You are on page 1of 12

Biological Journal of the Linnean Society (1996), 58: 471482

Conservation strategies: adaptation to stress


and the preservation of genetic diversity
P.A. PARSONS1
Faculty of Science and Technology, Griffith University, Nathan, Qld. 4111, Australia

Received 21 March 1995, accepted for publication 18 September 1995

The level of genetic diversity in free-living populations is not normally restrictive for conservation, since
it tends to be enhanced in stressed outlier populations. At the physiological level, this enhancement is
supported by the favouring of heterozygotes, especially when energy demands needed to adapt to stress
are high. Therefore ecophysiological considerations are important for conservation strategies, whereby
survival depends upon the metabolic potential of organisms to counter the energy cost of stress in their
environments. While abiotic stresses are primary, biotic stresses, in particular competition, can be
consolidated into this model as second-order effects. Irrespective of levels of genetic diversity, any species
can be incorporated into this approach to conservation. I therefore regard the monitoring of stress
response traits to be primary to the preservation of genetic diversity in developing conservation
strategies. In arriving at this conclusion, Fishers 1930 discussions of the environment and consequences
for adaptation, as presented in the The Genetical Theory of Natural Selection, play an initiating role.
1996 The Linnean Society of London

ADDITIONAL KEY WORDS: heterozygote advantage energy cost aridity sexual selection
metabolic potential competition fitness biodiversity.
CONTENTS
Introduction . . . . . . . . . . . . .
Heterozygote advantage, stress and sexual selection
Electrophoretic loci and stress . . . . . . .
Stress, metabolic potential and conservation . . .
Biotic stress, especially competition . . . . . .
Concluding comments . . . . . . . . . .
Acknowledgements . . . . . . . . . . .
References . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

471
473
475
476
477
479
480
480

INTRODUCTION

A key contribution to evolutionary biology in the long career of R.A. Fisher was
The Genetical Theory of Natural Selection published in 1930. This book had a long
gestation time, in excess of ten years. Sewall Wright, in reviewing the book for the
1
Present address: Department of Genetics and Human Variation, La Trobe University, Bundoora, Vic. 3083,
Australia. Correspondence, PO Box 906 Unley, SA 5061 Australia.
A version of this paper comprised the Third Sir Ronald Fisher Lecture, University of Adelaide, on 8th March,
1995.

00244066/96/080471 + 12 $18.00/0

471

1996 The Linnean Society of London

472

P. A. PARSONS

Journal of Heredity, commented that it is certain to take rank as one of the major
contributions to the theory of evolution, and J.B.S. Haldane said it laid the
foundations of a new branch of science (Bennett, 1983). However, we should be
warned, since in the Preface Fisher comments that No effort of mine could avail to
make the book easy reading. This is certainly true, but equally no book has had a
greater influence in the development of modern evolutionary biology. In this paper,
I will show this remains true with respect to his discussion of the environment and
consequences for adaptation, in relation to conservation strategies.
In his book Fisher (1930) wrote: If therefore an organism be really in any high
degree adapted to the place it fills in the environment, this adaptation will be
constantly menaced by any undirected agencies liable to cause changes to either
party in the adaptation.
A major objective of conservation biology is to accommodate the above menace,
and so preserve biological resources at orders of complexity ranging from genes to
ecosystems. Conservation biology rated little in the late 1920s when The Genetical
Theory was being written, yet in the Dover Edition published in 1958, Fisher
considered that owing to the rapid changes which man is making in his
environment, there will be special difficulties in determining and interpreting
consequential fitness changes.
In consequence, I look here at the major emphases of conservation biology in the
light of the constantly changing environments to which free-living populations are
normally exposed.
In contrast, a major topic in conservation biology concerns genetic problems faced
by endangered populations. There have been extensive and continuing discussions
on the maintenance of genetic variation to preserve evolutionary potential (e.g.
Frankel & Soule, 1981). Recent examples continue this trend (e.g. Prober & Brown,
1994; Raijmann et al., 1994). On the other hand, some species show extremely low
levels of genetic variation. A celebrated example is the endangered cheetah, Acinonyx
jubatus (OBrien, 1994), although other species of terrestrial carnivores have even
lower levels and are almost completely homozygous (Merola, 1994).
However, many conservation strategies are devised to avoid high levels of
homozygosity, because of the well-documented phenomenon of inbreeding depression, whereby fitness falls as inbreeding increases. A useful monitor of fitness is
fluctuating morphological asymmetry, FA, since it can be used comparatively across
taxa. FA tends to increase under genetic stress such as inbreeding in a typically
outbred species, and under environmental stresses such as extreme temperatures
which tend to occur at some species borders (e.g. Zakharov, 1989; Parsons, 1990).
Yet, FA measures appear not to support the hypothesis that the cheetah is suffering
genetic stress from inbreeding (e.g. Kieser & Groeneveld, 1991). This implies that the
largely homozygous genetic constitution of the cheetah does not threaten the survival
of this species (Merola, 1994). Extrapolating more widely, levels of genetic variation
appear to be species-specific, so that the level characteristic of a particular species will
reflect a balance between selection favouring heterozygosity and any historical events
depleting genetic variation. Furthermore, the heterozygosities of parents and
offspring are necessarily correlated (Mitton et al., 1993) which would help to maintain
this balance.
The exceedingly low genetic variation in cheetahs directs attention to causes other
than the depletion of genetic variation as the underlying cause of extinction. For
instance, ecological factors such as predation, the abandonment of cubs by mothers,

