still su⁄cient to restore the mass tra- patterns of compensation. In juvenile Atlantic sal-
jectory (Zhu et al. 2001). mon, a 3-week period of unseasonably low tempera-
In the pioneering study using individually housed tures in early summer induced full compensation,
¢sh, Miglavs and Jobling (1989a,b) imposed 8 weeks but the compensatory growth was not displayed until
of restricted feeding at a level close to maintenance the following autumn (Maclean and Metcalfe 2001).
ration on 1-year-old Arctic charr, Salvelinus alpinus Taxonomically close species may have di¡erent
(Salmonidae). On re-alimentation, the deprived charr propensities to show compensatory responses. In
showed a compensatory response, but failed to reach contrast to Atlantic salmon, two studies failed to
the mass of control ¢sh before the compensatory demonstrate compensatory growth in brown trout,
response abated. Minnows fed a maintenance ration Salmo trutta (Salmonidae). One-year-old brown trout
for 16 days did achieve full compensation on being that had been partially deprived of food between June
returned to satiation feeding, but the compensatory and November failed to show compensatory growth
response declined earlier than that shown by min- when fed to satiation from November to March (Pir-
nows experiencing total deprivation for 16 days honen and Forsman 1998). In this case, the low tem-
(Russell and Wootton1992). peratures in the winter period and the associated
The e¡ects of a single period of deprivation on low appetite may have restricted the scope for detect-
group-housed ¢sh have been more variable and not able compensation. Age 0þ juvenile brown trout
always consistent within taxa. Sockeye salmon fry, maintained at reduced rations for 1 month did not
Oncorhynchus nerka (Salmonidae), showed compen- show growth compensation after release in their
satory growth after 1^3 weeks of starvation and natal stream in two consecutive years (A¤lvarez 2002).
reached similar masses to controls within 8 weeks Growth rates were apparently not constrained after
of ad libitum refeeding. After 4 weeks of starvation, release. Further studies on the capacity of brown
mortality rates were high and survivors were not able trout to show compensatory growth are currently in
to catch up with the mass of control ¢sh within the progress (Hayward, personal communication).
8 weeks of recovery period (Bilton and Robins 1973). In cyprinids, Wieser et al. (1992) noted an inverse
Rainbow trout, with an initial mass of about10 g, still relationship between the length of starvation period
showed compensatory growth following 13 weeks of and the subsequent compensatory growth of group-
starvation (Weatherley and Gill 1981). Dobson and housed, early juveniles of three species of cyprinid
Holmes (1984) also reported compensatory growth (Scardinius erythrophthalmus, Leuciscus cephalus and
in rainbow trout subjected to total deprivation for Chalcalburnus chalcoides mento). Juvenile roach, Ruti-
3 weeks followed by 3 weeks of feeding. Nicieza and lus rutilus (Cyprinidae), housed at temperatures that
Metcalfe (1997) evoked compensatory growth in ranged from 4 to 33 8C and starved for 3 weeks,

# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190 159


Compensatory growth in ¢shes M Ali et al.

showed compensatory growth in the following normoxic conditions after 75 days under hypoxic
3 weeks of ad libitum feeding (van Dijk et al. 2002). conditions (Foss and Imsland 2002). The study was
Even at 4 and 12 8C, the roach showed a compensa- terminated after 21 days of the normoxic conditions,
tory growth response, although before the period of and it was unclear whether the compensatory res-
starvation, they had not grown at these low tempera- ponse would have persisted for longer.
tures. One example in which no growth compensa- Studies in which individually housed ¢sh have
tion was observed is a study on carp, Cyprinus carpio been exposed to cycles of deprivation and refeeding
(Cyprinidae). Schwarz et al. (1985) studied the e¡ects have provided some of the most compelling evidence
of protein or energy restriction on growth perfor- for compensatory growth in teleosts. Most studies of
mance of carp. Both restricted groups grew more compensatory growth have used ¢xed periods of
slowly than the control, but both showed positive deprivation and refeeding. However, a study of 0þ
growth rates. The duration of the restriction period Lepomis cyanellus  L. macrochirus (Centrarchidae)
was determined by the time taken to reach a target hybrids used cycles of deprivation and refeeding
mass set in the experiment so that treatment groups (Hayward et al.1997) in which the deprivation periods
and controls started the re-alimentation period at were ¢xed at 2, 4, 6, 10 or 14 days, but the length of
approximately the same mass. At the end of re-ali- the post deprivation period was determined by the
mentation period, there were no di¡erences in car- duration of the hyperphagic response that followed
cass composition between protein restricted and each period of deprivation.With this protocol, ¢sh in
control groups, but the energy-restricted group was some of the groups experiencing cyclical phases of
low in dry matter, fat and energy content. During re- feeding and deprivation grew more over a 105-day
alimentation, previously restricted groups had the period than control Lepomis fed ad libitum daily.
same growth rate as the control group, with no evi- This over-compensation was particularly high in the
dence of a compensatory growth response. The lack 2-day deprivation group (Fig. 3). This is the ¢rst
of a compensatory growth response may have demonstration of such a strong over-compensatory
re£ected inadequate rates of feeding in the re-alimen- response. Similar protocols applied to juvenile yellow
tation period, or the e¡ects of the deprivations may perch, Perca £avescens (Percidae), failed to evoke
not have been su⁄cient to trigger a compensatory over-compensation (Hayward and Wang 2001), al-
response. though full compensation was achieved by males
Marine species also have the capacity for compen-
satory growth. Atlantic cod showed full compensa-
tion when fed to satiation for 10 weeks after 8 weeks
of deprivation (Jobling et al. 1994). Cod that had been
lightest at a given length showed the highest growth
rates during the refeeding period. In the Pleuronecti-
formes, both full and partial compensation have been
evoked. During early spring, four groups of subadult
Alaska yellow¢n sole, Pleuronectes asper (Pleuronec-
tidae), were fasted for either 0, 2, 4 or 6 weeks at the
beginning of a 12-week experiment and then fed to
satiation. The groups of yellow¢n sole fed continu-
ously or fasted for 2 weeks gained the maximum
mass, 25 and 24%, respectively. Fishes fasted for
either 4 or 6 weeks exhibited signi¢cantly lower
mass gains of 16 and 15%, respectively, over the
12-week period, suggesting that P. asper had a limited
capacity for compensatory growth (Paul et al. 1995).
When 2-year-old turbot, Scophthalmus maximus
Figure 3 Growth over-compensation demonstrated by
(Pleuronectidae), were partially deprived but not individually housed Lepomis hybrids. , controls (fed to
starved of food for 41 days, they showed full compen- satiation daily); &, ¢sh experiencing cycles of 2 days of
sation after being placed on a high ration for a further deprivation followed by a phase of hyperphagia of varying
34 days (Saether and Jobling 1999). Spotted wol⁄sh length. Means for nine ¢sh (redrawn from Hayward et al.
showed compensatory growth when returned to 1997).

160 # 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190


Compensatory growth in ¢shes M Ali et al.

imposed on individually housed minnows, three-


spined sticklebacks, gibel carp and Chinese long-
snout cat¢sh Leiocassis longirostris (Bagridae; Xie
et al., unpublished data; Wu et al. 2003). At the end of
the experiment, the sticklebacks had achieved full
compensation (Fig. 4a), but the gibel carp, the cat¢sh
and the minnows had shown only partial compensa-
tion (Fig. 4b). Such experimental protocols can evalu-
ate the capacity of ¢sh to bu¡er the e¡ects of
alternating periods of high and low food availability.
When three-spined sticklebacks were exposed to 1, 2
or 6 days of deprivation followed by 2 days of satia-
tion feeding for 56 days, the ¢sh showed a partial
compensatory response that increased in expression
over the length of the experiment. The sticklebacks
on the 2-day deprivation regime almost matched the
growth of the control, although fed at half the fre-
quency (Ali and Wootton 2001).
Partial compensatory growth has been reported
for group-housed ¢sh subjected to cycles of depriva-
tion and refeeding. In Arctic charr deprived for either
1, 1.5 or 3 weeks and then fed for the same length of
time in an experiment lasting for 24 weeks, the
deprived ¢sh showed partial compensatory growth.
After 6 weeks of unrestricted feeding at the end of
the experiment, there was no signi¢cant di¡erence
in the mean masses of the deprived ¢sh, irrespective
of the previous feeding pattern, although they were
smaller than the control ¢sh fed daily throughout
the experiment (Jobling et al. 1993). In contrast,
groups of cod experiencing cycles of 3-week feeding
and 3-week deprivation did not show a compensa-
tory growth response (Jobling et al.1994). Frequently,
symmetrical cycles of deprivation and refeeding do
produce partial compensation (e.g. Aranyakananda
Figure 4 E¡ect of four cycles of1-week deprivation and et al.1996; Christensen and McLean1998). The incon-
2 weeks of satiation refeeding on growth compensation. , sistency of results when using grouped ¢sh is illu-
controls; &, deprived. (a) Full compensation illustrated by strated by laboratory studies on the channel cat¢sh.
the three-spined stickleback, Gasterosteus aculeatus. Means When groups were exposed to three cycles, each con-
for six ¢sh. (b) Partial compensation illustrated bygibel carp, sisting of 3 days of deprivation and 11 days of satia-
Carassius auratus gibelo. Means for15 ¢sh. tion feeding, the deprived ¢sh showed full
compensation by the end of the 6 weeks (Gaylord
experiencing12 days of deprivation per cycle. Growth and Gatlin 2001). But in a comparable experiment in
over-compensation could not be evoked in theLepomis which cat¢sh experienced 3, 5 or 7 days of depriva-
hybrids when they were housed in groups of 10 ¢sh tion in each 14-day cycle, even the ¢sh experiencing
percontainer (Haywardetal.2000).Incontrast,groups the lowest level of deprivation failed to show compen-
of juvenile channel cat¢sh did showgrowth over-com- sation (Gaylord et al. 2001).
pensation when subjected to cycles of deprivation for The frequency of food provision (or inversely, the
1, 2, or 3 days and then to refeeding in the hyper- frequency and duration of food shortage) can have a
phagic phase (Chatakondi andYant 2001). decisive in£uence on ¢sh growth. If the periods of
In a comparative study, four cycles of1 week of total food deprivation are short and su⁄cient food is avail-
deprivation and 2 weeks of satiation feeding were able between the starvation episodes, a hyperphagic

# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190 161


Compensatory growth in ¢shes M Ali et al.

response when food is present can prevent measur- 1996). The pattern of compensation can vary
able growth depression; so, the growth patterns of between individuals depending on their behavioural
continuously fed and temporarily deprived ¢sh are status and body condition (Jobling and Koskela
virtually indistinguishable. In this case, behavioural 1996). In groups of rainbow trout on restricted
compensation occurs before feeding restriction can rations, there was a high interindividual variability
result in detectable growth depression. This has been in food intake. A dominance hierarchy probably gen-
demonstrated in immature three-spined sticklebacks erated this variability by regulating access to food.
for deprivation periods with a mean length of 3 days The consequence was highly heterogeneous growth
(Ali and Wootton 1998; Ali et al. 1998). No di¡erences amongst the restricted ¢sh.When the trout were pro-
in growth, food conversion e⁄ciency or body compo- vided with unrestricted rations, they became hyper-
sition were detected between rainbow trout held in phagic. Growth compensation was highest in trout
groups fed daily and three times per week. Trout that had shown the poorest growth under restricted
experiencing two lower rations, being fed twice or rations. This suggests that with unrestricted rations,
once per week, were signi¢cantly smaller than those the dominance hierarchy relaxed and the individuals
fed daily (Teskeredzic et al. 1995). Compensatory whose growth had been most suppressed could dis-
growth may not be evoked if the food restriction play high rates of consumption (Jobling and Koskela
exceeds a certain severity (Wilson and Osbourn 1996).
1960; Ryan 1990). Small, juvenile sockeye salmon
starved for 1^3 weeks caught up with control ¢sh
Ontogenetic changes
over 7 weeks of refeeding, but ¢sh starved for longer
did not (Bilton and Robins1973). Comparisons between studies are di⁄cult because of
Some aspects of the factors controlling the magni- di¡erences in experimental protocols, species and
tude of the compensatory response remain obscure. ontogenetic stages. Overall, the studies suggest that
There is a limited understanding of how the compen- there are levels of deprivation that are too small to
satory process depends on the factor evoking growth evoke a compensatory response, levels that will pro-
depression. Growth reductions associated with tem- voke a full compensatory response and levels so
perature change can elicit compensatory growth severe that full recovery cannot be achieved. These
(Mortensen and DamsgQrd 1993; Chmilevskii 1994; levels probably change with ontogentic stage and
Nicieza and Metcalfe 1997; Maclean and Metcalfe state of sexual maturation. There have been few stu-
2001; Purchase and Brown 2001). However, some dies that have examined ontogenetic changes in the
studies have reported important di¡erences between capacity for compensatory growth within a species.
the growth patterns observed after a period of low Because of their small size, ¢rst-feeding teleosts
temperatures and a period of food shortage. The time are usually at a higher risk of starvation than when
lag between cessation of the cause of growth depres- theyare older and bigger (Miller et al.1988).The distri-
sion and the start of the compensatory response bution of planktonic prey of larval ¢sh is often patchy
seems to be greater for temperature than for food- in space and time (Mackas et al. 1985; Owen 1989), so
induced growth arrest (Nicieza and Metcalfe 1997; young ¢sh may encounter alternating periods of
Maclean and Metcalfe 2001). Di¡erent forms of abundant and scarce food (Letcher et al. 1996). If the
growth reduction can a¡ect other processes (e.g. period of food deprivation is su⁄ciently long, 4^
reproductive investment, storage, hormonal control) 25 days, depending on species and ¢sh size (Miller
di¡erentially, which suggests that there may be di¡er- et al. 1988), young ¢sh die as a result of starvation.
ences in both the temporal pattern of compensation Letcher et al. (1996) investigated the e¡ect of the order
and the ¢nal outcome in terms of body size. of feeding days that preceded starvation on times to
death from starvation and whether there is a size
dependence in time to 50% mortalityand to mass loss
In£uence of social factors
before starvation occurs. The time to starvation in
Density and social interactions also have conse- young yellow perch was insensitive to short-term
quences for compensatory growth. In juvenile tilapia, (4 days) £uctuations in food density. The propor-
Oreochromis niloticus (Cichlidae), relative growth tional mass loss up to starvation was independent of
losses caused by crowding conditions were rapidly previous maximum ¢sh mass, although ¢sh
eliminated when the ¢sh were returned to lower, con- exposed to intermittent food availability starved at
trol densities (Vera-Cruz and Mair 1994; Basiao et al. relatively heavier masses than ¢sh in constant prey

162 # 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190


Compensatory growth in ¢shes M Ali et al.

environments. Letcher et al. (1996) argued that the equal or superior performers in other phases. In the
existence of growth compensation suggested a winter £ounder Pleuronectes americanus (Pleuronec-
strong selection pressure for strategies to overcome tidae) slower growing larvae grew faster as juveniles
short periods of starvation. (Chambers et al. 1988; Bertram et al. 1993). These
Compensatory growth could be relatively common negative correlations between larval and juvenile
during larval development because environmental growth rates are noteworthy because rearing condi-
uncertainty can cause a delay in the time of ¢rst feed- tions that a¡ect growth (temperature, density and
ing that is susceptible to further compensation. Dab- food) had no in£uence on this response. The compen-
rowski et al. (1986) starved groups of coregonid larva satory growth reported in these studies was not
from the time of hatching, thus delaying the time of induced directly by changes in physical conditions.
¢rst feeding by 3^18 days. Delayed ¢rst feeding for Bertram et al. (1993) speculated that the convergence
13 days resulted in 50% mortality.With more moder- of growth trajectories was related to size-dependent
ate starvation (3^11 days), there was a positive corre- mortality and developmental processes. Fish with
lation between the length of delay (starvation) and longer developmental periods might have more e⁄-
subsequent growth rate during ad libitum refeeding. cient organs, which would result in more rapid
Compensatory growth in early feeding stages growth.
could o¡set di¡erences in size at hatching that re£ect Ontogenetic niche shifts provide an obvious poten-
di¡erences in egg size (e.g. Atlantic salmon; Thorpe tial for compensatory growth, as these are often
et al. 1984). Catch-up growth was observed in Clyde associated with an improvement in feeding condi-
herring larvae, Clupea harengus (Clupeidae), exposed tions. In brown trout, a shift from an invertebrate to
to a low ration (15 copepods larva1 day1) for a piscivorous diet can result in a re-acceleration of
10 days after yolk absorption followed by a high growth (Toledo 1996). Juvenile blue¢sh, Pomatomus
ration (80 copepods larva1 day1) for 3 weeks. After saltatrix (Pomatomidae), that delay the shift from
31 days of feeding, these larvae caught up with high- invertebrate feeding to piscivory grow initially at a
ration larvae in standard length and content of lower rate than piscivorous blue¢sh.Yet, they appar-
water-soluble protein, without an increase in mortal- ently experience a phase of increased consumption
ity. However, trypsin and trypsinogen content halved and growth rates (leading to partial compensation
in the re-alimented larvae compared to that of high- of their comparatively reduced growth) after the
ration ¢sh (Pedersen et al. 1990). In contrast, Baltic change to a ¢sh diet (Buckel et al. 1998).The seaward
herring exposed to two sequences of food restriction migration of anadromous ¢shes is often assumed to
for 5 days and re-alimentation for 10 days failed to be an escape from a highly food-limited breeding
show compensatory growth and experienced higher habitat to a more favourable growing habitat (Gross
mortality than controls (Pedersen 1993). Larger eggs 1987; Gross et al. 1988; McDowall 2001). Those ¢sh
in herring confer a higher survival potential on the that experienced the poorest growth in freshwater
resulting larvae (Blaxter and Hempel 1963). The abil- could exhibit fast growth at the sea. Although obser-
ity to exhibit catch-up growth found in the Clyde lar- vations of negative correlations between smolt length
vae (Pedersen et al.1990) may be linked to the higher and length increment of Atlantic salmon during the
resources in the form of organic matter in these large ¢rst growing season at sea support this idea (Skilbrei
larvae compared to larvae from Baltic stock (Peder- 1990; Nicieza 1993; Nicieza and Bran‹a 1993a), asses-
sen 1993). The protocols for restriction and refeeding sing the incidence of such interstage compensation
were di¡erent, so the lack of agreement may re£ect requires further research including manipulative
methodology. Some species do seem to lack compen- ¢eld experiments in di¡erent taxonomic groups.
satory abilities at early stages of development. For
instance, delayed ¢rst feeding, low temperature and
Seasonal variation
darkness reduced the intake of food and slowed down
somatic growth in Dover sole, Solea solea (Soleidae), An experiment on grouped juvenile Atlantic salmon
but there was no evidence of subsequent compensa- suggested that the nature of the compensatory
tory growth (Lagardere1989). response may change seasonally (Metcalfe et al.
Major ontogenetic changes in physiology, mor- 2002). The ¢sh were deprived of food in summer and
phology and behaviour can facilitate self-regulation autumn for periods that were su⁄cient to reduce
of growth patterns. Individuals that are poor perfor- the fat reserves of the ¢sh to similar levels. Because
mers during one phase of their life cycle can become of lower temperatures in autumn, the deprivation

# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190 163


Compensatory growth in ¢shes M Ali et al.

period required was longer. On refeeding, the sum- ment (Jobling et al. 1993). Irrespective of feeding
mer group showed compensatory growth both in regime, maturing males were signi¢cantly smaller
length and in fat when compared with control ¢sh. than immatures when the experiment terminated.
In contrast, the autumn ¢sh recovered their fat con- The existing information indicates that, among
tent, but their growth in length did not di¡er from marine species, the e¡ects of reduced food availabil-
control ¢sh. ity on growth and reproduction are similar to those
reported for freshwater or anadromous species.
One-year-old farmed Atlantic cod were subjected to
Sexual maturation and reproduction
periods of short-term starvation that lasted 3, 6, or
Environmental conditions and nutritional status 9 weeks with alternating periods of 3-week starva-
may have strong modifying e¡ects on maturation tion and 1-week feeding showed partial compensa-
and fecundity in ¢sh (Wootton 1982, 1998). Conver- tion (Karlsen et al. 1995. In the most restrictive
sely, the allocation of energy to gonad production treatment (9-week starvation over a period of
and increased activity associated with reproduction 11 weeks), body mass at maturity was only about
(e.g. migration, defense, access to mates) results in 60% of control ¢sh. Although controls had signi¢-
growth arrest. However, there have been no systema- cantly higher fecundities than ¢sh starved for
tic experimental studies on compensatory growth in 9 weeks, these di¡erences were related to body size
relation to reproduction, and no coherent picture and relative fecundities were similar for females in
has emerged. both groups. However, reduced growth and liver sizes
In some species, like Oreochromis mossambicus in feed-restricted groups did not result in lower pro-
(Cichlidae), low temperatures reduce growth rates portions of maturing ¢sh. In a similar experiment,
and delay gonad development. Transfer to optimal Jobling et al. (1994) observed full compensation in
temperatures induces growth compensation, but male and female cod after 12 weeks of re-alimenta-
maturation is still delayed in comparison to ¢sh that tion and reported an increased investment in repro-
did not experience the lowered temperatures (Chmi- duction in male cod previously subjected to a
levskii1994). In salmonids, the physiological decision low-energy diet.
to mature may be dependent on some measures of Apart from an obvious trade-o¡ between somatic
the rate of storage or turnover of surplus energy growth and reproduction, the interpretation of these
exceeding a threshold value during a limited season results is complex. Compensatory growth may use
(Thorpe 1986). The value for this threshold is not energy and nutrients that otherwise would be avail-
known, nor is the precise season or its duration, able for reproduction, thus causing a delay in age at
although there are some clues to both. For example, maturity. If food restriction does result in a subse-
lipid levels in winter and spring may in£uence the quent lower rate of maturation despite a compensa-
initiation and progression of maturation of salmon tory restoration of lipid levels or body mass, this may
parr (Rowe et al.1991). A series of studies investigated re£ect an e¡ect of priority in allocation to growth
the e¡ect of food restriction on the precocial matura- and storage. However, if there are body size or condi-
tion of male Atlantic salmon (Rowe and Thorpe tion thresholds for maturation (Thorpe 1986), com-
1990; Thorpe et al. 1990; Rowe et al. 1991). These pensatory growth would favour an earlier
experiments found evidence of growth compensa- attainment of the required threshold and so a lower
tion. However, some schedules of deprivation and age at maturity (Fig. 5). Experiments that compare
refeeding did reduce the proportion of maturing maturation rates of ¢sh that have di¡erent reactions
males. E¡ects on the proportion of maturing females to a period of growth depression would be useful in
were more pronounced. Similarly, Reimers et al. de¢ning the relationships between growth compen-
(1993) starved second sea winter farmed Atlantic sal- sation and reproduction. A potential problem for
mon through February and March. Starved ¢sh such experiments is having adequate controls
showed partial compensation in growth, but the inci- because of the di⁄culty in obtaining compensating
dence of maturation was reduced by 48% among and noncompensating individuals from a group of
females and 32% among males. In Arctic charr, animals that have shared the same growth history.
cycles of alternating periods of deprivation and feed- There are two ways by which ¢sh can mitigate the
ing over a 24-week period followed by 6 weeks of e¡ects of growth depression on reproductive output.
feeding produced no signi¢cant di¡erences in the Apparently, the simplest form is through growth
proportion of mature males at the end of the experi- compensation. For example, female three-spined

164 # 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190


Compensatory growth in ¢shes M Ali et al.

sticklebacks experiencing regular or irregular inter-


vals between feeding did not di¡er in mean date of
¢rst spawning and fecundity from females fed daily
(Ali andWootton1999); in this case, females compen-
sated for days without food by showing hyperphagia
on days they were fed. The second scenario is more
puzzling and refers to a direct compensation of the
potential reduction in reproductive output asso-
ciated with a smaller body size. An example is the
study by Siems and Sikes (1998) on the e¡ects of pre-
dictable and unpredictable food availability on male
and female fathead minnows, Pimephales promelas
(Cyprinidae). Males that were switched from an
unpredictable to a predictable food supply at an age
of 75 days showed compensatory growth after the
switch. Females on the unpredictable food supply
were smaller than females on a predictable supply
(i.e. they had not exhibited full compensation), but
even so the former started spawning at the same age
and had the same mean production of eggs, which
indicates a direct compensation of the reproductive
output by increasing reproductive investment per
unit of somatic mass.

Compensatory growth of body components

Figure 5 Potential problems in analysing the The pattern of growth compensation may vary
consequences of compensatory growth on reproductive between body components. In rainbow trout, refeed-
parameters. For clarity, we assumed linear growth and time- ing following a period of restriction was associated
constrained reproduction, and used the rate of maturation with rapid restoration of the relative sizes of organs
as response variable. T1 andT2 de¢ne the start and (Weatherley and Gill1981,1987). Both short- (3 weeks)
termination of a period of growth depression, andT3 and long-term (13 weeks) starvation entirely utilised
indicates the end of the phase of compensation. R and R0 visceral fat. Long-term starvation also reduced gut
represent alternative time-windows for maturation. (a) dry mass signi¢cantly. Despite the in£uence of star-
Compensating (broken line) and noncompensating
vation on the proportions of body components, dur-
(thick line) individuals have experienced di¡erent
ing re-alimentation, rainbow trout attained normal
conditions which could have direct or indirect (e.g. if
maturation rate is size dependent) e¡ects on maturation
proportions or even over-compensation of heart,
rate. The e¡ect of the compensatory process on the rate of liver, gonad, gut and visceral fat. Skin and carcase
maturation can only be evaluated if decision to mature dry masses were lower in controls and rainbow trout
occurs afterT3 (case R0); (b) after a period of growth previously exposed to short-term starvation than in
depression, some individuals show compensatory growth those previously subjected to long-term starvation.
(broken line) whereas others do not compensate In O. niloticus,Takagi (2001) reported compensatory
(thick line) and the control growth rate (dotted line). growth in acellular bone when ¢sh were refed for
Comparison of compensating and noncompensating 15 days after 15 days of starvation, but did not detect
individuals to detect direct in£uence of compensatory rates compensatory growth in total body mass.
on the rate of maturation is justi¢ed at both R and
R0. If the decision to start maturation takes place at any time
betweenT2 andT3, allocation and size (only if size Proximate causes of compensatory growth
dependence is assumed) e¡ects can be evaluated. AfterT3,
di¡erences can be attributed to size e¡ects or Several factors could contribute to the compensatory
impairment of the reproductive/hormonal function growth observed during re-alimentation after a per-
associated with a previous episode of very fast iod of relative or total food deprivation, or other
growth. causes of growth depression. These include a rate of

# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190 165


Compensatory growth in ¢shes M Ali et al.

food consumption greater than that of continuously


fed controls (hyperphagia), enhanced growth e⁄-
ciency, reduced metabolic costs and reduced expen-
diture on locomotion. Most experimental evidence
suggests that growth compensation operates by
adjustments in food intake of growth-depressed ani-
mals. In part, this evidence is biased by the dispropor-
tionate number of studies investigating the e¡ects of
growth depression on feeding rates compared to
other responses that might be involved in the com-
pensatory process. However, while almost all studies
measuring feeding rates have found that hyperpha-
gia makes an important contribution to the accelera-
tion of growth, a substantial proportion of studies
investigating other variables have failed to show a
signi¢cant contribution of these in growth compen-
sation and, when detected, their contribution is
minor compared to the hyperphagic e¡ect. The avail- Figure 6 Hyperphagia shown by gibel carp in
able evidence suggests that hyperphagia is the main re-alimentation phase following1or 2 weeks of total food
mechanism involved in the compensatory response, deprivation (symbols as in Fig. 2). Means for nine ¢sh.
although increased conversion e⁄ciencies or beha-
vioural adjustments might play a role. Russell and Wootton 1992; Hayward et al. 1997; Ali
and Wootton 1998a). Hyperphagic Arctic charr had a
food intake of approximately1.5^2.0% live body mass
Hyperphagia
day1, whereas in the same period, controls con-
Hyperphagia is, by far, the most common mechanism sumed approximately 1.0^1.4% body mass day1
of growth compensation, and its occurrence is not (Miglavs and Jobling 1989a,b). A hyperphagic
limited to the existence of previous episodes of star- response was demonstrated in European minnows
vation or food shortage (e.g. Nicieza and Metcalfe (Russell and Wootton 1992) and in gibel carp (Xie
1997). Hyperphagia is a rate of food consumption sig- et al. 2001; Fig. 6). In Atlantic salmon, hyperphagia
ni¢cantly higher than that shown by ¢sh that have followed periods of food restriction (Bull et al. 1996)
been continuously on an ad libitum ration. Because or unseasonably low temperatures (Nicieza and
the rate of consumption is allometrically size depen- Metcalfe1997). A raised food intake was also reported
dent (Wootton 1998), comparisons have to be made in groups of re-alimented juvenile Piaractus brachy-
for ¢sh of the same size or, alternatively, these must pomus L., a South American characoid (Koppe et al.
be based on size-corrected measurements. Hyper- 1993), and channel cat¢sh (Kim and Lovell1995).
phagia may allow an animal to achieve the same or An exception is a complex experiment on grouped
even a higher cumulative food intake than an indivi- rainbow trout, which failed to detect hyperphagia
dual that has had continuous access to food, and thus during compensatory growth following 21 days of
to achieve the same size. It may result from increased deprivation (Boujard et al. 2000). This study sug-
meal frequency, increased size of meals or a combina- gested that, in rainbow trout, hyperphagia was a
tion of both (Russell and Wootton1993). response to food restriction whereas total depriva-
tion generated an increase in growth e⁄ciency
during refeeding. The experiment included a prede-
Responses to free access to food after a period
privation period of 34 days, during which separate
of growth depression
groups of trout experienced three di¡erent ration
Several studies on ¢sh noted that food consumption levels factored with two di¡erent diets, one high
is elevated following starvation or deprivation (Tyler energy and one low energy.
and Dunn 1976; Jobling 1982; Dobson and Holmes There is evidence that ¢sh can maximise their
1984). Studies on individually housed ¢sh suggested intake rates during recovery after a period of
a hyperphagic response analagous to that observed depressed growth. The main e¡ect of the intensity
in mammals and birds (Miglavs and Jobling 1989a,b; of the growth depression on the subsequent

166 # 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190


Compensatory growth in ¢shes M Ali et al.

hyperphagia is on the duration of the hyperphagic


phase, rather than a higher maximum rate of con-
sumption (Russell and Wootton 1992; Ali and Woot-
ton 2001; Xie et al. 2001; Zhu et al. 2001; Fig. 6). This
suggests that the maximum daily rates observed dur-
ing hyperphagia represent a rate close to the maxi-
mum possible rate at which the gut can process food.
The £exibility of feeding rates seems to be large
enough to allow for a rapid acceleration of growth
rates. For example, three-spined sticklebacks fed at a
mean interval of 3 days showed hyperphagia, con-
suming as much as 21% of their body mass on the
day when food became available (Ali and Wootton
1998a); this ¢gure contrasts with a consumption rate
of 9^11% in sticklebacks fed ad libitum (Wootton et al.
1980). Similarly, the size-adjusted consumption of
hybrid sun¢sh (Lepomis cyanellus  L. macrochirus)
subjected to di¡erent schedules of food restriction
and re-alimentation was more than double that of
controls (Hayward et al. 1997). Although absorption
e⁄ciencies decrease with increasing feeding rates,
the acceleration of growth associated with an
increase in intake rates of 50^90% cannot be
achieved through other mecahnisms of compensa-
tion. The amount of energy that can be saved by redu-
cing nonfeeding activity would account for a tiny
proportion of the total energy that can be allocated
to the process of growth compensation. Individual
ranges of variation in digestive e⁄ciency are rela-
tively narrow (e.g. compared to passage time or feed-
ing rate; Nicieza et al. 1994b), which suggests that
the scope for growth rate acceleration associated
with the maximisation of absorption e⁄ciencies
might be negligible in comparison with the potential
e¡ects of increasing feeding rates.
Although the occurrence of hyperphagic
responses after a period of growth depression is a
generalised phenomenon, observations on the daily
rate of consumption during the period of compensa-
tory growth reveal that there is considerable inter-
speci¢c variation in the temporal pattern of the
hyperphagia (Hayward et al. 1997; Ali et al. 2001;
Hayward andYang 2001;Wu et al. 2002; Fig. 7). These
patterns suggest important interspeci¢c di¡erences
in the regulation of the hyperphagic response, and

Figure 7 Inter-speci¢c di¡erences in patterns of daily food


compensation during period of hyperphagia after1 week of
deprivation; , controls; &, deprived ¢sh. (a) Three-spined
stickleback (means for six ¢sh); (b) gibel carp (means for15
¢sh); (c) Chinese longsnout cat¢sh (means for15 ¢sh).

# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190 167


Compensatory growth in ¢shes M Ali et al.

thereby in the process of growth self-regulation. The ambiguous. It could be an indication that the com-
analysis of interspeci¢c variation in compensatory pensation of structural growth (sensu Broekhuizen
patterns is important because it can improve our et al. 1994) takes priority over allocation to storage,
understanding of the evolutionary signi¢cance of but may be a direct by-product of di¡erences in diet
compensatory growth. At present, the scarcity of composition. In rainbow trout, the energy content of
comparative studies precludes a deeper insight. the diet supplied in both the pre- and postdeprivation
period had no e¡ect on daily energy intake during
the compensatory growth phase (Boujard et al.
E¡ects of food quality on hyperphagic response
2000).
Like other vertebrates, ¢sh regulate, at least partly,
food intake to meet energy needs (Rozin and Mayer
Physiological basis of hyperphagia: consumption,
1961; Lee and Putnam 1973; Page and Andrews 1973;
absorption, and evacuation rates
Cho et al. 1976; Grove et al. 1978; Lovell 1979; Marais
and Kissil1979; Fletcher1984). Lovell (1979) fed chan- Rate of food consumption in ¢sh is related to the rate
nel cat¢sh, Ictalurus punctatus, at three digestible at which the gut is ¢lled and evacuated (Brett 1971;
energy levels (230, 290, and 350 kcal 100 g1) and Elliott1975a,b; Grove et al.1985; Singh and Srivastava
recorded elevated food intakes amongst cat¢sh fed 1985; Russell and Wootton 1993). Under some condi-
the low-energy diet. Rozin and Mayer (1961) and tions of temperature and food availability, intake
Grove et al. (1978) manipulated the energy contents rates will be limited by the maximal rate at which
of feed byaddition of kaolin to the diet of gold¢sh, Car- food can be digested. An elevation of the maximal
assius auratus (Cyprindae), and rainbow trout, rate (i.e. a reduction of the passage time of food
respectively. In both studies, increased intake of low- through the digestive tract) would contribute to a
energy food compensated for nutrient dilution. greater hyperphagic response. Whether the rate of
Furthermore, gold¢sh presented with large then gut evacuation can change directly in response to a
small pellets alternately adjusted the number of pel- period of growth depression is not known. The di⁄-
lets consumed to achieve an energy balance (Rozin culty rests on disentangling the potential direct e¡ect
and Mayer 1961). Self-selection of macronutrients and the decrease in passage time associated with
(protein, fat and carbohydrate) by ¢sh, including increased ingestion (Elliott 1972; Nicieza et al.
gold¢sh and rainbow trout, has been demonstrated 1994b). In the European minnows, there was no sig-
by the use of demand-feeders (Sanchez-Vazquez et al. ni¢cant di¡erence between the rates of foregut eva-
1998, 1999). A study of the e¡ect of periods of food cuation of the ¢rst satiation meal fed after 4 or
deprivation on nutrient self-selection by sea bass, 16 days of starvation (Russell and Wootton 1993).
Dicentrachus labrax, imposed periods of 6 and 15 days There was no evidence that the increase in food con-
of total food deprivation separated by a period of sumption that followed a deprivation period (Russell
8 days of demand feeding (Aranda et al. 2001). After and Wootton 1992) was associated with changes in
both periods of deprivation, the bass showed hyper- the rate of gut evacuation. A greater absorption e⁄-
phagia but the deprivation had only minor e¡ects on ciency during compensatory growth (Russell 1991)
diet selection; the intake of protein was slightly ele- did not result from a di¡erence in the rate of proces-
vated for 2 days when feeding resumed. Therefore, sing of ingested food in the foregut, although there
although the intensity of hyperphagia can be some- was some evidence of a longer retention time in the
what reduced if high-energy food is available, there hindgut. Studies incorporating a short (48 h), pre-
is no evidence of compensation mediated by beha- prandial starvation had no detectable e¡ect on subse-
vioural decisions in£uencing diet selection. quent evacuation rates in Atlantic salmon (Talbot
If the regulation of intake rates maintains steady et al. 1984). In brown trout, a starvation period of up
energy inputs, growth compensation should be inde- to 7 days prior to feeding did not a¡ect evacuation
pendent of energy content of the diet. Atlantic cod, rates, although periods of 10 days or more did reduce
¢sh fed with either high-energy or low-energy food the rate (Elliott 1972). Similarly, a decreased rate of
after a period of starvation showed full growth com- evacuation occurred in Pleuronectes platessa (Pleuro-
pensation by 12^18 weeks (Jobling et al. 1994). Cod nectidae) after 30 days of starvation (Goddard1974).
fed high-energy diets tended to have higher levels of In European minnows, the decline in gut contents
lipids in the liver at the end of the experiment (Jobling after a meal is correlated with return of appetite
et al. 1994). The interpretation of these results is (Russell and Wootton 1993), a correlation also found

168 # 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190


Compensatory growth in ¢shes M Ali et al.

in other teleosts (Fletcher 1984; Grove et al. 1985;


