ArticlePDF AvailableLiterature Review

Heat-shock proteins, molecular chaperones, and the stress response: Evolutionary and Ecological Physiology

Authors:

Abstract and Figures

Molecular chaperones, including the heat-shock proteins (Hsps), are a ubiquitous feature of cells in which these proteins cope with stress-induced denaturation of other proteins. Hsps have received the most attention in model organisms undergoing experimental stress in the laboratory, and the function of Hsps at the molecular and cellular level is becoming well understood in this context. A complementary focus is now emerging on the Hsps of both model and nonmodel organisms undergoing stress in nature, on the roles of Hsps in the stress physiology of whole multicellular eukaryotes and the tissues and organs they comprise, and on the ecological and evolutionary correlates of variation in Hsps and the genes that encode them. This focus discloses that (a) expression of Hsps can occur in nature, (b) all species have hsp genes but they vary in the patterns of their expression, (c) Hsp expression can be correlated with resistance to stress, and (d) species' thresholds for Hsp expression are correlated with levels of stress that they naturally undergo. These conclusions are now well established and may require little additional confirmation; many significant questions remain unanswered concerning both the mechanisms of Hsp-mediated stress tolerance at the organismal level and the evolutionary mechanisms that have diversified the hsp genes.
Content may be subject to copyright.
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
Annu. Rev. Physiol. 1999. 61:243–82
Copyright
c
1999 by Annual Reviews. All rights reserved
HEAT-SHOCK PROTEINS,
MOLECULAR CHAPERONES,
AND THE STRESS RESPONSE:
Evolutionary and Ecological
Physiology
Martin E. Feder
Department of Organismal Biology and Anatomy and Committee on Evolutionary
Biology, University of Chicago, 1027 East 57th Street, Chicago, Illinois 60637;
e-mail: m-feder@uchicago.edu
Gretchen E. Hofmann
Department of Biology, University of New Mexico, Albuquerque, NM 87131;
e-mail: ghofmann@unm.edu
KEY WORDS: hsp, temperature, protein denaturation and folding, inducible tolerance,
environmental gradients
ABSTRACT
Molecular chaperones, including the heat-shock proteins (Hsps), are a ubiquitous
feature of cells in which these proteins cope with stress-induced denaturation
of other proteins. Hsps have received the most attention in model organisms
undergoing experimental stress in the laboratory, and the function of Hsps at
the molecular and cellular level is becoming well understood in this context. A
complementary focus is now emerging on the Hsps of both model and nonmodel
organismsundergoingstressinnature,ontherolesofHspsinthestressphysiology
of whole multicellular eukaryotes and the tissues and organs they comprise, and
on the ecological and evolutionary correlates of variation in Hsps and the genes
that encode them. This focus discloses that (a) expression of Hsps can occur
in nature, (b) all species have hsp genes but they vary in the patterns of their
expression, (c) Hsp expression can be correlated with resistance to stress, and
(d) species’ thresholds for Hsp expression are correlated with levels of stress
that they naturally undergo. These conclusions are now well established and
may require little additional confirmation; many significant questions remain
243
0066-4278/99/0315-0243$08.00
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
244 FEDER & HOFMANN
unanswered concerning both the mechanisms of Hsp-mediated stress tolerance
at the organismal level and the evolutionary mechanisms that have diversified the
hsp genes.
INTRODUCTION
Although heat-shock proteins (Hsps) first achieved notoriety as gene products
whose expression is induced by heat and other stresses (1,2), discoveries of the
pastdecade haveshiftedthefocusof research to understanding therolesofHsps
as molecular chaperones (3–5). As a result, Hsps, their close relatives, their
molecularpartners, andmanynewlydiscovered proteinsare nowknownto play
diverse roles, even in unstressed cells, insuccessful folding, assembly, intracel-
lular localization, secretion, regulation, and degradation of other proteins (6);
failure of these activities is thought to underlie numerous and important human
diseases (7). Nonetheless, many of the questions of the past either remain
unanswered, awaiting the development of appropriate experimental tools, or
can now be revisited with new insights gleaned from the emerging understand-
ing of molecular chaperones. This review seeks to facilitate the examination or
re-examination of Hsps as responses to natural stress in diverse organisms in-
habiting environments outside the laboratory, the function of Hsps in tolerance
of natural stresses, and ecological and evolutionary variation in the heat-shock
system. The review sequentially considers (a) the principal implications of
laboratory-based studies for ecological and evolutionary research on Hsps, (b)
expression of Hsps in nature, (c) covariation of Hsp expression with environ-
mental and biological gradients of stress intensity, (d) the consequences of Hsp
expression for fitness, and (e) evolutionary variation in Hsps and the genes
that encode them. The primary objective of this review is to redirect the focus
of evolutionary and ecological research on Hsps beyond the conclusions that
are now well-established and onto the many important questions that remain
unanswered.
STATE OF THE LITERATURE
Established Conclusions
The relevant literature on Hsps and molecular chaperones is huge, now com-
prising more than 12,000 references. Even a review of the relevant reviews
is difficult. For this reason, we begin by describing several well established
conclusions and cite a fewof the manyexcellent recent reviews at diverse levels
of sophistication (4–6, 8–11).
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
HEAT-SHOCK PROTEINS 245
The genes encoding Hsps (hsps) are highly conserved and occur in every
species in which they have been sought. Many of these genes and their prod-
ucts can be assigned to families on the basis of sequence homology and typical
molecular weight (6): hsp110, hsp100, hsp90, hsp70, hsp60, hsp40, hsp10,
and small hsp families. Gething (6) recognizes 7 additional families and 12
genes/proteins for which families have not yet been described. In eukaryotes,
many families comprise multiple members that differ in inducibility, intracel-
lular localization, and function.
Hsps function as molecular chaperones; i.e. they interact with other pro-
teins and, in so doing, minimize the probability that these other proteins will
interact inappropriately with one another. Hsps recognize and bind to other
proteins when these other proteins are in non-native conformations, whether
due to protein-denaturing stress or because the peptides they comprise have not
yet been fully synthesized, folded, assembled, or localized to an appropriate
cellular compartment. Binding and/or release of these other proteins is often
regulated by association with and/or hydrolysis of nucleotides. Typically, Hsps
function as oligomers, if not as complexes of several different chaperones,
co-chaperones, and/or nucleotide exchange factors. Interaction with chaper-
ones is variously responsible for (a) maintaining Hsps’ partner proteins in a
folding-competent, folded, or unfolded state; (b) organellar localization, im-
port, and/or export; (c) minimizing the aggregation of non-native proteins; and
(d) targeting non-native or aggregated proteins for degradation and removal
from the cell. Presumably, the last two functions are most important in coping
with environmental stress.
Not all Hsps are stress-inducible, but those that are respond to a variety
of stresses, including extremes of temperature, cellular energy depletion, and
extreme concentrations of ions, other osmolytes, gases, and various toxic sub-
stances. Activation of various intracellular signaling pathways results in Hsp
expression. All known stresses, if sufficiently intense, induce Hsp expression.
Accordingly, Hsps are equally well termed stress proteins, and their expression
is termed the stress response. A common aspect of these inducing stresses
is that they result in proteins having non-native conformations (12), which is
consistent with the function of Hsps as molecular chaperones.
Implications of the Published Literature for Ecological
and Evolutionary Studies of Hsps
Space limitations necessitate that we choose among numerous equally valu-
able references in preparing this review. To present both the breadth and depth
of research relevant to the evolutionary and ecological physiology of the heat-
shock response, we have compiled a near-comprehensive bibliography of
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
246 FEDER & HOFMANN
that literature, which is available electronically (13) on the World Wide Web
in the Supplemental Materials Section of the main Annual Reviews site
(http://www.AnnualReviews.org).
A first implication of this massive literature is that the Hsp field has long ago
concluded its exploratory phase. Showing that an as-yet-unexamined species
expresses Hsps in response to heat or other stresses no longer has any particular
novelty.
Second, much of the work on Hsps outside the laboratory or in nonmodel
organisms was undertaken before the molecular diversity of Hsps and their
functionas molecular chaperones was obvious. In the interim, the experimental
tools for examining Hsps and the standards for such examinations have both
advanced considerably. As a result, much of the earlier work on evolutionary
and ecological physiology of Hsps regrettably either does not withstand current
scrutiny or contributes little to issues of current interest. Several issues are
obvious:
1. Many of the apparently singular Hsps of previous years, often detected
by one-dimensional electrophoresis and autoradiography, are now known
to represent entire families of Hsps, often with (a) discrete distributions
within the cell (e.g. cytoplasmic-nuclear, mitochondrial, chloroplast, or en-
doplasmic reticulum), (b) different degrees of inducibility (constitutively
expressed, constitutively expressed but increasing during or after stress, ex-
clusively inducible), (c) differing kinetics of induction and removal from
the cell, and (d) differing tissue specificity. Representing this diversity as
a single Hsp or two Hsps (“constitutive” and “inducible”) through use of
nonspecific probes or lysates of whole organisms and organs can obscure
phenomena of great significance (e.g. compare Refs. 14 and 15 with 16).
This problem is sometimes remediable only with great difficulty. Often,
highly specific probes are available only for standard model organisms, par-
ticularly at the level of proteins, and great care must be taken in applying
these probes to non-standard organisms (17).
2. Inducible stress tolerance is increasingly understood to result from numer-
ous molecular mechanisms, of which Hsps are collectively only one. Other
mechanisms include synthesis of osmotic stress protectants such as polyols
andtrehalose, modificationsofthesaturationofcellmembranelipids(home-
oviscousadaptation), compensatory expression of isozymes or allozymes of
significant enzymes, metabolic arrest, radical scavengers (superoxide dis-
mutase, glutathione system, cytochrome P450), and so on. Accordingly,
the unambiguous attribution of stress tolerance to Hsps in general or to
any specific Hsp requires more than correlative evidence (18–21). Increas-
ingly, proof resulting from genetic or direct experimental manipulation is
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
HEAT-SHOCK PROTEINS 247
becoming the standard for establishing the functional or evolutionary sig-
nificance of Hsps. Again, this rising standard is often met only with great
difficulty in ecological and evolutionary physiological studies, for many of
the techniques for genetic and experimental manipulation are not readily
applicable to the more ecologically and evolutionarily interesting species.
HSP-INDUCING STRESS IN NATURE
AND NATURAL INDUCTION OF HSPS
Depending on their geographic locale, organisms in nature risk exposure to
temperatures ranging from 100
to more than 100
C, and comparable ex-
tremes of chemical and gas concentration, food and water availability, hydro-
static pressure, radiation, and toxic substances of human origin. Seemingly,
Hsp expression should be a common occurrence in nature. In reality, how-
ever, movement and/or other behaviors may often enable organisms to avoid
Hsp-inducing stress in nature by exploiting equable microhabitats in otherwise
stressful environments (22–24). Also, biochemical specializations other than
Hsps may stabilize many organisms (or particular stages of their life cycles) so
that environmental extremes are not particularly stressful.
Even equable environments can contain Hsp-inducing microhabitats, and
even mild stresses can induce Hsps when multiple stresses act in combina-
tion. For these reasons, we can assume neither the presence nor the absence
of Hsp expression in nature; for that matter, we still do not know in any com-
prehensive sense whether wild organisms routinely, occasionally, or seldom
express inducible Hsps. A growing body of evidence, however, establishes that
at least in some circumstances and taxa, Hsp induction is not solely a laboratory
phenomenon.
One caution in evaluating the subsequent account is that organisms in na-
ture seldom undergo only one stress at a time. For example, an insect larva
undergoing natural heat stress in a rotting fruit may simultaneously experience
intense ultraviolet radiation, desiccation, and diverse alcohols and aldehydes,
among other stresses. This situation differs from that in most laboratory exper-
imentation, which involves one or a few stresses and makes attribution of Hsp
expression to a particular stress in nature more complicated.
Aquatic Temperature Stress
Due to the physical characteristics of water, the aquatic environment can be ex-
tremelystressful to its inhabitants.In general, the high specific heat and thermal
conductivity of water ensure that the majority of aquatic organisms will have
body temperatures equivalent to that of their surroundings. Furthermore, the
relative thermal homogeneity of aquatic environments can frustrate behavioral
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
248 FEDER & HOFMANN
avoidance of thermal extremes. Some aquatic ectotherms nonetheless inhabit
thermally equable habitats or waters with enough thermal diversity to enable
behavioral thermoregulation; our focus is on those species that do not.
In the aquatic environment, habitual exposureto Hsp-inducing thermal stress
may be most common in sessile organismsthat occur in shallow, stagnant water
(e.g. ponds, tidepools, swamps, tidal flats) or in the intertidal zone. Corals, for
example, routinely undergo thermal stress that results in bleaching, during
which the corals’ endosymbionts die. Even modest increases in water tempera-
ture of 1-2
C can bleach corals; these temperatures also induce Hsp expression
in several species (25,26). Marine intertidal invertebrates undergo even larger
increases in body temperatures during tidal emersion (27–30). For example,
during aerial exposure intertidal mussels’ body temperatures exceed seawater
temperaturesbymorethan20
C(31), resultinginHspexpression(29).Asimilar
phenomenonoccurs in encystedbrine shrimp (Artemia) embryos (32, 33). Even
relatively mobile aquatic ectotherms such as fish may undergo heat shock in
nature (34,35). For example, gobiid fishes of the genus Gillichthys can become
trapped in shallow water, which is heated by the sun. Summer-acclimatized
fish have higher levels of Hsp90 in brain tissue than do winter-acclimatized
fish (34). In addition, the threshold Hsp induction temperature for one species,
G. mirabilis, is significantly higher in summer than in winter. These data sug-
gest that seasonal variation in water temperature can alter the heat-shock re-
sponse. More exotic venues for aquatic thermal stress include thermal effluents
ofpowerplants,hydrothermalvents,andthermalhotsprings, inwhichtemepra-
tures can exceed 100
C (see Hsps of Archaea).
Terrestrial Temperature Stress
Unlike aquatic environments, terrestrial environments often offer diverse heat
sourcesand sinks and retreats that organismscan exploittoavoidthermal stress.
Thus, natural thermal stress and accompanying Hsp expression in terrestrial
environments typically involve limitations in mitigating thermal extremes by
movement and conflicts between thermoregulation and other needs. Salaman-
ders, for example, which ordinarily maintain cool temperatures in nature, can
inadvertently retreat beneath small sunlit rocks that become warm enough to
induce hsp70 mRNA expression (36); by abandoning these rocks to find cooler
retreats they may risk immediate desiccation or even warmer temperatures.
The least equivocal case for routine exposure to Hsp-inducing temperature
stress is for plants, which cannot change location except as seeds or pollen and
can be limited in their ability to adjust heat exchange with the environment
(37). Thus, plants in nature can become extremely hot (38). By inference, the
entirerange of plant heat-shock responses (8, 39) should manifestthemselvesin
nature.Indeed, asmall number ofcasestudies document naturalHspexpression
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
HEAT-SHOCK PROTEINS 249
(40–45),whichcanbegreatestattimesofdayorinregionsofanindividualplant
at which temperatures are highest (46). Plant species can differ dramatically,
however, in both the magnitude and diversity of the particular Hsps that are
expressed during days with especially warm weather (41). Plants should also
be prone to natural cold stress (47), which ought to induce expression of Hsps
(48–51).
Not surprisingly, therefore, many of the cases of natural thermal stress in an-
imals on land involve animals that live inside or on plants (e.g. 52). Drosophila
larvae and pupae encounter temperatures exceeding 40
C if the necrotic fruit
they infest is in the sun, and express Hsp70 in response (53,54). Presumably,
other animals that cannot escape or offset intense solar heat loads will also
express Hsps in nature; this hypothesis awaits systematic study. A unique case
concerns desert ants, Cataglyphis, which voluntarily undergo body tempera-
tures of >50
C, presumably to escape predators (55). The concentration of
Hsp70 family members increases in this species before it naturally encounters
high temperatures, as if in anticipation (56).
Terrestrial vertebrates are often especially effective in escaping heat stress,
but both they and invertebrates are occasionally hyperthermic during intense
physical activity, fever, or to conserve water. In birds and mammals, such hy-
perthermia activates HSF (the heat-shock transcription factor) and increases
the level of Hscs (Hsp cognates, constitutively-expressed Hsps) and Hsps
(57–59). Natural hypothermia of animals can be far more conspicuous than
natural heat stress, involving diapause, overwinteringin exposed sites, hiberna-
tion, and sometimes outright freezing. Diverseinsects expressHsps in response
tocoldshockor during overwinteringin diapause, although theidentityofthese
Hsps, their tissue specificity, and their developmental regulation vary greatly
(60–64). Some euthermic rodents express 70-kDa Hsps in response to cold
ambient temperatures, possibly in tandem with nonshivering thermogenesis
(65), and ground squirrels (Spermophilus) increase Hsp70 family members and
ubiquitin-protein conjugates during hibernation (66).
In summary, laboratory studies of the heat-shock response often have pro-
ceeded far in advance of fieldwork that establishes an ecological context for
their interpretation. Documentation of both natural thermal stress and Hsp ex-
pression in nature can provide this context, and a small but growing number of
field studies demonstrate that such documentation is feasible.
Inducing Stresses Other than Temperature
Virtually every nonthermal stress can induce Hsps (10, 67). Rarely, however,
are these nonthermal stresses ecologically relevant; the literature in this area
typically focuses on chemical stressors, and the corresponding data are es-
sentially pharmacological. Even when the stress in question is ecologically
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
250 FEDER & HOFMANN
relevant, few studies of multicellular eukaryotes examine it in the field or in
intact tissues and organisms. Some exceptional work, however, concerns plants
and brine-shrimp (Artemia). The resurrection plant, a desert species, expresses
Hsps in vegetative tissues during water stress; this expression is thought to
contribute to desiccation tolerance (68). Similarly, rice seedlings express two
proteins in the Hsp90 family upon exposure to water stress and elevated salinity
(69). Embryosof thebrineshrimp, oneofthemost hypoxia-tolerantmetazoans,
contain largequantities of p26, a molecular chaperone hypothesized to stabilize
proteins during long bouts of anaerobic dormancy (see Development). Clearly,
additional evolutionary physiological research in this area is sorely needed.
A recurrent theme is that thermal stress and these alternative stressors often
result in different patterns of Hsp expression, indicating a diversity of regula-
tory mechanisms. Examples include variation in the expression of Hsp70 and
ubiquitin in the Drosophila central nervous system under anoxia (70), and in
protein expression during osmotic shock in isolated fish gill cells (71).
Bioindicators
Owing to its responsiveness to diverse forms of stress, the heat-shock response
hasundergonewidespreadapplication in biomonitoring andenvironmental tox-
icology (72–75). In many cases, Hsps are especially useful biomarkers because
their induction is much more sensitive to stress than traditional indices such
as growth inhibition. The use of Hsps as biomarkers is most widespread in
aquatic toxicology. Most of the literature demonstrates elevated Hsp levels or
induction of Hsps under laboratory conditions and then proposes Hsps as a
potential indicator of pollutants or toxins in the environment. For example,
exposing freshwater sponges to pollutants extracted from river water elevates
Hsp70 levels, which increase still further when thermal stress is also imposed
(76). Additional examples of Hsp expression in aquatic toxicology concern ro-
tifers, (77), marine sponges (78), amphipods (79), polychaetes (80), mollusks
(81–84), and fish (85–87). Other applications purposefully deploy organisms
in potentially polluted aquatic systems as biosensors (88,89).
Intheterrestrialenvironment,whereheavymetalcontaminationandpesticide
or herbicide accumulation can be critical problems, common soil organisms
suchasinvertebrates(90)areusefulHsp-biomonitorsoftoxicants. Forexample,
centipedes (Lithobius) collected from near a smelter had higher Hsp70 levels
than those collected from unpolluted areas (91). Potentially, combinations of
heavy metals can induce such distinctive patterns of Hsp expression in soil
nematodes that these patterns can become diagnostic fingerprints for specific
toxicants in soils (92).