CONSERVATION STRATEGIES, STRESS AND DIVERSITY

473

and abiotic perturbations may play a more direct role in their conservation than does
low genetic variation (Caro & Laurenson, 1994). This appears to imply that genetic
variation may be the ultimate limiting factor determining whether free-living
populations can adapt to changing environments, but that ecological factors
including responses to stress may provide more proximate explanations.
Emphasizing ecological factors, Hoffmann & Parsons (1991) summarized a diffuse
literature on abiotic stress in living and fossil populations, and concluded that
responses to stress underlie much evolutionary change (and extinctions). Furthermore, inadequate nutrition is usual in free-living populations. Because of this,
White (1993) argues that animals normally struggle to exist in a hostile environment.
Consequently, many organisms are born but few survive due to a combination of
physical stresses, principally of climatic origin, interacting with and causing
nutritional stress.

HETEROZYGOTE ADVANTAGE, STRESS AND SEXUAL SELECTION

Under laboratory environments that are demonstrably extreme, heterozygotes


derived from inbred strains tend to be favoured with respect to fitness measures
(Parsons, 1959; Barnett & Coleman, 1960). Such experimental data imply that
heterozygote advantage should be a feature of polymorphisms in natural populations
under extreme environments (Parsons, 1971). Conversely, inbreeding depression
should be more pronounced in a stressful environment. For instance, in Drosophila
melanogaster, inbreeding effects can be increased when organisms are exposed to
stressful temperatures and intense competition (Miller, 1994). Similarly, survival
following heat shock declined with increasing inbreeding in D. buzzati (Dahlgaard,
Krebs & Loeschcke, 1995). Furthermore, in the white-footed mouse, Peromyscus
leucopus noveboracensis, inbreeding had an especially detrimental effect on the
survivorship of mice when released into nature (Jimenez et al., 1994). In the wild,
inbreeding depression was expressed during the abiotic stress of severe winter
population bottlenecks in song sparrows, Melospiza melodia (Keller et al., 1994).
Turning to plants, Van Treuren et al. (1993) and Latta & Ritland (1994)
summarized examples indicating that inbreeding depression is influenced by the
environments in which plants are grown. They concluded that it was accentuated in
stressful environments, for example field-sown progeny under competition, and in
habitats to which plants are poorly adapted (see also Schmitt & Gamble, 1990;
Montalvo, 1994). These results are consistent with earlier laboratory results showing
an enhancement of inbreeding depression for growth rate under temperature
extremes in Arabidopsis thaliana (Langridge, 1962) and Zea mays (McWilliam &
Griffing, 1965). Numerous additional studies of selection differentials in plant
populations discussed in a conservation context, provide support for the hypothesis
that viability selection favours heterozygotes, especially in small populations under
stress (Lesica & Allendorf, 1992). Further examples in plants and animals are
reviewed by Mitton (1993a).
In contrast, most laboratory experiments are carried out under relatively benign
circumstances, but in spite of these limitations, the level of heterozygosity of
individual organisms in populations tends to correlate with fitness measures, in
particular growth rate and developmental stability (Mitton & Grant, 1984). Enzyme
loci influencing metabolism and contributing to the amount of energy available for