Singh and Srivastava 1985). Because of this relation-
ship, a reduction in evacuation rates would be
expected to result in a corresponding delay in the
return of appetite. However, a period of food restric-
tion usually evokes a subsequent compensatory rise
in consumption (Tyler and Dunn 1976; Godin 1981;
Jobling 1982; Miglavs and Jobling 1989b; Ali and
Wootton 1998). A more rapid return of appetite may
cause the hyperphagia following food deprivation in
O. kisutch (Dunbrack 1988). An increase in the rate
of gut evacuation and more rapid return of appetite
have been implicated in the increased consumption
of low-energy diets by ¢sh (Grove et al.1978).
There is evidence for structural changes in the gut, Figure 8 Hypothetical responses of standard metabolic
which increase its capacity, when ¢sh are exposed to rate (SMR) to starvation. (a) Metabolic rate decreases during
intermittent feeding (Carter et al. 2001). Although starvation towards a physiological stable minimum. (b) The
these long-term changes probably cannot account adjustment of SMR is not immediate, but there is a time-lag
for the immediate hyperphagia seen in compensa- between deprivation and SMR decline. (c) SMR initially
increases as a result of hyperactivity and then declines
tory growth, they may account for the longer term
towards the stable level.
changes in hyperphagia shown when ¢sh are
exposed to cycles of feeding and deprivation (Ali and
Wootton 2001). and glycogenolytic enzymes in swimming muscles;
(iii) adaptation: stabilisation of whole body metabolic
rate around reduced level. If deprivation persists,
Physiological basis of compensatory growth:
then there is increasing replacement of lipid by pro-
metabolic rate
tein as the major metabolic fuel; (iv) recovery: rapid
Metabolic rates of ¢shes may decrease during food increase in rates of oxygen consumption and growth,
restriction (Beamish 1964; Love 1970, 1980; Blaxter and compensatory growth spurts correlate positively
and Ehrlich 1974; Jobling 1980; Johnston 1981; Du with the length of the starvation period. Therefore, it
Preez et al.1986a,b; Du Preez 1987;Wieser et al.1992). is crucial in studies on the e¡ects of starvation on
Brown (1946, 1957) reported that when the quantity standard metabolic rate to control ¢sh activity; other-
of food given to brown trout was reduced towards wise, it would be di⁄cult to distinguish between a
the maintenance level, the ¢sh ¢rst lost mass but reduction in metabolic rate associated with an accli-
soon became ‘acclimatised’ to the lower ration levels matisation e¡ect and a re-adjustment after the initial
and then gained mass. Similar results were obtained increase of activity during the stress phase (Wieser
by Hepher et al. (1983) for red tilapia, Oreochromis sp. 1991; Fig. 8). Nevertheless, a metabolic adjustment to
(Cichlidae), and suggest an adjustment in the meta- conditions of nutritional stress has been clearly
bolic rate. After an acclimatisation period, the meta- demonstrated. In juvenile cyprinids, both routine
bolic rate of tilapia was regulated, as loss of body and standard rates of metabolism can decrease by
mass was higher during the initial week of fasting 30^40% after 2 days of deprivation in relation to
but mass declined more slowly in subsequent weeks. more recently fed individuals (Wieser et al.1992).
In these studies, metabolic rate during the initial per- If, during restriction, there is an ‘acclimatisation
iod of starvation was higher than during the subse- period’ before metabolic rate declines, there could
quent period, indicating an acclimation response to also be a comparable time lag upon refeeding before
fasting. However, Wieser et al. (1992) suggested four metabolic rate reaches levels normal for feeding ¢sh.
phases in response to deprivation and subsequent This would increase the scope for growth during the
refeeding: (i) stress: characterised by a state of hyper- initial stages of re-alimentation of ¢sh. The monitor-
activity (i.e. search for food); (ii) transition: with con- ing of oxygen consumption during starvation and
tinued deprivation, reduction in routine rate of refeeding of juveniles of three cyprinid species, Leu-
respiration associated with reduction in locomotion ciscus cephalus, Chalcalburnus chalcoides mento and
and decline in the activity of some major glycolytic Scardinius erythrophthalmus, revealed only brief

# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190 169


Compensatory growth in ¢shes M Ali et al.

suppression of metabolic rate (compared to controls) dual can incur. An important, yet unresolved,
during the initial stages of re-alimentation, although question is how standard metabolic rates react dur-
compensatory growth was observed (Wieser et al. ing a phase of compensatory growth? This issue is
1992). In this study, mass-speci¢c maintenance and relevant because it can provide useful information
routine rates of metabolism rapidly increased to about the costs and risks associated with rapid
levels normal for feeding ¢sh, suggesting that their growth, and an understanding of the relationships
suppression could not account for observed growth between resting metabolism, maximum rate of meta-
compensation. Food restriction in juvenile carp also bolism and metabolic scope (e.g. Priede 1985; Cutts
led to biochemical changes that probably reduced et al. 2002).
metabolic rate (Bastrop et al. 1991). Activities of
NADPþ-speci¢c dehydrogenases (G6PDH, 6PGDH,
Growth e⁄ciency
IDH, ME, G1DH) implicated in lipogenesis and in
amino acid metabolism were depressed after food Dobson and Holmes (1984) suggested that compensa-
deprivation. Temporal priority of enzyme recovery tory growth in ¢sh could be attributed to an
on re-alimentation was observed with the level of increased e⁄ciency of food utilisation. Studies on
enzymes recovering at di¡erent rates. individually housed ¢sh demonstrated improved
Energy expenditure during starvation of ¢sh can conversion e⁄ciency, based on growth in mass and
be reduced by decreased locomotor activity (Beamish mass of food consumed on re-alimentation in Arctic
1964; Blaxter and Ehrlich 1974; Wieser et al. 1992). charr, European minnows and the three-spined
Reduced activity during refeeding could contribute sticklebacks (Miglavs and Jobling 1989a,b; Russell
to compensatory gain by increasing the proportion and Wootton 1992; Jobling et al.1994; Zhu et al. 2001).
of the energy available to growth. However, hyper- Koppe et al. (1993) reported that food conversion e⁄-
phagia is usually associated with a higher level of ciency, and protein and energy utilisation of re-
foraging activity; so, from a cost-bene¢t perspective, alimented ¢sh were superior to controls fed at satia-
only a reduction of nonfeeding activity could be tion levels in groups of Piractus brachypomus. In rain-
regarded as a bene¢cial e¡ect. Although swimming bow trout, one study found that compensatory
activity was not measured directly, Wieser et al. growth was entirely caused by an increase in growth
(1992) found that both standard and routine meta- e⁄ciency, with no hyperphagic response (Boujard
bolic (i.e. the metabolism of the ¢sh performing nor- et al.2000). In contrast, Hayward et al. (1997) detected
mal, spontaneous movements in the absence of no overall di¡erences in growth e⁄ciencies between
other stimuli) rates increased upon refeeding, sug- controls and groups on cycles of starvation and
gesting that activity also had increased. Juvenile refeeding in Lepomis hybrids. Lepomis experiencing
roach showed a reduction in swimming activity dur- cycles of 14 days of deprivation did have high growth
ing 3 weeks of starvation, but activity returned to e⁄ciencies during the phases of compensatory
control levels as soon as refeeding started (van Dijk growth that occurred on refeeding. In channel cat-
et al. 2002). In the sable¢sh (Anoplopoma ¢mbria; Ana- ¢sh, there was no di¡erence between control and
plopomatidae), compensatory growth seems to be starved ¢sh during 18-week trials (Kim and Lovell
mediated by a shift in resource allocation, which 1995). Growth e⁄ciency did not di¡er between con-
results in a reduced physiological capacity for forced trol and starved group after refeeding for 7 weeks in
swimming (Sogard and Olla 2002). The compensa- rainbow trout (Speare and Arsenault1997) or in carp
tory ability of this species is reduced (Sogard and Olla that had been fed diet reduced in protein or energy
2002), which has been interpreted as a result of the content during the deprivation period (Plank et al.
¢sh having high routine consumption rates that can- 1984).
not be exceeded. In the same study, walleye pollocks If growth e⁄ciency was higher during compensa-
(Theragra chalcogramma; Gadidae) exhibited a strong tory growth, this can be demonstrated by comparing
compensatory response by greatly increasing their the growth of re-alimented ¢sh with controls when
consumption rates compared to control ¢sh (Sogard both groups are fed identical rations. Again, because
and Olla 2002). of the potential inverse correlation between con-
In ectotherms, standard metabolic rates are a mea- sumption rate and conversion e⁄ciency (Kinne
sure of the resting metabolism and represent the 1960; Pandian 1967; Solomon and Bra¢eld 1972;
minimum cost of maintenance that a healthy, inac- Elliott1976;Talbot 1985), evaluating the contribution
tive, unstressed, unfed and nonreproductive indivi- of changes in food conversion e⁄ciency to growth

170 # 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190


Compensatory growth in ¢shes M Ali et al.

compensation requires adequate control of variation was in£uenced by starvation in juvenile African cat-
in intake rates. ¢sh, Clarias gariepinus (Clariidae).
Changes in lipid levels take place before and during
compensatory growth. Decreases in liver and viscera
Proximate composition, metabolites and
as proportions of body mass occurred in Arctic charr
endocrines during food deprivation
in response to food restriction for 8 weeks and star-
and recovery
vation for 8, 12 and 16 weeks (Miglavs and Jobling
If compensatory growth is to be initiated on refeed- 1989a). Starvation (for 16 weeks) resulted in signi¢-
ing, changes in the physiological state of ¢sh during cant decreases in the lipid contents of the carcase
the deprivation period must signal a discrepancy and viscera. Charr restricted for 8 weeks showed a
between the current state and the state of ¢sh that marked decrease in the proportion of body energy
have not experienced that deprivation. It is unclear present as lipid. The lipid energy:protein energy ratio
which, if any, of the changes observed during depri- decreased from 1.16 in the initial sample to 0.80 after
vation provide the error signals for compensatory 8 weeks of restriction. Restoration of satiation feed-
growth when food becomes freely available. ing was followed by signi¢cant increases in liver and
In mammals, there are distinctive neural systems, viscera as per cent of body mass. After 8 weeks of re-
with their associated peptides, that regulate appetite alimentation, there were higher percentages of lipid
for fat, protein and carbohydrate (Hoebel1997).Appe- and lower percentages of water in the liver, compared
tite in ¢sh is probably under comparable multifactor- to continuously fed controls. By the end of re-alimen-
ial control (Fletcher 1984; Le Bail and Boeuf 1997; tation, whole body ratios of lipid energy:protein
DePedro and Bjornsson 2001), although the details energy had risen to1.33, similar to the1.30 of controls.
probably di¡er in some respects from endotherms. However, in terms of total body mass, only partial
The control of appetite after a period of food depriva- compensation had been achieved.
tion has both short-term and longer term compo- Rainbow trout starved for 3 weeks showed a sig-
nents (Carter et al. 2001). The short-term e¡ects ni¢cant change in carcase composition (Quinton
partly re£ect the immediate e¡ects of re-¢lling and and Blake 1990); a reduction in fat and an increase
emptying the alimentary canal, including sensory in moisture and protein concentration had occurred
and hormonal feedbacks, together with short-term compared to controls. No di¡erences were evident
changes in circulating nutrients. The longer term after a subsequent re-alimentation period lasting for
component would re£ect the error signalled by phy- 3 weeks. In gibel carp, the ratio of fat to lean body
siological state. Changes in proximate composition mass was signi¢cantly reduced by 2 weeks of starva-
of the ¢sh during deprivation and refeeding may o¡er tion, but restored within 2 weeks of satiation feeding
clues as to the control of appetite during the hyper- (Xie et al. 2001).
phagic phase, although such changes have rarely Golden perch, Macquaria ambigua (Percichthidae),
been studied in the context of compensatory growth. a carnivorous Australian freshwater ¢sh, experi-
ences marked seasonal hydrological changes that
a¡ect food availability in its natural habitat. It
Lipids
showed declines in hepatosomatic index (HIS) and
Fish, whose metabolism is largely based on lipids and depletion of liver lipid, protein and glycogen over
proteins (Jobling1994), store lipids in the liver, viscera 60 days of starvation. Then, levels were maintained
and muscle. The detailed distribution between these over a longer period of starvation. Following deple-
body components does vary between species (Love tion of lipid stores in liver, lipid contained in perivisc-
1970). Lipids are broken down early in starvation eral adipose tissue was utilised along with epaxial
(e.g. Anguilla anguilla (Anguillidae), Larsson and muscle glycogen. Muscle was not mobilised until
Lewander (1973); Dicentrarchus labrax, Stirling after depletion of hepatic and perivisceral reserves.
(1976); Esox lucius (Esocidae), Ince and Thorpe 1976; Hepatic energy reserves were rapidly restored on
rainbow trout, Jezierska et al. 1982; Oreochromis nilo- refeeding. Levels of protein and lipid in the muscula-
ticus, Satoh et al. 1984). Lipids often constitute the ture did not change in response to food deprivation.
main energy source for maintenance of ¢sh during No signi¢cant increase in water content of the muscle
over-wintering fasts (Weatherley and Gill 1987; Bull of starved ¢sh was observed over 210 days of the
and Metcalfe 1997). Zamal and Ollevier (1995) experiment. The water content in the tissue of fed ¢sh
reported that the relative composition of fatty acids did decrease as lipid accumulated. Liver was used as

# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190 171


Compensatory growth in ¢shes M Ali et al.

an initial source of endogenous energy in response to that the growth compensation in the refeeding per-
food deprivation (Collins and Anderson1995). iod was entirelycaused byan improvement in growth
Periods of food deprivation imposed on slow-grow- e⁄ciency and not hyperphagia.
ing, juvenile Atlantic salmon in winter signi¢cantly In mammals, the brain neuropeptides galanin (sti-
reduced fat levels in comparison to control ¢sh. Dur- mulatory) and leptin (inhibitory) have been impli-
ing the refeeding period, the food-restricted groups cated in the control of an appetite for lipid (Hoebel
restored their fat levels to levels similar to those of 1997). Their roles in teleosts are still to be clari¢ed
control ¢sh (Metcalfe and Thorpe 1992; Bull et al. (Hoebel 1997; Le Bail and Boeuf 1997; DePedro
1996; Bull and Metcalfe 1997). Both slow- and fast- and Bjornsson 2001), although Silverstein and
growing salmon exhibited a rapid, full re-adjustment Plisetskaya (2000) detected no e¡ect of injection of a
of their mass:length ratios, regardless of the degree fragment of rat leptin into the brain of channel cat¢sh
of compensation achieved for absolute mass or length on appetite.
(Nicieza and Metcalfe 1997); such rapid re-adjust-
ments seem to be related to changes in lipid levels.
Protein and RNA:DNA ratios
When deprived, the salmon responded by exhibiting
a hyperphagic response (Metcalfe and Thorpe 1992; In ¢sh, muscle RNA concentrations, because of the
Bull et al. 1996; Bull and Metcalfe 1997; Nicieza and role of RNA in protein synthesis, re£ect di¡erent
Metcalfe 1997). This protective response was at least somatic growth rates induced by di¡erent levels of
in part responsible for the restoration of body fat lost feedings (Buckley 1979; Westerman and Holt 1988;
during the food deprivation. The extent of the esti- Bastrop et al.1991). In juvenile Atlantic salmon, RNA
mated lipid de¢cit incurred during a period of food concentrations and the RNA:DNA ratio responded
restriction primarily a¡ected the duration of the to three levels of reduced ration representing 0, 20
hyperphagic response that occurred when food was and 50% of the control level imposed for a maximum
once again available, rather than intensity of feeding of 8 days (Arndt et al. 1996). By day 4 of restriction,
(Bull and Metcalfe1997). the starved group had signi¢cantly lower RNA con-
Arctic charr showed an inverse relationship centrations than control ¢sh. By day 8, all four ration
between food intake (% body mass day1) and car- levels were re£ected in RNA concentrations. For sal-
case lipid (% wet mass) and visceral lipid (% wet mon starved or fed a 20% ration for 4 days, the RNA
mass; Jobling et al. 1993). Koppe et al. (1993) reported levels recovered close to control levels after 4 days of
a correlation between food intake and visceral fat refeeding, although for the starved ¢sh, the
contents in juvenile Piaractus brachypomus. A com- RNA:DNA ratio was a better indicator of refeeding.
parison between temporal changes in food intake Ornthine decarboxylase (ODC, the enzyme that plays
and body composition suggested that the level of food a rate-limiting role in polyamine synthesis, which
intake in re-alimented ¢sh decreased, as visceral fat re£ects protein synthesis) activity also dropped shar-
contents approached the level of controls fed at satia- ply during food restriction. ODC activity increased
tion level. Peter (1979) indicated the possibility of a rapidly after feeding was returned to the level of the
similar mechanism in other species of ¢sh (see also control group and showed some evidence of over-
Machiels and van Dam 1987). In channel cat¢sh, ¢sh compensation in the refeeding period (Arndt et al.
were manipulated by diet composition to reach di¡er- 1996). Unfortunately, the length of this promising
ent levels of fat content and then fed the same diet. experiment was too short for compensatory growth
In the ¢rst 2 weeks following a switch to the same to be clearly displayed. In Arctic charr subject to
diet, the low-fat cat¢sh ate more than the high-fat ¢sh regimes of food restriction and re-alimentation,
(Silverstein and Plisetskaya 2000). These results are RNA:DNA ratios in muscle and liver correlated with
in accord with the expectations of a theory of appetite the speci¢c growth rate (Miglavs and Jobling 1989b).
regulation based on maintenance of energy reserves However, the growth rates during the compensatory
within narrowly de¢ned set point limits (Jobling and growth phase were higher than those that would have
Johansen 1999). One result not compatible with the been predicted from the RNA:DNA ratios observed.
model was the observation that, in rainbow trout In carp, food deprivation reduced cell size of liver
subjected to 1, 11 or 21 days of deprivation, growth (estimated indirectly), but not of white muscle. This
and feed intake in a refeeding period of 10 days were reduction was associated with a decrease in soluble
not related to lipid levels at the end of deprivation and total protein concentrations (Bastrop et al.1991).
(Boujard et al. 2000). This study was also unusual in Starved carp had lower white muscle RNA and liver

172 # 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190


Compensatory growth in ¢shes M Ali et al.