Some aspects of the stress response, however, present problems for the use
of Hsps as biomarkers in environmental toxicology. Because so many different
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
HEAT-SHOCK PROTEINS 251
stresses induce Hsps, investigators may be unable to attribute changes in Hsp
expression to any particular environmental stress. Organisms in the field often
undergo multiple stresses simultaneously, the interaction of which can yield
significantHsp expressionevenwhenno single monitoredtoxicantis at harmful
levels. Conversely, Hsps induced by another stress can enhance tolerance of
a toxicant whose presence is being monitored. Laboratory studies support the
difficulty of teasing apart environmental factors and attributing Hsp induction
to a single stressor. For example, freshwater sponges exhibit greater tolerance
of pollutants following a sublethal heat stress (76). Among the vertebrates,
diseased fish have elevated levels of Hsps in their tissues, and disease-related
expressionmayinterferewiththeuseofHspsasabiomarker(93).Thus, because
numerous factors can induce Hsp expression and stress tolerance, the utility of
Hsps as biomarkers of environmental toxins may be limited.
ENVIRONMENTAL AND BIOLOGICAL CORRELATES
OF THE HEAT-SHOCK RESPONSE
AND HEAT-SHOCK PROTEINS
Many investigators view correlations of organismal traits (e.g. Hsp expression)
and environmental or biological variables (e.g. level of environmental stress,
developmentalstage,distinctiveroleinaparasiticorsymbioticrelationshipwith
another species) as prima facie evidence of biological adaptation, and thus have
actively sought such correlations in terms of the heat-shock response. While
the probative value of such evidence in establishing adaptation has met with
skepticism (94), in this section we consider the evidence for such correlations,
whatever their meaning.
Variation in the Stress Response Along Environmental
Gradients of Stress
To understand how Hsps result in stress tolerance at the organismal level, many
investigators have characterized the stress response along gradients that occur
in nature. One central question is whether organisms from environments with
littlestress havea differentorreducedstress response compared with organisms
from environments with much stress. Little and much stress might correspond
to the center and edge of a species’ range, low versus high elevation, xeric and
hot versus mesic and cool climate, temperate versus tropical/polar latitude, low
versus high intertidal, and so on. In general, the resulting data support a corre-
lation among Hsp expression, stress tolerance, and gradients of environmental
stress. Thesegradients havereceivedunevenattention, however, and theirstudy
has yielded mixed results. Comparative studies across many degrees of latitude
have not produced the same results as studies of gradients on smaller scales
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
252 FEDER & HOFMANN
(e.g. diurnal or microclimatic variation in stress). Currently, not a single study
has examined the stress response over the entire geographical distribution of
a species; thus, whether species at the extremes of their distributions have an
augmented heat-shock response is yet to be determined.
The majority of multi-species comparative studies focus on three aspects of
the stress response: the minimum (threshold) and maximum temperatures at
whichHspsareexpressedand/orarepresentincells,Hspconcentrationsincells,
and the diversity of the specific Hsps that are expressed. Except for the work on
threshold and maximum temperatures, much of this literature is a hodgepodge
of disconnected studies that are seldom comparable because of methodological
differencesand permitfewconclusionsotherthanthat speciesvaryinthedetails
of their stress response. Whether this variation has environmental correlates is
uncertain. A rare and exemplary exception is the work of Bosch and colleagues
on species of Hydra (95,96); below we discuss this and other similar work.
In general, the threshold temperature for Hspinduction is correlated withthe
typical temperatures at which species live, with thermophilic species having
a higher threshold than psychrophilic species. For example, a relatively cold-
water, northern species of mussel (Mytilus trossulus) has a lower threshold for
Hsp70 expession than its congener, M. galloprovincialis, a warm-water species
with a more southern distribution (97). Limpet species that occur in the upper
regions of the intertidal (Lottia digitalis and Lottia pelta) induce Hsps at 3–5
C
higher than the threshold for limpets that occur lower in the intertidal (Tectura
scutum) (AL Haag & GE Hofmann, unpublished data). Subtidal species of
the marine snail Tegula exhibit much the same pattern (98). Aggregate ex-
pression of Hsp70 family members (17) occurs at 3–4
C higher in Drosophila
melanogaster than in D. ambigua, a fruit fly of Palearctic origin (56). The
same study reports a similar pattern for the desert ant Cataglyphis and Formica
polyctena, a red wood ant from a temperate climate. One remarkable outcome
oftheCataglyphisstudyisthat Hspsynthesis inthedesert antscontinues attem-
peratures up to 45
C, whereas temperatures above 39
C inhibit Hsp synthesis
in the temperate species. A similar pattern (although not as extreme) is evident
for desert and non-desert Drosophila (22). These results suggest that translation
itself may have an upper thermal maximum that varies among species adapted
to different temperature environments.
Antarctic organisms represent a special case of psychrophily because the
temperatures they experience are both extremely cold and extremely stable.
In combination, do these conditions result in the evolutionary loss of a heat-
shock response? In the subtidal alga, Plocamium cartilagineum, heat-inducible
hsp70 and ubiquitin transcription still occur, although the threshold is a spec-
tacularly cold 5
C (99). Antarctic yeast species express Hsps at much lower
temperatures than does Saccharomyces, and at least one species lacks inducible
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
HEAT-SHOCK PROTEINS 253
thermotolerance (100, 101). In Antarctic fish, the picture is not as clear. Al-
though a broadly cross-reactive anti-Hsp70 antibody can detect isoforms of
Hsp70 in various tissues of the fish Trematomus bernacchii, heat shock tempe-
ratures from 6 to 10
C do not induce additional Hsp70 accumulation (GE
Hofmann, unpublished data). A member of the hsp70 gene family is present in
twoAntarcticfishspecies, T.bernacchiiandNototheniacoriiceps(ACWhitmer
& GE Hofmann, personal communication), and has been sequenced in Antarc-
tic fish species (102). At the other extreme, some hyperthermophilic Archaea
require temperatures in excess of 100
C to induce Hsp expression (see Hsps of
Archaea).
Thermal stress gradients can be seasonal as well as geographic. In some
cases, both Hsp expression and thermotolerance increase during warm seasons.
The intertidal mussel Mytilus californianus displays significantly different Hsp
induction profiles in summer than in winter, and summer-acclimatized mussels
induce Hsps at a threshold temperature that is 6
C higher than the threshold
in winter-acclimatized mussels (103). However, whether the accentuated Hsp
expression in mussels in summer results in greater thermotolerance at the or-
ganismal level is unknown. Fish (34, 35) and intertidal invertebrates (31) also
vary seasonally in Hsp expression.
In addition to work we cite elsewhere, other studies examine geographical
gradients in fish (104, 105, 105a), maize (106), reptiles (107), and Drosophila
(108); intertidal gradients in limpets (27); diurnal temperature change in Droso-
phila (109) and intertidal mussels (29); diurnal variation in spruce trees (46);
and seasonal variation in insects (110).
One issue for future consideration is whether Hsps in general are specialized
to function at higher temperatures than other proteins (especially enzymes),
and whether homologous Hsps of species from various thermal environments
have corresponding variation in thermostability of function (111). For exam-
ple, that an Hsp’s resistance to thermal denaturation varies according to the
thermal niche of the species that expresses it has been demonstrated for only
a single Hsp, alpha-crystallin (112). Another issue is how differing thresholds
of Hsp expression have evolved, whether through mutations in HSF, general
thermostability of proteins and cells (113), or some other mechanism (114).
The Parasitic Environment
The roles of Hsps in host-parasite interactions have received considerable at-
tention from both clinical and biological perspectives, with the majority of the
research in two general categories. First, from the perspective of the host,
Hsps expressed by invading parasites are potent antigens that elicit an immune
response (115–117); parasites’ Hsps are thus potentially useful in generating
vaccines (118). From the perspective of the parasite, the synthesis of Hsps is a
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
254 FEDER & HOFMANN
cellular defense mechanism that enables the parasite to live in different thermal
environments throughout its life-cycle (119). Parasites that infect mammalian
and avian hosts can undergo profound changes in temperature during the tran-
sition to these hosts (with internal temperatures of 37
C or above) from ec-
totherm hosts or free-living stages. Induction of Hsps commonly accompanies
thistransition. Numerous studies havedemonstrateddevelopmentallyregulated
expression of Hsps in parasites; expression differs throughout the life-cycle
both quantitatively and in the types of Hsps that temperature change induces.
For example, mRNA transcripts for hsp70 and hsp83 homologues increase up
to 100-fold as Trypanosoma brucei leaves the tsetse fly and enters a mam-
malian host (120). Aquatic snails release cercariae of the parasitic helminth,
Schistosoma mansoni, into freshwater; cercariae penetrate human skin and de-
velopintoadultworms,eventuallycausinglivercirrhosis. Thecercariaeexpress
two heat-inducible proteins that are not present in other stages (121).
Parasites that have an insect as the invertebrate vector have received much
attention with regard to the developmentally regulated expression of Hsps. Ex-
amples include parasitic nematodes (122); cestode parasites (123); the malarial
organism Plasmodium (124); Borrelia burgdorferi, the etiological agent of
Lyme disease (125); the protist Leishmania (126); Trypanosoma cruzi (127);
and Theileria (128).
Some parasite life cycles do not involve an animal vector; a free-living stage
of the parasite occurs in water or soil and enters the host. In several cases,
induction of Hsps accompanies the transition from the environment into the
host. Inthefungal parasiteHistoplasmacapsulatum, thetemperature shiftupon
infection of a mammalian host cues both the transformation from a mycelial
form to a budding yeast morphology and the expression of Hsps (129,130).
Eimeria, an intestinal parasite of numerous animals, expresses Hsp90 during
infective life stages. Eimeria parasites are particularly interesting because this
genus infects diverse hosts with correspondingly diverse body temperatures
(e.g. marine fish, poultry, and cattle). However, specificity of infection is high
at the species level, e.g. cattle are the exclusive host for Eimeria bovis (131).
Whether the heat-shock response of Eimeria co-evolved with its speciation into
these hosts is an open question.
Finally, the heat stress that infective life cycle stages of parasites experience
is as diverse as their hosts. In nature, parasites of ectotherms can encounter
dramatic shifts in temperature when their hosts’ body temperature varies, as has
been reported for parasites of reptiles, fish, and intertidal organisms (132,133).
Symbiosis
Just as Hsps may play an important role in parasitism, in which one species
maintains a close but antagonistic relationship with others, they also function
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
HEAT-SHOCK PROTEINS 255
in symbiosis, in which interspecific relationships can be equally close but not
adversarial. Perhapsthe mostgeneral example of this point concerns mitochon-
dria and chloroplasts, which evolved from endosymbionts that colonized other
cells early in the history of life. These organelles often require proteins that
are encoded in the nuclear genome and synthesized by the host cell, and hence
must be imported into the organelle. Hsps play diverse roles in this process in
mitochondria. A cytoplasmic Hsp70 family member maintains peptides in an
unfolded conformation, which enables the peptides topass through pores inthe
mitochondrial membrane; a mitochondrial Hsp70 is part of theprotein machin-
ery that imports the peptide; and the Hsp60/Hsp10 apparatus participates in the
folding of the imported protein (134). Several groups of primitive eukaryotes
contain still other endosymbiotically derived organelles, the hydrogenosome
and the nucleomorph, whose Hsps share a characteristic sequence with those
of mitochondria and proteobacteria (135–138). The Hsp sequence similarities
have been used to suggest that hydrogenosomes may derivefrom mitochondria,
share a common origin with mitochondria, or represent independent coloniza-
tions of early eukaryotic cells (135–137, 139).
Asidefromendosymbioticallyderivedorganelles,thebest-studiedsymbioses
concern bacterial endosymbionts that infect insects, including aphids, flies,
ants, and cockroaches. Aphids, for example, harbor the bacterium Buchnera in
specializedcells(bacteriocytes)within a distinctive structure in the body cavity,
the bacteriome (140). The bacteria express a protein, symbionin, at especially
high levels, and this protein is a member of the GroEL (Hsp60) family. Other
bacterial chaperones, including GroES (Hsp10) and DnaK (Hsp70), are also
present at high levels (140). A similar phenomenon is evident in tsetse flies
(141).
The function and significance of these high Hsp levels is enigmatic. The
Hsps apparently are not a response of the endosymbionts to a novel (and there-
fore stressful) host environment, as the Buchnera/aphid symbiosis is 150–250
million years old. The endosymbionts, however, havebeen evolving at an espe-
ciallyhigh rate; thus, theelevatedmolecularchaperones could be compensating
for decreased protein stability due to the accumulation of numerous amino acid
substitutions (142,143). Nonetheless, the bacteria themselves can mount a
strong heat-shock response when their host undergoes stress (144). Other rel-
evant symbioses include X-bacteria in the symbiosomes of Amoeba (e.g. 145),
Bradyrhizobium and Rhizobium in the root nodules of nitrogen-fixing plants
(146), and the zooxanthellae component of corals (26, 30). A recurrent theme
is that the endosymbionts modify the amount and/or diversity of Hsps present
in the symbiosis. In some cases, this modification is thought to contribute to
the maintenance of the endosymbionts within the host, and in others to the
augmentation of the heat-shock response of the symbiosis as a whole. Finally,
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
256 FEDER & HOFMANN
Wolbachia, a bacterial endosymbiont that infects millions of arthropod species,
both interferes with the mating of infected and uninfected hosts and can alter
their constitutive expression of Hsp70 and Hsp90 family members (147). Sim-
ulated natural heat stress can diminish this reproductive interference, possibly
by overriding the symbiont’s effect on the host Hsps.
Development
Many species exhibit characteristic and distinctive patterns of Hsp expression
(ornonexpression)duringthe variousstagesofdevelopment,includinggameto-
genesis, embryogenesis, and metamorphosis (e.g. 148–151). These patterns are
oftenconsistentwithenhancedstressresistanceindevelopmentalstagesthaten-
counter unusual levels of environmental stress or during circumstances such as
dormancyand diapause (see below). In other cases, developmental programs of
Hsp expression ensue even in the absence of any obvious environmental stress.
One common pattern is that one or more Hsps are not expressed in the initial
phases of embryogenesis (152–156) or late in gametogenesis (157–161), pos-
sibly because Hsps can be harmful to developing cells (see Deleterious Aspects
of Hsps). Parental provision of Hsps or hsp mRNAs can sometimes override
gametic or embryonic absence of Hsp expression (162,163); in other cases this
absence presumably poses a significant problem for continued development
in the face of stress (164). Stress not only can kill vulnerable developmental
stages outright, but also can produce lasting damage to surviving organisms,
such as the phenocopying of genetic defects; Hsps may minimize such defects
(165,166).
Adaptational analyses of the developmental expression of Hsps are diverse.
Some plant seeds presumably must endure extremes of heat, desiccation, and
other stresses before germinating, and some must germinate under especially
challenging conditions. However, although seeds clearly undergo developmen-
tally regulated expression of Hsps and embryos can express Hsps in response
to environmental stress (167–169), few investigators have considered whether
these patterns of expression are amplified or modified in species and ecotypes
that naturally encounter especially challenging stress regimes (68, 170). Our
stateofknowledgeis similar for fungal spores, which expressparticularHsps in
a developmentally regulated program (e.g. 171,172). Several interesting case
studies are available for animals, although a general pattern is yet to emerge.
In the most spectacular example, encysted brine-shrimp (Artemia) embryos
undergo developmental arrest, in which they may survive for years without
environmental water or oxygen. The encysted embryos accumulate enormous
concentrations (15% of total protein) of a small Hsp (173–177) and treha-
lose (178), and suppress ubiquitination of damaged proteins (179). Non-adult
D. melanogaster infest necrotic fruit, which can become extremely warm if it
is sunlit (53,54); this species mounts a massive heat-shock response, which
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
HEAT-SHOCK PROTEINS 257
is greatest in the developmental stages that presumably undergo the most ex-
posure to natural heat stress (18,19,180). Other flies overwinter while at a
particular developmental stage, and undergo considerable Hsp expression in
response to cold (60–63). Later in development, ubiquitin may assist in the
degeneration of flight muscles that are no longer needed after nuptial flights of
insects (181). Finally, the temperature threshold for expression of Hsps may
itself undergo modification; e.g. the threshold decreases in mammalian testis,
in which gametogenesis normally occurs at lower temperatures than in the core
of the body (182,183).
Aging and Senescence
As mammals age, damage to proteins progressively accumulates, and both the
ability to express Hsps (e.g. Hsp or hsp mRNA levels after a standard exposure
to heat or other stress) and stress tolerance deteriorate (184,185). Moreover,
individualHspscan become less able to mitigate the effectsof stress on proteins
as mammals age (186). In ecological and evolutionary terms, whether similar
Hsp-aging relationships are important or even evident in wild organisms is
unknown,althoughtheserelationshipsoccurindiversespeciesinthelaboratory:
Drosophila (187,188), nematodes (189–191), and Daphnia (192).
These findings have provoked great interest in how Hsps potentially affect
senescence and lifespan. A unifying hypothesis in the foregoing work is that
proteindamage, due primarily to oxidation/free radical activity, gradually accu-
mulates during the life of a cell or organism and can lead to death if unabated;
Hsps and other molecular stress responses ordinarily can mitigate this dam-
age to some extent, and the decreasing expression of Hsps with age therefore
contributes to mortality. If this hypothesis is correct, then treatments that both
reduce damage to protein and increase Hsp expression (e.g. heat shock) should
prolong life. In nematodes (Caenorhabditis elegans), some single-gene muta-
tions that increase lifespan are associated with increased thermotolerance, but
through as-yet-undescribed mechanisms (189–191). In Drosophila, heat shock
extends lifespan (193), and this extension is enhanced in flies transformed with
additional copies of the hsp70 gene (194). Nutrient deprivation can also extend
lifeinrodents,presumablybyreducingthemetabolicrateandconsequently,ox-
idative damage to proteins; starvation, however, variously increases, decreases,
or has no effect on Hsp expression (195–199).
FITNESS CONSEQUENCES OF HSP EXPRESSION
General Issues and Beneficial Aspects of Hsps
Understandingthe consequences of variationin Hspsandthestress response for
Darwinian fitness requires a detailed appreciation of the mechanisms by which
Hsps mitigate the impact of stress on individuals in natural populations. These
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
258 FEDER & HOFMANN
mechanisms are becoming well understood at the level of model proteins with
whichHspscaninteract, butareprogressivelylesswellunderstoodatthelevelof
the cell, tissue, organ, and whole organism. At the level of the model protein,
various stresses clearly either directly or indirectly result in conformational
change, and Hsps typically promote the reacquisition or maintenance of the
native structure and function by minimizingthe tendencyof non-native proteins
to interact inappropriately (200,201). In cells, stress-induced conformational
change, protein aggregation, and molecular chaperoning of model proteins are
also well established (200–202), and many cellular components differin stress-
tolerant and stress-intolerant cells (67).
Two primary issues impede the linkage of variation in these well-established
mechanisms and phenomena to variation in the fitness of individual complex
multicellular eukaryotes. First, is the variation in sensitivity to stress among
cells, cell types, tissues, organs, and organisms attributable to a small number
of critical lesions, especially sensitive targets of stress and functions of specific
Hsps in protecting or repairing these lesions/targets? Or is variationin sensitiv-
ity to stress an aggregate function of a widespread and diverse impact of stress
on cellular structures, with Hsps mitigating multiple lesions in diverse ways
(21)? The former alternative may be more analytically tractable than the latter.
Second, given that cells and organisms may have multiple Hsps in each Hsp
family, multiple Hsp families, and multiple non-Hsp mechanisms of stress mit-
igation, howcan we unambiguously establish the contributionor importance of
anyparticularHsp, Hsp family,ormechanism in the complexcell, tissue, organ,
or organism? Much of the published literature on the functional consequences
of Hsp expression for whole organisms or the cells they comprise runs afoul of
these issues. Literally thousands of studies report correlations between Hsp ex-
pression, diverse biological functions in the face of stress, and stress tolerance,
but these typically conclude that their findings are at best consistent with a role
of one or more Hsps in stress tolerance. Evaluating the roles of single factors
in complex systems is an ongoing challenge in most areas of the biological
sciences, and the heat-shock field largely has not yet deployed counterparts of
the solutions that other fields have developed.