474

P. A. PARSONS

development and growth, show the most significant positive associations with
heterozygosity (Mitton, 1993b). This is clearest under extreme conditions when the
energy demands from the environment would be highest. Since one direct effect of
stress is to increase the expenditure of metabolic energy (Odum, Finn & Franz,
1979), this suggests that heterozygotes may have the maximum genetic potential to
withstand the energy costs of any stress.
In the coot clam, Mulinia lateralis, there is a negative correlation between growth
rate and routine metabolic costs as the level of heterozygosity increases (Garton,
Koehn & Scott, 1984). In mussels, Mytilus edulis, heterozygotes have lower energy
requirements than homozygotes apparently because of greater efficiency in protein
synthesis, which releases energy for ingestion and absorption (Koehn & Bayne,
1989). These observations suggest that heterozygous individuals with a high level of
metabolic efficiency can maintain growth and reproduction under a wider range of
environmental conditions than individuals with a lower level of efficiency, because
they can continue to grow and reproduce as stress increases under resource
limitations. Examples of heterozygote advantage under energy stress include the
juvenile manure worm, Eisenia foetida, under limited food and low moisture (Diehl,
1988), Mulinia lateralis, under temperature and salinity stress (Scott & Koehn, 1990),
and the oldfield mouse, P. polionotus, under low food quality (Teska, Smith & Novak,
1990). Similarly, based on a recent literature review covering many additional
examples, Hawkins (1995) concluded that heterozygote advantage is environmentally dependent becoming clearest under conditions of food limitation or other
stresses.
Turning to stress from parasites, in feral rock doves, Columbia livia, lice reduced
feather and host body mass, and increased thermal conductance and metabolic rate
indicating an energy cost. This is exacerbated in a deteriorating abiotic environment
during winter (Booth, Clayton & Block, 1993). Similarly de Lope et al. (1993) found
that the ectoparasitic house martin bug, Oeciacius hirundinis, had larger negative effects
on the reproduction of its host, Delichon urbica, when nutritional conditions were poor
during the second compared with the first clutch in the season. In the lizard,
Sceloporus occidentalis, Schall & Sarni (1987) found that the time males spend in social
behaviours is reduced when infected with the malarial parasite, Plasmodium mexicanum,
and furthermore infected males perch more often in shade. Hence, the energy cost
from parasites reduces social behaviours and stressful microhabitats are avoided.
Under these stressful situations, selection should favour heterozygotes. For example,
a direct relationship between disease resistance and multilocus heterozygosity occurs
in the rainbow trout, Oncorhynchus mykiss, following exposure to bacterial gill disease,
a potentially lethal epizootic in freshwater fishes (Ferguson & Drahushchak, 1990).
If natural populations experience recurrent stresses, heterozygous advantage
should therefore maintain genetic variation. There are a number of additional outlier
examples, where the energy cost from stress is demonstrably extreme.
(1) In a population of Soay Sheep on Hirta, St. Kilda, Scotland, population
crashes occur periodically, primarily following winter starvation. At these times
of nutritional stress, heterozygotes at the adenosine deaminase locus had the
lowest parasite burden, and so were least likely to die (Gulland et al., 1993).
(2) Aridity stress plays a major role in adaptation of the mole rat, Spalax ehrenbergi,
in Israel and neighbouring locations. Average heterozygosity increases as
aridity stress and climatic unpredictability increase. In particular, in outlier

CONSERVATION STRATEGIES, STRESS AND DIVERSITY

475

steppe and desert habitats, where climates are most unpredictable and variable
than elsewhere in the distribution of mole rats, population sizes are typically
< 100, but genetic diversity is much higher than on any expectation based
upon history, demography, gene flow or inbreeding (Nevo, Filipucci & Beiles,
1994).
(3) Viewed energetically, mating is an extremely expensive process. Since there is
a limit to the amount of energy that can be assimilated from food (Weiner,
1992), there is an upper limit to total behavioural activity. In parallel, oxygen
consumption increases to meet a high demand for ATP production, but the
maximum possible oxygen consumption is ultimately limiting (Bennett, 1991).
Consequently fitness, assessed by mating success, is often relatable to available
metabolic energy. Heterozygotes should be favoured at times of high energy
demands during mating, and this has been found in a number of species,
especially insects and fish (Thornhill & Gangestad, 1993).
On top of the cost involved in mating, some organisms develop and maintain
energetically costly sexual ornaments. Fisher (1930) argued that without any
check, a runaway process is expected and consequently the sexual ornament is
continuously exaggerated in dimensions. However, in some birds, sexual
ornamentation is restricted to the breeding season, indicating a substantial
energetic cost. This suggests a tradeoff, whereby the compounded energy cost
underlying the development and maintenance of ornaments of increasing size, is
countered by the cost of abiotic stress, especially when resources are limiting.
Since secondary sexual traits are extremely exaggerated versions of ordinary
traits, they are close to their limits of production and maintenance. Thus,
elaboration of secondary sexual traits can bring individuals to the energetic brink,
when stress-resistant genes should be favoured (Parsons, 1995a).
Under such extreme energetic circumstances, viability selection will favour
metabolic efficiency to enable survival. These are circumstances when the size of
ornaments should be associated with heterozygote advantage in natural
populations. For instance, in bighorn sheep, Ovis canadensis, at year 7 of life, which
is around the time of onset of breeding, 21% of variation in horn volume of rams
was attributable to an association with heterozygosity, but in young rams this
association was not apparent. Therefore, when energy demands from the
development and maintenance of horns and from the mating process itself are
high, heterozygote advantage is maximal (Fitzsimmons, Buskirk & Smith,
1995).
These examples suggest that under highly stressed situations in free-living
populations, genetic variation is likely to be enhanced. This is in accord with a
substantial body of laboratory data indicating increased mutation, recombination,
developmental variability, and phenotypic variability as stresses approach levels
where extinctions become a possibility (Parsons, 1987; Hoffmann & Parsons,
1991).