RNA concentration than fed ¢sh, indicating reduced matic index, and protein and lipid contents indicated
protein synthesis under food restriction. (Bastrop that the muscle did not completely recover from
et al.1991). Mendez and Wieser (1993) subjected juve- 2 months of fasting, although an overshoot of muscle
nile roach, 280^460 mg in fresh mass, to 7-,14-, 22-, glycogen was observed (Blasco et al.1992a,b). In gold-
29-, 36- and 49-day starvation regimes at 20 8C to ¢sh (Chavin andYoung1970) and Opsanus tau: Batra-
study the e¡ects of deprivation and refeeding on choididae (Tashima and Cahill 1968), blood glucose
metabolic events. No e¡ect on protein g1 dry mass levels were also maintained for long periods during
was observed. Enzyme activities were una¡ected in food restriction.
the 7-day group. In the14-day group, enzyme activity The role of glucose in regulating appetite in tele-
dropped during the ¢rst week but remained constant osts is unclear. In comparison to mammals, the ¢sh
thereafter. In the 22-day group, enzyme activity have a poor ability to regulate plasma glucose
declined continuously over starvation. There was a (Mommsen1998). Carter et al. (2001) suggest that this
slow recovery of enzyme activity upon refeeding. In makes it unlikely that glucose plays an important role
the 29-day group, activity of one enzyme (glutamate- in the regulation of food intake in ¢sh. The lowest
pyruvate transaminase) increasedduring deprivation. blood glucose levels in Rooseveltiella nattererri (Char-
Activityof this enzyme declined on refeeding. acidae) preceded the period of greatest food intake
In mammals, growth hormone releasing factor (Bellamy 1968). In contrast, Katsuwanus (Scombri-
(GHRF) plays a role in regulating the appetite for pro- dae) continued to feed when blood glucose levels
tein (Hoebel1997). Clearly, it would be of value to esti- were rising (Magnusson 1969). However, one study
mate the levels of the equivalent hormone in teleost has implicated plasma glucose levels in compensa-
during the periods of deprivation and compensatory tory growth. In rainbow trout deprived of food for 1,
growth (Le Bail and Boeuf1997). A role for neuropep- 11 or 21 days, the growth e⁄ciency during a 10-day
tideY (NPY) in stimulating appetite during the com- re-alimentation period was inversely related to
pensatory growth phase is suggested by the plasma glucose at the start of the re-alimentation
observation that injection of NPY into the third ven- period (Boujard et al. 2000).
tricle of the brain of channel cat¢sh stimulated appe-
tite (Silverstein and Plisetskaya 2000; DePedro and
Water content
Bjornsson 2001).
Starvation results in tissue hydration (Jobling 1980;
Weatherley and Gill 1987; Miglavs and Jobling 1989a;
Carbohydrates
Quinton and Blake1990). An increase in tissue hydra-
Carbohydrates play onlya minor role as storage mate- tion plays a role in the limitation of the loss or even
rials in ¢sh (Mommsen1998). Carp had similar levels the maintenance of wet body mass during starvation
of blood glucose and liver glycogen when starved or (Love 1970). In juvenile roach, Rutilus rutilus L., it
fed for 22 days (Nagai and Ikeda 1971). When carp took 36 days of deprivation before tissue water con-
were subjected to 2 months of fasting and 12 days of tent was higher than that of controls (Mendez and
refeeding, the ¢rst phase (until day 8 of fasting) was Wieser 1993). Whether the level of tissue hydration
characterised by a reduction in the hepatosomatic plays any direct role in appetite regulation in teleosts
index mainly because of glycogen mobilisation. A is not known. In ¢sh with swim bladders, changes in
transitory increase in plasma glucose and lactate the speci¢c mass of the body would require adjust-
suggested an initial increase in energy demand. No ments in the volume of the bladder to attain neutral
changes were produced in the percentage of glycogen buoyancy, which might signal changes in body com-
and protein in muscles, but the musclosomatic index position.
and the total body muscle protein decreased. Stabili-
sation of liver glycogen content, and plasma glucose
Endocrines
and lactate levels decreased muscle protein levels,
and a reduction in the rate of body mass loss charac- Growth is under endocrine control (Mommsen1998);
terised the second phase (from day 8 of fasting). Total however, the role of hormones in compensatory
protein content in whole muscle decreased by 22%, growth is little understood. Temporal changes in
similar to the ¢rst phase. Twelve days of refeeding growth, plasma thyroid hormones (triiodothyronine,
produced recovery of plasma glucose levels. Liver gly- T3, and tetraiodothyronine,T4), cortisol, growth
cogen completely recovered. In contrast, muscloso- hormone (GH) and nonesteri¢ed fatty acid (NEFA)

# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190 173


Compensatory growth in ¢shes M Ali et al.

concentrations, hepatic T3 content and hepatic 50 - 1999). This suggests that compensatory changes in
monodeiodinase activity were measured in rainbow appetite and growth may re£ect temporary, but unu-
trout subjected to a sustained fast for up to 8 weeks, sually high, levels of stimulation of the GHRF^GH^
and during an 8-week refeeding period. Di¡erences IGF axis. Once the time trajectories of the compensa-
in growth rate between fed and fasted groups were tory growth response have been adequately de¢ned,
evident after 2 weeks, but signi¢cant mass loss by the concomitant changes in levels of GHRF, GH,
fasted groups was not evident until between 4 and GHR and IGF can also be de¢ned. However, the de¢-
6 weeks into the fast. Liver protein content was cits that evoke compensatory changes in this endo-
depressed inthe fasted ¢shwithin 2 days of food depri- crine axis for growth control are obscure.
vation. Interrenal activity suggested a depressed pitui-
tary^interrenal axis in fasted animals.Within1 day of
Modelling compensatory growth
refeeding, plasma cortisol levels had increased to
levels characteristic of controls. The status of thyroid There are numerous models of ¢sh growth in the lit-
hormones indicated by plasma thyroid hormone level, erature (Wootton 1998). These range from complex
liver T3 content, hepatic 50 -monodeiodinase (MD) bioenergetic models, which partition the total meta-
activity and thyroid epithelial cell height also indicat- bolic expenditure into many terms, corresponding
ed a depressed pituitary^thyroid axis in fasted to expenditures associated with locomotion, tissue
animals, with recovery to levels of the fed animals synthesis, tissue maintenance and digestion (e.g.
within 1 week. Despite compensatory changes in Kitchell et al. 1977; From and Rasmussen 1984), to
accumulation of reserves (as indicated bya compensa- the simple von Bertalan¡y (1960) model. However,
tory increase in hepatosomatic index), there were no testing such models often reveals signi¢cant discre-
apparent compensatory changes in any of the endo- pancies between predicted and observed growth
crine variables evident during the refeeding period rates made at low or £uctuating rations (Cui and
(Farbridge and Leatherland 1992a,b; Farbridge et al. Wootton 1989; Ruohonen and Ma«kinen 1992).
1992; Leatherland and Farbridge1992). Whitledge et al. (1998) found that a detailed bioener-
In channel cat¢sh, Gaylord et al. (2001) recorded a getics model was not a good predictor of compensa-
decline in plasma T3 and T4 with food deprivation tory growth or consumption in Lepomis hybrids,
for as few as 3 days, with a subsequent rapid recovery although the model performed well for control ¢sh
on refeeding. However, in this experiment, there was fed on a daily basis. In contrast, a simple, empirical
no evidence of hyperphagia during the refeeding regression model provided good predictions of the
period. growth of three-spined sticklebacks experiencing
In carp fasted for 2 months followed by 12 days of cycles of feeding and deprivation (Ali and Wootton
refeeding, plasma insulin levels decreased twofold 2001).
and plasma glucagon threefold over the ¢rst 8 days Most of the models, including models that describe
of fasting and remained low in the second phase energy £ows into and out of the ¢sh in detail, treat
(from day 8 onwards of fasting).Twelve days of refeed- all internally stored energy as interchangeable, with
ing produced a higher daily growth rate than in con- the instantaneous state of ¢sh characterised by a sin-
trols and a recovery of plasma insulin and glucagon gle variable, normally total dry mass or energy con-
levels (Blasco et al.1992a,b). tent. A ¢xed mass^length relationship is usually
In salmonids, exogenous treatment with GH can assumed. However, this assumption of a constant
induce increased appetite and growth rates in excess mass^length relationship may not be valid if changes
of those normally observed (McLean et al. 1992; Le in food availability over days or weeks cause an
Bail et al. 1993; Johnsson and Bjo«rnsson 1994; alteration in the nutritional status of the ¢sh. The
Johnsson et al. 1996, 1999). This mimics compensa- assumption that two conspeci¢cs of equal mass, one
tory growth. The e¡ect of GH on growth rates is at short and fat and the other long and thin, will grow
least partly mediated through the production of insu- at identical rates may also be invalid (Broekhuizen
lin-like growth factor (IGF) by the liver as a response et al.1994).
to circulating GH. This response depends on the pre-
sence of growth hormone receptors (GHR) in the liver
Structural and storage model
cells. At least in some species, low food levels inhibit
the stimulation of IGF by GH, while levels of GH in Broekhuizen et al. (1994) formulated a simple, physio-
the blood plama increase (Perez-Sanchez and Le Bail logically based model to investigate the phenomenon

174 # 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190


Compensatory growth in ¢shes M Ali et al.

of compensatory growth. The model is based upon Nicieza and Metcalfe (1997) found that compensa-
two key assumptions. The ¢rst is that an individual tory growth continued after the deprived ¢sh had
¢sh partitions net assimilate between two tissue attained the same length^mass relationship as con-
types: those which can and those which cannot be trol ¢sh. This suggests that compensatory growth is
remobilised once laid down. The second is that the more than just the readjustment of the ratio of
individual modulates its behaviour and physiology reserve to structural tissue. In cod, the early phase
in response to the instantaneous ratio of mobilisable of compensatory growth was characterised by
and nonmobilisable tissues. The model indicated that growth of swimming muscle rather than more
the temporal pattern of growth trajectories observed obviously reserve tissue such as liver or visceral fat
during recovery need not indicate the presence of a (Jobling et al. 1994). In the three-spined sticklebacks,
complex memory of environmental conditions. The preliminary estimates of the characteristics of com-
trajectories can be produced by an instantaneous pensatory growth did not suggest that the growth of
response to a simple measure of nutritional well- reserve material (lipids) had precedence over protein
being, the ratio of the mass of mobilisable reserve tis- growth (Zhu et al. 2001). The time resolution of the
sue to that of structural tissue. The model assumes studies on cod and sticklebacks may have been too
that, as this ratio falls, the ¢rst response is that of coarse to detect short-term shifts in the rate of
‘hunger’ or increased appetite. When reserves have growth of reserve and structural material. Strikingly,
fallen far enough to indicate a signi¢cant risk of star- when compensatory growth in rainbow trout was
vation, the animal reduces its basal metabolic rate, entirely a consequence of increased growth e⁄-
thus lowering the risk, but at the cost of a decrease ciencyand not hyperphagia, no e¡ect of lipid reserves
in foraging potential. This model was developed spe- on appetite was detected (Boujard et al. 2000).
ci¢cally to address the question of how compensa-
tory growth might be regulated, and the other
Model based on speci¢c growth rate
details were not incorporated. Although the model
illustrates some general principles and a subset of Hubbell (1971) developed a model of growth regula-
the speci¢c responses in one species group (salmo- tion in animals based on control theory. It assumed
nids), it is not a complete model of ¢sh growth and tis- that individuals had an optimal growth trajectory. If
sue allocation pattern. the growth rate falls so that the individual grows
A related model is the lipostat of Jobling and Johan- slower than the optimal trajectory, compensatory
sen (1999). This argues that appetite is regulated in changes are induced that reduce the error between
relation to lipid levels, so hyperphagia and the conse- the optimum and achieved growth rate. The model
quent compensatory growth cease as lipid levels does not seek to identify the mechanisms by which
reach the optimal levels. In a study of grouped post- the control is achieved, but it does provide a theoreti-
smolt Atlantic salmon, ¢sh that had been deprived of cal framework within which such mechanisms could
food showed hyperphagia and higher growth rates. be sought. This model was used to interpret the com-
However, their compensatory response abated once pensatory growth patterns of the European minnows
the deprived salmon had attained the same body (Russell and Wootton1992).
composition as that of the control ¢sh, although the The multifactorial control of appetite (Hoebel1997;
deprived ¢sh were smaller than controls and so had Le Bail and Boeuf 1997; DePedro and Bjornsson
only shown partial compensation in mass (Johansen 2001) suggests that the control of compensatory
et al. 2001). Further evidence compatible with a lipo- growth may be equally complex and may depend on
stat model of appetite control was obtained by manip- the e¡ect that the period of deprivation has had on
ulating the lipid content of juvenile Atlantic salmon body composition. There may be di¡erent trajectories
by providing high rations of either high-fat or low-fat of recovery for the carbohydrate, lipid and protein
feed (Johansen et al. 2002). This avoided the use of a components of the body. The roles of hyperphagia
period of food deprivation to reduce lipid levels, which and increased growth e⁄ciency may also depend on
confounds the e¡ects of deprivation per se with the whether the preceding food deprivation was total or
e¡ects of body lipid levels. It is unclear whether a lipo- partial (Boujard et al. 2000). If so, this complexity
stat model is relevant when growth depression is needs to be re£ected in the models. There is a need to
induced by unseasonably low temperatures, rather integrate the general control theory framework
than food deprivation (Nicieza and Metcalfe 1997; developed by Hubbell (1971) with models of the con-
Maclean and Metcalfe 2001). trol of appetite such as provided by DePedro and

# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190 175


Compensatory growth in ¢shes M Ali et al.

Bjornsson (2001) for the e¡ects of neuropeptides and Juvenile Atlantic salmon rely heavily on lipid
hormones and Carter et al. (2001) for short-, medium- reserves to avoid starvation during their freshwater
and long-term physiological control. phase, drawing upon reserves during winter
(Egglishaw and Shackley 1977; Gardiner and Geddes
1980) when food supply is inadequate and unpredict-
Functional significance of compensatory
able. During this time, fat stores and appetite are
growth
regulated in relationship to body condition. Size-
Sibly and Calow (1986) argued that, on theoretical dependent mortality suggests a function for compen-
grounds, compensatory growth is likely to be advan- satory growth at this stage (Metcalfe and Thorpe
tageous only under certain environmental condi- 1992; Bull et al. 1996). In juvenile Atlantic salmon on
tions. These include seasonality associated with a slow-growing trajectory, an accelerated depletion
changes in the suitability of the environment and of fat reserves in early winter led to a hyperphagic
conditions in which individuals of reduced size response, which replenished losses (Metcalfe and
experienced higher risks. Thorpe 1992). Appetite fell to a low level once lipid
levels had been restored. This is an example of a com-
pensatory response that seems to relate solely to the
Size-dependent mortality
storage component of the body.
Mortality rates in aquatic systems are inversely
related to size (McGurk 1986; Sogard 1997). In ¢sh
Size-dependent prey selection
populations, mortality over some seasons can be
inversely related to size, partly because smaller indi- Fish are gape-limited predators, and feeding habits
viduals accumulate relatively smaller energy are strongly in£uenced by ¢sh size (Dunbrack and
reserves (Schultz and Conover1999). Menidia menidia Dill 1983; Macchiusi and Baker 1991). As they grow,
display size-dependent over-wintering mortality the increase in size of the mouth allows them to take
(Conover 1992). Over-wintering mortality of young- larger size prey. This often leads to large ontogenetic
of-the-year yellow perch and white perch, Morone changes in diet (Gerking1994;Wootton1998). A shift
americana (Moronidae), predominantly a¡ects the to a larger and more pro¢table prey may result in an
smallest individuals (Post and Evans 1989; Johnson acceleration in growth, for example when ¢sh switch
and Evans 1990, 1991). This pattern appears to be from a diet of invertebrates to one of ¢sh (e.g. Gorman
common among teleosts (Hunt 1969; Oliver et al. and Nielson 1982; Buckel et al.1998). A phase of com-
1979; Toneys and Coble 1979; Adams et al. 1982; Con- pensatory growth may allow individuals to reach
over and Ross1982; Henderson et al.1988). Often sus- the size at which the ontogentic diet switch can be
ceptibility to predators is also higher for the smaller made. A failure to make that switch may result in a
¢sh, even when the larger ones are more prone to risk low body size throughout life.
exposure to predators (Johnsson 1993; but see Litvak
and Leggett 1992). In brook trout, Salvelinus fontina-
Size-dependent ontogenetic changes
lis, postreproductive, over-wintering survival of
females increased with body size (Hutchings 1994). During ontogeny, there may be critical transitions
When ¢sh with a small body size do su¡er higher that depend on an individual reaching a threshold
rates of mortality, there is clearly an advantage if size. Compensatory growth would allow an indivi-
compensatory growth can allow some recovery of dual to reach that threshold size, despite experien-
body size after a period of growth depression (Letcher cing a prior period of growth depression. Juvenile,
et al. 1996). In a ¢eld experiment with juvenile Atlan- anadromous salmonids undergo smolti¢cation as
tic salmon, the juveniles were stocked at di¡erent the precursor to their migration to the sea. In some
sizes in late spring, but no di¡erences in size were species, there is a narrow time window during spring
detected in the following autumn (Letcher and when this process occurs. Marine survival is maxi-
Terrick 2001). Estimates of the survival rates sug- mised by the juveniles reaching a critical minimum
gested that the result was not a consequence of size prior to seaward migration (e.g. Parker 1971;
size-dependent mortality over the summer. Letcher Bilton et al. 1982; Healey 1982; Henderson and Cass
and Terrick (2001) proposed that the convergence 1991; Lundqvist et al.1994). The capacity for compen-
of size by autumn was a result of compensatory satory growth shown by juvenile salmon can
growth. increase the chance of an individual achieving

176 # 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190


Compensatory growth in ¢shes M Ali et al.

that critical minimum at an early age (Nicieza and The nature of compensatory growth may di¡er
Metcalfe 1997). In Atlantic salmon, juveniles show from long-term growth and be less metabolically
di¡erent growth trajectories that result in di¡erences expensive. This would allow a transient period of
in age at smolti¢cation. The faster growing group unusually high growth rates. A careful analysis of
showed a stronger compensatory response than the the nature of the material laid down during compen-
slower growing group. Nicieza and Metcalfe (1997) satory growth, in association with a detailed energy
suggest that growth, including any compensatory budget, would identify this type of growth. However,
growth, of the faster-growing group would be tar- neither juvenile three-spined sticklebacks nor rain-
geted at reaching the critical size for smolti¢cation, bow trout found showed unusual patterns of lipid or
whereas that of the slower-growing group would be protein deposition during compensatory growth
targeted at a level of storage. (Boujard et al. 2000; Zhu et al. 2001).
Priede (1985) has used an analogy with machines
to argue that organisms operating at maximum rates
Size-dependent reproductive success
pay a cost in high mortalities. This suggests that
Fecundity in female teleosts is a function of body size long-term growth rates represent an optimal solution
(Wootton 1998). A disadvantage of small size is an to the trade-o¡ between growth rate and mortality
absolute lower fecundity. In ¢rst-spawning cod, that maximises ¢tness (Sibly and Calow 1986). More
9 weeks of starvation was not fully compensated for rapid growth entails physiological adjustments that
in the subsequent period of refeeding. On spawning, would increase mortality under adverse conditions.
the smaller females had lower absolute fecundities, For example, ¢sh growing faster would be expected
but fecundity per unit of body mass did not re£ect to have higher oxygen consumption (Brown and
nutritional history (Karlsen et al.1995). Thus, fecund- Cameron 1991; Houlihan et al. 1993; Jobling 1993)
ity showed neither over- nor under-compensation. and thereby a higher risk of mortality when oxygen
The e¡ects of timing of periods of food deprivation concentration is lowered. There are no reports of
and any subsequent period of growth compensation increased mortality rates in groups of ¢sh under-
on fecundity are yet to be adequately de¢ned, despite going compensatory growth. Curiously, a study on
their potential importance in determining the egg channel cat¢sh found that groups of ¢sh fed intermit-
production by the spawning stock of a population. tently and showing growth compensation were less
Although the relationship between fecundity and susceptible to infection by the pathogenic bacteria
size seems general in teleosts, in some species, large Edwardsiella ictaluri than continuously fed controls
size also confers a breeding advantage on males. This (Chatakondi and Yant 2001).
advantage may accrue when the mating system High rates of growth may be associated with a
involves sperm competition or territorial defence greater risk of developmental abnormalities (Leamy
(Parker 1992). In such species, growth compensation and Atchley 1985; Sibly and Calow 1986; Arendt
would be advantageous to males. 1997). This potential cost of compensatory growth is
likely to be most obvious in the early life-history
stages when both growth and developmental rates
Costs of compensatory growth
are high. In a comparison of two populations of
Compensatory growth with its associated changes in pumpkinseed sun¢sh, Lepomis gibbosus (Centrachi-
appetite presents an evolutionary puzzle. The advan- dae), ¢sh from the population with the faster post-
tages of large body size suggest that individual ¢sh hatching growth showed a delay in the start of the
should grow at the maximum rate physiologically ossi¢cation of the cranial bones (Arendt and Wilson
possible to obtain those advantages (Sibly and Calow 2000). Comparisons both within and between popu-
1986; Arendt and Wilson 1997; Arendt 1997). How- lations of pumpkinseed sun¢sh found an inverse
ever, rates during compensatory growth exceed relationship between the strength of body scales and
those shown by continuously fed ¢sh, indicating that rate of growth (Arendt et al. 2001). An experimental
the latter are consuming food and growing at rates study on immature three-spined sticklebacks in
less than the rate that is maximally possible. This which growth rate was manipulated by di¡erences
suggests that there are costs associated with com- in ration and temperature assessed developmental
pensatory growth that must be traded o¡ against stability by estimating the £uctuating asymmetry of
its bene¢ts (Arendt 1997; Metcalfe and Monaghan ¢n and skeletal traits (Robinson and Wardop 2002).
2001). Sticklebacks with the higher growth rate had higher

# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190 177


Compensatory growth in ¢shes M Ali et al.

levels of £uctuating asymmetry, suggesting a devel- (DamsgQrd and Dill 1998). Rainbow trout exogen-
opmental cost of rapid growth. In Atlantic salmon, ously treated with GH, which induced an increased
the prevalence and severity of coronary arterio- appetite, showed increased boldness in the presence
sclerosis has been associated with high growth rates, of a model heron (Jo«nsson et al. 1996), as did trans-
especially those associated with the marine phase of genic Atlantic salmon with enhanced growth rates
the salmon life cycle (Saunders et al. 1992). None of (Abrahams and Sutterlin 1999). Fast-growing Atlan-
these studies explicitly analysed the costs of compen- tic silversides were more vulnerable to predation by
satory growth. piscivorous ¢shes (Lankford et al. 2001). Hyperphagia
Conditions leading to rapid growth can in£uence may also induce higher levels of intraspecifc aggres-
muscle cellularity and development, and conse- sion as individuals compete for food, with associated
quently function. Christiansen et al. (1992) noted that costs in terms of higher rates of energy expenditure
rapidly growing Arctic charr sometimes showed a and increased risk of mortality. Food-deprived Atlan-
distinctive pattern of damage in their white swim- tic salmon juveniles did show increased levels of
ming muscle. Unusually fast-growing coho salmon, aggression during their compensatory growth phase
which had been genetically manipulated by the (Nicieza and Metcalfe 1997). Fast-growing juvenile
insertion of copies of a gene for growth hormone, Atlantic salmon were more aggressive than slow-
showed a trade-o¡ between growth and swimming growing ¢sh, but they were also more vulnerable to
performance (Farrell et al. 1997). In Menidia menidia, being attacked (Nicieza and Metcalfe1999).
interpopulation comparisons and manipulations of
growth rates within populations indicated that there
Compensatory growth as a by-product of
was a trade-o¡ between growth rate and swimming
flexible growth and dynamic trade-offs
performance (Billerbeck et al. 2001): faster growing
silversides had a poorer prolonged and burst swim- Compensatory growth can be interpreted as a conse-
ming performance. A cost in locomotion perfor- quence of a dynamic process of optimising the alloca-
mance has also been described in the sable¢sh, tion of resources between structural, storage and
whose reduced compensatory ability appears to be gonadal growth, maintenance and energy-consum-
mediated bya diversion of energy otherwise allocated ing activities such as foraging and predator avoid-
to forced swimming (Sogard and Olla 2002). ance (Sibly and Calow 1986; Nicieza and Metcalfe
One study has explicitly linked compensatory 1997). The extent of the development of a compensa-
growth with delayed but measurable growth costs tory growth response should re£ect the balance
(Morgan and Metcalfe 2001). Grouped juvenile between the costs associated with increased growth
Atlantic salmon that had shown full compensation rates and the bene¢ts of regaining a growth trajec-
in autumn after a period of food deprivation showed tory and an appropriate balance between structural
signi¢cantly lower growth rates and lipid reserves and reserve materials.
the following spring than continuously fed controls, Compensatory growth shows a diversity of forms.
although food was freely available. The study also In some conditions, it simply restores lost reserves,
suggested a cost to future reproduction. The male sal- or it may restore the growth trajectory in length or
mon that had shown compensatory growth also had mass, or it may consist of a combination of reserve
lower subsequent rates of sexual maturation in the and somatic growth. Several hypotheses address this
following autumn. However, female three-spined diversity (Metcalfe et al. 2002). First, the model of
sticklebacks exposed to alternating short-term peri- Broekhuizen et al. (1994) is a nutritional state hypoth-
ods of deprivation and refeeding in the month before esis, predicting a consistent compensatory response
the breeding season started spawning at the same to a given de¢cit in the ratio of storage to structural
time, at the same size and had the same fecundity materials. Second, there may be physiological con-
at ¢rst spawning as females fed ad libtum (Ali and straints on the compensatory response that can be
Wootton1999). mounted, so di¡erent responses occur depending on
The hyperphagia usually associated with compen- the environmental circumstances. An example
satory growth may also impose behavioural costs. would be low temperatures that depress appetite to
Increased foraging may expose the animal to a such an extent that a compensatory growth response
greater risk of predation. Foraging coho salmon juve- cannot be mounted. A third hypothesis is that com-
niles showing compensatory growth were bolder in pensatory responses are adaptive and vary faculta-
the presence of a potential predator than control ¢sh tively according to ontogenetic stage, nutritional

178 # 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190


Compensatory growth in ¢shes M Ali et al.

state of the individual and time of the year. Metcalfe type occurs when initially smaller individuals do
et al. (2002) interpret their study of seasonal di¡er- eventually reach a certain length by showing, but
ences in the compensatory response of juvenile not exceeding, the growth rates of initially larger
Atlantic salmon as a possible example of this faculta- individuals, whereas ‘true’ compensation refers to a
tive response hypothesis. higher growth rate. Only one study on ¢sh presented
Compensatory growth is often treated as a discrete a comprehensive analysis of the problems in the
process, isolated from the theoretical framework ‘determination, essence and reasons of the growth
developed to explain the £exibility of growth trajec- compensation’ (Zivkov 1982). He suggested that
tories and how these are a¡ected by a trade-o¡ growth compensation must not be regarded as some-
between growth and mortality rates (e.g. Ludwig thing di¡erent from ‘normal’ growth and noted that
and Rowe 1990; Houston et al. 1993; Werner and understanding the nature of compensatory growth
Anholt 1993; Abrams et al. 1996; Houston and requires understanding the more general process of
McNamara 1999). While this narrow approach has growth regulation. His view that growth compensa-
proved useful in identifying generalised characteris- tion and growth depensation constitute the two ends
tics of thegrowthtrajectories (e.g. submaximalgrowth (convergence and divergence of individual growth
rates), a wider view is needed to understand the trajectories) of the process of growth regulation has
evolutionary mechanisms that gave rise to and main- not been discussed, although these two processes
tain growth compensation in natural populations. are commonly observed in ¢shes.This view implicitly
It is frequently assumed that growth rates are considers that selection might be acting not on
directly determined (as opposed to just ‘in£uenced’) growth compensation itself but on the continuous
by environmental factors including temperature, risk adjustment of growth rates. A wider view of growth
of predation or food availability (Ro¡ 1992). Compen- compensation goes beyond the partial compensa-
satory growth provides evidence that growth resili- tion^full compensation dichotomy by considering
ence persists without the direct in£uence of ‘negative compensation’. Fish often grow at a lower
environmental conditions. There is not a single rate than that imposed by direct, existing environ-
growth rate associated with a given temperature, mental conditions (chie£y, temperature and food
meal size, risk or time, but the rate also depends on availability), which provides evidence of growth costs
individual state (e.g. size, energy reserves). The bal- and suggests that there is an optimal growth trajec-
ance of costs and bene¢ts of rapid growth can vary tory. Positive deviations from this growth curve
with individual size. Individuals would bene¢t by would be penalised, and under some circumstances,
being able to vary growth rates with some indepen- growth rates lower than the‘normal’, i.e. submaximal
dence from their environment (Case 1978; Sibly and growth rates, should be expected. To our knowledge,
Calow 1986; Metcalfe et al. 2002). This suggests that no studies have determined yet whether or not this
compensatory growth itself might be adaptive. negative compensation can occur. Experimental
Compensatory growth might just re£ect the exis- research in this line is currently being developed
tence of trade-o¡s betweengrowthand mortalityrates (A.G. Nicieza and D. Alvarez, unpublished data).
and the state-dependent £exibility of growth. Some
theoretical studies have stressed the importance of
Future directions for research
recognising endogenously controlled variation in
growth rates (Abrams et al. 1996), which adjusts The study of compensatory growth as de¢ned in this
growth trajectories to modi¢cations in existing condi- review has been almost entirely based on laboratory
tions (e.g. photoperiod, predation risk, competition). experiments. Little is known on the incidence, conse-
The type of endogenous control promoting compensa- quences and importance of compensatory growth in
tory growth as discussed here is much more subtle. It natural populations. Only two studies have consid-
has not been explicitly considered in these theore- ered geographical variation in the compensatory
tical studies, despite their incorporation of time con- growth response (Purchase and Brown 2001; Schultz
straints, which can be inherently associated with et al. 2002), and these lead to di¡erent conclusions.
individual state (for example, when certain sizes or Populations do show compensatory responses in
energylevels must be achieved bycertaintimes). growth rates, but these do not necessarily depend on
Some investigators have made a distinction the form of growth compensation discussed here. In
between ‘nonreal’ and ‘real’ growth compensation many cases, they may simply represent a response to
(see Zivkov 1982, 1996; Zivkov et al. 1999). The ¢rst improved food availability, and not involve rates

# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190 179


Compensatory growth in ¢shes M Ali et al.

greater the usual long-term rates. Most of the The taxonomic coverage of studies dealing with
reported information is based on indirect methods compensatory growth is still limited. Comparative
like the back-calculation of growth from scales or studies involving closely related species di¡ering in
otoliths, and there is a bias towards species relevant ecological or life-historical characteristics can con-
to commercial ¢sheries (Beacham 1981; Moores and tribute to an understanding of the functional signi¢-
Winters 1982; Nicieza and Bran‹a 1993a,b; Haweet cance of compensatory growth processes. Even
and Ozawa 1996). This does not provide conclusive groups that have attracted maximum attention have
evidence for the prevalence of true compensatory important taxonomic and ecological gaps. Among
responses. Studies monitoring the growth of indivi- salmonids, some widely distributed species like
dually marked ¢sh under manipulated and nonmani- brown trout have attracted little interest (for an
pulated conditions could provide considerable exception, see Pirhonen and Forsman 1998). In that
information on the extent and costs of compensatory case, it is not possible to determine whether the fail-
growth in the wild (Letcher and Terrick 2001). ure to demonstrate compensatory growth re£ects a
The consequences of ignoring compensatory true absence or simply lack of information.
growth for general studies of ¢sh growth have There is a need to develop experimental protocols
received little attention. For example, the existence that can be used to compare the capacity for compen-
of compensatory growth can put at risk the suitabil- satory growth for di¡erent ontogenetic stages within
ity of usual condition indices (Raikova-Petrova and species and between species. Imposing ¢xed periods
Zivkov 1998), growth back-calculation methods of deprivation may lead to di¡erent degrees of growth
(Zivkov1996) and usual growth parameters for com- depression. The alternative is to set a target change
paring ¢sh growth (Zivkov et al.1999).Vertical studies in body mass before re-alimentation is imposed
comparing older and younger individuals make up a (Schwarz et al. 1985). The protocols should de¢ne the
body of evidence for optimising selection in natural range of growth depression that evokes growth com-
populations. Yet, the restrictive assumption that pensation and the degree of compensation that is
growth is not compensatory must be met for such achieved. The need for comparative studies extends
methods to be valid (Travis 1989). The ecological to the di¡erent factors provoking growth losses that
implications cannot be ignored. The indirect e¡ects can be reversed. Most studies on compensatory
of growth compensation are simply not known. growth in ¢sh have examined responses to changes
Reinhardt and Healey (1998) have speculated that in food availability, with only a few studies on the
the presence of predators may serve to depress e¡ects of food quality, temperature or ¢sh density.
growth rates of the largest individuals in a cohort No studies have tested whether ¢sh compensate
through risk-avoidance behaviour, which indirectly growth losses imposed through predator avoidance,
would favour better feeding opportunities of the although research is being conducted to test this idea
smallest ¢sh, less responsive to predator presence. (D. Alvarez and A.G. Nicieza, unpublished data).
Most of the studies on compensatory growth in ¢sh Whichever mechanisms operate, compensatory
are focused on overall size (generally length or mass), growth represents a temporal, within-individual
whereas there is little understanding of how the com- level of phenotypic plasticityand a source of variation
pensatory growth processes can a¡ect morphology that is often overlooked in comparative studies and
(Bell 1974; Emerson 1986). This is a major gap in our theoretical analysis of growth and developmental
knowledge because ¢sh morphology can in£uence strategies (Nicieza and Metcalfe 1997). If shown to be
swimming performance (Hawkins and Quinn 1996; a widespread response in natural populations, it
Webb et al. 1996), feeding e⁄ciency (Wainwright could have important implications for the dynamics
1987; Norton 1991; Wainwright and Richard 1995) of populations and hence communities because of
and escape from predators (Taylor and McPhail 1985; its role in bu¡ering populations against the e¡ects of
Swain 1992). At least, two simple, but important, environmental variations (Wootton 1998). Targeted
questions still deserve investigation: (i) Do all body growth also has important implications at the evolu-
dimensions maintain their relative growth over tionary level. First, it reduces or eliminates some of
the compensatory period (i.e. does compensatory the variance that would otherwise be available
growth entail morphological costs)? (ii) Can environ- for selection acting on ‘¢nal’ size (Riska et al. 1984).
mentally induced deviations from the ‘normal’shape Second, it could reduce serial genetic autocorrela-
be corrected by di¡erential, compensating growth of tion, allowing di¡erent segments of development
the di¡erent body parts? and growth to respond more independently to

180 # 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190


Compensatory growth in ¢shes M Ali et al.

age-speci¢c selection pressures (Riska 1986). Even if Aranda, A., Sanchez-Vazquez, F.J. and Madrid, J.A. (2001)
it proves to have only a minor role in natural popula- E¡ect of short-term fasting on macronutrient self-selec-
tions, compensatory changes in appetite and growth tion in sea bass. Physiology and Behaviour 73,105^109.
Aranyakananda, P., Moore, N. and Singhagraiwan,T. (1996)
may be important in developing optimal patterns of
E¡ect of feeding frequency on compensatory growth of
feeding in aquaculture facilities (Gaylord and Gatlin
Asian sea bass, Lates calcarifer. Arri Newsletter 3,11^13.
2001).
Arendt, J.D. (1997) Adaptive intrinsic growth rates: An inte-
gration across taxa. Quarterly Review of Biology 72,
Acknowledgements 149^177.
Arendt, J.D. andWilson, D.S. (1997) Optimistic growth: com-
M. Ali was supported by a scholarship from the Gov- petition and an ontogenetic niche-shift select for rapid
ernment of Pakistan. R. J.Wootton acknowledges the growth in pumpkinseed sun¢sh (Lepomis gibbosus).
support of the Royal Society of London ^ Chinese Evolution 51,1946^1954.
Academy of Sciences Exchange award.We would like Arendt, J.D. and Wilson, D.S. (2000) Population di¡erences
to thank the late Prof. Yibo Cui for his comments on in the onset of cranial ossi¢cation in pumpkinseed (Lepo-
an earlier draft of this review and the constructive mis gibbosus), a potential cost of growth. Canadian Journal
comments of two anonymous referees. of Fisheries and Aquatic Sciences 57,351^356.
Arendt, J.D.,Wilson, D.S. and Starck, E. (2001) Scale strength
as a cost of rapid growth in sun¢sh. Oikos 93,95^100.
References
Arndt, S.K.A., Benfy, T.J. and Cunjak, R.A. (1996) E¡ect of
Abrahams, M.V. and Sutterlin, A. (1999) The foraging and temporary reductions in feeding on protein synthesis
antipredator behaviour of growth-enhanced transgenic and energy storage of juvenile Atlantic salmon. Journal
Atlantic salmon. Animal Behaviour 58,933^942. of Fish Biology 49, 257^276.
Abrams, P.A., Leimar, O., Nylin, S. and Wiklund, C. (1996) Atchley, R.W. (1984) Ontogeny, timing of development, and
The e¡ect of £exible growth rates on optimal sizes and genetic variance-covariance structure. American Natur-
development times in a seasonal environment. American alist123,519^540.
Naturalist147,381^395. Basiao, Z.U., Doyle, R.W. and Arago, A.L. (1996) A statistical
Adams, S.M., McLean, R.B. and Hu¡man, M.M. (1982) Struc- power analysis of the ‘internal reference’ technique for
turing of a predator population through temperature- comparing growth and growth depensation of tilapia
mediated e¡ects on prey availability. Canadian Journal of strains. Journal of Fish Biology 49, 277^286.
Fisheries and Aquatic Sciences 39,1175^1184. Bastrop, R., Spangenberg, R. and Ju«rss, K. (1991) Biochem-
Ali, M. and Wootton, R.J. (1998) Do random £uctuations in ical adaptations of juvenile carp (Cyprinus carpio L.) to
the intervals between feeding a¡ect growth rate in juve- food deprivation. Comparative Biochemistry and Physiol-
nile three-spined sticklebacks? Journal of Fish Biology ogy 98,143^149.
53,1006^1014. Beacham,T.D. (1981) Variability in growth during the ¢rst 3
Ali, M. and Wootton, R.J. (1999) Coping with resource varia- years of life of cod (Gadus morhua) in the southern Gulf
bility: e¡ect of constant and variable intervals between of St. Lawrence. CanadianJournal of Zoology 59,614^620.
feeding on reproductive performance at ¢rst spawning Beamish, F.W.H. (1964) In£uence of starvation on standard
of female three-spined sticklebacks, Gasterosteus aculea- and routine oxygen consumption. Transactions of the
tus L. Journal of Fish Biology 55, 211^220. American Fisheries Society 50,153^164.
Ali, M. and Wootton, R.J. (2001) Capacity for growth com- Bell, G. (1974) The reduction of morphological variation in
pensation in juvenile three-spined sticklebacks ex- natural populations of smooth newt larvae. Journal of
periencing food deprivation. Journal of Fish Biology 58, Animal Ecology 43,115^128.
1531^1544. Bellamy, D. (1968) Metabolism of the red piranha (Roosevel-
Ali, M., Przybylski, M. and Wootton, R.J. (1998) Do random tiella nattererri) in relation to feeding behaviour. Com-
£uctuations in daily ration a¡ect the growth rate of juve- parative Biochemistry and Physiology 25,343^347.
nile three-spined sticklebacks? Journal of Fish Biology von Bertalan¡y, L. (1960) Principles and theory of growth.
52, 223^229. In: Fundamental Aspects of Normal and Malignant Growth
Ali, M., Cui,Y., Zhu, X. and Wootton, R.J. (2001) Dynamics of (ed W. Nowinski). Elsevier, Amsterdam, pp.137^252.
appetite in three ¢sh species (Gasterosteus aculeatus, Bertram, D.F., Chambers, R.C. and Leggett,W.C. (1993) Nega-
Phoxinus phoxinus and Carassius auratus gibelio) after tive correlations between larval and juvenile growth-
feed deprivation. Aquaculture Research 32, 443^450. rates in winter £ounder- implications of compensatory
A¤lvarez, D. (2002) Implicaciones del comportamiento y ¢siolo- growth for variation in size-at-age. Marine Ecology Pro-
g|¤ a individuales en el ciclo de vida y dina¤mica poblacional de gress in Series 96, 209^215.
la trucha comu¤n (Salmo trutta L.). Doctoral Thesis, Uni- Billerbeck, J.M., Lankford,T.E. Jr and Conover, D. (2001) Evo-
versity of Oviedo, Spain,109pages. lution of intrinsic growth and energy acquisition rates.

# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190 181


Compensatory growth in ¢shes M Ali et al.

Part I. Trade-o¡s with swimming performance in Meni- Buckley, L.J. (1979) Relationships between RNA-DNA ratio,
dia menidia. Evolution 55,1863^1872. prey density, and growth rate in Atlantic cod (Gadus mor-
Bilton, H.T. and Robins, G.L. (1973) The e¡ects of starvation hua) larvae. Journal of Fisheries Research Board of Canada
and subsequent feeding on survival and growth of Fulton 36,1497^1502.
Channel Sockeye Salmon fry (Oncorhynchus nerka). Jour- Bull, C.D. and Metcalfe, N.B. (1997) Regulation of hyperpha-
nal of the Fisheries Research Board of Canada 30,1^5. gia in response to varying energy de¢cits in over-winter-
Bilton, H.T., Alderdice, D.F. and Schnute, J.T. (1982) In£uence ing juvenile Atlantic salmon. Journal of Fish Biology 50,
of time and size at release of juvenile coho salmon 498^510.
(Oncorhynchus kisutch) on returns at maturity. Canadian Bull, C.D., Metcalfe, N.B. and Mangel, M. (1996) Seasonal
Journal of Fisheries and Aquatic Sciences 39, 426^447. matching of foraging e¡ort to anticipated energy require-
Bjo«rnsson, B., Sigurthorsson, G., Hemre, G.I. and Lie, O. ments in anorexic juvenile salmon. Proceedings of the
(1992) Growth rate and feed conversion factor on young Royal Society of London B 263,13^18.
halibut (Hippoglossus hippoglossus L.) fed six di¡erent Carter, C., Houlihan, D., Kiessling, A., Medalo, F. and Jobling,
diets. Fiskeridirektoratets Skrifter Ernaering 5, 25^35. M. (2001) Physiological e¡ects of feeding. In: Feed Intake
Blasco, J., Fernandez, J. and Gutierrez, J. (1992a) Variations in in Fish (eds D. Houlihan, T. Boujard, T. and M. Jobling),
tissue reserves, plasma metabolites and pancreatic hor- Blackwell Scienti¢c, Oxford, pp. 297^331.
mones during fasting in immature carp (Cyprinus carpio). Case,T.J. (1978) On the evolution and adaptive signi¢cance of
Comparative Biochemistry and Physiology A-Physiology postnatal growth rates in terrestrial vertebrates. Quar-
103,357^363. terly Review of Biology 53, 243^279.
Blasco, J., Fernandez, J. and Gutierrez, J. (1992b) Fasting and Chambers, R.C. Leggett, W.C. and Brown, J.A. (1988) Varia-
refeeding in carp, Cyprinus carpio L. ^ the mobilization of tion in and among early life history traits of laboratory-
reserves plasma metabolites and hormone variations. reared winter £ounder Pseudopleuronectes americanus.
Journal of Comparative Physiology B-Biochemical Systemic Marine Ecology Progress in Series 47,1^15.
and Environmental Physiology162,539^546. Chatakondi, N.G. and Yant, R.D. (2001) Application of com-
Blaxter, J.H.S. and Ehrlich, K.F. (1974) Change in behaviour pensatory growth to enhance production in channel cat-
during starvation of herring and plaice larvae. In: The ¢sh Ictalurus punctatus. Journal of World Aquaculture
Early Life History of Fish (ed J.H.S. Blaxter), Springer Society of 32, 278^285.
Verlag, Heidelberg, pp.575^588. Chavin, W. and Young, J.E. (1970) Factors in the determina-
Blaxter, J.H.S. and Hempel, G. (1963) The in£uence of egg tion of normal serum glucose levels of gold¢sh, Carass-
size on herring larvae (Clupea harengus L.). Journal du Con- sius auratus L. Comparative Biochemistry and Physiology
seil international pour l’Exploration de la Mer 28, 211^240. 33,629^653.
Boujard, T., Burel, C., Medale, F., Haylor, G. and Moisan, A. Chmilevskii, D.A. (1994) Vliyanie ponizhennoj temperatury
(2000) E¡ect of past nutritional history and fasting on na on oogenez tilapii Oreochromis mossambicus 2. Vozde-
feed intake and growth in rainbow trout Oncorhynchus jstvie na ryb v vozraste dvadtsati dvukh sutok posle vylu-
mykiss. Aquatic Living Resources13,129^137. pleniya.Voprosy Ikhtiologii 34,675^680. [In Russian].
Brett, J.M. (1971) Satiation time, appetite and maximum food Chmilevskii, D.A. (1996) The in£uence of a low temperature
intake of sockeye salmon (Oncorhynchus nerka). Journal on the oogenesis of tilapia Oreochromis mossambicus 4.
of Fisheries Research Board of Canada 28, 409^415. E¡ect on ¢sh 106 days after their hatching. Journal of
Broekhuizen, N., Gurney,W.S.C., Jones, A. and Bryant, A.D. Ichthyology 36,615^620.
(1994) Modelling compensatory growth. Functional Ecol- Chmilevskii, D.A. (1998) The in£uence of low temperature
ogy 8,770^782. on the growth of Oreochromis mossambicus. Journal of
Brown, M.E. (1946) The growth of brown trout (Salmo trutta Ichthyology 38,86^92.
Linn.) II The growth of two-year-old trout at constant Cho, C.Y. Slinger, S.J. and Bayley, H.S. (1976) In£uence of level
temperature of 11.5 8C. Journal of Experimental Biology and type of dietary protein and of level of feeding on feed
22,130^144. utilization by rainbow trout. Journal of Nutrition 106,
Brown, M.E. (1957) Experimental studies on growth. In:The 1547^1556.
Physiology of Fishes, Vol. I (ed. M.E. Brown), Academic Christensen, S.M. and McLean, E. (1998) Compensatory
Press, NewYork, pp.361^400. growth in Mozambique tilapia (Oreochromis mossambi-
Brown, C.R. and Cameron, J.N. (1991) The relationship cus) fed a sub-optimal diet. Ribarstvo 56,3^19.
between speci¢c dynamic action (SDA) and protein Christiansen, J.S. Martinez, I. Jobling, M. and Amin, A.B.
synthesis rates in the channel cat¢sh. Physiological Zool- (1992) Rapid somatic growth and muscle damage in a
ogy 64, 298^309. salmonid ¢sh. Basic and Applied Myology 2, 235^239.
Buckel, J.A., Letcher, B.H. and Conover, D.O. (1998) E¡ects Collins, A.L. and Anderson, T.A. (1995) The regulation of
of a delayed onset of piscivory on the size of age-0 blue- endogenous energy stores during starvation and refeed-
¢sh. Transactions of American Fisheries Society 127, ing in the somatic tissues of the golden perch. Journal of
576^587. Fish Biology 47,1004^1015.

182 # 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190


Compensatory growth in ¢shes M Ali et al.

Conover, D.O. (1992) Seasonality and the scheduling of life- Egglishaw, H.J. and Shackley, P.E. (1977) Growth, survival
history at di¡erent latitudes. Journal of Fish Biology 41, and production of juvenile salmon and trout in a Scottish
161^178. stream. Journal of Fish Biology11,647^672.
Conover, D.O. and Ross, M.R. (1982) Patterns in seasonal Elliott, J.M. (1972) Rates of gastric evacuation in brown trout,
abundance, growth and biomass of the Atlantic silver- Salmo trutta L. Freshwater Biology 2,1^18.
side, Menidia menidia, in a New England estuary. Elliott, J.M. (1975a) Weight of food and time required to sati-
Estuaries 5, 275^286. ate brown trout, Salmo trutta L. Freshwater Biology 5,
Cui, Y. and Wootton, R.J. (1989) Bioenergetics of growth of 51^64.
a cyprinid, Phoxinus phoxinus (L.): development and Elliott, J.M. (1975b) Number of meals in a day, maximum
testing of growth model. Journal of Fish Biology 34, weight of food consumed in a day and maximum rate of
47^64. feeding for brown trout Salmo trutta. Freshwater Biology
Cutts, C.J. Metcalfe, N.B. and Taylor, A.C. (2002) Juvenile 5, 287^303.
Atlantic salmon (Salmo salar) with relatively high stan- Elliott, J.M. (1976) Energy losses in the waste products of
dard metabolic rates have smal metabolic scopes. Func- brown trout (Salmo trutta L.). Journal of Animal Ecology
tional Ecology16,74^78. 45,561^580.
Dabrowski, K.Takashima, F. and Strussmann, C. (1986) Does Emerson, S.B. (1986) Heterochrony and frogs: the relation-
recovery growth occur in larval ¢sh Bulletin of the of ship of a life history trait to morphological form. Ameri-
Japanese Society of Science 52,1869. can Naturalist127,163^183.
DamsgQrd, B. and Arnesen, A.M. (1998) Feeding, growth Farbridge, K.J. and Leatherland, J.F. (1992a) Plasma growth-
and social interactions during smolting and seawater hormone levels in fed fasted rainbow trout (Oncor-
acclimation in Atlantic salmon, Salmo salar L. Aquacul- hynchus mykiss) are decreased following handling stress.
ture168,7^16. Fish Physiology and Biochemistry10,67^73.
DamsgQrd, B. and Dill, L.M. (1998) Risk-taking behaviour in Farbridge, K.J. and Leatherland, J.F. (1992b) Temporal
weight compensating coho salmon, Oncorhynchus changes in plasma thyroid-hormone, growth-hormone
kisutch. Behavioural Ecology 9, 26^32. and free fatty-acid concentrations, and hepatic 50 -mono-
DePedro, N. and Bjornsson, B.T. (2001) Regulation of food deiodinase on refeeding in rainbow trout (Oncorhynchus
intake by neuropeptides and hormones. In: Feed Intake mykiss). Fish Physiology and Biochemistry10, 245^257.
in Fish (eds D. Houlihan,T. Boujard and M. Jobling), Black- Farbridge, K.J. Flett, P.A. and Leatherland, J.F. (1992)Temporal
well Scienti¢c, Oxford, pp. 297^331. e¡ects of restricted diet and compensatory increased diet-
van Dijk, P.L.M. Staaks, G. and Hardewig, I. (2002) The e¡ect ary intake on thyroid-function, plasma growth-hormone
of fasting and refeeding on temperature preference, levels and tissue lipid reserves of rainbow trout (Oncor-
activity and growth of roach, Rutilus rutilus. Oecologia hynchus mykiss). Aquaculture104,157^174.
130, 496^504. Farrell, A.P. Bennett, W.B. and Devlin, R.H. (1997) Growth-
Dobson, S.H. and Holmes, R.M. (1984) Compensatory enhanced transgenic salmon can be inferior swimmers.
growth in the rainbow trout, Salmo gairdneri Richard- CanadianJournal of Zoology 75,335^337.
son. Journal of Fish Biology 25,649^656. Ferreri, C.P. and Taylor,W.W. (1996) Compensation in indivi-
Du Preez, H.H. (1987) Laboratory studies on the oxygen con- dual growth-rates and its in£uence on lake trout popula-
sumption of the marine teleost, Lichia amia (Linnaeus, tion-dynamics in the Michigan waters of Lake Superior.
1758). Comparative Biochemistry and Physiology 88A, Journal of Fish Biology 49,763^777.
523^532. Fletcher, D.A. (1984) The physiological control of appetite
Du Preez, H.H. McLachlan, A. and Marais, J.K.F. (1986a) Oxy- in ¢sh. Comparative Biochemistry and Physiology 78A,
gen consumption of a shallow water teleost, the spotted 617^628.
grunter, Pomadasys commersonii (Lace¤pe'de, 1802). Com- Foss, A. and Imsland, A.K. (2002) Compensatory growth in
parative Biochemistry and Physiology 84A,61^70. the spotted wol⁄sh Anarhichas minor (Olafsen) after a
Du Preez, H.H. Strydom,W. and Winter, P.E.D. (1986b) Oxy- period of limited oxygen supply. Aquaculture Research
gen consumption of two teleosts Lithognathus mormyrus 33,1097^1101.
(Linnaeus, 1758) and Lithognathus lithognathus (Cuvier, From, J. and Rasmussen, G. (1984) A growth model, gastric
1830) (Teleosti: Sparidae). Comparative Biochemistry and evacuation and body composition in rainbow trout,
Physiology 85A,313^331. Salmo gairdneri Richardson,1836. Dana 3,61^139.
Dunbrack, R.L. (1988) Feeding of juvenile coho salmon Gardiner, W.R. and Geddes, P. (1980) The in£uence of body
(Oncorhynchus kisutch): maximum appetite, sustained composition on the survival of juvenile salmon. Hydro-
feeding rate, appetite return, and body size. Canadian biologia 69,67^72.
Journal of Fisheries and Aquatic Sciences 45,1191^1196. Gaylord,T.G. and Gatlin,D.M. III (2001) Dietary protein and
Dunbrack, R.L. and Dill, L.M. (1983) A model of size depen- energy modi¢cations to maximize compensatory
dent surface feeding in a stream dwelling salmonid. growth of channel cat¢sh (Ictalurus punctatus). Aquacul-
Environmental Biology of Fishes 8, 203^216. ture194,337^348.

# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190 183


Compensatory growth in ¢shes M Ali et al.

Gaylord, T.G. MacKenzie, D.S. and Gatlin,D.M. III (2001) Henderson, M.A. and Cass, A.J. (1991) E¡ect of smolt size on
Growth performance, body composition and plasma smolt-to-adult survival for Chilko Lake sockeye salmon
thyroid hormone status of channel cat¢sh (Ictalurus (Oncorhynchus nerka). Canadian Journal of Fisheries and
punctatus) in response to short-term feed deprivation Aquatic Sciences 48,988^994.
and refeeding. Fish Physiology and Biochemistry 24, Henderson, P.A., Holmes, R.H.A. and Bamber, R.N. (1988)
73^79. Size-selective overwintering mortality in the sand smelt,
Gerking, S.D. (1994) Feeding Ecology of Fish. Academic Press, Atherina boyeri Risso, and its role in population regula-
London. tion. Journal of Fish Biology 33, 221^233.
Gerking, S.D. and Raush, R.R. (1979) Relative importance of Hepher, B. Liao, I.C. Cheng, S.H. and Hsieh, C.S. (1983) Food
size and chronological age in the life programme of utilization by tilapia-e¡ect of diet composition, feeding
¢shes. Ergebnis Limnologie1979,181^194. level and temperature on utilization e⁄ciency for main-
Goddard, J.S. (1974) An x-ray investigation of the e¡ects of tenance and growth. Aquaculture 32, 255^275.
starvation and drugs on intestinal motility in the plaice, Hoebel, B.G. (1997) Neuroscience and appetitive behaviour
Pleuronectes platessa L. Icthyologica 6, 49^58. research: 25 years. Appetite 29,119^133.
Godin, J.-G.J. (1981) E¡ect of hunger on the daily pattern of Houlihan, D.F. Mathers, E.M. and Foster, A. (1993) Biochem-
feeding rates in juvenile pink salmon (Oncorhynchus gor- ical correlates of growth rate in ¢sh. In: Fish Ecophysiol-
buscha). Journal of Fish Biology19,63^71. ogy (eds J.C. Rankin and F.B. Jensen), Chapman & Hall,
Gorman, G.C. and Nielson, L.A. (1982) Piscivory by stocked London, pp. 45^71.
brown trout (Salmo trutta) and its impact on the non- Houston, A.I. and McNamara, J.M. (1999) Models of Adaptive
game ¢sh community of Bottom Creek, Virginia. Cana- Behaviour. Cambridge University Press, Cambridge.
dianJournal of Fisheries and Aquatic Sciences 39,862^869. Houston, A.I. McNamara, J.M. and Hutchinson, J.M.C. (1993)
Gross, M.R. (1987) The evolution of disdromy in ¢shes. Amer- General results concerning the trade-o¡ between gain-
ican Fisheries Society of Symposium1,14^25. ing energy and avoiding predation. Philosophical Transac-
Gross, M.R. Coleman, R.M. and McDowall, R.M. (1988)Aqua- tions of the Royal Society of London B 341,375^397.
tic productivity and the evolution of diadromous ¢sh Hubbell, S.P. (1971) Of sowbugs and systems: the ecological
migration. Science 239,1291^1293. bioenergetics of a terrestrial isopod. In: Systems Analysis
Grove, D.J. Lozoides, L. and Nott, J. (1978) Satiation amount, and Simulation in Ecology (ed. B.C. Patten), Academic
frequency of feeding and gastric emptying rate in Salmo Press, London, pp. 269^323.
gairdneri. Journal of Fish Biology12,507^516. Hunt, R.L. (1969) Overwinter survival of wild ¢ngerling
Grove, D.J. Mortezuma, M.A. Flett, H.R.J. Foot, J.S.Watson,T. brook trout in Lawrence Creek,Wisconsin. Journal of the
and Flowerdew, M.W. (1985) Gastric emptying and the Fisheries Research Board of Canada 26,1473^1483.
return of appetite in juvenile turbot, Scophthalmus maxi- Hutchings, J.A. (1994) Age-speci¢c and size-speci¢c costs of
mus L. fed on arti¢cial diets. Journal of Fish Biology 26, reproduction within population of brook trout Salvelinus
339^354. fontinalis. Oikos 70,12^20.
Haweet, A. and Ozawa, T. (1996) Age and growth of ribbon Ince, B.W. and Thorpe, A. (1976) The e¡ects of starvation and
¢sh Trichiurus japonicus in Kagoshima Bay, Japan. Fish- force-feeding on the metabolism of the Northern Pike,
eries Science 62,529^533. Esox lucius L. Journal of Fish Biology 8,79^88.
Hawkins, D.K. and Quinn, T.P. (1996) Critical swimming Jezierska, B. Hazel, J.R. and Gerking, S.D. (1982) Lipid mobili-
velocity and associated morphology of juvenile coastal zation during starvation in the rainbow trout, Salmo
cutthroat trout (Oncorhynchus clarki clarki), steelhead gairdneri Richardson, with attention to fatty-acids. Jour-
trout (Oncorhynchus mykiss), and their hybrids. Canadian nal of Fish Biology 21,681^692.
Journal of Fisheries and Aquatic Sciences 53,1487^1496. Jobling, M. (1980) E¡ects of starvation on proximate chemi-
Hayward, R.S. and Wang, N. (2001) Failure to induce over- cal composition and energy utilization of plaice, Pleuro-
compensation of growth in maturing yellow perch. Jour- nectes platessa L. Journal of Fish Biology17,325^334.
nal of Fish Biology 59,126^140. Jobling, M. (1982) Some observations on the e¡ects of feed-
Hayward, R.S. Noltie, D.B. and Wang, N. (1997) Use of com- ing frequency on the food intake and growth of plaice,
pensatory growth to double hybrid sun¢sh growth Pleuronectes platessa L. Journal of Fish Biology 20,
rates. Transactions of American Fisheries Society of 126, 431^444.
316^322. Jobling, M. (1993) Bioenergetics: feed intake and energy par-
Hayward, R.S.Wang, N. and Noltie, D.B. (2000) Group hold- titioning. In: Fish Ecophysiology (eds J.C. Rankin and F.B.
ing impedes compensatory growth of hybrid sun¢sh. Jensen), Chapman & Hall, London, pp.1^44.
Aquaculture183, 299^305. Jobling, M. (1994) Fish Bioenergetics. Chapman & Hall,
Healey, M.C. (1982) Timing and relative intensity of size London.
selective mortality of juvenile chum salmon (Oncor- Jobling, M. and Johansen, S.J.S. (1999) The lipostat, hyper-
hynchus keta) during early sea life. Canadian Journal of phagia and catch-up growth. Aquaculture Research 30,
Fisheries and Aquatic Sciences 39,952^957. 473^478.