One conspicuous and major exception includes techniques and approaches,
primarily drawn from molecular biology and genetics, that allow the manipu-
lation of individual Hsps or specific genes that encode them. In rare instances,
a species naturally may have an unusual genetic system (203) or a diminished
suite of Hsps (95,204) that accomplishes the same end; also, several chemical
compounds may specifically inhibit one or more Hsps (e.g. 205,206). The gen-
eral implication of the resulting work (Table 1) is that, even in whole organisms
or the cells they comprise, variation in single Hsps can be consequential for
fitness. Some specific implications are as follows: Individual Hsps can have
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
HEAT-SHOCK PROTEINS 259
Table 1 Phenotypes of multicellular eukaryotes, and the cells and tissues that they comprise,
for which Hsps are necessary and/or sufficient
a
Protein Phenotype
Hsp10 Cellular: tolerance of ischemia (no phenotype) (308); tolerance of ischemia when
co-expressed with Hsp60 (308)
Hsp27 Cellular: resistance to chemotherapeutic drugs (309); resistance to hydrogen
peroxide (310,311); resistance to hydrogen peroxide (no phenotype) (312);
resistance to ultraviolet radiation (no phenotype) (312); resistance of tumor
cells to monocytes (311); sensitivity to lymphokine-activated killer cells (no
phenotype) (311); tolerance of hyperthermia (312–316); resistance to tumor
necrosis factor (317) (310); tolerance of ischemia (318); resistance of actin
polymers to cytochalasin (314); accelerated nuclear protein aggregation (319);
accelerated decline of thermal radiosensitization (319)
Crystallin Cellular: tolerance of hyperthermia (320,321); tolerance of ischemia (318);
resistance to tumor necrosis factor (310); resistance to hydrogen peroxide (310)
Hsp60 Cellular: tolerance of hyperthermia (no phenotype) (322,323); tolerance of
ischemia (no phenotype) (308,322, 323); tolerance of ischemia when
co-expressed with Hsp10 (308)
Hsp65 Cellular: tumor regression (324); loss of tumorigenicity (325)
Tissue/organ: regression of malignant tumors (324)
Hsp70 Cellular: tolerance of hyperthermia (326) (316, 322, 323, 327–342); tolerance of
ischemia/hypoxia (322, 323, 340, 343–345); recovery from translational and
transcriptional inhibition following heat shock (335); regulation of heat-shock
response (331,346, 347); tolerance of endotoxin (348); reduced protein
denaturation upon heat exposure (349); tumorigenicity (350); cell proliferation
(351,164); resistance to hydrogen peroxide (311); resistance of tumor cells to
monocytes (311); sensitivity to lymphokine-activated killer cells (no phenotype)
(311); escape from drug-induced cell cycle arrest (352); protein glycosylation
(353); tolerance of ultraviolet radiation (354); apoptosis (351,355,356);
resistance to apoptosis (no phenotype) (328,329)
Tissue/organ: recovery of contractility after ischemia (345, 357–359); reduction
in myocardial infarct size (345,359); reduction of hyperthermic damage to
midgut (221); resistance of heart to ischemic injury (357–359); resistance of
hippocampus to ischemic injury (360)
Organismal: tolerance of hyperthermia (18–20, 109, 156, 203, 221, 224, 278, 361);
growth and development (222); regulation of heat-shock response (361);
persistence in nature (no phenotype) (277)
Hsc70 Organismal: tolerance of hyperthermia (203)
Hsp72 Cellular: apoptosis (no phenotype) (362); protection against heat-induced nuclear
protein aggregation (319); protection against hypoxia (363); protection against
thermal radiosensitization (319)
Tissue/organ: reduction in myocardial infarct size (364)
Grp78 Cellular: protein secretion (229–231)
(Continued)
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
260 FEDER & HOFMANN
Table 1 (Continued)
Protein Phenotype
Hsp90 Cellular: tolerance of hyperthermia (322,323,328, 329,340); tolerance of
ischemia (no phenotype) (322, 323, 340); apoptosis (362); apoptosis
(no phenotype) (328, 329); cell proliferation and cell cycle control (365);
glucocorticoid receptor function (205)
Hsp100 Organismal: host infection in Leishmania (126)
Hsp101 Cellular: tolerance of hyperthermia (366)
Many Hsps Cellular: recovery of cell proliferation after heat shock (367); recovery from
chromosome damage after heat shock (367–369); tolerance of hyperthermia
(370, 371); tolerance of ischemia (336)
HSF Organismal: oogenesis and development (372); thermotolerance (372, 373)
a
In all cited work, specific Hsps have undergone experimental or natural manipulation.
pleiotropic effects, interacting with multiple systems in diverseways. Findings
from manipulations of individual Hsps usually (but not always) are consistent
with the outcome of correlative studies (see above). Finally, despite the huge
body of work on Hsps and the growing use of manipulative techniques, we
have remarkably little physiological insight into exactly how the activity of
Hsps culminates in the enhanced stress tolerance of multicellular eukaryotes
and the cells and tissues that they comprise.
Interestingly, one clear conclusion that correlative studies have yielded is
that Hsps cannot account for the entirety of inducible stress tolerance (207–
217). Indeed, some component of inducible stress tolerance may be unrelated
to protein synthesis in general (214,218, 219).
Deleterious Aspects of Hsps
The many advantages of the heat-shock response suggest that natural selection
should maximize the expression of Hsps. By contrast, the genes encoding Hsps
have not undergone unlimited amplification in the genome, and the Hsps them-
selves are subject to strict autoregulation by multiple molecular mechanisms
(220). These contrary findings suggest that Hsps can have both positive and
negativeimpacts onfitness, and that natural selectionmayhave acted to balance
these impacts in setting the level of Hsp expression. For example, while small
to moderate increases in Hsp70 levels enhance inducible thermotolerance in
Drosophila, large increases in Hsp70 levels actually decrease thermotolerance
(221); evolution thus may favor an intermediate level of Hsp70. A common
theme in related work is that high levels of Hsps may be especially detrimental
to cells or developmental stages in which cell growth and division proceed at
high rates. Drosophila larvae transformed with extra copies of the hsp70 gene
have greater larva-to-adult mortality and slower development than do control
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
HEAT-SHOCK PROTEINS 261
larvae; these strain differences are proportional to the number of Hsp-inducing
heatshocks administered to the larvae(222). LarvaenaturallyvaryinginHsp70
expression display a similar pattern (223). Drosophila cells engineered to ex-
press Hsp70 constitutively at first grow more slowly than control cells, but
subsequently resume control growth rates once the Hsp70 is sequestered from
the cytoplasm (164); indeed, Drosophila embryos remove Hsp70 from their
cells rapidly after heat shock (224). A yeast strain that cannot express Hsp104
grows faster than its wild-type counterpart on some media (171). More gen-
erally, most animal species that have been studied do not mount a heat-shock
responseduringearlystagesofembryogenesis(seeDevelopment),whenprotein
synthesis may be especially intense.
These negative effects may have at least two nonexclusive explanations
(222,225, 226): First, Hsps at high concentration could be toxic, directly in-
terfere with ongoing processes in the cell, or otherwise alter function to the
detriment of the cell (220). Second, the synthesis and degradation of Hsps
could consume an intolerably large fraction of a cell’s or organism’s nutrient
and energy stores, and/or occupy so large a fraction of the synthetic/catabolic
apparatusthattheprocessingofotheressentialbiomoleculessuffers(226–228).
Consistent with the first explanation, cellular sequestration of Hsp70 is corre-
lated with the resumption of proliferation in cells constitutively expressing this
protein (164). Also, overexpression of an Hsp70 family member inhibits pro-
tein secretion and reduction increases secretion in mammalian cells in culture
(229–231); excess amounts of another Hsp70 family member canpromote pro-
tein aggregation in vitro (M Borrelli & J Lepock, personal communication);
and Hsp70 can perturb the normal structures of nascent polypeptides (232).
Tests of the second explanation have manipulated the costs of or resources
for Hsp expression. Growth of corn in nitrogen-rich soil increases the syn-
thesis of Hsps in response to a standard heat shock (233); in plants grown in
nitrogen-poor soil, other proteins may be catabolized to supply amino acids
for synthesis of Hsps (234). These findings suggest that Hsp synthesis can be
nitrogen-limited in plants. Starvation reduces the expression of Hsp 70 family
members in mice (195). In Drosophila larvae, by contrast, co-expression of
β-galactosidase and Hsps has no greater cost for growth and development than
does expression of Hsps alone (225). Further study of this apparent trade-off of
the benefits and disadvantages of Hsp expression, moreover, has the potential
to link evolutionary and mechanistic views of this problem that heretofore have
been separate (222,225).
MICROEVOLUTIONARY VARIATION IN HSPS
Hspsareroutinelytoutedasadaptationsthataroseandaremaintainedvianatural
selection for stress resistance. Origin and maintenance of a trait by selection
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
262 FEDER & HOFMANN
require that it vary within populations, and that this intra-population variation
have a genetic basis and affect the Darwinian fitness of individuals. Here we
ask whether Hsps, the genes that encode them, and the factors that modify
their expression display such patterns of variation and undergo stabilizing or
directional selection in response to environmental stress.
First, not all intrapopulation and intraspecific variation results from genetic
differences. For example, seasonal acclimatization and temperature acclima-
tion in the laboratory can alter the minimum temperature at which Gillichthys,
a gobiid fish, expresses an Hsp90 family member (34,235). Seasonal acclima-
tization likewise affects Hsp70 levels in mussels (Mytilus) (103), and routine
culture temperature affects the magnitude and temporal pattern of Hsp expres-
sion in HeLa cells (236). Such changes may stem from alterations in the cellu-
lar environment that modify the activation of HSF (113,182, 236,237). These
changes, however, are not universal; laboratory thermal acclimation does not
alter the thermal sensitivity of Hsp expression in fish hepatocytes in culture
(238), Drosophila larvae (19), and mussels (103).
Evenwhenacclimationandseasonalchangearecontrolled,however, individ-
uals within a population or species may varyin Hsp expressionand/or the genes
that determine it. Relevant research has examined this issue on two levels: di-
rectsequencevariationandrestrictionfragmentlengthpolymorphisms(RFLPs)
(240). Thesequenceof hsp70 variesamongstrainsof the parasiteTrypanosoma
(241,242) andthenematode C. elegans (243), and among conspecificsforsome
but not all of the mammalian hsp70 family members (240), as does that of the
3
0
untranslated region of hsp27 in normotensive and hypertensive rats (244).
RFLPs consistent with intraspecific variation either in the hsp genes or flanking
regions are detectable in the hsp60 and hsp70 of the spirochete Borrelia (245),
in multiple hsp70 family members of mammals (240,246, 247), and in several
plant species (248–250). One putative instance of intraspecific variation in hsp
copy number concerns D. melanogaster, in which at least five nearly identical
copies of hsp70 occur at two chromosomal loci. At locus 87A7, two copies are
arranged as an inverted repeat (251,252); at 87C1, two copies flank a region
containing at least one additional copy (252,253) plus numerous α/β repeats,
which encode heat-inducible mRNAs of no proven function (254–257). Up
to five additional hsp70 copies have been reported, with copy number varying
among strains and time of year (253,255, 258–261). However, these reports
either cannot exclude that such variation is actually in intergenic regions or that
the reports are for Drosophila cells in culture or mutagenized laboratory strains
rather than wild or even wild-type strains. The organization of the two chromo-
somal loci reportedly variesamong natural populations (253). A less equivocal
instance of evolutionary change in gene copy number concerns Arabidopsis,in
which ecotypes vary inthe number of ubiquitin-encoding repeats (262). For all
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
HEAT-SHOCK PROTEINS 263
of the foregoing reports, the functional significance of intraspecific variation
awaits elucidation or direct verification.
Hspexpressionalsoexhibitsgeneticvariationamongindividualsofaspecies;
often,thisvariationiscorrelatedwithstressresistance (248,250, 263–271). For
example, isofemale lines of Drosophila founded from a single wild population
differ more than twofold in Hsp70 expression; this variation is correlated with
thermotolerance and is heritable (180,223). Similarly, in the pathogenic fun-
gus Histoplasma, naturally temperature-insensitive strains express more hsp70
mRNA and do so at lower temperatures than in a temperature-sensitive strain
(272). In humans, fibroblasts isolated from desert-dwelling Turkmen express
more Hsps and have greater thermotolerance than fibroblasts from residents of
more equable climates (273); presumably, however, these peoples do not differ
in body temperature.
Given that the patterns of variation necessary for natural selection occur
within species, that selection can alter Hsp expression is not surprising. Lab-
oratory evolution at high temperatures paradoxically lowers Hsp70 expression
and inducible thermotolerance in Drosophila (19,274), and selection for re-
sistance to hyperthermic paralysis alters both the hsp68 promoter and the hsr-
omega locus in Drosophila (254,275). Additional findings relevant to natural
selection and its underlying genetic basis come from closely related species,
some so similar that they hybridize. In the fish species Poeciliopsis monacha
and P. lucida, an unusual genetic system permits the generation of hemi-
clonal lines in which the paternal genome varies against a constant mater-
nal genome. Hemiclonal thermotolerance was most strongly related to Hsc70
and only secondarily to Hsp70 levels, in a pattern consistent with straightfor-
ward Mendelian inheritance of parental genotypes and adaptation to the local
thermal environment. By contrast, the heat-shock response of interspecific hy-
brids of tomato (Lycopersicon) is not intermediate to the parental responses
(276). Non-hybridizing congeners often exhibit a correlation among the ac-
tual or inferred incidence of thermal stress in their environment, heat-shock
response, and stress tolerance. Such data are now available for diverse an-
imals (see Variation in the Stress Response Along Environmental Gradients
of Stress).
A particular problem with such species comparisons is that the interpreta-
tion of the observed patterns is readily confounded by phylogenetic and sta-
tistical issues. A more general problem with both laboratory evolution and
species comparisons is that they describe only a supposed correlation of the
heat-shock genes/proteins of interest with evolution and seldom can establish
the importance of the genes/proteins of interest to evolutionary process and
outcome (19,20). Study of free-ranging organisms (277) with hsp transgenes
(e.g. 109,278) may contribute much to resolving these problems.
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
264 FEDER & HOFMANN
THE EVOLUTIONARY HISTORY OF HSPS
AND THE GENES THAT ENCODE THEM
Hsps are among the most ancient and highly conserved of all proteins. Homo-
logues of Hsps occur in every species in which they have been sought, and in
all kingdoms of living things. Thus, Hsps represent a remarkable example of
molecular “descent with modification” at the levels of gene sequence, genomic
organization, regulation of gene expression, and protein structure and function.
So clear are the patterns of descent and modifications that they can be used to
establish the evolutionary origins and the phylogenetic affinities of the major
groups of organisms.
Hsps of Archaea and Exceptional Prokaryotes
The Archaea or archaebacteria are the most extremophilic and most primi-
tive organisms. The heat-shock response of extremophilic Archaea and nonar-
chaeal extremophiles occurs at remarkably high temperatures (279), e.g. 88
C
in Sulfolobus (279) and >100
C in the hyperthermophilic species designated
ES4, a heterotrophic sulfur reducer isolated from a deep-sea hydrothermal vent
(280). The archaeal genome encodes homologues of most Hsps represented
in other prokaryotes and eukaryotes (279), as well as their consensus pro-
moter sequences (281). Notably, the archaeal Hsp60 homologues assemble
into a dual ring-like structure, termed a rosettasome or thermosome, that re-
sembles the structure that the chaperones GroEL and GroES form in bacteria
(279,282, 283). The archaeal structures have ATPase activity and can bind de-
natured proteins (282). At least some Archaea and Eubacteria differ in the
number of monomers that comprise these structures (284). Surprisingly, the
archaeal Hsp60s (e.g. TF55 of Sulfolobus) most closely resemble not a bac-
terial homologue, but the eukaryotic protein TCP1, which assembles into the
t-complex polypeptide 1 ring complex (TRiC) in the cytosol (285, 286). Pre-
viously, Hsp60 homologues were thought to be absent from the eukaryotic cy-
tosol. AgrowingbodyofevidencesuggeststhatTCP1/TRiCandGroEL/GroES
play comparable roles in their respective organisms and cellular compartments
(287–289). Meanwhile, Trent (290) has suggested that the primary function of
TF55 may be cytoskeletal, with molecular chaperoning a secondary or derived
function.
Genes encoding Hsps are present even in the smallest of genomes. These
include the genomes of mycoplasmas (291) and the nucleomorph, the vestigial
nucleusofaphototrophiceukaryoticendosymbiontincryptomonadalgae(138).
Thesection on symbiosis (see above)reviews the distributionof Hsps in various
other organelles of endosymbiotic origin. Apparently, the problem of protein
folding is ancient and ubiquitous, necessitating molecular chaperones in these
diverse cases.
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
HEAT-SHOCK PROTEINS 265
Large-Scale Evolution of hsp Genes
The extraordinarily conserved nature of hsp genes (292) has facilitated their
cloning, sequencing, and comparison in diverse organisms; their evolution is
now becoming understood in detail. Gupta and colleagues have undertaken the
most extensive surveys of hsp sequences, with a particular focus on organisms
deemedcriticaltounderstandingtherelationshipsofmajortaxa(292–297). The
interpretations resulting from these comparisons relate to hypotheses about
(a) the origin of eukaryotic cells, the eukaryotic nucleus, and endoplasmic
reticulum (292,296); (b) polyphyletic versus monophyletic origin of the major
bacterial groups (292–294); and (c) the validity of the three-domain (Archaea,
bacteria, and eukaryotes) dogma (292,294). Whereas these interpretations are
controversial, if not revolutionary, and therefore have not received universal
acceptance, theynonetheless clearly illustrate howcomparativeanalyses of hsp
genes may address fundamental issues in evolutionary history.
On a less grand scale, hsp gene families represent superb case studies of how
one or a small number of primitive genes can diversify to encode a suite of
compartment- and function-specific proteins. One of many examples is dnaK,
a single gene in Archaea and bacteria that has become the complex multi-
genehsp70 familiesofSaccharomyces(298), Drosophila (253,299), and Homo
(300). Another example concerns the small Hsps, which evolution recruited
to become a major component of the lens of the eye: alpha crystallin (301).
A growing body of work examines the discrete evolutionary events by which
thesechangesmay have occurred, includinggeneduplication/conversionevents
(302), retrotransposition, horizontal exchange of genomes, and others. New
technologies promise to advance this work exponentially.
Discrete Examinations of Molecular Evolution
Ideally,a complete study of theevolutionary physiologyofHspsmight examine
how the following co-evolve as populations or how closely related species
enter environments in which they face novel stresses: sequence (both coding
and non-coding) of the gene(s) for a particular Hsp, regulation of hsp gene
expression, the role and importance of the Hsp in stress resistance, and the
intensities and durations of stress that the populations and species actually face.
Much of the evolutionary physiological investigation of Hsps fails to attain this
admittedly ambitious goal for one or more reasons: (a) The species under study
aretoodistantlyrelatedtoreconstructthefunctional,environmental,andgenetic
events during their divergence; (b) molecular biology, manipulative genetics,
physiology, and environmental assessment are not all possible for the species
in question; or (c) the breadth of the techniques and approaches necessary to
perform such research is too daunting for a single research program. Two case
studies exemplify both how this goal could be approached and how far the field
has yet to go to attain it.
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
266 FEDER & HOFMANN
The coelenterate Hydra oligactis and several of its congeners are the only
multicellular eukaryotes reported not to express Hsps in response to heat shock
and other stresses. Other congeners (e.g. Hydra attenuata and H. magnipapil-
lata) have a well-developed stress response; these and other data for puta-
tive ancestors of Hydra suggest an evolutionary loss of Hsp expression in
H. oligactis (95, 96). Physiologically and ecologically, H. oligactis is defi-
cient in inducible stress tolerance and disappears from certain habitats in na-
ture during periods of stress (95). Subsequent work suggests that, at least
for Hsp70, the loss has occurred due to mutations that affect the stability of
hsp70 mRNA, as H. oligactis hasan hsp70 gene and expresses a heat-inducible
hsp70 mRNAin quantities similar to that in the heat-tolerant H. magnipapillata
(96).
Dipteran insects and their ancestors have undergone an evolutionary prolif-
eration of hsp70 genes. Mosquitoes and Drosophila share a distinctly arranged
duplication of the inducible hsp70 gene (303), suggesting that this prolifer-
ation predates the original diversification of the Diptera. Within the genus
Drosophila, all groups other than the melanogaster subgroup of species ap-
parently retain the primitive copy number of two (108,304,305). Within the
melanogaster subgroup, all species examined to date have four copies except
for D. melanogaster (253), which has at least five hsp70 copies in its haploid
genome(seeMicroevolutionaryVariationinHsps). Curiously,D.melanogaster
expresses no Hsp100 family members, which are critical for thermotolerance
in other organisms (204,306). The proliferation of hsp70 copies is correlated
with the ecological and biogeographic distribution of Drosophila species (19).