ELECTROPHORETIC LOCI AND STRESS

To what extent does genetic variation at specific loci underlie adaptation to


changing conditions? There is much evidence that gene frequencies involving

476

P. A. PARSONS

electrophoretic variants can be correlated with temperature and other variables at


the geographic level. Some of these studies have demonstrated phenotypic
differences under extreme conditions reducible to the gene locus level (e.g. Piazza,
Menozzi & Cavalli-Sforza, 1981; Watt, 1985; Riddoch, 1993).
At the phosphoglucose isomerase, PGI, locus, the more anodal allozyme/isozyme
is favoured under stressful conditions including high temperature, high salinity,
anoxia and desiccation in data covering a wide range of animal and plant taxa. This
suggests at least one locus in natural populations of major importance in determining
resistance to stress (Riddoch, 1993). This particular situation is not surprising, since
in vitro biochemical studies of PGI allozymes suggest that this enzyme, which
catalyzes a metabolic reaction and regulates flux through glycolysis, is a direct target
of selection of such stresses (Watt, 1985).
Therefore, direct associations of stress with particular electrophoretic loci can
occur. In addition, the PGI locus provides the clue that it may be important to
consider the effect of stress on the energy balance of alternative genotypes, or more
broadly, at the level of the metabolic potential of organisms. Aridity, or desiccation
stress in the laboratory, appears to fulfil this role. In contrast, climatic stresses are
more typically expressed in terms of temperature shifts than aridity. However,
hydrological extremes of drought and deluge can have long and short term effects
that appear more direct and local as selective agents than temperature (Parsons,
1995b), so that aridity is the main stress under discussion in this paper.

STRESS, METABOLIC POTENTIAL AND CONSERVATION

Nevo et al (1994) found that molecular genetic data in mole rats, S. ehrenbergi, show
parallel patterns across habitats with data at the nuclear and mitochondrial DNA
levels as well as the organismic level. Ecological factors with an emphasis on aridity
stress underlie these patterns. The convergence of the organismic and molecular
levels to give parallel patterns, suggests that much of the total metabolic potential of
mole rats is assayed which is reflected at the organismic level, and is revealed by
desiccation stress as an environmental probe. This would be compatible with the
target of selection of desiccation stress at the level of energy carriers.
Furthermore, the inference that stress resistance is associated with metabolic
potential implies that in extreme environments, the preservation of organisms having
maximum metabolic potential to withstand the energy costs of stress should be at a
premium. Therefore, stress applied at the organismic level is likely to have
evolutionary consequences at the physiological and molecular levels. In this case, the
connection is indirect, since an extensive coverage of 36 electrophoretic loci and
nuclear and mitochondrial DNA, should involve sufficient of the complete metabolic
potential of organisms, that associations become apparent across levels of
organization from the molecular to the organismic.
This suggests that the survival of a species in its environment depends upon
possessing the metabolic potential to survive the energy costs of stress of that
environment. Under the stressful scenario, the habitats of organisms can be
expressed in terms of an interaction between stress intensity, magnitude of
environmental fluctuations, and energy available from resources as a first
approximation (Parsons, 1994a). The interaction of stress of various types causing
energy costs with energy gained from resources is a central tradeoff. Therefore, as

CONSERVATION STRATEGIES, STRESS AND DIVERSITY

477

conditions deviate from optimal, energy costs increase (Porter & Gates, 1969), so that
physical conditions can limit the occurrence of organisms to particular habitats. This
approach implies that the distribution and abundance of organisms should, in
principle, be relatable to energy balances (Hall, Stanford & Hauer, 1992), derived
from the costs of various abiotic (and biotic) stresses interacting with gains from
resources.
The difficulty in using the preservation of genetic variation as a primary strategy
in conservation follows from the point that natural selection acts primarily at the
organismic level. The approach outlined here is based upon the potential for
adaptation of organisms to the abiotic stresses to which they are normally exposed in
free-living conditions. This is a reductionist model (Parsons, 1992) based upon a
primary target of selection at the level of energy carriers. The extent to which the
model is reductionist to the gene level in the sense of Williams (1985) needs further
study, although the analysis of Riddoch (1993) is suggestive.
Since environmental perturbations increase energy costs, adaptation to a
permanent increase in stress should involve selection for reduced energy costs often
via a lowered metabolic rate, with a range of concomitant effects for life-history
traits. This has been observed experimentally in D. melanogaster by selecting for
desiccation resistance (Hoffmann & Parsons, 1989, 1991). Furthermore, reduced
fecundity and behavioural activity are direct indicators of a fall in fitness following
selection. The cost of this process therefore reduces the potential for further
adaptation, and a limit is likely at species borders when the cost can become totally
restrictive (Parsons, 1991). It is under these outlier conditions that heterozygotes may
be favoured, because of their greater efficiency of protein synthesis which releases
energy for ingestion and absorption. This effect should be maximal under the most
extreme environmental conditions when energy demands are highest. Therefore,
under the variably stressful environments of free-living populations, there should
normally be a premium on the conservation of genetic diversity. Hawkins (1995)
argues that under circumstances of heterogeneous and stressful environments, the
metabolic potential of multi-locus heterozygotes tends to ensure their superiority. In
this way adaptation to extreme environments can be enhanced (Koehn & Bayne,
1989).
In comparisons among species, levels of genetic variation range from predominantly homozygous at one limit, so that the favouring of heterozygotes under stress
cannot be universal. In contrast, irrespective of the level of variation, I argue that an
assessment of traits for stress resistance is an important primary strategy in
conservation biology for all species. This leads to direct predictions concerning the
survival of species during periods of environmental change. Unfortunately, this
phenotypic approach is more time consuming than studies of genetic diversity, but
there is a strong case for parallel experiments based on stress response traits of
ecological significance whenever this is feasible.