184 # 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190


Compensatory growth in ¢shes M Ali et al.

Jobling, M. and Koskela, J. (1996) Interindividual variations Karlsen, O. Holm, J.C. and Kjesbu, O.S. (1995) E¡ects of peri-
in feeding and growth in rainbow trout during restricted odic starvation on reproductive investment in 1st-time
feeding and in subsequent period of compensatory spawning Atlantic cod (Gadus morhua L). Aquaculture
growth. Journal of Fish Biology 49,658^667. 133,159^170.
Jobling, M. Jorgensen, E.H. and Siikavuopio, S.I. (1993) The Kim, M.K. and Lovell, R.T. (1995) E¡ect of feeding regimes on
in£uence of previous feeding regime on the compensa- compensatory weight gain and body tissue changes in
tory growth response of maturing and immature Arctic channel cat¢sh, Ictalurus punctatus in ponds. Aquacul-
charr, Salvelinus alpinus. Journal of Fish Biology 43, 409^ ture135, 285^293.
419. Kindschi, G.A. (1988) E¡ect of intermittent feeding on
Jobling, M. Meloy, O.H. dos Santos, J. and Christiansen, B. growth of rainbow trout, Salmo gairdneri Richardson.
(1994) The compensatory growth response of the Atlan- Aquaculture and Fisheries Management19, 213^215.
tic cod: e¡ects of nutritional history. Aquaculture Interna- Kinne, O. (1960) Growth, food intake and food conversion in
tional 2,75^90. an euryplastic ¢sh exposed to di¡erent temperatures
Johal, M.S. and Tandon, K.K. (1986) Phenomenon of growth and salinities. Physiological Zoology 33, 288^317.
compensation in Cyprinus carpio. IndianJournal ofEcology Kitchell, J.F. Stewart, D.J. and Weininger, D. (1977) Appli-
13,350^352. cations of bioenergetics model to yellow perch (Perca £a-
Johansen, S.J.S. Ekli, M. Stanges, B. and Jobling, M. (2001) vescens) and walleye (Stizostedion vitreum vitreum).
Weight gain and lipid deposition in Atlantic salmon, Journal of Fisheries Research Board of Canada 34, 1922^
Salmo salar, during compensatory growth: evidence for 1935.
lipostatic regulation? Aquaculture Research 32,963^974. Koppe,W. Pockrandt, J. Meyer-Burgdor¡, K.H. and Gunther,
Johansen, S.J.S. Ekli, M. and Jobling, M. (2002) Is there lipo- K.D. (1993) E¡ects of realimentation after a period of
static regulation of feed intake in Atlantic salmon Salmo restricted feeding on food intake, growth and body com-
salar L.? Aquaculture Research 33,515^524. position in Piaractus brachypomus (Cuvier 1818), a South
Johnson,T.B. and Evans, D.O. (1990) Size-dependent winter American characoid ¢sh. In: Fish Ecotoxicology and Eco-
mortality of young-of-the-year white perch: climate physiology (eds T. Braunbeck, W. Hanke and H. Segner),
warming and invasion of the Laurentian Great Lakes. VCH, NewYork, pp. 263^268.
Transactions of the American Fisheries Society 119, 301^ Lagardere, F. (1989) In£uence of feeding conditions and tem-
313. perature on growth rate and otolith increment deposi-
Johnson, T.B. and Evans, D.O. (1991) Behaviour, energetics, tion of larval Dover sole (Solea solea L.). Rapports Process:
and associated mortality of young-of-the-year white Verbaux Reunion Conseil International Pour l’Exploration
perch (Morone americana) and yellow perch (Perca £aves- de la Mer191,390^399.
cens) under simulated winter conditions. Canadian Jour- Lankford, T.E., Jr, Billerbeck, J.M. and Conover, D.O. (2001)
nal of Fisheries and Aquatic Scicences 48,672^680. Evolution of intrinsic growth and energy acquisition
Johnsson, J.I. (1993) Big and brave: size selection a¡ects fora- rates. II. Trade-o¡s with vulnerability to predation. Evo-
ging under risk of predation in juvenile rainbow trout, lution 55,1873^1881.
Oncorhynchus mykiss. Animal Behaviour 45,1219^1225. Larsson, A. and Lewander, K. (1973) Metabolic e¡ects of
Johnsson, J.I. and Bjo«rnsson, B.T. (1994) Growth hormone starvation in the eel, Anguilla anguilla L. Comparative
increases growth rate, appetite and dominance in juve- Biochemistry and Physiology 44A,367^374.
nile rainbow trout, Oncorhynchus mykiss. Animal Beha- Le Bail, P.-Y. and Boeuf, G. (1997) What hormones may
viour 47,177^186. regulate food intake in ¢sh? Aquatic Living Resources 10,
Johnsson, J.I. Petersson, E.J. Jo«nsson, E. Bjo«rnsson, B.T. and 371^379.
Ja«rvi,T. (1996) Domestication and growth hormone alter Le Bail, P.-Y. Perez-Sanchez, J. Yao, K. and Maisse, G. (1993)
antipredator behaviour and growth patterns in juvenile E¡ect of GH treatment on salmonid growth, study of the
brown trout, Salmo trutta. Canadian Journal of Fisheries variability of the response. In: Aquaculture, Fundamental
and Aquatic Sciences 53,1546^1554. and Applied Research (eds B. Lalhou and P.Vitiello), AGU,
Johnsson, J.I. Petersson, E.J. Jo«nsson, E. Ja«rvi,T. and Bjo«rns- Washington, DC, pp.137^197.
son, B.Th (1999) Growth hormone-induced e¡ects on Leamy, L. and Atchley, W. (1985) Directional selection and
mortality, energy status and growth: a ¢eld study on developmental stability: evidence from £uctuating
brown trout. Functional Ecology13,514^522. asymmetric of morphometric characters in rats. Growth
Johnston, I.A. (1981) Quantitative analysis of muscle break- 49,8^18.
down during starvation in the marine £at¢sh Pleuro- Leatherland, J.F. and Farbridge, K.J. (1992) Chronic fasting
nectes platessa. Cell andTissue Research 214,369^386. reduces the response of the thyroid to growth-hormone
Jo«nsson, E. Johnsson, J.I. and Bjo«rnsson, B.Th (1996) Growth and TSH, and alters the growth hormone-related
hormone increases predation exposure of rainbow changes in hepatic 50 - monodeiodenase activity in rain-
trout. Proceedings of the Royal Society of London B 263, bow trout, Oncorhynchus mykiss. General and Compara-
647^651. tive Endocrinology 87,342^353.

# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190 185


Compensatory growth in ¢shes M Ali et al.

Lee, D.J. and Putnam, C.B. (1973) The response of rainbow McGurk, M.D. (1986) Natural mortality of marine pelagic
trout to varying protein/energy ratios in a test diet. Jour- eggs and larvae: role of spatial patchiness. Marine Ecol-
nal of Nutrition103,916^922. ogy: Progress Series 34, 227^242.
Letcher, B.H. and Terrick,T. (2001) E¡ects of developmental McMillan, D.N. and Houlihan, D.F. (1992) Protein synthesis
stage at stocking on growth and survival of Atlantic sal- in trout liver is stimulated by both feeding and fasting.
mon fry. North American Journal of Fisheries Management Fish Physiology and Biochemistry10, 23^34.
21,102^110. McLean, E. Teskeredzic, E. Donaldson, E.M. et al. (1992)
Letcher, B.H. Rice, J.A. and Crowder, L.B. (1996) Size-depen- Accelerated growth of coho salmon Oncorhynchus
dent e¡ects of continuous and intermittent feeding on kisutch following sustained release of recombinant por-
starvation time and mass loss in starving yellow perch cine somatotropin. Aquaculture103,377^387.
larvae and juveniles. Transactions of the American Fish- Mendez, G. andWieser,W. (1993) Metabolic responses to food
eries Society125,14^26. deprivation and refeeding in juveniles of Rutilus rutilus
Litvak, M.K. and Leggett,W.C. (1992) Age and size-selective (Teleostei: Cyprinidae). Environmental Biology of Fishes
predation on larval ¢shes: the bigger-is-better hypothesis 36,73^81.
revisited. Marine Ecology: Progress Series 81,13^24. Metcalfe, N.B. and Monaghan, P. (2001) Compensation for a
Love, R.M. (1970) The Chemical Biology of Fishes. Academic bad start: grow now, pay later.Trends in Ecology andEvolu-
Press, London. tion16, 255^260.
Love, R.M. (1980) The Chemical Biology of Fishes, Vol. 2. Metcalfe, N.B. andThorpe, J.E. (1992) Anorexia and defended
Academic Press, London. energy levels in over-wintering juvenile salmon. Journal
Lovell, R.T. (1979) Factors a¡ecting voluntary food con- of Animal Ecology 61,175^181.
sumption of the channel cat¢sh. Proceedings of Annual Metcalfe, N.B., Bull, C.D. and Mangel, M. (2002) Seasonal
Conference S.E. Association of Fisheries andWildlife Agency variation in catch-up growth reveals state-dependent
33,563^571. somatic allocations in salmon. Evolutionary Ecology
Ludwig, D. and Rowe, L. (1990) Life-history strategies for Research 4,871^881.
energy gain and predator avoidance under time con- Miglavs, I. and Jobling, M. (1989a) The e¡ect of feeding
straints. American Naturalist135,686^707. regime on proximate body composition and patterns of
Lundqvist, H. McKinnell, S. Fa«ngstam, H. and Berglund, I. energy deposition in juvenileArctic charr, Salvelinus alpi-
(1994) The e¡ect of time, size and sex on recapture rates nus. Journal of Fish Biology 35,1^11.
and yield after river releases of Salmo salar smolts. Aqua- Miglavs, I. and Jobling, M. (1989b) E¡ects of feeding regime
culture121, 245^257. on food consumption, growth rates and tissue nucleic
Luquet, P. Oteme, Z.J. and Cisse, A. (1995) Evidence for com- acids in juvenile Arctic charr, Salvelinus alpinus, with
pensatory growth and its utility in the culture of Hetero- particular respect to compensatory growth. Journal of
branchus-longi¢lis. Aquatic Living Resources 8,389^394. Fish Biology 34,947^957.
Macchiusi, F. and Baker, R.L. (1991) Prey behaviour and size- Miller,T.J., Crowder, L.B., Rice, J.A. and Marshall, E.A. (1988)
selective predation by ¢sh. Freshwater Biology 25, Larval size and recruitment mechanisms in ¢shes:
533^538. toward a conceptual framework. Canadian Journal of
Machiels, M.A.M. and van Dam, A.A. (1987) A dynamic Fisheries and Aquatic Science 45,1657^1670.
simulation ^ model for growth of the African cat¢sh, Mommsen, T.P. (1998) Growth and metabolism. In: The
Clarias gariepinus (Burchell 1822). Part 3. The e¡ect of Physiology of Fishes, 2nd edn. (ed. D.H. Evans), CRC Press,
body composition on growth and feed intake. Aquacul- Boca Raton, pp.65^97.
ture 60,55^71. Monteiro, L.S. and Falconer, D.S. (1966) Compensatory
Mackas, D.L. Denman, K.L. and Abbott, M.R. (1985) Plank- growth and sexual maturity in mice. Animal Production
ton patchiness: biology in the physical vernacular. Bulle- 8,179^192.
tin of the of Marine Science 37,652^674. Moores, J.A. and Winters, G.H. (1982) Growth patterns in a
Maclean, A. and Metcalfe, N.B. (2001) Social status, access Newfoundland Atlantic herring (Clupea harengus haren-
to food and compensatory growth in juvenile Atlantic gus) stock. Canadian Journal of Fisheries and Aquatic
salmon. Journal of Fish Biology 58,1331^1346. Sciences 39, 454^461.
Magnusson, J.J. (1969) Digestion and food consumed by Morgan, I.J. and Metcalfe, N.B. (2001) Deferred costs of com-
skipjack tuna, Katsuwanus pelamis. Transactions of the pensatory growth after autumnal food shortage in juve-
American Fisheries Society 98,379^392. nile salmon. Proceedings of the Royal Society of London B
Marais, J.K.J. and Kissil, G.W. (1979) The in£uence of energy 268, 295^301.
level on the feed intake, growth, food conversion and Mortensen, A. and DamsgQrd, B. (1993) Compensatory
composition of Sparus aurata. Aquaculture17, 203^220. growth and weight segregation following light and tem-
McDowall, R.M. (2001) Anadromy and homing: two life-his- perature manipulation of juvenile Atlantic salmon
tory traits with adaptive synergies in salmonid ¢shes. (Salmo salar L.) and Arctic charr (Salvelinus alpinus L.).
Fish and Fisheries 2,78^85. Aquaculture114, 261^272.

186 # 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190


Compensatory growth in ¢shes M Ali et al.

Nagai, M. and Ikeda, S. (1971) Carbohydrate metabolism in Paul, A.J., Paul, J.M. and Smith, R.L. (1995) Compensatory
¢sh. I. E¡ects of starvation and dietary composition on growth in Alaska yellow¢n sole, Pleuronectes asper,
the blood glucose level and the hepatopancreatic glyco- following food deprivation. Journal of Fish Biology 46,
gen and lipid contents in carp. Bulletin of the of Japanese 442^448.
Society of Scienti¢c Fisheries 37, 404^409. Pedersen, B.H. (1993) Growth and mortality in young larval
Nicieza, A.G. (1993) Estrategias de desarrollo y reproduccio¤n en herring (Clupea harengus); e¡ect of repetitive changes in
el Salmo¤n Atla¤ntico Salmo salar L. PhD Thesis, University food availability. Marine Biology117,547^550.
of Oviedo, 206 pages. Pedersen, B.H., Ugelstad, I. and Hjelmeland, K. (1990) E¡ects
Nicieza, A.G. and Bran‹a, F. (1993a) Relationships among of transitory, low food supply in the early life of larval
smolt size, marine growth, and sea age at maturity of herring (Clupea harengus) on mortality, growth and
Atlantic salmon (Salmo salar) in Northern Spain. Cana- digestive capacity. Marine Biology107,61^66.
dianJournal of Fisheries and Aquatic Science 50,1632^1640. Perez-Sanchez, J. and Le Bail, P.-Y. (1999) Growth hormone
Nicieza, A.G. and Bran‹a, F. (1993b) Compensatory growth axis as a marker of nutritional status and growth per-
and optimum size in one-year-old smolts of Atlantic formance in ¢sh. Aquaculture177,117^128.
salmon (Salmo salar). In: Production of Juvenile Atlantic Peter, R.E. (1979) The brain and feeding behaviour. In: Fish
Salmon, Salmo salar, in Natural Waters (eds R.J. Gibson Physiology, Vol. VIII (eds W.S. Hoar, D.J. Randall and J.R.
and R.E. Cutting), Canadian Special Publications in Fish- Brett), Academic Press, London, pp.121^159.
eries and Aquatic Sciences No.118, 225^237. Pirhonen, J. and Forsman, L. (1998) E¡ect of prolonged feed
Nicieza, A.G. and Metcalfe, N.B. (1997) Growth compensa- restriction on size variation, feed consumption, body
tion in juvenile Atlantic salmon: Responses to depressed composition, growth and smolting of brown trout, Salmo
temperature and food availability. Ecology 78, 2385^ trutta. Aquaculture162, 203^217.
2400. Plank, J., Kirchgessner, M. and Schwarz, F.J. (1984) Ein£uss
Nicieza, A.G. and Metcalfe, N.B. (1999) Costs of rapid einer Futterrestriktion und ^realimentation auf Wach-
growth: the risk of aggression is higher for fast-growing stum, Futteraufwand und Schlachtkorperzusammenset-
salmon. Functional Ecology13,793^800. zung von Karpfen (Cyprinus carpio). Zeitschrift Fur
Nicieza, A.G., Reyes-Gavila¤n, F.G. and Bran‹a, F. (1994a) Dif- Tierphysiologie 53,138^149. [In German]
ferentiation in juvenile growth and bimodality patterns Post, J.R. and Evans, D.O. (1989) Size dependent overwinter-
between northern and southern-populations of Atlantic ing mortality of young-of-the-year yellow perch (Perca
salmon (Salmo salar L). Canadian Journal of Zoology 72, £avescens): laboratory, in situ enclosure, and ¢eld experi-
1603^1610. ments. Canadian Journal of Fisheries and Aquatic Sciences
Nicieza, A.G., Reiriz, L. and Bran‹a, F. (1994b) Variation in 46,1958^1968.
digestive performance between geographically disjunct Priede, I.G. (1985) Metabolic scope in ¢shes. In: Fish Ener-
populations of Atlantic salmon: countergradient in pas- getics New Perspectives (eds P. Tytler, and P. Calow),
sage time and digestion rate. Oecologia 99, 243^251. Croom-Helm, London, pp.33^64.
Norton, S.F. (1991) Capture success and diet of cottid ¢shes: Purchase, C.F. and Brown, J.A. (2001) Stock-speci¢c changes
the role of predator morphology and attack kinematics. in growth rates, food conversion e⁄ciencies, and energy
Ecology 72,1807^1819. allocation in response to temperature change in juvenile
Oliver, J.D., Holeton, G.F. and Chua, K.F. (1979) Overwinter Atlantic cod. Journal of Fish Biology 58,36^52.
mortality of ¢ngerling smallmouth bass in relation to Qian, X., Cui, Y., Xiong, B. and Yang, Y. (2000) Compen-
size, relative energy stores, and environmental tempera- satory growth, feed utilization and activity in gibel carp,
ture. Transactions of the American Fisheries Society 108, following feed deprivation. Journal of Fish Biology 56,
130^136. 228^2232.
Owen, R.W. (1989) Microscale and ¢nescale variations of Quinton, J.C. and Blake, R.W. (1990) The e¡ect of feed cycling
small plankton in coastal and pelagic environments. and ration level on the compensatory growth response
Journal of Marine Research 47,197^240. in rainbow trout, Oncorhynchus mykiss. Journal of Fish
Page, G.W. and Andrews, J.W. (1973) Interactions of dietary Biology 37,33^41.
levels of protein and energy on channel cat¢sh (Ictalurus Raikova-Petrova, G.N. and Zivkov, M.T. (1998) Growth self-
punctatus). Journal of Nutrition103,1339^1346. regulation: a reason for the variability of ¢sh condition
Pandian,T.J. (1967) Transformation of food in the ¢sh Mega- indices. International Revue of Hydrobiology 83,599^602.
lops cyprinoides. Part II. In£uence of quantity of food. Reimers, E., Kjorrefjord, G. and Stavostrand, M. (1993) Com-
Marine Biology1,107^109. pensatory growth and reduced maturation in second
Parker, G.A. (1992) The evolution of sexual size dimorphism sea winter farmed Atlantic salmon following starvation
in ¢sh. Journal of Fish Biology 41 (Suppl. B),1^20. in February and March. Journal of Fish Biology 43,
Parker, R.R. (1971) Size-selective predation among juvenile 805^810.
salmonid ¢shes in a British Columbia inlet. Journal of the Reinhardt, U.G. and Healey, M.C. (1998) Predation risk as
Fisheries Research Board of Canada 28,1503^1510. an opportunity for compensatory growth in juvenile

# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190 187


Compensatory growth in ¢shes M Ali et al.

coho salmon? North Paci¢c Anadromous Fish Commission after restricted feeding, in turbot Scophthalmus maximus
Bulletin1, 403. L. Aquaculture Research 30,647^653.
Ricker,W.E. (1969) E¡ects of size-selective mortality, produc- Sanchez-Vazquez, F.J., Yamamoto, T., Akiyama, T., Madrid,
tion and yield. Journal of the Fisheries Research Board of J.A. and Tabata, M. (1998) Selection of macronutrients
Canada 26, 479^541. by gold¢sh operating self-feeders. Physiology and Beha-
Ricker,W.E. (1975) Computation and interpretation of biolo- viour 65, 211^218.
gical statistics of ¢sh populations. Bulletin of the of the Sanchez-Vazquez, F.J., Yamamoto, T., Akiyama, T., Madrid,
Fisheries Research Board of Canada191,1^382. J.A. and Tabata, M. (1999) Macronutrient self-selection
Riska, B. (1986) Some models for development, growth, and through demand-feeders in rainbow trout. Physiology
morphometric correlation. Evolution 46,1303^1311. and Behaviour 66, 45^51.
Riska, B., Atchley, W.R. and Rutledge, J.J. (1984) A genetic- Satoh, S., Takeuchi, T. and Watanabe, T. (1984) Studies on
analysis of targeted growth in mice. Genetics107,79^101. nutritive-value of dietary lipid in ¢sh. 28. E¡ects of star-
Robinson, B.W. and Wardrop, S.L. (2002) Experimentally vation and environmental-temperature on proximate
manipulated growth rate in threespine sticklebacks: and fatty-acids compositions of Tilapia nilotica. Bulletin
assessing trade o¡s with developmental stability. Envir- of the of theJapanese Society of Scienti¢c Fisheries Research
onmental Biology of Fishes 63,67^78. 50,79^84.
Ro¡, D.A. (1992) The Evolution of Life Histories: Theory and Saunders, R.L., Farrell, A.P. and Knox, D.E. (1992) Pro-
Analysis. Chapman & Hall, London. gression of coronary arterial lesions in Atlantic
Rowe, D.K. and Thorpe, J.E. (1990) Suppression of matura- salmon (Salmo salar) as a function of growth rate. Cana-
tion in male Atlantic salmon (Salmo salar L.) parr by dian Journal of Fisheries and Aquatic Sciences 49,
reduction in feeding and growth during spring months. 878^884.
Aquaculture 86, 291^313. Schultz, E.T. and Conover, D.O. (1999) The allometry of
Rowe, D.K., Thorpe, J.E. and Shanks, A.M. (1991) Role of fat energy reserve depletion: test of a mechanism for size-
stores in the maturation of male Atlantic salmon (Salmo dependent winter mortality. Oecologia119, 474^483.
salar L.) parr. Canadian Journal of Fisheries and Aquatic Schultz, E.T., Lankford, T.E. and Conover, D.O. (2002) The
Sciences 48, 405^413. covariance of routine and compensatory juvenile
Rozin, P.N. and Mayer, J. (1961) Regulation of food intake growth rates over a seasonality gradient in a coastal ¢sh.
in the gold¢sh. American Journal of Physiology 201, Oecologia133,501^509.
968^974. Schwarz, F.J., Plank, J. and Kirchgessner, M. (1985) E¡ects of
Rueda, F.M., Martinez, F.J., Zamora, S., Kentouri, M. and protein or energy restriction with subsequent realimen-
Divanach, P. (1998) E¡ect of fasting and refeeding on tation on performance of carp (Cyprinus carpio L.). Aqua-
growth and body composition of red porgy, Pagrus pagrus culture 48, 23^33.
L. Aquaculture Research 29, 447^452. Sibly, R.M. and Calow, P. (1986) Physiological Ecology of
Ruohonen, K. and Ma«kinen,T. (1992) Validation of rainbow Animals. Blackwell Scienti¢c, Oxford.
trout (Oncorhynchus mykiss) growth model under Fin- Siems, D.P. and Sikes, R.S. (1998) Tradeo¡s between growth
nish circumstances. Aquaculture105,353^362. and reproduction in response to temporal variation in
Ruohonen, K., Kettunen, J. and King, J. (2001) Experimental food supply. Environmental Biology of Fishes 53,319^329.
design in feeding experiments. In: Feed Intake in Fish Silverstein, J.T. and Plisetskaya, E.M. (2000) The e¡ects of
(eds D. Houlihan, T. Boujard and M. Jobling), Blackwell NPY and insulin on food regulation in ¢sh. American
Scienti¢c, Oxford, pp. 88^107. Zoologist 40, 296^308.
Russell, N.R. (1991) The dynamics of appetite and growth con- Singh, R.P. and Srivastava, A.K. (1985) Satiation time,
trol in the minnow (Phoxinus phoxinus L.). PhD Thesis, gastric evacuation and appetite revival in Heteropneustes
University of Wales, 224 pages. fossilis (Block) (Siluriformes: Pisces). Aquaculture 49,
Russell, N.R. and Wootton, R.J. (1992) Appetite and growth 307^313.
compensation in the European minnow, Phoxinus phoxi- Skilbrei, O.T. (1990) Compensatory sea growth of male
nus (Cyprinidae) following short term of food restriction. Atlantic salmon, Salmo salar L. which previously mature
Environmental Biology of Fishes 34, 277^285. as parr. Journal of Fish Biology 37, 425^435.
Russell, N.R. and Wootton, R.J. (1993) Satiation, digestive Smith, R.R. (1987) Methods of controlling growth of steel-
tract evacuation and return of appetite in the European head. Progressive Fish Culturist 49, 248^252.
minnow, Phoxinus phoxinus (Cyprinidae) following short Sogard, S.M. (1997) Size-selective mortality in the juvenile
periods of pre-prandial starvation. EnvironmentalBiology stage of teleost ¢shes: a review. Bulletin of the of Marine
of Fishes 38,385^390. Science 60,1129^1157.
Ryan,W.J. (1990) Compensatory growth in cattle and sheep. Sogard, S.M. and Olla, B.L. (2002) Contrasts in the capacity
Nutritional Abstract Review of Series B 60,653^664. and underlying mechanisms for compensatory growth
Saether, B.-S. and Jobling, M. (1999) The e¡ects of ration level in two pelagic marine ¢shes. Marine Ecology: Progress
on feed intake and growth, and compensatory growth Series 243,165^177.

188 # 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190


Compensatory growth in ¢shes M Ali et al.

Solomon, D.J. and Bra¢eld, A.E. (1972) The energetics of feed- salar, in pumped seawater tanks by restricted food intake.
ing, metabolism and growth of perch (Perca £uviatilis L.). Aquaculture 86,315^326.
Journal of Animal Ecology 41,699^718. Toledo, M.M. (1996) Ciclos de vida y estrategias reproductivas
Speare, D.J. and Arsenault, G.J. (1997) E¡ects of intermittent de la trucha comu¤n (Salmo trutta L.). PhD Thesis, Univer-
hydrogen peroxide exposure on growth and columnaris sity of Oviedo, 203 pages.
disease prevention of juvenile rainbow trout (Oncor- Toneys, M.L. and Coble, D.W. (1979) Size-related, ¢rst winter
hynchus mykiss). CanadianJournal of Fisheries and Aquatic mortality of freshwater ¢shes. Transactions of the Ameri-
Science 54, 2653^2658. can Fisheries Society108, 415^419.
Stirling, H.P. (1976) E¡ect of experimental feeding and star- Travis, J. (1989) The role of optimizing selection in natural
vation on the proximate composition of the European populations. Annual Review of Ecology and Systematics
bass, Dicentrachus labrax. Marine Biology 34,85^91. 20, 279^296.
Swain, D.P. (1992) Selective predation for vertebral pheno- Tyler, A.V. and Dunn, R.S. (1976) Ration, growth and mea-
type in Gasterosteus aculeatus: reversal in the direction sures of somatic and organ condition in relation to the
of selection at di¡erent larval sizes. Evolution 46, 998^ meal frequency in winter £ounder, Pseudopleuronectes
1013. americanus, with hypotheses regarding population
Takagi, Y. (2001) E¡ects of starvation and subsequent re- homeostasis. Journal of Fisheries Research Board of Canada
feeding on formation and resorption of acellular bone 33,63^75.
in tilapia, Oreochromis niloticus. Zoological Science 18, Underwood, A.J. (1997) Experiments in Ecology ^ Their Logical
623^629. Design and Analysis Using Analysis Of Variance. Cam-
Talbot, C. (1985) Laboratory methods in ¢sh feeding and bridge University Press, Cambridge.
nutritional studies. In: Fish Energetics New Perspectives Vera-Cruz, E.M. and Mair, G.C. (1994) Conditions for e¡ective
(eds P. Tytler and P. Calow), London: Croom-Helm, pp. androgen sex reversal in Oreochromis niloticus (L.). Aqua-
125^154. culture122, 237^248.
Talbot, C., Higgins, P.J. and Shanks, A.M. (1984) E¡ects of Wainwright, P.C. (1987) Biomechanical limits to ecological
pre-and post-prandial starvation on meal size and eva- performance: molluscs-crushing by the Caribbean hog-
cuation rate of juvenile Atlantic salmon, Salmo salar L. ¢sh, Lachnolaimus maximus (Labridae). Journal of Zoology
Journal of Fish Biology 25,551^560. 213, 283^297.
Tandon, K.K. and Johal, M.S. (1983a) Study on age and Wainwright, P.C. and Richard, B.A. (1995) Predicting pat-
growth of Tor putitora (Hamilton) as evidenced by scales. terns of prey use from morphology of ¢shes. Environmen-
Indian Journal of Fisheries 30,171^175. tal Biology of Fishes 44,97^113.
Tandon, K.K. and Johal, M.S. (1983b) Occurrence of the phe- Wang, Y., Cui, Y., Yang, Y.X. and Cai, F.S. (2000) Compensa-
nomenon of growth compensation in Indian major tory growth in hybrid tilapia, Oreochromis mossambicus
carps. Indian Journal of Fisheries 30,180^182.  O. niloticus, reared in seawater. Aquaculture 189,
Tanner, J.M. (1963) Regulation of growth in size in mam- 101^108.
mals. Nature199,845^850. Weatherley, A.H. and Gill, H.S. (1981) Recovery growth fol-
Tashima, L. and Cahill, G.F. (1968) E¡ects of insulin in the lowing periods of restricted rations and starvation in
toad¢sh, Opsanus tau. General and Comparative Endocri- rainbow trout, Salmo gairdneri Richardson. Journal of
nology11, 262^271. Fish Biology18,195^208.
Taylor, E.B. and McPhail, J.D. (1985) Burst swimming and Weatherley, A.H. and Gill, H.S. (1987) The Biology of Fish
size-related predation of newly emerged coho samon Growth. Academic Press, London.
Oncorhynchus kisutch. Transactions of the American Fish- Webb, P.W., LaLiberte, G.D. and Schrank, A.J. (1996) Does
eries Society114,546^551. body and ¢n form a¡ect maneuverability of ¢sh traver-
Teskeredzic, Z., Teskeredzic, E., Tomec, M., Hacmanjek, M. sing vertical and horizontal slits? Environmental Biology
and Mclean, E. (1995) The impact of restricted rationing of Fishes 46,7^14.
upon growth food conversion e⁄ciency and body-com- Werner, E.E. and Anholt, B.R. (1993) Ecological conse-
position of rainbow-trout. Water Science and Technology quences of the trade-o¡ between growth and mortality
31, 219^223. rates mediated by foraging activity. American Naturalist
Thorpe, J.E. (1986) Age at ¢rst maturity in Atlantic Salmon, 142, 242^272.
Salmo salar: freshwater period in£uences and con£icts Westerman, M.E. and Holt, G.S. (1988) The RNA-DNA
with smolting. Canadian Special Publication of the of Fish- ratio: measurement of nucleic acids in larval Sciaenops
eries and Aquatic Science 89,7^14. ocellatus. Contributions in Marine Science 30 (Suppl.),
Thorpe, J.E., Miles, M.S. and Keay, D.S. (1984) Developmental 117^124.
rate, fecundity and egg size in Atlantic salmon, Salmo Whitledge, G.W., Hayward, R.S., Nolte, R.B. and Wang, N.
salar L. Aquaculture 43, 289^305. (1998) Testing bioenergetics models under feeding
Thorpe, J.E.,Talbot, C., Miles, M.S. and Keay, D.S. (1990) Con- regimes that elicit compensatory growth. Transactions of
trol of maturation in cultured Atlantic salmon, Salmo theAmerican Fisheries Society127,740^746.

# 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190 189


Compensatory growth in ¢shes M Ali et al.

Wieser, W. (1991) Limitations of energy acquisition and feed deprivation: temporal patterns in growth, nutrient
energy use in small poikilotherms: evolutionary implica- deposition, feed intake and body composition. Journal of
tions. Functional Ecology 5, 234^240. Fish Biology 58,999^1009.
Wieser,W., Krumschnabel, G. and Ojwang-Okwor, J.P. (1992) Zamal, H. and Ollevier, F. (1995) E¡ect of feeding and lack of
The energetics of starvation and growth after refeeding food on the growth, gross biochemical and fatty acid
in juveniles of three cyprinid species. EnvironmentalBiol- composition of juvenile cat¢sh. Journal of Fish Biology
ogy of Fishes 33,63^71. 46, 404^414.
Wilson, P.N. and Osbourn, D.F. (1960) Compensatory growth Zhu, X., Cui,Y., Ali, M. and Wootton, R.J. (2001) Comparison
after undernutrition in mammals and birds. Biological of compensatory growth responses of juvenile three-
Review 35,324^363. spined stickleback and minnow following similar food
Wootton, R.J. (1982) Environmental factors in ¢sh reproduc- deprivation protocols. Journal of Fish Biology 58, 1149^
tion. In: Reproductive Physiology of Fish (eds C.J.J. Richter 1165.
and H.J.T. Goos), Pudoc,Wageningen, pp. 201^219. Zhu, X.,Wu, L., Cui,Y.,Yang,Y. andWootton, R.J. (2003) Com-
Wootton, R.J. (1998) Ecology of Teleost Fishes, 2nd edn. pensatory growth in three-spined stickleback in relation
Kluwer, Dordrecht. to feed-deprivation protocols. Journal of Fish Biology 62,
Wootton, R.J., Allen, J.R.M. and Cole, S.J. (1980) E¡ect of body 195^205.
weight and temperature on the maximum daily food Zivkov, M.T. (1982) On the e¡ect and nature of growth com-
consumption of Gasterosteus aculeatus L. and Phoxinus pensation of ¢sh.Vestnik Ceskoslovenske Spolecnosti Zoolo-
phoxinus (L.): selecting an appropriate model. Journal of gicke 46,142^160.
Fish Biology17,695^705. Zivkov, M.T. (1996) Critique of proportional hypotheses and
Wu, L., Xie, S., Zhu, X., Cui,Y. and Wootton, R.J. (2002) Feed- methods for back-calculation of ¢sh growth. Environmen-
ing dynamics in ¢sh experiencing cycles of feed depriva- tal Biology of Fishes 46,309^320.
tion: a comparison of four species. Aquaculture Research Zivkov, M.T. and Raikova-Petrova, G.N. (1991) Growth com-
33, 481^489. pensation of pike-perch, Stizostedion lucioperca (L.), in
Wu, L., Xie, S., Cui, Y. and Wootton, R.J. (2003) E¡ects of the Batak and Ovcharitsa dams. Comptes Rendus de l’Aca-
cycles of feed deprivation on growth and food consump- damie Bulgare des Science 44, 45^58.
tion of immature three-spined sticklebacks and Eur- Zivkov, M.T., Trichkova, T.A. and Raikova-Petrova, G.N.
opean minnows. Journal of Fish Biology 62,184^194. (1999) Biological reasons for the unsuitability of growth
Xie, S., Zhu, X., Cui, Y., Lei, W., Yang, Y. and Wootton, R.J. parameters and indices for comparing ¢sh growth.
(2001) Compensatory growth in the gibel carp following Environmental Biology of Fishes 54,67^76.

190 # 2003 Blackwell Publishing Ltd, F I S H and F I S H E R I E S, 4, 147^190

Вам также может понравиться