Whereas most Drosophila species have small geographic ranges or narrow
ecological niches, two of the melanogaster subgroup species (simulans and
melanogaster) have cosmopolitan distributions, and a third (yakuba) is eco-
logically diverse throughout sub-Saharan Africa.
CONCLUSION
Ecological and evolutionary physiological analysis of heat-shock proteins may
be nearing the end of its initial descriptive phase. Although accounts of spec-
tacular levels of Hsp-mediated stress resistance and exceptional consequences
of Hsp expression will continue to be newsworthy, the major patterns of Hsp
expression in multicellular eukaryotes are becoming so obvious that additional
descriptive work is becoming increasingly difficult to justify. Clearly, how-
ever, major questions remain unanswered. How the activities of Hsps at the
molecularlevel culminate in organismalstress tolerance and how thehsp genes,
their regulation, the function of the proteins they encode, and the environments
faced by the organisms in which they occur all co-evolve are but two of the
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
HEAT-SHOCK PROTEINS 267
unresolved issues reviewed here. The perspective of evolutionary physiology
can make significant contributions to the resolution of these and other issues.
By placing results in actual environmental contexts, by assessing phenotypes
of Hsps in the context of whole multicellular organisms, and by characterizing
extant and historical variation in Hsps in natural populations and taxa, evolu-
tionary physiologists can complement and extenda spectacular area of research
that has been largely restricted to the molecular/cellular levels in the laboratory.
By the same token, insights and techniques that laboratory-based investigators
provide promise to continue to revolutionize the ecological and evolutionary
study of Hsps. These approaches are both logical partners and necessary com-
plements to one another (22,307). Our understanding of Hsps has much to gain
from the continued if not expanded synergy of these approaches.
A
CKNOWLEDGMENTS
We thank BR Bettencourt, K Chavez, UT D’OBrador, AP Nguyen, J Meredith-
Patla, and AC Whitmer for bibliographic assistance and/or editorial advice,
and Susan Lindquist for founding the bibliographic database that made this re-
view possible. Research was supported by National Science Foundation grants
97-23298 and 97-23063. Finally and most importantly, we beg the indulgence
of numerous investigators whom space limitations precluded us from citing
directly here; their contributions have made the field what it is today. We urge
allreaderscontemplatingresearch inthisareatoscan thecompletebibliography
(13) available online at http://www.AnnualReviews.org.
Visit the Annual Reviews home page at
http://www.AnnualReviews.org
Literature Cited
1. Ritossa F. 1996. Discovery of the heat
shock response. Cell Stress Chaperones
1:97–98
2. Lindquist S. 1986. The heat-shock re-
sponse. Annu. Rev. Biochem. 55:1151–
91
3. Gething MJ, Sambrook J. 1992. Protein
folding in the cell. Nature 355:33–45
4. Morimoto RI, Tissieres A, Georgopou-
los C, eds. 1994. Heat Shock Pro-
teins: Structure, Function and Regu-
lation. Cold Spring Harbor, NY: Cold
Spring Harbor Lab. Press
5. Hartl FU. 1996. Molecular chaper-
ones in cellular protein folding. Nature
381:571–80
6. Gething MJ, ed. 1997. Guidebook to
Molecular Chaperones and Protein-
Folding Catalysts. Oxford, UK: Oxford
Univ. Press
7. Thomas PJ, Qu BH, Pedersen PL. 1995.
Defective protein folding as a basis of
human disease. Trends Biochem. Sci.
20:456–59
8. Boston RS, Viitanen PV, Vierling E.
1996. Molecular chaperones and pro-
tein folding in plants. Plant Mol. Biol.
32:191–222
9. Feige U, Morimoto RI, Yahara I, Polla
BS, eds. 1996. Stress-Inducible Cellular
Responses. Basel: Birkh¨auser
10. Feder ME, Parsell DA, Lindquist SL.
1995.The stress responseand stresspro-
teins. In Cell Biology of Trauma, ed. JJ
Lemasters, C Oliver, pp. 177–91. Boca
Raton, FL: CRC
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
268 FEDER & HOFMANN
11. Nover L, ed. 1991. Heat Shock Re-
sponse. Boca Raton, FL: CRC
12. Somero GN. 1995. Proteins and temper-
ature. Annu. Rev. Physiol. 57:43–68
13. Feder ME, Hofmann GE. 1998. Evo-
lutionary and ecological physiology of
heat-shock proteins and the heat-shock
response: a comprehensive bibliogra-
phy. http://www.AnnualReviews.org
14. Stephanou G, Alahiotis SN, Christo-
doulou C, Marmaras VJ. 1983. Adap-
tation of Drosophila melanogaster to
temperature. Heat-shock proteins and
survival in Drosophila melanogaster.
Dev. Genet. 3:299–308
15. Alahiotis SN, Stephanou G. 1982. Tem-
perature adaptation of Drosophilapopu-
lations. The heat shock proteins system.
Comp. Biochem. Physiol. 73B:529–33
16. Palter KB, Watanabe M, Stinson L,
Mahowald AP, Craig EA. 1986. Ex-
pression and localization of Drosophila
melanogaster hsp70 cognate proteins.
Mol. Cell Biol. 6:1187–203
17. Hightower LE. 1995. Desert ants. Sci-
ence 268:1417
18. Feder ME. 1998. Engineering candi-
date genes in studies of adaptation: the
heat-shock protein Hsp70 in Drosophila
melanogaster. Am. Nat. In press
19. Feder ME, Krebs RA. 1998. Natural and
genetic engineering of thermotolerance
in Drosophila melanogaster. Am. Zool.
38:503–17
20. Feder ME, Krebs RA. 1997. Ecologi-
cal and evolutionary physiology of heat-
shock proteins and the stress response
in Drosophila: complementary insights
from genetic engineering and natural
variation. In Stress, Adaptation, and
Evolution, ed. R Bijlsma, V Loeschcke,
pp. 155–73. Basel: Birkh¨auser
21. Feder ME. 1996. Ecological and evo-
lutionary physiology of stress proteins
and the stress response: the Drosophila
melanogaster model. In Animals and
Temperature: Phenotypic and Evolu-
tionary Adaptation, ed. IA Johnston, AF
Bennett, pp. 79–102. Cambridge, UK:
Cambridge Univ. Press
22. Huey RB, Bennett AF. 1990. Physiolog-
ical adjustments to fluctuating thermal
environments: an ecological and evolu-
tionary perspective.In Stress Proteins in
BiologyandMedicine, ed.RIMorimoto,
ATissieres,CGeorgopoulos,pp.37–59.
Cold Spring Harbor, NY: Cold Spring
Harbor Lab. Press
23. Huey RB. 1991. Physiological conse-
quences of habitat selection. Am. Nat.
137:S91–115
24. Bartholomew GA. 1964. The roles of
physiology and behaviour in the mainte-
nance of homeostasis in the desert envi-
ronment. In Homeostasis and Feedback
Mechanisms, ed. GMHughes, pp. 7–29.
Cambridge, UK: Cambridge Univ.Press
25. Sharp VA, Brown BE, Miller D. 1997.
Heat shock protein (HSP 70) expression
in the tropical reef coral Goniopora dji-
boutiensis. J. Therm. Biol. 22:11–19
26. Hayes RL, King CM. 1995. Induction
of 70-kD heat shock protein in scler-
actinian corals by elevated temperature:
significance for coral bleaching. Mol.
Mar. Biol. Biotechnol. 4:36–42
27. Sanders BM, Hope C, Pascoe VM, Mar-
tin LS. 1991. Characterization of the
stress protein response in two species
of Collisella limpets with different
temperature tolerances. Physiol. Zool.
64:1471–89
28. Sanders BM, Pascoe VM, Nakagawa
PA, Martin LS. 1992. Persistence of the
heat-shock response over time in a com-
mon Mytilus mussel. Mol. Mar. Biol.
Biotechnol. 1:147–54
29. Hofmann GE, Somero GN. 1996. Pro-
tein ubiquitination and stress protein
synthesis in Mytilus trossulus occurs
during recovery from tidal emersion.
Mol. Mar. Biol. Biotechnol. 5:175–84
30. Sharp VA, Miller D, Bythell JC, Brown
BE. 1994. Expression of low molecu-
lar weight HSP 70 related polypeptides
fromthesymbioticsea anemoneAnemo-
nia viridis Forskal in response to heat
shock. J. Exp. Mar. Biol.Ecol. 179:179–
93
31. Hofmann GE, Somero GN. 1995. Ev-
idence for protein damage at environ-
mental temperatures: seasonal changes
in levels of ubiquitin conjugates and
hsp70 in the intertidal mussel Mytilus
trossulus. J. Exp. Biol. 198:1509–18
32. MillerD,McLennan AG.1988.Theheat
shock response of the cryptobiotic brine
shrimp Artemia. I. A comparison of the
thermotolerance of cysts and larvae. J.
Therm. Biol. 13:119–24
33. MillerD,McLennan AG.1988.Theheat
shock response of the cryptobiotic brine
shrimp Artemia. II. Heat shock proteins.
J. Therm. Biol. 13:125–34
34. Dietz TJ,Somero GN. 1992. The thresh-
old induction temperature of the 90-
kDa heat shock protein is subject to ac-
climatization in eurythermal goby fishes
(genus Gillichthys). Proc. Natl. Acad.
Sci. USA 89:3389–93
35. Fader SC, Yu Z, Spotila JR. 1994. Sea-
sonal variation in heat shock proteins
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
HEAT-SHOCK PROTEINS 269
(hsp70) in stream fish under natural con-
ditions. J. Therm. Biol. 19:335–41
36. Near JC, Easton DP, RutledgePS, Dick-
inson DP, Spotila JS. 1990. Heat shock
protein70 geneexpression inintact sala-
manders Eurycea bislineata in response
to calibrated heat shocks and to high
temperatures encountered in the field. J.
Exp. Zool. 256:303–14
37. Gates DM. 1980. Biophysical Ecology.
New York: Springer-Verlag
38. Kee SC, Nobel PS. 1986. Concomitant
changes in high-temperature tolerance
and heat-shock proteins in desert succu-
lents. Plant Physiol. 80:596–98
39. Nagao RT, Kimpel JA, Key JL. 1990.
Molecular and cellular biology of
the heat-shock response. Adv. Genet.
28:235–74
40. Nguyen HT, Joshi CP, Klueva N, Weng
J, Hendershot KL, Blum A. 1994. The
heat-shock response and expression of
heat-shock proteins in wheat underdiur-
nal heat stress and field conditions. Aust.
J. Plant Physiol. 21:857–67
41. Hamilton EW, Heckathorn SA, Downs
CA,Schwarz TE, ColemanJS, Hallberg
RL. 1996. Heat shock proteins are pro-
duced by field-grown naturally occur-
ring plants in the summer in the temper-
ate northeast U.S. Bull. Ecol. Soc. Am.
77, Suppl. Part 2:180 (Abstr.)
42. Kimpel JA, Key JL. 1985. Presence of
heat shock mRNAs in field grown soy-
beans. Plant Physiol. 79:672–78
43. Hernandez LD, Vierling E. 1993. Ex-
pression of low molecluar weight heat-
shock proteins under field conditions.
Plant Physiol. 101:1209–16
44. Hendershot KL, Weng J, Nguyen HT.
1992. Induction temperature of heat-
shock protein synthesis in wheat. Crop
Sci. 32:256–61
45. Burke JJ, Hatfield JL, Klein RP, Mullet
JE. 1985. Accumulation of heat shock
proteins in field-grown cotton. Plant
Physiol. 78:394–98
46. Colombo SJ, Timmer VR, Colclough
ML, Blumwald E. 1995. Diurnal varia-
tioninheat toleranceandheatshock pro-
tein expression in black spruce (Picea
mariana). Can. J. Forest Res. 25:369–
75
47. Morris GJ, Coulson G, Meyer MA,
McLellan MR, Fuller BJ, et al. 1983.
Cold shock—a widespread cellular re-
action. Cryo-Letters 4:179–92
48. Danyluk J, Rassart E, Sarhan F. 1991.
Gene expression during cold and heat
shock in wheat. Biochem. Cell Biol.
69:383–91
49. Krishna P, Sacco M, Cherutti JF, Hill
S. 1995. Cold-induced accumulation of
hsp90 transcripts in Brassica napus.
Plant Physiol. 107:915–23
50. Neven LG, Haskell DW, Guy CL,
Denslow N, Klein PA, et al. 1992. Asso-
ciation of 70-kilodalton heat-shock cog-
nate proteins with acclimation to cold.
Plant Physiol. 99:1362–69
51. Van Berkel J, Salamini F, Gebhardt C.
1994. Transcripts accumulating during
cold storage of potato (Solanum tubero-
sum L.) tubers are sequence related to
stress-responsive genes. Plant Physiol.
104:445–52
52. Layne JR. 1991. Microclimate variabil-
ity and the eurythermic natural of gold-
enrod gall fly (Eurosta solidaginis) lar-
vae(Diptera: Tephritidae). Can. J. Zool.
69:614–17
53. Feder ME, Blair N, Figueras H. 1997.
Natural thermal stress and heat-shock
protein expression in Drosophila larvae
and pupae. Funct. Ecol. 11:90–100
54. Feder ME. 1997. Necrotic fruit: a novel
model system for thermal ecologists. J.
Therm. Biol. 22:1–9
55. Wehner R, Marsh AC, Wehner S. 1992.
Desert ants on a thermal tightrope. Na-
ture 357:586–87
56. Gehring WJ, Wehner R. 1995. Heat
shockproteinsynthesis andthermotoler-
ance in Cataglyphis, an ant from the Sa-
hara desert. Proc. Natl. Acad. Sci. USA
92:2994–98
57. Locke M, Noble EG. 1995. Stress pro-
teins: the exercise response. Can. J.
Appl. Physiol. 20:155–67
58. Brown IR, Rush SJ. 1996. In vivo ac-
tivation of neural heat shock transcrip-
tion factor HSF1 by a physiologically
relevant increase in bodytemperature. J.
Neurosci. Res. 44:52–57
59. Di YP, Repasky EA, Subjeck JR. 1997.
DistributionofHSP70, protein kinaseC,
and spectrin is altered in lymphocytes
during a fever-like hyperthermia expo-
sure. J. Cell. Physiol. 172:44–54
60. Joplin KH, Denlinger DL. 1990. Devel-
opmental and tissue specific control of
the heat shock induced 70 kDa related
proteins in the flesh fly, Sarcophaga
crassipalpis. J. Insect Physiol. 36:239–
49
61. Joplin KH, Yocum GD, Denlinger DL.
1990. Cold shock elicits expression of
heat shock proteins in the flesh fly Sar-
cophaga crassipalpis. J. Insect Physiol.
36:825–34
62. Yocum GD, Joplin KH, Denlinger DL.
1991. Expression of heat shock proteins
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
270 FEDER & HOFMANN
in response to high and low temperature
extremes in diapausing pharate larvae of
the gypsy moth Lymantria dispar. Arch.
Insect Biochem. Physiol. 18:239–50
63. Lee RE, Dommel RA, Joplin KH, Den-
linger DL. 1995. Cryobiology of the
freeze-tolerant gall fly Eurosta solidagi-
nis: overwintering energetics and heat
shock proteins. Climate Res. 5:61–67
64. Denlinger DL, Lee RE, Yocum GD,
KukalO.1992.Roleofchillinginthe ac-
quisitionofcoldtoleranceandthecapac-
itation to express stress proteins in dia-
pausingpharatelarvaeofthegypsymoth
Lymantriadispar.Arch. Insect Biochem.
Physiol. 21:271–80
65. MatzJM,LaVoiKP,MoenRJ,BlakeMJ.
1996. Cold-induced heat shock protein
expression in rat aorta and brown adi-
pose tissue. Physiol. Behav. 60:1369–74
66. Sills NS, Gorham DA, Carey HV. 1998.
Stress protein expression in a mam-
malian hibernator. FASEB J. 12:A379
67. Li GC, Nussenzweig A. 1996. Thermo-
tolerance and heat shock proteins: pos-
sible involvement of Ku autoantigen in
regulatingHsp70expression. SeeRef. 9,
pp. 121–37
68. Alamillo J, Almoguera C, Bartels D,
Jordano J. 1995. Constitutive expres-
sion of small heat shock proteins
in vegetative tissues of the resurrec-
tion plant Craterostigma plantagineum.
Plant Mol. Biol. 29:1093–99
69. Pareek A, Singla SL, Kush AK, Grover
A.1997. DistributionpatternsofHSP 90
protein in rice. Plant Sci. 125:221–30
70. Ma E, Haddad GG. 1997. Anoxia reg-
ulates gene expression in the cen-
tral nervous system of Drosophila
melanogaster. Brain Res. Mol. Brain
Res. 46:325–28
71. Kultz D. 1996. Plasticity and stressor
specificity of osmotic and heat shock
responses of Gillichthys mirabilis gill
cells. Am. J. Physiol. 271:C1181–93
72. de Pomerai D. 1996. Heat-shock pro-
teins as biomarkers of pollution. Hum.
Exp. Toxicol. 15:279–85
73. Sanders BM, Dyer SD. 1994. Cellular
stress response. Environ. Toxicol. Chem.
13:1209–10
74. Sanders BM. 1993. Stress proteins in
aquatic organisms: an environmental
perspective. Crit.Rev.Toxicol.23:49–75
75. Ryan JA, Hightower LE. 1996. Stress
proteins asmolecular biomarkers for en-
vironmental toxicology. See Ref. 9, pp.
411–24
76. Mueller WEG, Koziol C, Kurelec B,
Dapper J, Batel R, Rinkevich B. 1995.
Combinatory effects of temperature
stressandnonionic organicpollutantson
stressprotein (hsp70)gene expressionin
the freshwater sponge Ephydatia fluvi-
atilis. Environ.Toxicol.Chem. 14:1203–
8
77. Cochrane BJ, Irby RB, Snell TW. 1991.
Effects of copper and tributylin on
stress protein abundance in the rotifer
Brachionus plicatilis. Comp. Biochem.
Physiol. 98C:385–90
78. Krasko A, Scheffer U, Koziol C, Pancer
Z, Batel R, et al. 1997. Diagnosis of
sublethal stress in the marine sponge
Geodia cydonium: application of the
70 kDa heat-shock protein and a novel
biomarker, the Rab GDP dissociation
inhibitor, as probes. Aquat. Toxicol.
37:157–68
79. Werner I, Nagel R.1997. Stress proteins
HSP60 and HSP70 in 3 species of am-
phipods exposed to cadmium, diazinon,
dieldrin and fluoranthene. Environ. Tox-
icol. Chem. 16:2393–403
80. Ruffin P, Demuynck S, Hilbert JL,
Dhainaut A. 1994. Stress protein in the
polychaete annelid Nereis diversicolor
induced by heat shock or cadmium ex-
posure. Biochimie 76:423–27
81. SteinertSA, PickwellGV.1988. Expres-
sion of heat shock proteins and metal-
lothionein in mussels exposed to heat
stress and metal ion challenge. Mar. En-
viron. Res. 24:211–14
82. Veldhuizen Tsoerkan MB, Holwerda
DA, van der MastCA, ZandeeDI. 1991.
Synthesis of stress proteins under nor-
mal and heat shockconditions ingill tis-
sue of sea mussels (Mytilus edulis) af-
terchronic exposureto cadmium.Comp.
Biochem. Physiol. 100C:699–706
83. Nascimento IA, Dickson KL, Zimmer-
man EG. 1996. Heat shock protein re-
sponse to thermal stress in the Asi-
atic clam, Corbicula fluminea. J. Aquat.
Ecosystem Health 5:231–38
84. Sanders BM, Martin LS, Howe SR,
Nelson WG, Hegre ES, Phelps DK.
1994. Tissue-specific differences in ac-
cumulation of stress proteins in Mytilus
edulis exposed toa range of copper con-
centrations. Toxicol. Appl. Pharmacol.
125:206–13
85. Ryan JA, Hightower LE. 1994. Evalua-
tion of heavy-metal ion toxicity in fish
cells using a combined stress protein
and cytotoxicity assay. Environ. Toxicol.