BIOTIC STRESS, ESPECIALLY COMPETITION

Fisher (1930) in Chapter VI of The Genetical Theory wrote: Any environmental


heterogeneity which requires special adaptations, which are either irreconcilable or
difficult to reconcile, will exert upon the cohesive powers of the species a certain
stress. He goes on to argue that speciation can occur under sufficiently strong

478

P. A. PARSONS

selection in terms of geographical distance. It is a remarkably modern statement,


since Wilson (1992) argues that differences between species originally originate as
traits that adapt them to the environment. He was emphasizing droughts, sea level
changes, and island environments.
However, in Chapter II Fisher (1930) comments: Probably more important than
the changes in climate will be the evolutionary changes in associated organisms. As
each organism increases in fitness, so will its enemies and competitors increase in
fitness.
This statement parallels Darwins (1859) emphasis on biotic interactions, as being
dominant in underlying evolutionary change. In the one book, therefore, Fisher
discusses the environment in a way that represents the transition from Darwin to the
harsher world I am now assuming. Can these approaches be reconciled? An answer
comes from energy considerations. This means incorporating biotic stresses, in
particular competition, into the context of the energy costs of abiotic stresses.
Witter (1995) recorded that in the European starling, Sturnus vulgaris, an
experimental increase in the level of competition at feeding sites decreases the
number of prey items which starlings carry back to the nest. A simple interpretation
is reduced fitness consequent upon the energy cost of competition. Indeed, there are
situations indicating that combinations of biotic and abiotic stress can push the
physiological state of organisms closer to lethality that can abiotic stress alone. An
example occurs in suspension-feeding venerid bivalves where a history of crowding
increased susceptibility to sedimentation stress (Peterson & Black, 1988). This means
that the energy cost of competition can decrease the stress resistance of organisms,
so reducing the range of physical conditions under which habitats could be
occupied.
However, the heritability of stress resistance can be very high, while that of
competitive ability is often barely distinguishable from zero (Mather & Cooke, 1962;
Parsons, 1973). This suggests that competitive ability can be viewed as a secondorder effect compared with stress as factors underlying evolutionary change.
This conclusion implies that the development of competitive relationships can only
be expected in abiotically rather benign habitats. In such restricted circumstances,
the energy cost of accommodating abiotic stress is relatively low, and a premium on
energy efficiency could select for diversification to exploit differing resources
(Parsons, 1994a), perhaps assisted by competition. Similarly, laboratory conditions
where temperature is controlled, represent a situation where energy costs from
environmental perturbations should be low compared with unstable field situations.
Under these conditions, intraspecific competition has been demonstrated in insects
such as Lucilia cuprina and Drosophila (e.g. Nicholson, 1957; Mueller, 1988).
In any case, this approach shows that competition can be accommodated into
energy budgets (Parsons, 1996), as can abiotic stress, so that the varying approaches
to the environment, as presented by Darwin, Fisher and those mentioned in this
paper, can be reconciled. Competition becomes almost a special case in the longer
term, since severe abiotic perturbations occur unpredictably and sporadically on
time scales ranging from the seasonal to El Nino events, and ultimately on the
geological scale. Under such stressful circumstances, complex communities appear
likely to be particularly vulnerable. In an increasingly stressed world, conservation
strategies appear to be most realistically developed assuming substantial stress, so
that biotic effects and interactions can normally be assumed to be relatively
unimportant.