Chem. 13:1231–40
86. Dyer SD, Brooks GL, Dickson KL,
Sanders BM, Zimmerman EG. 1993.
Synthesis and accumulation of stress
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
HEAT-SHOCK PROTEINS 271
proteins in tissues of arsenite-exposed
fathead minnows Pimephales promelas.
Environ. Toxicol. Chem. 12:913–24
87. Vijayan MM, Pereira C, Forsyth RB,
Kennedy CJ, Iwama GK. 1997. Han-
dling stress does not affect the expres-
sionof hepatic heat shockprotein 70 and
conjugation enzymes in rainbow trout
treated with beta-naphthoflavone. Life
Sci. 61:117–27
88. Van Dyk TK, Majarian WR, Kon-
stantinov KB, Young RM, Dhurjati
PS, LaRossa RA. 1994. Rapid and
sensitive pollutant detection by induc-
tionofheatshockgene-bioluminescence
gene fusions. Appl. Environ. Microbiol.
60:1414–20
89. Veldhuizen Tsoerkan MB, Holwerda
DA,deBontAM,SmaalAC, ZandeeDI.
1991. A field study on stress indices in
the sea mussel, Mytilus edulis: applica-
tion of the “stress approach” in biomon-
itoring. Arch. Environ. Contam. Toxicol.
21:497–504
90. Kohler HR, Triebskorn R, Stocker W,
KloetzelPM,AlbertiG.1992.The70kD
heat shock protein (hsp 70) in soil inver-
tebrates: a possible tool for monitoring
environmental toxicants. Arch. Environ.
Contam. Toxicol. 22:334–38
91. Pyza E, Mak P, Kramarz P, Laskowski
R.1997. Heat-shockproteins (Hsp70)as
biomarkers in ecotoxicological studies.
Ecotoxicol. Environ. Safety 38:244–51
92. Stringham EG, Candido EPM. 1994.
Transgenic hsp16-lacZ strains of the
soil nematode Caenorhabditis elegans
as biological monitors of environmen-
tal stress. Environ. Toxicol. Chem. 13:
1211–20
93. Forsyth RB, Candido EPM, Babich SL,
Iwama GK. 1997. Stress protein ex-
pression in coho salmon with bacterial
kidney disease. J. Aquat. Anim. Health
9:18–25
94. Rose MR, Lauder GV, eds. 1996. Adap-
tation. New York/London: Academic
95. BoschTC,KrylowSM,BodeHR,Steele
RE. 1988. Thermotolerance and synthe-
sis of heat shock proteins: These re-
sponses are present in Hydra attenuata
butabsentinHydraoligactis.Proc. Natl.
Acad. Sci. USA 85:7927–31
96. Gellner K, Praetzel G, Bosch TC. 1992.
Cloning and expression of a heat-
inducible hsp70 gene in two species of
Hydra which differ in their stress re-
sponse. Eur. J. Biochem. 210:683–91
97. Hofmann GE, Somero GN. 1996. In-
terspecific variation in thermal denat-
uration of proteins in the congeneric
mussels Mytilus trossulus and M. gal-
loprovincialis: evidence from the heat-
shock response and protein ubiquitina-
tion. Mar. Biol. 126:65–75
98. Tomanek L, Somero GN. 1997. The ef-
fect of temperature on protein synthe-
sis in snails of the genus Tegula from
the sub- and intertidal zone. Am. Zool.
37:188A
99. Vayda ME, Yuan ML. 1994. The heat
shock response of an Antarctic alga is
evident at 5 degrees C. Plant Mol. Biol.
24:229–33
100. Berg GR, Inniss WE, Heikkila JJ. 1987.
Stress proteins and thermotolerance in
psychrotrophic yeasts from Arctic envi-
ronments. Can. J. Microbiol. 33:383–89
101. DeegenaarsML, WatsonK. 1997. Stress
proteins and stress tolerance in an
Antarctic, psychrophilic yeast, Candida
psychrophila. FEMS Microbiol. Lett.
151:191–96
102. Carratu L, Maresca B. 1997. Evolution-
aryadaptationofhsp70geneinAntarctic
fish. Exp. Biol. Online 2:C5.1 (Abstr.)
103. Roberts DA, Hofmann GE, Somero
GN. 1997. Heat-shock protein expres-
sion in Mytilus californianus: acclima-
tization (seasonal and tidal-height com-
parisons) and acclimation effects. Biol.
Bull. 192:309–20
104. Norris CE, diIorio PJ, Schultz RJ,
Hightower LE. 1995. Variation in heat
shockproteins within tropical and desert
species of poeciliid fishes. Mol. Biol.
Evol. 12:1048–62
105. White CN, Hightower LE, Schultz RJ.
1994. Variation in heat-shock proteins
among species of desert fishes (Poe-
ciliidae, Poeciliopsis). Mol. Biol. Evol.
11:106–19
105a. Maresca B, Patriarcha E, Goldenberg C,
Sacco M. 1988. Heat shock and cold
adaptation in Antarctic fishes: a molec-
ular approach. Comp. Biochem. Physiol.
90B:623–29
106. Ristic Z, Williams G, Yang G, Martin B,
Fullerton S. 1996. Dehydration, damage
to cellular membranes, and heat-shock
proteins in maize hybrids from different
climates. J. Plant Physiol. 149:424–32
107. Ulmasov KA, Shammakov S, Karaev K,
EvgenevMB. 1992. Heat shock proteins
and thermoresistance in lizards. Proc.
Natl. Acad. Sci. USA 89:1666–70
108. Konstantopoulou I, Drosopoulou E,
Scouras ZG. 1997. Variations in the
heat-induced protein pattern of several
Drosophila montium subgroup species
(Diptera: Drosophilidae). Genome 40:
132–37
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
272 FEDER & HOFMANN
109. Feder ME, Carta˜no NV, Milos L, Krebs
RA, Lindquist SL. 1996. Effect of engi-
neering Hsp70 copy number on Hsp70
expression and tolerance of ecologically
relevant heat shock in larvae and pu-
pae of Drosophila melanogaster. J.Exp.
Biol. 199:1837–44
110. Nath BB, Lakhotia SC. 1989. Heat-
shock response in a tropical Chirono-
mus: seasonal variation in response and
theeffectofdevelopmentalstageandtis-
sue typeon heat shock protein synthesis.
Genome 32:676–86
111. Hofmann GE. 1996. Molecular chaper-
one activity of the stress protein Hsc70
purified from an eurythermal goby,
Gillichthys mirabilis. Am. Zool. 36:36A
112. McFall-Ngai M, Horwitz J. 1990. A
comparative study of the thermal stabil-
ity of the vertebrate eye lens: Antarctic
fish to the desert iguana. Exp. Eye Res.
50:703–9
113. Clos J, Rabindran S, Wisniewski J, Wu
C. 1993. Induction temperature of hu-
man heat shock factor is reprogrammed
inaDrosophilacellenvironment.Nature
364:252–55
114. Carratu L, Franceschelli S, Pardini CL,
Kobayashi GS, Horvath I, et al. 1996.
Membrane lipid perturbation modifies
the set point of the temperature of heat
shock response in yeast. Proc. Natl.
Acad. Sci. USA 93:3870–75
115. Maresca B, KobayashiGS. 1994. Hsp70
in parasites: as an inducible protective
protein and as an antigen. Experientia
50:1067–74
116. Kaufmann SH. 1992. The cellular im-
mune response to heat shock proteins.
Experientia 48:640–43
117. Polla BS. 1991. Heat shock proteins in
host-parasite interactions. Immunol. To-
day 12:A38–41
118. Newport GR. 1991. Heat shock proteins
as vaccine candidates. Semin. Immunol.
3:17–24
119. Tsuji N, Ohta M, Fujisaki K. 1997. Ex-
pression of a 70-kDa heat-shock-related
protein duringtransformation from free-
living infective larvae to the parasitic
stage in Strongyloides venezuelensis.
Parasitol. Res. 83:99–102
120. Van der Ploeg LH, Giannini SH, Cantor
CR. 1985. Heat shock genes: regulatory
role for differentiation in parasitic pro-
tozoa. Science 228:1443–46
121. Neumann S, Ziv E, Lantner F, Schechter
I. 1993. Regulation of HSP70 gene ex-
pression during the life cycle of the par-
asitic helminth Schistosoma mansoni.
Eur. J. Biochem. 212:589–96
122. vanLeeuwenMA.1995.Heat-shockand
stressresponse ofthe parasiticnematode
Haemonchus contortus. Parasitol. Res.
81:706–9
123. Ernani FP, Teale JM. 1993. Release of
stress proteins from Mesocestoides corti
isa brefeldin A-inhibitable process: evi-
denceforactiveexportofstress proteins.
Infect. Immun. 61:2596–601
124. Syin C, Goldman ND. 1996. Cloning of
a Plasmodium falciparum gene related
to the human 60-kDa heat shock pro-
tein. Mol. Biochem. Parasitol. 79:13–
19
125. Cluss RG, Boothby JT. 1990. Ther-
moregulation of protein synthesis in
Borrelia burgdorferi. Infect. Immun.
58:1038–42
126. Hubel A, Krobitsch S, Horauf A, Clos J.
1997. The Leishmania major Hsp100 is
required chiefly in the mammalian stage
ofthe parasite.Mol. Cell.Biol. 17:5987–
95
127. Giambiagi-de Marval M, Souto-Padron
T, Rondinelli E. 1996. Characterization
and cellular distribution of heat-shock
proteins HSP70 and HSP60 in Try-
panosomacruzi.Exp.Parasitol.83:335–
45
128. Daubenberger C, Heussler V, Gobright
E, Wijngaard P, Clevers HC, et al. 1997.
Molecular characterisation of a cognate
70 kDa heat shock protein of the pro-
tozoan Theileria parva. Mol. Biochem.
Parasitol. 85:265–69
129. Shearer GJ, Birge CH, Yuckenberg
PD, Kobayashi GS, Medoff G. 1987.
Heat-shock proteins induced during the
mycelial-to-yeast transitions of strains
of Histoplasma capsulatum. J. Gen. Mi-
crobiol. 133:3375–82
130. Maresca B. 1995. Unraveling the secrets
of Histoplasma capsulatum. A model
to study morphogenic adaptation during
parasite host/host interaction. Verh. K.
Acad. Geneeskd. Belg. 57:133–56
131. ClarkTG,Abrahamsen MS, White MW.
1996. Developmental expression of heat
shock protein 90 in Eimeria bovis. Mol.
Biochem. Parasitol. 78:259–63
132. Tirard CT, Grossfeld RM, Volety AK,
Chu FLE. 1995. Heat shock proteins of
the oyster parasite Perkinsus marinus.
Dis. Aquat. Organisms 22:147–51
133. Ulmasov KA, Ovezmukhammedov A,
KaraevKK, EvgenevMB. 1988. Molec-
ular mechanisms of adaptation to hy-
perthermia in higher organisms. III. In-
duction of heat-shock proteins in two
Leishmania species. Mol. Biol. 22:
1583–89
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
HEAT-SHOCK PROTEINS 273
134. Ryan MT, Naylor DJ, Hoj PB, Clark
MS, Hoogenraad NJ. 1997. The role of
molecular chaperones in mitochondrial
protein import and folding. Int. Rev. Cy-
tol. 174:127–93
135. Germot A, Philippe H, Le Guyader H.
1997. Evidence for loss of mitochondria
in microsporidia from a mitochondrial-
type HSP70 in Nosema locustae. Mol.
Biochem. Parasitol. 87:159–68
136. Bui ET, Bradley PJ, Johnson PJ. 1996.
A common evolutionary origin for mi-
tochondria and hydrogenosomes. Proc.
Natl. Acad. Sci. USA 93:9651–56
137. Sogin ML. 1997. Organelle origins:
energy-producingsymbiontsinearlyeu-
karyotes? Curr. Biol. 7:R315–17
138. Hofmann CJ, Rensing SA, Hauber MM,
Martin WF, Muller SB, et al. 1994. The
smallest known eukaryotic genomes en-
code a protein gene: towards an under-
standingofnucleomorphfunctions.Mol.
Gen. Genet. 243:600–4
139. Germot A, Philippe H, Le Guyader
H. 1996. Presence of a mitochondrial-
type 70-kDa heat shock protein in Tri-
chomonas vaginalis suggests a very
early mitochondrial endosymbiosis in
eukaryotes. Proc. Natl. Acad. Sci. USA
93:14614–17
140. Baumann P, Moran NA, Baumann
L. 1997. The evolution and genetics
of aphid endosymbionts. BioScience
47:12–20
141. Aksoy S. 1995. Molecular analysis of
the endosymbionts of tsetse flies: 16S
rDNA locus and over-expression of a
chaperonin. Insect Mol. Biol. 4:23–29
142. Moran NA, Von Dohlen CD, Baumann
P. 1995. Faster evolutionary rates in en-
dosymbiotic bacteria than in cospeciat-
inginsecthosts. J.Mol. Evol. 41:727–31
143. Moran NA. 1996. Accelerated evolu-
tion and Muller’s rachet in endosymbi-
otic bacteria. Proc. Natl. Acad. Sci. USA
93:2873–78
144. Morioka M, Ishikawa H. 1992. Mutual-
ism based on stress: selective synthesis
and phosphorylation of a stress protein
byanintracellularsymbiont.J.Biochem.
111:431–35
145. Hong HK, Choi JY, Ahn TI. 1994.
Molecularbiologicalstudieson theheat-
shock responses in Amoeba proteus:I.
Detection of heat-shock proteins. Ko-
rean J. Zool. 37:554–64
146. Choi EY, Ahn GS, Jeon KW. 1991. El-
evated levels of stress proteins associ-
ated with bacterial symbiosis in Amoeba
proteus and soybean root nodule cells.
Biosystems 25:205–12
147. Feder ME, Karr TL. 1997. Evolution-
arily significant consequences of the
heat shock response for Drosophila and
its endosymbiont Wolbachia. Am. Zool.
37:8A
148. Dix DJ. 1997. Hsp70 expression and
function during gametogenesis. Cell
Stress Chaperones 2:73–77
149. Lin JC, Song CW. 1993. Heat shock
gene expression and development. I. An
overview of fungal, plant, and poikilo-
thermic animal developmental systems.
Dev. Genet. 14:1–5
150. Winter J, Sinibaldi R. 1991. The expres-
sion of heat shock protein and cognate
genes during plant development. Results
Prob. Cell Differ. 17:85–105
151. Mosser DD, Duchaine J, Bourget L,
Martin LH. 1993. Heat shock gene
expression and development. II. An
overview of mammalian and avian de-
velopmental systems. Dev. Genet. 14:
87–91
152. Heikkila JJ, Ohan N, Tam Y, Ali A.
1997. Heat shock protein gene expres-
sion during Xenopus development. Cell.
Mol. Life Sci. 53:114–21
153. Muller WU, Li GC, GoldsteinLS. 1985.
Heat does not induce synthesis of heat
shock proteins or thermotolerance in the
earliest stage of mouseembryo develop-
ment. Int. J. Hypertherm. 1:97–102
154. Edwards JL, Ealy AD, Monterroso
VH, Hansen PJ. 1997. Ontogeny of
temperature-regulated heat shock pro-
tein 70 synthesis in preimplantation
bovine embryos. Mol. Reprod. Dev. 48:
25–33
155. Dura JM. 1981. Stage dependent syn-
thesis of heat shock induced pro-
teins in early embryos of Drosophila
melanogaster. Mol. Gen. Genet. 184:
381–85
156. Welte MA, Tetrault JM, Dellavalle RP,
Lindquist SL. 1993. A new method for
manipulating transgenes: engineering
heat tolerance in a complex, multicel-
lular organism. Curr. Biol. 3:842–53
157. Gagliardi D, Breton C, Chaboud A,
VergneP, Dumas C. 1995. Expression of
heat shock factor and heat shock protein
70 genes during maize pollen develop-
ment. Plant Mol. Biol. 29:841–56
158. Hendrey J, Kola I. 1991. Thermolabil-
ity of mouse oocytes is due to the lackof
expressionand/or inducibilityof Hsp70.
Mol. Reprod. Dev. 28:1–8
159. Curci A, Bevilacqua A, Mangia F. 1987.
Lack of heat-shock response in pre-
ovulatory mouse oocytes. Dev. Biol.
123:154–60
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
274 FEDER & HOFMANN
160. Curci A, Bevilacqua A, Fiorenza MT,
Mangia F. 1991. Developmental regu-
lation of heat-shock response in mouse
oogenesis: identification of differen-
tially responsive oocyte classes during
Graafianfollicledevelopment.Dev.Biol.
144:362–68
161. Zakeri ZF, Welch WJ, Wolgemuth DJ.
1990. Characterization and inducibility
of hsp 70 proteins in the male mouse
germ line. J. Cell Biol. 111:1785–92
162. Bedard PA, Brandhorst BP.1986. Trans-
lational activation of maternal mRNA
encoding the heat-shock protein hsp90
during sea urchin embryogenesis. Dev.
Biol. 117:286–93
163. Gordon S, Bharadwaj S, Hnatov A, Ali
A, Ovsenek N. 1997. Distinct stress-
inducible and developmentally regu-
lated heat shock transcription factors in
Xenopus oocytes. Dev. Biol. 181:47–63
164. Feder JH, Rossi JM, Solomon J,
Solomon N, Lindquist S. 1992. The
consequences of expressing Hsp70 in
Drosophilacellsatnormaltemperatures.
Genes Dev. 6:1402–13
165. Mitchell HK, Moller G, Petersen NS,
Lipps-Sarmiento L. 1979. Specific pro-
tection from phenocopy induction by
heat shock. Dev. Genet. 1:181–92
166. WelteMA, Duncan I, LindquistS. 1995.
The basis for a heat-induced develop-
mental defect: defining crucial lesions.
Genes Dev. 9:2240–50
167. Wehmeyer N, Hernandez LD, Finkel-
stein RR, Vierling E. 1996. Synthesis of
small heat-shock proteins is part of the
developmentalprogram oflateseedmat-
uration. Plant Physiol. 112:747–57
168. Duck N, McCormick S, Winter J. 1989.
Heat shock protein Hsp70 cognate gene
expression in vegetative and reproduc-
tive organs of Lycopersicon esculentum.
Proc. Natl. Acad. Sci. USA 86:3674–78
169. Coca MA, Almoguera C, Jordano J.
1994. Expression of sunflower low-
molecular-weight heat-shock proteins
during embryogenesis and persistence
after germination: localization and pos-
sible functionalimplications. Plant Mol.
Biol. 25:479–92
170. Helm KW, Petersen NS, Abernethy RH.
1989. Heat-shock response of germinat-
ing embryos of wheat: effectsof imbibi-
tion time and seed vigor. Plant Physiol.
90:598–605
171. Sanchez Y, Taulien J, Borkovich KA,
Lindquist S. 1992. Hsp104 is required
for tolerance to many forms of stress.
EMBO J. 11:2357–64
172. Silva AM, Juliani MH, da Costa JJ,
Bonato MC. 1987. Acquisition of ther-
motolerance during development of
Blastocladiella emersonii. Biochem.
Biophys. Res. Commun. 144:491–98
173. Liang P, Amons R, Clegg JS, MacRae
TH. 1997.Molecularcharacterization of
a small heat shock/alpha-crystallin pro-
tein in encysted Artemia embryos. J.
Biol. Chem. 272:19051–58
174. Liang P, Amons R, Macrae TH, Clegg
JS. 1997. Purification, structure and in
vitro molecular-chaperone activity of
Artemia p26, a small heat-shock/alpha-
crystallin protein. Eur. J. Biochem. 243:
225–32
175. Jackson SA, Clegg JS. 1996. Ontogeny
of low molecular weight stress protein
p26 during early development of the
brine shrimp, Artemia franciscana. J.
Exp. Biol. 200:467–75
176. Clegg JS, Jackson SA, Liang P, MacRae
TH. 1995. Nuclear-cytoplasmic translo-
cations of protein p26 during aerobic-
anoxictransitionsinembryos ofArtemia
franciscana. Exp. Cell Res. 219:1–7
177. Clegg JS, Jackson SA, Warner AH.
1994. Extensive intracellular transloca-
tions of a major protein accompany
anoxia in embryos of Artemia francis-
cana. Exp. Cell Res. 212:77–83
178. Clegg JS, Jackson SA. 1992. Aerobic
heat shock activates trehalose synthe-
sis in embryos of Artemia franciscana.
FEBS Lett. 303:45–47
179. AnchordoguyTJ, Hand SC.1994. Acute
blockage of the ubiquitin-mediated pro-
teolyticpathwayduringinvertebratequi-
escence. Am. J. Physiol. 267:R895–900
180. Krebs RA, Feder ME, Lee J. 1998.
Heritability of expression of the 70-
kD heat-shock protein in Drosophila
melanogaster and its relevance to the
evolution of thermotolerance. Evolution
52:841–47
181. Davis WL, Jacoby BH, Goodman DB.
1994. Immunolocalization of ubiquitin
indegeneratinginsectflightmuscle.His-
tochem. J. 26:298–305
182. Sarge KD, Bray AE, Goodson ML.
1995. Altered stress response in testis.
Nature 374:126
183. SargeKD.1995.Malegermcell-specific
alteration in temperature set point of the
cellular stress response. J. Biol. Chem.
270:18745–48
184. Holbrook NJ, Udelsman R. 1994. Heat
shock protein gene expression in re-
sponse to physiologic stress and aging.
See Ref. 4, pp. 577–93
185. Lee YK, Manalo D, Liu AY. 1996. Heat
shock response, heat shock transcrip-
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
HEAT-SHOCK PROTEINS 275
tion factor and cell aging. Biol. Signals
5:180–91
186. Shpund S, Gershon D. 1997.Alterations
inthechaperoneactivityofHSP70inag-
ing organisms. Arch. Gerontol. Geriatr.
24:125–31
187. Wheeler JC, Bieschke ET, Tower J.
1995. Muscle-specific expression of
Drosophila hsp70 in response to aging
and oxidative stress. Proc. Natl. Acad.
Sci. USA 92:10408–12
188. Marin R, Valet JP, Tanguay RM. 1993.
Heat shock induces changes in the
expression and binding of ubiquitin
in senescent Drosophila melanogaster.
Dev. Genet. 14:78–86
189. Lithgow GJ, White TM, Hinerfeld DA,
Johnson TE. 1994. Thermotolerance of
a long-lived mutant of Caenorhabditis
elegans. J. Gerontol. 49B:270–76
190. Lithgow GJ, White TM, Melov S, John-
son TE. 1995. Thermotolerance and
extended life-span conferred by single-
gene mutations and induced by ther-
mal stress. Proc. Natl. Acad. Sci. USA
92:7540–44
191. LithgowGJ. 1996. Invertebrate geronto-
logy: the age mutations of Caenorhab-
ditis elegans. BioEssays 18:809–15
192. Bond JA, Gonzalez CRM, Bradley BP.
1993. Age-dependent expression of pro-
teins in the cladoceran Daphnia magna
undernormal and heat-stress conditions.
Comp. Biochem. Physiol. 106B:913–
17
193. Khazaeli AA, Tatar M, Pletcher SD,
Curtsinger JW. 1997. Heat-induced
longevity extension in Drosophila.I.
Heat treatment, mortality, and thermo-
tolerance. J. Gerontol. 52A:B48–52
194. Tatar M, Khazaeli AA, Curtsinger JW.
1997.Chaperoningextendedlife.Nature
390:30
195. Dhahbi JM, Mote PL, Tillman JB, Wal-
ford RL, Spindler SR. 1997. Dietary
energy tissue-specifically regulates en-
doplasmicreticulum chaperone geneex-
pression in the liver of mice. J. Nutr.
127:1758–64
196. Heydari AR, Conrad CC, RichardsonA.
1995. Expression of heat shock genes in
hepatocytes is affected by age and food
restriction in rats. J. Nutr. 125:410–18
197. Pahlavani MA, Harris MD, Moore SA,
Richardson A. 1996. Expression of heat
shock protein 70 in rat spleen lympho-
cytes is affected by age but not by food
restriction. J. Nutr. 126:2069–75
198. Lu Q, WallrathLL, Granok H, Elgin SC.
1993. Expression of heat shock protein
70isalteredbyageanddietatthelevelof
transcription. Mol. Cell Biol. 13:2909–
18
199. Masoro EJ. 1996. Possible mecha-
nisms underlying the antiaging actions
of caloric restriction. Toxicol. Pathol.
24:738–41
200. ParsellDA, LindquistS.1993. Thefunc-
tion of heat-shock proteins in stress tol-
erance: degradation and reactivation of
damaged proteins. Annu. Rev. Genet.
27:437–96
201. Parsell DA, Lindquist S. 1994. Heat
shock proteins and stress tolerance. See
Ref. 9, pp. 457–94
202. Bensaude O, Bellier S, Dubois MF, Gi-
annoni F, Nguyen VT. 1996. Heat-shock
induced protein modifications and mod-
ulation of enzyme activities. See Ref. 9,
pp. 199–219
203. diIorio PJ, Holsinger K, Schultz RJ,
Hightower LE. 1996. Quantitative evi-
dence that both Hsc70 and Hsp70 con-
tribute to thermal adaptation in hybrids
of the livebearing fishes Poeciliopsis.
Cell Stress Chaperones 1:139–47
204. Feder ME, Lindquist SL. 1992. Evolu-
tionary loss of a heat shock protein. Am.
Zool. 32:51A (Abstr.)
205. Whitesell L, Cook P. 1996. Stable and
specific binding of heat shock protein
90 by geldanamycin disrupts glucocor-
ticoid receptor function in intact cells.
Mol. Endocrinol. 10:705–12
206. Elia G, Santoro MG. 1994. Regula-
tion of heat shock protein synthesis by
quercetin in human erythroleukaemia
cells. Biochem. J. 300:201–9
207. Bonham-Smith PC, Kapoor M, Bewley
JD. 1987. Establishment of thermotol-
erance in maize by exposure to stresses
other than a heat shock does not require
heatshockproteinsynthesis.PlantPhys-
iol. 85:575–80
208. Xiao CM, Mascarenhas JP. 1985. High
temperature-inducedthermotolerance in
pollen tubes of Tradescantia and heat-
shock proteins. Plant Physiol. 78:887–
90
209. VanBogelen RA, Acton MA, Neidhardt
FC. 1987. Induction of the heat shock
regulon does not produce thermotoler-
ance in Escherichia coli. Genes Dev.
1:525–31
210. Yocum GD, Denlinger DL. 1992. Pro-
longed thermotolerance in the flesh fly
Sarcophaga crassipalpis does not re-
quire continuous expression or persis-
tence of the 72 kDa heat-shock protein.
J. Insect Physiol. 38:603–9
211. Boon-Niermeijer EK, Tuyl M, van de
Scheur H. 1986. Evidence for two states
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
276 FEDER & HOFMANN
of thermotolerance. Int. J. Hypertherm.
2:93–105
212. Smith BJ, Yaffe MP. 1991. Uncoupling
thermotolerance from the induction of
heat shock proteins. Proc. Natl. Acad.
Sci. USA 88:11091–94
213. Easton DP, Rutledge PS, Spotila JR.
1987. Heat shock protein induction and
induced thermal tolerance are indepen-
dent in adult salamanders. J. Exp. Zool.
241:263–67
214. DingleyF,MaynardSmithJ. 1968.Tem-
perature acclimatization in the absence
of protein synthesis of Drosophila sub-
obscura. J. Insect Physiol. 14:1185–
94
215. Finnell RH, Van Waes M, Bennett
GD, Eberwine JH. 1993. Lack of con-
cordance between heat shock proteins
and the development of tolerance to
teratogen-induced neural tube defects.
Dev. Genet. 14:137–47
216. Fisher B, Kraft P, Hahn GM, Ander-
son RL. 1992. Thermotolerance in the
absence of induced heat shock proteins
in a murine lymphoma. Cancer Res.
52:2854–61
217. Watson K, Dunlop G, Cavicchioli R.
1984. Mitochondrial and cytoplasmic
protein syntheses are not required for
heat shock acquisition of ethanol and
thermotolerance in yeast. FEBS Lett.
172:299–302
218. Widelitz RB, Magun BE, Gerner EW.
1986. Effects of cycloheximide on ther-
motolerance expression, heat shock pro-
tein synthesis, and heat shock protein
mRNA accumulation in rat fibroblasts.
Mol. Cell Biol. 6:1088–94
219. Jozwiak Z, Leyko W. 1992. Role of
membranecomponentsinthermalinjury
of cells and development of thermotol-
erance. Int. J. Radiat. Biol. 62:743–56
220. Lindquist S.1993. Autoregulation of the
heat-shock response. In Translational
Regulation of Gene Expression 2, ed. J
Ilan, pp. 279–320. New York: Plenum
221. Krebs RA, Feder ME. 1998. Hsp70 and
larval thermotolerance in Drosophila
melanogaster: Howmuchisenough and
when is more too much? J. Insect Phys-
iol. 44:1091–1101
222. Krebs RA, Feder ME. 1997.Deleterious
consequences of Hsp70 overexpression
in Drosophila melanogaster larvae. Cell
Stress Chaperones 2:60–71
223. Krebs RA, Feder ME. 1997. Natural
variation in the expression of the heat-
shock protein Hsp70 in a population of
Drosophila melanogaster, and its cor-
relation with tolerance of ecologically
relevant thermal stress. Evolution 51:
173–79
224. Welte MA. 1994. Thermotolerance in
Drosophila embryos: the role of hsp70
and the basis for a specific phenocopy.
Ph.D. thesis. Univ. Chicago. 230 pp.
225. Krebs RA, Feder ME. 1998. Ex-
perimental manipulation of the cost
of thermal acclimation in Drosophila
melanogaster. Biol. J. Linn. Soc. 63:
593–601
226. Hoffmann AA. 1995. Acclimation: in-
creasing survival at a cost. Trends Ecol.
Evol. 10:1–2
227. Calow P. 1991. Physiological costs of
combating chemical toxicants: ecologi-
cal implications. Comp. Biochem. Phys-
iol. 100C:3–6
228. Koehn RK, Bayne BL. 1989. Towards a
physiological and genetical understand-
ing of the energetics of the stress re-
sponse. Biol. J. Linn. Soc. 37:157–71
229. Dorner AJ, Krane MG, Kaufman RJ.
1988. Reduction of endogenous GRP78
levels improves secretion of a heterol-
ogous protein in CHO cells. Mol. Cell
Biol. 8:4063–70
230. Dorner AJ, Wasley LC, Kaufman RJ.
1992. Overexpression of GRP78 miti-
gates stress induction of glucose regu-
lated proteins and blocks secretion of
selective proteins in Chinese hamster
ovary cells. EMBO J. 11:1563–71
231. Dorner AJ, Kaufman RJ. 1994. The lev-
els of endoplasmic reticulum proteins
and ATP affect folding and secretion of
selective proteins. Biologicals 22:103–
12
232. Ryan C, Stevens TH, Schlesinger MJ.
1992. Inhibitory effects of HSP70 chap-
erones on nascent polypeptides. Protein
Sci. 1:980–85
233. Heckathorn SA, Poeller GJ, Coleman
JS, Hallberg RL. 1996. Nitrogen avail-
ability alters patterns of accumulation
of heat stress-induced proteins in plants.
Oecologia 105:413–18
234. Heckathorn SA, Poeller GJ, Coleman
JS, Hallberg RL. 1996. Nitrogen avail-
ability and vegetative development
influence the response of ribulose 1,5-
bisphosphate carboxylase/oxygenase,
phosphoenolpyruvate carboxylase, and
heat-shock protein content to heat
stress in Zea mays L. Int. J. Plant Sci.
157:588–95
235. Dietz TJ, Somero GN. 1993. Species-
and tissue-specific synthesis patterns for
heat-shock proteins hsp70 and hsp90 in
several marine teleost fishes. Physiol.
Zool. 66:863–80
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
HEAT-SHOCK PROTEINS 277
236. Abravaya K, Phillips B, Morimoto RI.
1991. Attenuation of the heat shock re-
sponse in HeLa cells is mediated by the
releaseofboundheatshocktranscription
factor and is modulated by changes in
growth and in heat shock temperatures.
Genes Dev. 5:2117–27
237. Sarge KD, Cullen KE. 1997. Regula-
tion of hsp expression during rodent
spermatogenesis. Cell. Mol. Life Sci.
53:191–97
238. Koban M,Graham G, Prosser CL.1987.
Induction of heat-shock protein synthe-
sis in teleost hepatocytes: effects of
acclimation temperature. Physiol. Zool.
60:290–96
239. Deleted in proof
240. Favatier F, Bornman L, Hightower LE,
Gnther E, Polla BS. 1997. Variation in
hsp gene expression and Hsp polymor-
phism: Do they contribute to differ-
ential disease susceptibility and stress
tolerance? Cell Stress Chaperones 2:
141–55
241. Engman DM, Sias SR, Gabe JD, Donel-
son JE, Dragon EA. 1989. Comparison
of HSP70 genes fromtwo strains of Try-
panosoma cruzi. Mol. Biochem. Para-
sitol. 37:285–87
242. Engman DM, Reddy LV, Donelson
JE, Kirchhoff LV. 1987. Trypanosoma
cruzi exhibits inter- and intra-strain
heterogeneity in molecular karyotype
and chromosomal gene location. Mol.
Biochem. Parasitol. 22:115–23
243. Snutch TP, Baillie DL. 1984. A high de-
gree of DNA strain polymorphism asso-
ciated with the major heat shock gene
in Caenorhabditis elegans. Mol. Gen.
Genet. 195:329–35
244. Hamet P, Kaiser MA, Sun Y, Page V,
Vincent M, et al. 1996. HSP27 locus
cosegregates with left ventricular mass
independentlyofblood pressure. Hyper-
tension 28:1112–17
245. Wallich R, Helmes C, Schaible UE, Lo-
bet Y, Moter SE, et al. 1992. Evaluation
of genetic divergence among Borrelia
burgdorferi isolates by use of OspA, fla,
HSP60, and HSP70 gene probes. Infect.
Immun. 60:4856–66
246. Hamet P, Kong D, Pravenec M, Kunes J,
Kren V, et al. 1992. Restriction fragment
length polymorphism of hsp70 gene, lo-
calizedintheRT1complex,isassociated
with hypertension in spontaneously hy-
pertensiverats. Hypertension19:611–14
247. Grosz MD, Skow LC, Stone RT. 1994.
An AluI polymorphism at the bovine 70
kD heat-shock protein-1 (HSP70-1) lo-
cus. Anim. Genet. 25:196
248. Jorgensen JA, Nguyen HT. 1995. Ge-
netic analysis of heat shock proteins in
maize. Theor. Appl. Genet. 91:38–46
249. Ottaviano E, Sari Gorla M, Pe E, Frova
C. 1991. Molecular markers, RFLPs
and Hsps for the genetic dissection of
thermotolerance in maize. Theor. Appl.
Genet. 81:713–19
250. Dimascio JA, Sweeney PM, Dan-
neberger TK, Kamalay JC. 1994. Anal-
ysis of heat shock response in perennial
ryegrass using maize heat shock protein
clones. Crop Sci. 34:798–804
251. Goldschmidt-Clermont M. 1980. Two
genes for the major heat-shock protein
of Drosophila melanogaster arrangedas
an inverted repeat. Nucleic Acids Res.
8:235–52
252. Ish-HorowiczD, Pinchin SM. 1980. Ge-
nomic organization of the 87A7 and
87C1 heat-induced loci of Drosophila
melanogaster. J. Mol. Biol. 142:231–45
253. Leigh-Brown AJ, Ish-Horowicz D.
1981. Evolution of the 87A and 87C
heat-shock loci in Drosophila. Nature
290:677–82
254. McKechnie SW, Halford MM, McColl
G, Hoffmann AA. 1998. Both allelic
variation and expression of nuclear and
cytoplasmictranscriptsofhsr-omegaare
closely associated with thermal pheno-
type in Drosophila. Proc. Natl. Acad.
Sci. USA 95:2423–28
255. Ish-Horowicz D, Pinchin SM, Schedl
P, Artavanis-Tsakonas S, Mirault ME.
1979. Genetic and molecular analysis of
the 87A7 and 87C1 heat-inducible loci
of D. melanogaster. Cell 18:1351–58
256. Lis JT, Ish-Horowicz D, Pinchin SM.
1981. Genomic organization and tran-
scription of the alpha beta heat shock
DNA in Drosophila melanogaster. Nu-
cleic Acids Res. 9:5297–310
257. Lis JT, Prestidge L, Hogness DS. 1978.
A novel arrangement of tandemly re-
peated genes at a major heat shock site
in D. melanogaster. Cell 14:901–19
258. Mirault ME, Goldschmidt-Clermont M,
Artavanis-Tsakonas S, Schedl P. 1979.
Organization of the multiple genes for
the 70,000-dalton heat-shock protein in
Drosophila melanogaster. Proc. Natl.
Acad. Sci. USA 76:5254–58
259. CraigEA,McCarthyBJ,WadsworthSC.
1979. Sequence organization of two re-
combinant plasmids containing genes
forthe majorheat shock-inducedprotein
of D. melanogaster. Cell 16:575–88
260. Artavanis-Tsakonas S, SchedlP, Mirault
ME, Moran L, Lis J.1979. Genes forthe
70,000 dalton heat shock protein in two
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
278 FEDER & HOFMANN
cloned D. melanogasterDNAsegments.
Cell 17:9–18
261. HolmgrenR, Livak K,Morimoto R,Fre-
und R, Meselson M. 1979. Studies of
cloned sequences from four Drosophila
heat shock loci. Cell 18:1359–70
262. Sun CW, Griffen S, Callis J. 1997. A
model for the evolution of polyubiqui-
tin genes from the study of Arabidop-
sis thaliana ecotypes. Plant Mol. Biol.
34:745–58
263. Ristic Z, Gifford DJ, Cass DD. 1991.
Heat shock proteins in two lines of Zea
mays L. that differ in drought and heat
resistance. Plant Physiol. 97:1430–34
264. Colombo SJ, Colclough ML, Timmer
VR,BlumwaldE.1992.Clonal variation
in heat tolerance and heat shock protein
expressioninblackspruce.SilvaeGenet.
41:234–39
265. Fender SE, O’Connell MA. 1989. Heat
shockprotein expression inthermotoler-
ant and thermosensitive lines of cotton.
Plant Cell Rep. 8:37–40
266. Malayer JR, Hansen PJ. 1990. Differ-
ences between Brahman and Holstein
cows in heat-shock induced alterations
of protein synthesis and secretion by
oviducts and uterine endometrium. J.
Anim. Sci. 68:266–80
267. OtsukaY,TakanoTS,YamazakiT.1997.
Genetic variation in the expression of
the six hsp genes in the presence of
heat shock in Drosophila melanogaster.
Genes Genetic Syst. 72:19–24
268. JorgensenJA,WengJ,HoTHD,Nguyen
HT. 1992. Genotype-specific heat shock
proteinsin twomaizeinbreds.PlantCell
Rep. 11:576–80
269. Frova C, Gorla MS. 1993. Quantitative
expression of maize Hsps: genetic dis-
section and association with thermotol-
erance. Theor. Appl. Genet. 86:213–20
270. Weng J, Nguyen HT. 1992. Differences
inthe heat-shock responsebetween ther-
motolerant and thermosusceptible cul-
tivars of hexaploid wheat. Theor. Appl.
Genet. 84:941–46
271. Brown DC, Bradley BP, Tedengren M.
1995. Genetic and environmental regu-
lation of HSP70 expression. Mar. Env.
Res. 39:181–84
272. CarusoM, Sacco M, MedoffG,Maresca
B. 1987. Heat shock 70 gene is differ-
entially expressed in Histoplasma cap-
sulatum strains with different levels of
thermotoleranceandpathogenicity.Mol.
Microbiol. 1:151–58
273. Lyashko VN, Vikulova VK, Chernikov
VG, Ivanov VI, Ulmasov KA, et al.
1994. Comparison of the heat shock re-
sponse in ethnically and ecologically
differenthuman populations.Proc.Natl.
Acad. Sci. USA 91:12492–95
274. Bettencourt BR, Feder ME, Cavicchi S.
1997. Laboratory evolution of Hsp70
expression in Drosophila melanogaster:
functional consequences and molecular
bases. Am. Zool. 37:189A (Abstr.)
275. McColl G, Hoffmann AA, McKechnie
SW. 1996. Response of two heat shock
genes to selection for knockdown heat
resistance in Drosophila melanogaster.