CONSERVATION STRATEGIES, STRESS AND DIVERSITY

479

CONCLUDING COMMENTS

I suggest that the emphasis of conservation biology on preserving genetic


variability be reviewed, in terms of two central points: (1) the need to assume a
realistically stressful environment for free-living populations, and (2) the energy costs
for organisms to survive and reproduce in this environment.
These are direct effects with no obvious consequences for genetic diversity, except
that under very extreme circumstances an enhancement is likely. Fisher (1930) did
not discuss energy budgets, although in developing the Fundamental Theory of
Natural Selection, he noted that fitness could be considered in thermodynamic
terms. This is the approach taken in a number of recent publications. For instance,
Van Valen (1991) argues that energy provides a secure foundation for fitness, and
removes various anomalies from a purely genetical approach, and Brown, Marquet
& Taper (1993) define fitness in energy terms as the rate that resources, in excess of
those required for growth and maintenance of the individual, can be harvested from
the environment and used for reproduction. More generally, Torres (1991) gives a
definition of fitness based upon the distances of a given individuals thermodynamic
parameters from their optimum values. In these terms, fitness will fall as the level of
stress increases, since stress has an energy cost.
Fisher (1930) in Chapter V of The Genetical Theory wrote: An evolutionary
consequence of some importance is that in general a smaller number of large species
must be increasing in number at the expense of a larger number of small species, the
continuous extinction of the latter setting a natural check to the excessive subdivision
of species which would ensue upon too fine and detailed specialization. Here he was
apparently assuming an equilibrium, which is now under threat, and specialist
species, which tend to be stress-sensitive, face extinction. Generalist more stressresistant species would be favoured (Parsons, 1994a,b). Regarding a shift towards
ecosystem maturity as a process of increasing information, increased environmental
perturbations would reverse this trend and biodiversity would fall, forcing the
community into an earlier stage of ecological succession (Margalef, 1968).
Experimental perturbations using copper as a stress had this effect on a freshwater
plankton community (Havens, 1994). Furthermore, there is some recent evidence
suggestive of this trend in free-living populations. Since the 1950s and especially since
1980 in tropical forests of the Americas, trends in deforestation and atmospheric
change appear to be leading towards lowered precipitation, increased seasonality,
and more extreme weather (Hartshorn, 1992). Abiotic stress therefore is increasing,
and consequently the slowest growing shade-tolerant trees and tropical forest
organisms with life-cycles tied to those trees appear particularly vulnerable (Phillips
& Gentry, 1994). This may be the start of a trend from specialist diversification to
generalist uniformity. Under these circumstances competitive relationships would be
particularly vulnerable. It is a likely trend at any time when the energy cost from
stress increases, and could now be a general trend.
Therefore, the need to monitor the effect of stress in free-living populations as a
primary approach in conservation biology is reinforced. Recent examples illustrative
of this approach include the interactive effects of ambient ozone and climate, in
particular moisture stress, in reducing the growth of mature loblolly pine, Pinus taeda,
trees (McLaughlin & Downing, 1995) and changes in the breeding cycles of three
amphibian species native to Britain consequent upon recent steadily increasing
winter and spring temperatures, (Beebee, 1995). In accord with Hoffmann & Parsons

480

P. A. PARSONS

(1991) environmental extremes are significant in these examples, which appear


indicative of future emphases.

ACKNOWLEDGEMENTS

It is impossible to acknowledge all the influences over the years that have led to
this paper. However, A.R.G. Owen was a most important and helpful mentor in
Cambridge and subsequently. R.W. Allard of the University of California at Davis
was central in helping me to appreciate that the environment of free-living
populations is much tougher than assumed by many evolutionary biologists. At the
University of Adelaide J.H. Bennett has been of great assistance. I am grateful to all
these people, as well as to many others, including R.J. Berry, W.F. Bodmer and
A.W.F. Edwards, but most of all to R.A. Fisher. Finally, J.B. Mitton was most helpful
in improving this paper.

REFERENCES
Barnett SA, Coleman EM. 1960. Heterosis in F1 mice in a cold environment. Genetical Research 1: 2538.
Beebee TJC. 1995. Amphibian breeding and climate. Nature 374: 219220.
Bennett AF. 1991. The evolution of activity capacity. Journal of Experimental Biology 160: 123.
Bennett JH. ed. 1983. Natural Selection, Heredity and Eugenics. Oxford: Clarendon Press.
Booth DT, Clayton DH, Block BA. 1993. Experimental demonstration of the energetic costs of parasitism in
free-ranging hosts. Proceedings of the Royal Society of London B 223: 125129.
Brown JH, Marquet PA, Taper ML. 1993. Evolution of body size: consequences of an energetic definition of
fitness. American Naturalist 142: 373384.
Caro TM, Laurenson MK. 1994. Ecological and genetic factors in conservation: a cautionary tale. Science 263:
485486.
Dahlgaard J, Krebs RA, Loeschcke V. 1995. Heat-shock tolerance and inbreeding in Drosophila buzzatii.
Heredity 74: 157163.
Darwin C. 1859. On the Origin of Species by means of Natural Selection. London: Murray.
de Lope F, Gonzalez
`
G, Perez JJ, Mller AP. 1993. Increased detrimental effects of ectoparasites on their bird
hosts during adverse environmental conditions. Oecologia 95: 234240.
Diehl WJ. 1988. Genetics of carbohydrate metabolism and growth in Eisenia foetida (Oligochaeta: Lumbricidae).
Heredity 61: 379387.
Ferguson MM, Drahushchak LR. 1990. Disease resistance and enzyme heterozygosity in rainbow trout.
Heredity 64: 413417.
Fisher RA. 1930. The Genetical Theory of Natural Selection. Oxford: Clarendon Press.
Fisher RA. 1958. The Genetical Theory of Natural Selection. 2nd ed. New York: Dover Publications.
Fitzsimmons NN, Buskirk SW, Smith MH. 1995. Population history, genetic variability, and horn growth in
bighorn sheep. Conservation Biology 9: 314323.
Frankel OH, Soule M. 1981. Conservation and Evolution. Cambridge University Press.
Garton DW, Koehn RK, Scott TM. 1984. Multiple-locus heterozygosity and physiological energetics of growth
in the coot clam, Mulinia lateralis, from a natural population. Genetics 108: 445455.
Gulland FMD, Albon SD, Pemberton JM, Moorcroft PR, Clutton-Brock TH. 1993. Parasite-associated
polymorphism in a cyclic ungulate population. Proceedings of the Royal Society of London B 254: 713.
Hall CAS, Stanford JA, Hauer FR. 1992. The distribution and abundance of organisms as a consequence of
energy balances along multiple environmental gradients. Oikos 65: 377390.
Hartshorn G. 1992. Possible effects of global warming on the biological diversity in tropical forests. In: Peters RL,
Lovejoy TE, eds. Global Warming and Biodiversity. New Haven: Yale Univ. Press, 137146.
Havens, KE. 1994. Experimental perturbations of a freshwater plankton community: a test of hypotheses
regarding the effects of stress. Oikos 69: 147153.
Hawkins AJS. 1995. Effects of temperature change on ectotherm metabolism and evolution: metabolic and
physiological interrelations underlying the superiority of multi-locus heterozygotes in heterogeneous environments. Journal of Thermal Biology 20: 2333.
Hoffmann AA, Parsons PA. 1989. An integrated approach to environmental stress tolerance and life-history
variation: desiccation tolerance in Drosophila. Biological Journal of the Linnean Society 37: 117136.