Genetics 143:1615–27
276. Fender SE, O’Connell MA. 1990. Ex-
pression of the heat shock response in
a tomato interspecific hybrid is not in-
termediate between the two parental re-
sponses. Plant Physiol. 93:1140–46
277. Gaugler R, Wilson M, Shearer P. 1997.
Field release and environmental fate of
a transgenic entomopathogenic nema-
tode. Biol. Control 9:75–80
278. Hashmi S, Hashmi G, Glazer I, Gau-
gler R. 1998. Thermal response of Het-
erorhabditis bacteriophora transformed
with the Caenorhabditis elegans hsp70
encodinggene.J.Exp.Zool.281:164–70
279. Trent JD. 1996. A review of acquired
thermotolerance, heat-shock proteins,
and molecular chaperones in Archaea.
FEMS Micro. Rev. 18:249–58
280. Holden JF, Baross JA. 1993. Enhanced
thermotolerance and temperature-in-
duced changes in protein composition in
the hyperthermophilic archaeon ES4. J.
Bacteriol. 175:2839–43
281. Hamilton PT, Reeve JN. 1985. Struc-
ture of genes andan insertionelement in
the methane producing archaebacterium
Methanobrevibacter smithii. Mol. Gen.
Genet. 200:47–59
282. Waldmann T, Nimmesgern E, Nitsch M,
Peters J, Pfeifer G, et al. 1995. The ther-
mosome of Thermoplasma acidophilum
and its relationship to the eukaryotic
chaperonin TRiC. Eur. J. Biochem. 227:
848–56
283. Phipps BM, Typke D, Heger R, Volker
S, Hoffmann A, et al. 1993. Struc-
ture of a molecular chaperone from a
thermophilic archaebacterium. Nature
361:475–77
284. Nitsch M, Klumpp M, Lupas A, Bau-
meister W. 1997. The thermosome: al-
ternating alpha and beta-subunits within
the chaperonin of the archaeon Ther-
moplasma acidophilum. J. Mol. Biol.
267:142–49
285. Lewis VA, Hynes GM, Zheng D,
Saibil H, Willison K. 1992. T-complex
polypeptide-1 is a subunit of a het-
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
HEAT-SHOCK PROTEINS 279
eromeric particle in the eukaryotic cy-
tosol. Nature 358:249–52
286. Trent JD, Nimmesgern E, Wall JS, Hartl
FU, Horwich AL. 1991. A molecular
chaperone from a thermophilic archae-
bacterium is related to the eukaryotic
proteint-complexpolypeptide-1.Nature
354:490–93
287. Hendrick JP, Hartl FU. 1995. The role of
molecularchaperonesin proteinfolding.
FASEB J. 9:1559–69
288. Eggers DK, Welch WJ, Hansen WJ.
1997. Complexesbetween nascent poly-
peptides and their molecular chaperones
in the cytosol of mammalian cells. Mol.
Biol. Cell 8:1559–73
289. Burston SG, Clarke AR. 1995. Molecu-
lar chaperones: physical and mechanis-
tic properties. Essays Biochem. 29:125–
36
290. Trent JD, Kagawa HK, Yaoi T, Olle
E, Zaluzec NJ. 1997. Chaperonin fila-
ments: the archaeal cytoskeleton? Proc.
Natl. Acad. Sci. USA 94:5383–88
291. Dascher CC, Poddar SK, Maniloff J.
1990. Heat shock response inmycoplas-
mas, genome-limited organisms. J. Bac-
teriol. 172:1823–27
292. Gupta RS, Singh B. 1994. Phylogenetic
analysis of 70 kD heat shock protein se-
quences suggests a chimeric origin for
the eukaryotic cell nucleus. Curr. Biol.
4:1104–14
293. Bustard K, Gupta RS. 1997. The se-
quences of heat shock protein 40
(DnaJ) homologs provideevidencefor a
close evolutionary relationship between
the Deinococcus thermus group and
Cyanobacteria. J. Mol. Evol. 45:193–
205
294. GuptaRS, Bustard K, FalahM, SinghD.
1997. Sequencing of heat shock protein
70 (DnaK) homologs from Deinococcus
proteolyticus and Thermomicrobium ro-
seum and their integration in a protein-
based phylogeny of prokaryotes. J. Bac-
teriol. 179:345–57
295. Gupta RS. 1995. Phylogenetic analysis
of the 90 kD heat shock family of pro-
teinsequencesand anexaminationofthe
relationship among animals, plants, and
fungi species. Mol. Biol. Evol. 12:1063–
73
296. Gupta RS, Aitken K, Falah M, Singh B.
1994. Cloning of Giardia lamblia heat
shock protein HSP70 homologs: im-
plications regarding origin of eukary-
otic cells and of endoplasmic reticulum.
Proc. Natl. Acad. Sci. USA 91:2895–
99
297. GuptaRS, SinghB. 1992. Cloningof the
HSP70genefromHalobacteriummaris-
mortui: relatedness of archaebacterial
HSP70 to its eubacterial homologs and
a model for the evolution of the HSP70
gene. J. Bacteriol. 174:4594–605
298. Boorstein WR, Ziegelhoffer T, Craig
EA. 1994. Molecular evolution of the
HSP70 multigene family. J. Mol. Evol.
38:1–17
299. Rubin DM, Mehta AD, Zhu J, Shoham
S, Chen X, et al. 1993. Genomic struc-
ture and sequence analysis of Droso-
phila melanogaster HSC70 genes. Gene
128:155–63
300. Tavaria M, Gabriele T, Kola I, Anderson
RL. 1996. A hitchhiker’s guide to the
human Hsp70 family. Cell Stress Chap-
erones 1:23–28
301. de Jong WW, Leunissen JA, Voorter
CE. 1993. Evolution of the alpha-
crystallin/small heat-shock protein fam-
ily. Mol. Biol. Evol. 10:103–26
302. Ohta T. 1994. Further examples of
evolution by gene duplication revealed
through DNA sequence comparisons.
Genetics 138:1331–37
303. Benedict MQ, Cockburn AF, Seawright
JA. 1993. The Hsp70 heat-shock gene
family of the mosquito Anopheles albi-
manus. Insect Mol. Biol. 2:93–102
304. Drosopoulou E, Konstantopoulou I,
Scouras ZG. 1996. The heat shock
genes in the Drosophila montium sub-
group: chromosomal localization and
evolutionary implications.Chromosoma
105:104–10
305. Pardali E, Feggou E, Drosopoulou
E, Konstantopoulou I, Scouras ZG,
Mavragani-Tsipidou P. 1996. The Afro-
tropical Drosophila montium subgroup:
Balbiani ring 1, polytene chromosomes,
and heat shock response of Drosophila
vulcana. Genome 39:588–97
306. SanchezY, Lindquist SL.1990. HSP104
required for induced thermotolerance.
Science 248:1112–15
307. Feder ME, Block BA. 1991. On the
future of physiological ecology. Funct.
Ecol. 5:136–44
308. Lau S, Patnaik N, Sayen MR, Mestril
R. 1997. Simultaneous overexpression
of two stress proteins in rat cardiomy-
ocytes and myogenic cells confers pro-
tection against ischemia-induced injury.
Circulation 96:2287–94
309. Huot J, Roy G, Lambert H, Chretien P,
Landry J. 1991. Increased survival af-
ter treatments with anticancer agents of
Chinesehamster cellsexpressing thehu-
man Mr 27,000 heat shock protein. Can-
cer Res. 51:5245–52
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
280 FEDER & HOFMANN
310. Mehlen P, Preville X, Chareyron P,
Briolay J, Klemenz R, Arrigo AP.
1995. Constitutive expression of human
hsp27, Drosophila hsp27, or human al-
pha B-crystallin confers resistance to
TNF- and oxidative stress-induced cy-
totoxicity in stably transfected murine
L929 fibroblasts. J. Immunol. 154:363–
74
311. JaattelaM,WissingD.1993.Heat-shock
proteins protect cells from monocytecy-
totoxicity: possible mechanism of self-
protection. J. Exp. Med. 177:231–36
312. Trautinger F, Kokesch C, Herbacek I,
Knobler RM, Kindas-Mugge I. 1997.
Overexpression of the small heat shock
protein, hsp27, confers resistance to hy-
perthermia, but not to oxidative stress
and UV-induced cell death, in a sta-
bly transfected squamous cell carci-
nomacell line. J.Photochem.Photobiol.
39B:90–95
313. Rollet E, Lavoie JN, Landry J, Tanguay
RM. 1992. Expression of Drosophila’s
27 kDa heat shock protein into
rodent cells confers thermal resis-
tance. Biochem. Biophys. Res. Commun.
185:116–20
314. Lavoie JN, Gingras-Breton G, Tanguay
RM, Landry J. 1993. Induction of Chi-
nese hamster HSP27 gene expression in
mouse cells confers resistance to heat
shock. HSP27 stabilization of the mi-
crofilament organization. J. Biol. Chem.
268:3420–29
315. LandryJ, ChretienP, LambertH, Hickey
E, Weber LA. 1989. Heat shock resis-
tance conferred by expression of the hu-
man HSP27 gene in rodent cells. J. Cell
Biol. 109:7–15
316. Wissing D,Jaattela M.1996. HSP27 and
HSP70 increase the survival of WEHI-S
cellsexposedtohyperthermia.Int.J.Hy-
pertherm. 12:125–38
317. Wang G, Klostergaard J, Khodadadian
M, Wu J, Wu TW, et al. 1996. Murine
cells transfected with human Hsp27
cDNA resist TNF-induced cytotoxicity.
J. Immunother. Emphasis Tumor. Im-
munol. 19:9–20
318. Martin JL, Mestril R, Hilal-Dandan
R, Brunton LL, Dillmann WH. 1997.
Small heat shock proteins and protec-
tion against ischemic injury in cardiac
myocytes. Circulation 96:4343–48
319. Stege GJ, Kampinga HH, Konings AW.
1995. Heat-induced intranuclear protein
aggregation and thermal radiosensitiza-
tion. Int. J. Radiat. Biol. 67:203–9
320. BlackburnR,GaloforoS,BernsCM,Ire-
land M, Cho JM, et al. 1996. Thermal
response in murine L929 cells lacking
alpha B-crystallin expression and alpha
B-crystallin expressing L929 transfec-
tants. Mol. Cell. Biochem. 155:51–60
321. vandenIjsselPR,OverkampP,KnaufU,
Gaestel M, de Jong WW. 1994. Alpha
A-crystallin confers cellular thermore-
sistance. FEBS Lett. 355:54–56
322. Heads RJ, Yellon DM, Latchman DS.
1995.Differentialcytoprotectionagainst
heat stress or hypoxia following expres-
sion of specific stress protein genes in
myogenic cells. J. Mol. Cell. Cardiol.
27:1669–78
323. Cumming DV, Heads RJ, Watson A,
Latchman DS, Yellon DM. 1996. Dif-
ferential protection of primary rat car-
diocytes by transfection of specific heat
stress proteins. J. Mol. Cell. Cardiol.
28:2343–49
324. Lukacs KV, Nakakes A, Atkins CJ,
Lowrie DB, Colston MJ. 1997. In vivo
gene therapy of malignant tumours with
heat shock protein-65 gene. Gene Ther.
4:346–50
325. Lukacs KV, Lowrie DB, Stokes RW,
Colston MJ. 1993. Tumor cells trans-
fected with a bacterial heat-shock gene
lose tumorigenicity and induce protec-
tion against tumors. J. Exp. Med. 178:
343–48
326. Wischmeyer PE, Musch MW, Madonna
MB, Thisted R, Chang EB. 1997.
Glutamine protects intestinal epithelial
cells: role of inducible HSP70. Am. J.
Physiol. 272:G879–84
327. Angelidis CE, Lazaridis I, Pagoulatos
GN. 1991. Constitutive expression of
heat-shock protein 70 in mammalian
cells confers thermoresistance. Eur. J.
Biochem. 199:35–39
328. Mailhos C, Howard MK, Latchman DS.
1994. Heat shock proteins hsp90 and
hsp70 protect neuronal cells from ther-
mal stress butnot fromprogrammed cell
death. J. Neurochem. 63:1787–95
329. Wyatt S, Mailhos C, Latchman DS.
1996. Trigeminal ganglion neurons are
protected by the heat shock proteins
hsp70 and hsp90 from thermal stress
but not from programmed cell death fol-
lowing nerve growth factor withdrawal.
Brain Res. Mol. Brain Res. 39:52–56
330. Uney JB, Staley K, Tyers P, Sofroniew
MV, Kew JN. 1994. Transfection with
hsp70i protects rat dorsal root ganglia
neurones and glia from heat stress. Gene
Ther. 1:S65
331. Solomon JM, Rossi JM, Golic K, Mc-
Garry T, Lindquist S. 1991. Changes in
Hsp70 alter thermotolerance and heat-
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
HEAT-SHOCK PROTEINS 281
shock regulation in Drosophila. New
Biol. 3:1106–20
332. Sato K, Saito H, Matsuki N. 1996.
HSP70 is essential to the neuroprotec-
tive effect of heat-shock. Brain Res.
740:117–23
333. Riabowol KT, Mizzen LA, Welch WJ.
1988. Heat shock is lethal to fibroblasts
microinjected with antibodies against
hsp70. Science 242:433–36
334. MestrilR,Giordano FJ,CondeAG,Dill-
mann WH. 1996. Adenovirus-mediated
gene transfer of a heat shock protein 70
(hsp 70i) protects against simulated is-
chemia. J. Mol. Cell. Cardiol. 28:2351–
58
335. LiuRY,LiX, LiL, Li GC.1992. Expres-
sion of human hsp70 in rat fibroblasts
enhances cell survival and facilitates re-
covery from translational and transcrip-
tional inhibition following heat shock.
Cancer Res. 52:3667–73
336. Nakata N, Kato H, Kogure K. 1993. In-
hibition of ischaemic tolerance in the
gerbil hippocampus by quercetin and
anti-heat shock protein-70 antibody.
NeuroReport 4:695–98
337. LeeYJ, Kim D, HouZZ, Curetty L, Bor-
relli MJ, Corry PM. 1993. Alteration of
heat sensitivity by introductionof hsp70
or anti-hsp70 in CHO cells. J. Therm.
Biol. 18:229–36
338. Johnston RN, Kucey BL. 1988. Com-
petitive inhibitionof hsp70gene expres-
sion causes thermosensitivity. Science
242:1551–54
339. Heads RJ, Latchman DS, Yellon DM.
1994. Stable high level expression of a
transfected human HSP70 gene protects
a heart-derived muscle cell line against
thermal stress. J. Mol. Cell. Cardiol.
26:695–99
340. Amin V, Cumming DV, Latchman DS.
1996. Over-expression of heat shock
protein70protectsneuronalcellsagainst
both thermal and ischaemic stress but
with different efficiencies. Neurosci.
Lett. 206:45–48
341. Khan NA, Sotelo J. 1989. Heat shock
stressisdeleterioustoCNSculturedneu-
rons microinjected with anti-HSP70 an-
tibodies. Biol. Cell 65:199–202
342. Li GC, Li LG, Liu YK, Mak JY, Chen
LL, Lee WM. 1991. Thermal response
of rat fibroblasts stably transfected with
the human 70-kDa heat shock protein-
encoding gene. Proc. Natl. Acad. Sci.
USA 88:1681–85
343. Williams RS, Thomas JA, Fina M,
German Z, Benjamin IJ. 1993. Hu-
man heat shock protein 70 (hsp70) pro-
tects murine cells from injury during
metabolic stress. J.Clin. Invest. 92:503–
8
344. Mestril R, Chi SH, Sayen MR, O’Reilly
K, Dillmann WH. 1994. Expression
of inducible stress protein 70 in rat
heart myogenic cells confers protection
against simulated ischemia-induced in-
jury. J. Clin. Invest. 93:759–67
345. Dillmann WH, Mestril R. 1995. Heat
shock proteins in myocardial stress. Z.
Kardiol. 4:87–90
346. Jacobs M, Andersen JB, Kontinen V,
Sarvas M. 1993. The DNA-binding ac-
tivityof thehuman heatshock transcrip-
tionfactoris regulatedin vivoby Hsp70.
Mol. Cell Biol. 13:5427–38
347. Ding XZ, Tsokos GC, Smallridge RC,
Kiang JG. 1997. Heat shock gene-
expression in HSP-70 and HSF1 gene-
transfected human epidermoid A-431
cells. Mol. Cell. Biochem. 167:145–52
348. Chi SH, Mestril R. 1996. Stable ex-
pression of a human HSP70 gene in a
rat myogenic cell line confers protec-
tion against endotoxin. Am. J. Physiol.
270:C1017–21
349. Han MY, Park YM. 1997. Reduced pro-
tein denaturation in thermotolerant cells
by elevated levels of HSP70. Korean J.
Pharmacol. 32:433–44
350. Jaattela M. 1995. Over-expression of
hsp70 confers tumorigenicity to mouse
fibrosarcoma cells. Int. J. Cancer 60:
689–93
351. WeiYQ,Zhao X,Kariya Y, Teshigawara
K, Uchida A. 1995. Inhibition of pro-
liferation and induction of apoptosis by
abrogation of heat-shock protein (HSP)
70expression intumor cells. CancerIm-
munol. Immunother. 40:73–78
352. Karlseder J, Wissing D, Holzer G, Orel
L, Sliutz G, et al. 1996. Hsp70 over-
expression mediates the escape of a
doxorubicin-induced G2 cell cycle ar-
rest. Biochem. Biophys. Res. Commun.
220:153–59
353. Henle KJ, Jethmalani SM, Li L, Li GC.
1997. Protein glycosylation in a heat-
resistant rat fibroblast cell model ex-
pressing human HSP70. Biochem. Bio-
phys. Res. Commun. 232:26–32
354. Simon MM, Reikerstorfer A, Schwarz
A, Krone C, Luger TA, et al. 1995. Heat
shock protein 70 overexpression affects
theresponsetoultravioletlightinmurine
fibroblasts. Evidence for increased cell
viability and suppressionof cytokinere-
lease. J. Clin. Invest. 95:926–33
355. Liossis SN, Ding XZ, Kiang JG,
Tsokos GC. 1997. Overexpression of
P1: ARK/spd P2: ARK/ary QC: ARS
December 22, 1998 14:39 Annual Reviews AR077-10
282 FEDER & HOFMANN
the heat shock protein 70 enhances
the TCR/CD3- and Fas/Apo-1/CD95-
mediated apoptotic cell death in Jurkat
T cells. J. Immunol. 158:5668–75
356. Mosser DD, Caron AW, Bourget L,
Denis-LaroseC,MassieB.1997.Roleof
the human heat shock protein hsp70 in
protection against stress-induced apop-
tosis. Mol. Cell. Biol. 17:5317–27
357. Suzuki K, Sawa Y, Kaneda Y, Ichikawa
H, Shirakura R, Matsuda H. 1997. In
vivo gene transfection with heat shock
protein 70 enhances myocardial toler-
ance to ischemia-reperfusion injury in
rat. J. Clin. Invest. 99:1645–50
358. PlumierJC,RossBM,CurrieRW,Ange-
lidis CE, Kazlaris H, et al. 1995. Trans-
genic mice expressing the human heat
shock protein 70 have improved post-
ischemic myocardial recovery. J. Clin.
Invest. 95:1854–60
359. Marber MS, Mestril R, Chi SH, Sayen
MR, Yellon DM, Dillmann WH. 1995.
Overexpression of the rat inducible
70-kD heat stress protein in a transgenic
mouse increases the resistance of the
heart to ischemic injury. J. Clin. Invest.
95:1446–56
360. Plumier JC, Krueger AM, Currie RW,
Kontoyiannis D, Kollias G, Pagoulatos
GN. 1997. Transgenic mice expressing
the human inducible Hsp70 have hip-
pocampal neurons resistant to ischemic
injury.Cell StressChaperones2:162–67
361. Lee JH, Schoffl F. 1996. An Hsp70 an-
tisense gene affects the expression of
HSP70/HSC70, the regulation of HSF,
and the acquisition of thermotolerance
in transgenic Arabidopsisthaliana. Mol.
Gen. Genet. 252:11–19
362. Galea-Lauri J, Richardson AJ, Latch-
man DS, Katz DR. 1996. Increased
heat shock protein 90 (hsp90) expres-
sion leads to increased apoptosis in the
monoblastoid cell line U937 following
inductionwithTNF-alphaandcyclohex-
imide: a possible rolein immunopathol-
ogy. J. Immunol. 157:4109–18
363. Nakano M, Mann DL, Knowlton AA.
1997. Blocking the endogenous in-
crease in HSP 72 increases suscepti-
bility to hypoxia and reoxygenation in
isolated adult feline cardiocytes. Circu-
lation 95:1523–31
364. Hutter JJ, Mestril R, Tam EK, Sievers
RE, Dillmann WH, Wolfe CL. 1996.
Overexpressionof heat shockprotein 72
in transgenic mice decreases infarct size
in vivo. Circulation 94:1408–11
365. Galea-Lauri J, Latchman DS, Katz DR.
1996. The role of the 90-kDa heat shock
protein in cell cycle control and differ-
entiation of the monoblastoid cell line
U937. Exp. Cell Res. 226:243–54
366. Schirmer EC, Lindquist S, Vierling E.
1994. An Arabidopsis heat shock pro-
tein complements a thermotolerance de-
fect in yeast. Plant Cell 6:1899–909
367. Kutskova IUA, Mamon LA. 1996. Con-
sequences of exposure to extreme con-
ditions in somatic cells of Drosophila
melanogaster under conditions of dis-
turbed synthesis of heat shock proteins.
Genetika 32:1406–16
368. Mamon LA, Kutskova YA. 1993.
The role of the heat-shock proteins
in recovery of high temperature in-
duced damages of mitotic chromosomes
in Drosophila melanogaster. Genetika
29:604–12
369. Mamon LA, Kutskova YA. 1993. The
role of the heat-shock proteins in recov-
ery of cell proliferation following high
temperature treatment of Drosophila
melanogaster. Genetika 29:791–98
370. Koishi M, Hosokawa N, Sato M, Nakai
A, Hirayoshi K, et al. 1992. Quercetin,
aninhibitorofheatshockproteinsynthe-
sis, inhibits the acquisitionofthermotol-
erance in a human colon carcinoma cell
line. Jpn. J. Cancer Res. 83:1216–22
371. Lee YJ, Curetty L, Hou ZZ, Kim
SH, Kim JH, Corry PM. 1992. Effect
of pH on quercetin-induced suppres-
sion of heat shock gene expression and
thermotolerance development in HT-29
cells. Biochem. Biophys. Res. Commun.
186:1121–28
372. Jedlicka P, Mortin MA, Wu C. 1997.
Multiple functions of Drosophila heat
shock transcription factor in vivo.
EMBO J. 16:2452–62
373. Lee JH, Hubel A, Schoffl F. 1995. Dere-
pressionof theactivityofgenetically en-
gineered heat shock factor causes con-
stitutive synthesisof heatshock proteins
and increased thermotolerance in trans-
genic Arabidopsis. Plant J. 8:603–12
... These bacteria reside in the host's intestinal tract, epidermis and various organs, and jointly regulate each other's life activities. The number of symbiotic bacteria even far exceeds the number of cells of the organism itself, and the diversity and the functions of symbiotic bacteria are vast, ranging from digestion and absorption [2][3][4] to detoxification, metabolism [5,6], controlling the growth and reproduction of the host [7,8], and adaptation to the environment [9][10][11]. ...
... Organisms residing in deep-sea hydrothermal regions characterized by high pressure, hypoxia, high levels of toxic substances, sulfides, and heavy metals, as well as being dark, are doubly symbiotic with two gammaproteobacterial endosymbionts: a sulfur oxidizer and a methane oxidizer [62,63]. High temperatures can denature proteins, causing physiological and developmental problems [10]. Endosymbiosis with heat shock tolerance can reduce the impact of high temperature on the survival and development of the host [11]. ...
Article
Full-text available
Background Karst caves serve as natural laboratories, providing organisms with extreme and constant conditions that promote isolation, resulting in a genetic relationship and living environment that is significantly different from those outside the cave. However, research on cave creatures, especially Opiliones, remains scarce, with most studies focused on water, soil, and cave sediments. Results The structure of symbiotic bacteria in different caves were compared, revealing significant differences. Based on the alpha and beta diversity, symbiotic bacteria abundance and diversity in the cave were similar, but the structure of symbiotic bacteria differed inside and outside the cave. Microorganisms in the cave play an important role in material cycling and energy flow, particularly in the nitrogen cycle. Although microbial diversity varies inside and outside the cave, Opiliones in Beijing caves and Hainan Island exhibited a strong similarity, indicating that the two environments share commonalities. Conclusions The karst cave environment possesses high microbial diversity and there are noticeable differences among different caves. Different habitats lead to significant differences in the symbiotic bacteria in Opiliones inside and outside the cave, and cave microorganisms have made efforts to adapt to extreme environments. The similarity in symbiotic bacteria community structure suggests a potential similarity in host environments, providing an explanation for the appearance of Sinonychia martensi in caves in the north.
... By binding to heat stress elements (HSEs) in the promoters of heat stress-responsive genes, HSFs regulate the expression of heat shock proteins (HSPs), which function as molecular chaperones to prevent protein denaturation [8,9]. Additionally, HSPs regulate protein folding, accumulation, localization, and degradation, and are believed to play significant roles in coping with various environmental stresses [10]. Furthermore, HSFs can regulate reactive oxygen species (ROS)-scavenging enzymes, such as ascorbate peroxidase (APX) and catalase (CAT) [11]. ...
Article
Full-text available
Background The heat shock transcription factor (HSF) plays a crucial role in the regulatory network by coordinating responses to heat stress as well as other stress signaling pathways. Despite extensive studies on HSF functions in various plant species, our understanding of this gene family in garlic, an important crop with nutritional and medicinal value, remains limited. In this study, we conducted a comprehensive investigation of the entire garlic genome to elucidate the characteristics of the AsHSF gene family. Results In this study, we identified a total of 17 AsHSF transcription factors. Phylogenetic analysis classified these transcription factors into three subfamilies: Class A (9 members), Class B (6 members), and Class C (2 members). Each subfamily was characterized by shared gene structures and conserved motifs. The evolutionary features of the AsHSF genes were investigated through a comprehensive analysis of chromosome location, conserved protein motifs, and gene duplication events. These findings suggested that the evolution of AsHSF genes is likely driven by both tandem and segmental duplication events. Moreover, the nucleotide diversity of the AsHSF genes decreased by only 0.0002% from wild garlic to local garlic, indicating a slight genetic bottleneck experienced by this gene family during domestication. Furthermore, the analysis of cis-acting elements in the promoters of AsHSF genes indicated their crucial roles in plant growth, development, and stress responses. qRT-PCR analysis, co-expression analysis, and protein interaction prediction collectively highlighted the significance of Asa6G04911. Subsequent experimental investigations using yeast two-hybridization and yeast induction experiments confirmed its interaction with HSP70/90, reinforcing its significance in heat stress. Conclusions This study is the first to unravel and analyze the AsHSF genes in garlic, thereby opening up new avenues for understanding their functions. The insights gained from this research provide a valuable resource for future investigations, particularly in the functional analysis of AsHSF genes.
... The various types of hsp are differentiated into five families based on their molecular weights. Among these families, hsp70 translocases proteins across cellular membranes and protects neurons from apoptosis [46,47], while hsp90 assists cortisol by transmitting signals from glucocorticoid receptors [48]. In this study, the expressions of hsp70 and hsp90 were significantly upregulated in the acute and chronic thermal-treated groups compared to the control groups. ...
Article
Full-text available
Affected by the continuously rising temperature, thermal stress leads to a delinked growth rate and resistance to stress in cultured largemouth bass (Micropterus salmoides, LMB) in China. Identification of LMB with better thermal resistance will benefit the breeding of new varieties. However, there has been limited reporting on the evaluation to identify LMB with better thermal resistance. LMB consists of the northern LMB (Micropterus salmoides salmoides, NLMB) and the Florida LMB (Micropterus salmoides floridanus, FLMB). Due to their different geographical distributions, it has been suggested that FLMB exhibit better thermal resistance compared to NLMB. In this study, NLMB and FLMB were subjected to thermal stress for 3 h (acute) and 60 d (chronic) at 33 °C, respectively. Subsequently, the variations of 12 candidate biomarkers between NLMB and FLMB were analyzed. Exposure to acute thermal stress significantly increased plasma cortisol, blood glucose, and lactate levels; activities of superoxide dismutase (SOD), glutathione peroxidase (GPX), catalase (CAT), glucose kinase (GK), pyruvate kinase (PK), lactate dehydrogenase (LDH), and glucose 6 phosphatase (G6Pase); and the expressions of hsp70 and hsp90 in both NLMB and FLMB (p < 0.05). Compared to NLMB, FLMB exhibited a lower plasma cortisol level and a higher expression of hsp90 under acute thermal stress (p < 0.05). Exposure to chronic thermal stress significantly increased plasma cortisol and blood glucose levels, as well as activities of GK, PK, LDH, and G6Pase, as well as expressions of hsp70 and hsp90 in both NLMB and FLMB (p < 0.05). Additionally, FLMB showed a lower expression of hsp70 compared to NLMB (p < 0.05). In conclusion, our results showed that LMB with lower plasma cortisol level and higher expression of hsp90 under acute thermal stress, as well as lower expression of hsp70 under chronic thermal stress were suggested to have better thermal resistance. Our study provides valuable information for identifying and breeding LMB varieties with better thermal resistance in the future.
... Thermal adaptation is key for organisms to survive and reproduce; this involves genetic and epigenetic changes affecting gene expression and cellular function [18][19][20] . Changes in gene expression help organisms maintain energy homeostasis, adapt to heat stress and activate pathways involved in response to heat shock, oxidative stress, unfolded proteins and infection 21,22 . The SITT process may be an effective solution against climate change and disease. ...
Preprint
Full-text available
Brine shrimp Artemia franciscana is a commercially important species in aquaculture and is well adapted to harsh environmental conditions. In this study, an A. franciscana population selectively bred for induced thermotolerance over 12 generations (TF12) was subjected to transcriptomic analysis relative to the non-selective population (CF12) to determine the effect of selective breeding for induced thermotolerance (SITT). The transcriptomic response of CF12 to non-lethal heat stress (NLHS) from ITT within one generation was also investigated. A total of 232 and 218 differentially expressed genes (DEGs) were upregulated and downregulated in TF12, respectively, whereas fewer DEGs were upregulated (114) and downregulated (180) in CF12 exposed to NLHS, suggesting a more complex mechanism of thermal adaptation in SITT than in ITT. The TF12 population exhibited a reduction in immune processes as indicated by the downregulation of several pattern-recognition receptors. Conversely, stress protein genes such as those of heat shock proteins (HSPs) and of antioxidant enzymes and genes involved in carbohydrate metabolism, were highly upregulated. ITT downregulated immune-related and growth-related genes, whereas genes involved in energy production and electron carrier activity were upregulated. Our findings suggest possible immunological and physiological mechanisms and molecular pathways involved in adaptation of A. franciscana to thermal stress.
... As a result of the imposition of a stress factor (biotic or abiotic), metabolic changes occur in the cells. The cell's response to stressful stimuli includes, among others, an increase in the expression of heat shock genes, and then the synthesis of heat shock proteins also called chaperones [31]. ...
Article
Full-text available
Meloidogyne hapla is one of the most important nematode pathogens. It is a sedentary, biotrophic parasite of plants that overwinters in the soil or in diseased roots. The development of M. hapla is temperature dependent. Numerous studies have been performed on the effect of temperature on the development of M. hapla, but only a few of them analyzed the heat shock protein (hsp) genes. The aim of the study was to perform expression profiling of eight hsp genes (Mh-hsp90, Mh-hsp1, Mh-hsp4, Mh-hsp6, Mh-hsp60, Mh-dnj19, Mh-hsp43, and Mh-hsp12.2) at two development stages of M. hapla, i.e., in eggs and second-stage juveniles (J2). The eggs and J2 were incubated under cold stress (5 °C), heat stress (35 °C, 40 °C), and non-stress (10 °C, 20 °C, and 30 °C) conditions. Expression profiling was performed by qPCR. It was demonstrated that only two genes, Mh-hsp60 and Mh-dnj19, have been upregulated by heat and cold stress at both development stages. Heat stress upregulated the expression of more hsp genes than cold stress did. The level of upregulation of most hsp genes was more marked in J2 than in eggs. The obtained results suggest that the Mh-hsp90 and Mh-hsp1 genes can be used as bioindicators of environmental impacts on nematodes of the Meloidogyne genus.
Article
Full-text available
Understanding how freshwater species respond to changes in salinity is essential for comprehending their distribution and ecological niches. We investigated the osmoregulatory mechanisms of the freshwater catfish Rhamdia quelen, which could justify occasional incursions and residence in brackish waters and their resistance to salt treatments in aquaculture. Specimens were exposed to freshwater (control) and dilute seawater (15 ‰) for 6 h. It was measured: plasma osmolality and Na⁺, K⁺, and Cl⁻ concentrations; gills and kidneys Na⁺/K⁺-ATPase (NKA), V(H⁺)-ATPase (VAT), and carbonic anhydrase (CA) enzymes activities; gills enzymatic immunolocalization and Heat Shock Protein (HSP70) abundance. Our results showed that R. quelen presented hypoosmotic/ionic regulation in 15 ‰. However, there was no significant effect of the increased salinity on enzymatic activities in the gills and kidneys, or in the abundance of HSP70. All enzymatic activities, except for CA, were higher in the kidneys compared to the gills. The immunolocalization of NKA and VAT was observed in ionocytes of the interlamellar region of the gills, and the pavement cells, while CA was detected in both ionocytes and red blood cells. The lack of change in the abundance of HSP70 may be attributed to the protein’s high constitutive abundance in the gill cells. The results indicate a complex regulatory network in response to high salinity, showing that brief periods of salty water exposure could be enough for the species to move or spread in brackish waters or survive aquaculture disease treatments.
Article
Anthropogenic threats such as water infrastructure, land-use changes, overexploitation of fishes and other biological resources, invasive species and climate change present formidable challenges to freshwater biodiversity. Historically, management of fish and fishery species has largely been based on studies of population- and community-level dynamics; however, the emerging field of conservation physiology promotes the assessment of individual fish health as a key management tool. Fish health is highly sensitive to environmental disturbances and is also a fundamental driver of fitness, with implications for population dynamics such as recruitment and resilience. However, the mechanistic links between particular anthropogenic disturbances and changes in fish health, or impact pathways, are diverse and complex. The diversity of ways in which fish health can be measured also presents a challenge for researchers deciding on methods to employ in studies seeking to understand the impact of these threats. In this review, we aim to provide an understanding of the pathway through which anthropogenic threats in freshwater ecosystems impact fish health and the ways in which fish health components impacted by anthropogenic threats can be assessed. We employ a quantitative systematic approach to a corpus of papers related to fish health in freshwater and utilize a framework that summarizes the impact pathway of anthropogenic threats through environmental alterations and impact mechanisms that cause a response in fish health. We found that land-use changes were the most prolific anthropogenic threat, with a range of different health metrics being suitable for assessing the impact of this threat. Almost all anthropogenic threats impacted fish health through two or more impact pathways. A robust understanding of the impact pathways of anthropogenic threats and the fish health metrics that are sensitive to these threats is crucial for fisheries managers seeking to undertake targeted management of freshwater ecosystems.
Article
Full-text available
Chaperonins are high molecular mass double-ring structures composed of 60-kDa protein subunits. In the hyperthermophilic archaeon Sulfolobus shibatae the two chaperonin proteins represent approximate to 4% of its total protein and have a combined intracellular concentration of >30 mg/ml. At concentrations greater than or equal to 0.5 mg/ml purified chaperonins form filaments in the presence of Mg2+ and nucleotides. Filament formation requires nucleotide binding (not hydrolysis), and occurs at physiological temperatures in biologically relevant buffers, including a buffer made from cell extracts. These observations suggest that chaperonin filaments may exist in vivo and the estimated 4600 chaperonins per cell suggest that such filaments could form an extensive cytostructure. We observed filamentous structures in unfixed, uranyl-acetate-stained S. shibatae cells, which resemble the chaperonin filaments in size and appearance. ImmunoGold (Janssen) labeling using chaperonin antibodies indicated that many chaperonins are associated with insoluble cellular structures and these structures appear to be filamentous in some areas, although they could not be uranyl-acetate-stained. The existence of chaperonin filaments in vivo suggests a mechanism whereby their protein-folding activities can be regulated. More generally, the filaments themselves may play a cytoskeletal role in Archaea.
Article
The precise shape of a protein is a crucial factor in its function. How do proteins become folded into the right conformation? Molecular chaperones and protein folding catalysts bind to developing polypeptides in the cytoplasm and ensure correct folding and transport. This Guidebook catalogues the latest information on nearly 200 of these molecules, including the important class of heat shock proteins; each entry is written by leading researchers in the field.
Chapter
Environmental temperature varies in time and space on timescales ranging from a few hours to long-term climate change. Organisms are therefore continually challenged to regulate and maintain functional capacities as their thermal environment changes. This volume brings together many of the leading workers in thermal biology, with backgrounds spanning the disciplines of molecular biology, cell biology, physiology, zoology, ecology and evolutionary biology, to discuss the responses of a wide range of species to temperature change at all scales of organization, ranging through the molecular, cellular, organismal, population and ecosystem levels. The volume provides an important and comprehensive contribution to the study of temperature adaptation, which, given the concern about global climate change, will provide much to interest a wide range of biologists.
Article
The objective of this book is to make analytical methods available to students of ecology. The text deals with concepts of energy exchange, gas exchange, and chemical kinetics involving the interactions of plants and animals with their environments. The first four chapters are designed to show the applications of biophysical ecology in a preliminary, sim­ plified manner. Chapters 5-10, treating the topics of radiation, convec­ tion, conduction, and evaporation, are concerned with the physical environment. The spectral properties of radiation and matter are thoroughly described, as well as the geometrical, instantaneous, daily, and annual amounts of both shortwave and longwave radiation. Later chapters give the more elaborate analytical methods necessary for the study of photosynthesis in plants and energy budgets in animals. The final chapter describes the temperature responses of plants and animals. The discipline of biophysical ecology is rapidly growing, and some important topics and references are not included due to limitations of space, cost, and time. The methodology of some aspects of ecology is illustrated by the subject matter of this book. It is hoped that future students of the subject will carry it far beyond its present status. Ideas for advancing the subject matter of biophysical ecology exceed individual capacities for effort, and even today, many investigators in ecology are studying subjects for which they are inadequately prepared. The potential of modern science, in the minds and hands of skilled investigators, to of the interactions of organisms with their advance our understanding environment is enormous.
Article
We describe a novel approach to evaluate quantitatively the amounts of denatured proteins in cells upon heat exposure. A thiol compound, diamide [azodicarboxylic acid bis (dimethylamide)] causes protein cross-linking with exposed sulfhydryl residues of denatured proteins. Since denatured proteins expose normally well-hidden sulfhydryl groups, these will be preferentially cross-linked by diamide. Thus diamide acts to 'trap' denatured proteins. We observed that protein aggregates (high molecular weight protein aggregates, HMA) appeared on SDS-polyacrylamide gels run under non-reducing conditions and that the amount of HMA can be quantified by scanning the gels using a gas flow counter. Heating cells followed by a fixed dose of diamide exposure resulted in HMA increases in a heat-dose dependent manner, demonstrating that the quantitation of HMA could ve as a measure of heat-denatured proteins. We compared thermotolerant and nontolerant cells and found decreased HMA in tolerant cells upon heat treatment. As an attempt to examine the kinetics of protein renaturation (or 'repair'), we measured the amounts of aggregates formed by the addition of diamide at various times after heat shock. Such experiments demonstrate an equally rapid disappearance of HMA in previously unheated and in thermotolerant cells. Levels of HMA in tolerant cells increased significantly after electroporation of HSP70 specific mAbs, suggesting an involvement of HSP70 in reducing HMA levels in thermotolerant cells upon heat exposure. Immunoprecipitation studies using anti-HSP70 antibody indicated an association of HSP70 with heat-denatured proteins. Our results suggest that heat induces protein denaturation, and that elevated level of HSP70 present in thermotolerant cells protects them by reducing the level of protein denaturation rather than by facilitating the 'repair' (or degradation) process.