CONSERVATION STRATEGIES, STRESS AND DIVERSITY

481

Hoffmann AA, Parsons PA. 1991. Evolutionary Genetics and Environmental Stress. Oxford: Oxford University
Press.
Jimenez JA, Hughes KA, Alaks G, Graham L, Lacy RC. 1994. An experimental study of inbreeding
depression in a natural habitat. Science 266: 271273.
Keller LF, Arcese P, Smith JNM, Hochachka WM, Stearns SC. 1994. Selection against inbred song
sparrows during a natural bottleneck. Nature 372: 356357.
Kieser JA, Groenevold HT. 1991. Fluctuating odontometric asymmetry, morphological variability, and genetic
monomorphism in the cheetah Acinonyx jubatus. Evolution 45: 11751183.
Koehn RK. Bayne BL. 1989. Towards a physiological and genetical understanding of the stress response.
Biological Journal of the Linnean Society 37: 157171.
Langridge J. 1962. A genetic and molecular basis for heterosis in Arabidopsis and Drosophila. American Naturalist 96:
527.
Latta R, Ritland K. 1994. The relationship between inbreeding depression and prior inbreeding among
populations of four Mimulus taxa. Evolution 48: 806817.
Lesica P, Allendorf FW. 1992. Are small populations of plants worth preserving? Conservation Biology 6:
135139.
Margalef R. 1968. Perspectives in Ecological Theory. Chicago: University of Chicago Press.
Mather K, Cooke P. 1962. Differences in competitive ability between genotypes of Drosophila. Heredity 17:
381407.
McLaughlin SB, Downing, DJ. 1995. Interactive effects of ambient ozone and climate measured on growth of
mature forest trees. Nature 374: 252254.
McWilliam JR, Griffing B. 1965. Temperature-dependent heterosis in maize. Australian Journal of Biological
Sciences 18: 569583.
Merola M. 1994. A reassessment of homozygosity and the case for inbreeding depression in the cheetah, Acrinonyx
jubatus: implications for conservation. Conservation Biology 8: 961971.
Miller PS. 1994. Is inbreeding depression more severe in a stressful environment? Zoo Biology 13: 195208.
Mitton JB. 1993a. Theory and data pertinent to the relationship between heterozygosity and fitness. In: Thornhill
NW, ed, The Natural History of Inbreeding and Outbreeding. Chicago: Chicago University Press, 1741.
Mitton JB. 1993b. Enzyme heterozygosity, metabolism and developmental variability. Genetica 89: 4765.
Mitton JB, Grant MC. 1984. Associations among protein heterozygosity, growth rate, and developmental
homeostasis. Annual Review of Ecology and Systematics 15: 479499.
Mitton JB, Schuster WSF, Cothran EG, DeFries JC. 1993. Correlation between the individual heterozygosity
of parents and their offspring. Heredity 71: 5963.
Montalvo AM. 1994. Inbreeding depression and maternal effects in Aquilegia caerulea, a partially selfing plant.
Ecology 75: 23952409.
Mueller LD. 1988. Evolution of competitive ability in Drosophila by density-dependent natural selection. Proceedings
of the National Academy of Sciences USA 85: 43834386.
Nevo E, Filippucci MG, Beiles A. 1994. Genetic polymorphisms in subterranean mammals (Spalax ehrenbergi
superspecies) in the Near East revisited: patterns and theory. Heredity 72: 465487.
Nicholson AJ. 1957. The self-adjustment of populations to change. Cold Spring Harbor Symposia on Quantitative Biology
22: 153173.
OBrien SJ. 1994. A role for molecular genetics in biological conservation. Proceedings of the National Academy of
Sciences USA 91: 57485755.
Odum EP, Finn JT, Franz EH. 1979. Perturbation theory and the subsidy stress gradient. BioScience 29:
349352.
Parsons PA. 1959. Genotype-environmental interactions for various temperatures in Drosophila melanogaster. Genetics
44: 13251333.
Parsons PA. 1971. Extreme-environment heterosis and genetic loads. Heredity 26: 579583.
Parsons PA. 1973. Behavioural and Ecological Genetics: Study in Drosphila. Oxford: Clarendon Press.
Parsons PA. 1987. Evolutionary rates under environmental stress. Evolutionary Biology 21: 311347.
Parsons PA. 1990. Fluctuating asymmetry: an epigenetic measure of stress. Biological Reviews 63: 131145.
Parsons PA. 1991. Evolutionary rates: stress and species boundaries. Annual Review of Ecology and Systematics 22:
116.
Parsons PA. 1992. Evolutionary adaptation and stress: the fitness gradient. Evolutionary Biology 26: 191223.
Parsons PA. 1994a. Habitats, stress, and evolutionary rates. Journal of Evolutionary Biology 7: 387397.
Parsons PA. 1994b. The energetic cost of stress. Can biodiversity be preserved? Biodiversity Letters 2: 1115.
Parsons PA. 1995a. Stress and limits to adaptation: sexual ornaments. Journal of Evolutionary Biology 8:
455461.
Parsons PA. 1995b. Evolutionary response to drought stress: conservation implications. Biological Conservation 73:
2127.
Parsons PA. 1996. Competition versus abiotic factors in variably stressful environments: evolutionary implications.
Oikos 75: 129132.
Peterson CH, Black R. 1988. Density-dependent mortality caused by physical stress interacting with biotic
history. American Naturalist 131: 257270.
Phillips OL, Gentry AH. 1994. Increasing turnover through time in tropical forests. Science 263: 954958.

482

P. A. PARSONS

Piazza A, Menozzi P, Cavalli-Sforza LL. 1981. Synthetic gene frequency maps of man and selective effects of
climate. Proceedings of the National Academy of Sciences USA 78: 26382642.
Porter WP, Gates DM. 1969. Thermodynamic equilibria of animals with environment. Ecological Monographs 39:
227244.
Prober SM, Brown AHD. 1994. Conservation of the grassy white box woodlands: population genetics and
fragmentation of Eucalyptus albens. Conservation Biology 8: 10031013.
Raijmann LEL, van Leeuwen NC, Kersten R, Oostermeijer JGB, den Nijs HCM, Menken SBJ. 1994.
Genetic variation and outcrossing rate in relation to population size in Gentiana pneumonanthe L. Conservation Biology
8: 10141026.
Riddoch BJ. 1993. The adaptive significance of electrophoretic mobility in phosphoglucose isomerase (PGI).
Biological Journal of the Linnean Society 50: 117.
Schall JJ, Sarni GA. 1987. Malarial parasitism and the behavior of the lizard, Sceloporus occidentalis. Copeia 1987
(1):8493.
Schmitt J, Gamble SE. 1990. The effects of distance from the parental site on offspring performance and
inbreeding depression in Impatiens capensis: a test of the local adaptation hypothesis. Evolution 44: 20222030.
Scott TM, Koehn RK. 1990. The effect of environmental stress on the relationship of heterozygosity to growth
rate in the coot clam Mulinia lateralis (Say). Journal of Experimental Marine Biology and Ecology 135: 109116.
Teska WR, Smith MH, Novak JM. 1990. Food quality, heterozygosity, and fitness correlates in Peromyscus
polionotus. Evolution 44: 13181325.
Thornhill R, Gangestad SW. 1993. Human facial beauty: averageness, symmetry and parasite resistance. Human
Nature 4: 237269.
Torres J-L. 1991. Natural selection and thermodynamic optimality. Il Nuovo Cimento 13: 177185.
Van Valen LM. 1991. Biotal evolution: a manifesto. Evolutionary Theory 10: 113.
van Treuren R, Bijlsma R, Ouborg NJ, van Delden W. 1993. The significance of genetic erosion in the
process of extinction. IV Inbreeding depression and heterosis effects caused by selfing and outcrossing in Scabiosa
columbaria. Evolution 47: 16691680.
Watt WB. 1985. Bioenergetics and evolutionary genetics: opportunities for new synthesis. American Naturalist 125:
118143.
Weiner J. 1992. Physiological limits to sustainable energy budgets in birds and mammals: ecological implications.
Trends in Ecology and Evolution 7: 384389.
White TCR. 1993. The Inadequate Environment: Nitrogen and the Abundance of Animals. Berlin: Springer-Verlag.
Williams GC. 1985. A defense of reductionism in evolutionary biology. Oxford Surveys in Evolutionary Biology 2:
127.
Wilson EO. 1992. The Diversity of Life. Cambridge, Massachusetts: Belknap Press.
Witter MS. 1995. The effect of intraspecific competition on prey load size in the European starling, Sturnus vulgnaris.
Animal Behaviour 49: 261264.
Zakharov VM. 1989. Future prospects for population phenogenetics. Soviet Science Reviews, Section F. Physiology and
General Biology Reviews 4: 179.

You might also like