Tải bản đầy đủ (.pdf) (14 trang)

Tài liệu Báo cáo khoa học: The heat shock factor family and adaptation to proteotoxic stress pdf

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (449.23 KB, 14 trang )

MINIREVIEW
The heat shock factor family and adaptation to
proteotoxic stress
Mitsuaki Fujimoto and Akira Nakai
Yamaguchi University School of Medicine, Ube, Japan
Introduction
All living organisms respond to elevated temperatures
by producing a set of highly conserved proteins,
known as heat shock proteins (HSP) [1]. This response
is called the heat shock response, and is a universal
mechanism of protection against proteotoxic stress,
including heat shock and oxidative stress. In Escheri-
chia coli, heat shock genes are under the control of a
specific transcription factor, r32, which directs the
core RNA polymerase to promoters [2]. In eukaryotes,
the heat shock response is regulated mainly at the level
of transcription by heat shock factors (HSFs) [3]. Heat
shock genes, such as HSP110, HSP90, HSP70, HSP40
and HSP27, contain heat shock elements (HSEs) com-
posed of at least three inverted repeats of the highly
conserved consensus sequence nGAAn in the proximal
promoter region [4]. Here we call them ‘classical heat
shock genes’, which encode major HSPs or molecular
chaperones. Heat shock triggers the conversion of an
HSF1 monomer in a metazoan species that is nega-
tively regulated by HSPs into a trimer that binds to
Keywords
evolution; heat shock; protein homeostasis;
protein-misfolding disorder; transcription
factor; vertebrate
Correspondence


Akira Nakai, Department of Biochemistry
and Molecular Biology, Yamaguchi
University School of Medicine, Minami-
Kogushi 1-1-1, Ube 755-8505, Japan
Fax: 81 836 22 2315
Tel: 81 836 22 2214
E-mail:
(Received 10 May 2010, revised 7 July
2010, accepted 23 July 2010)
doi:10.1111/j.1742-4658.2010.07827.x
The heat shock response was originally characterized as the induction of a
set of major heat shock proteins encoded by heat shock genes. Because
heat shock proteins act as molecular chaperones that facilitate protein fold-
ing and suppress protein aggregation, this response plays a major role in
maintaining protein homeostasis. The heat shock response is regulated
mainly at the level of transcription by heat shock factors (HSFs) in eukary-
otes. HSF1 is a master regulator of the heat shock genes in mammalian
cells, as is HSF3 in avian cells. HSFs play a significant role in suppressing
protein misfolding in cells and in ameliorating the progression of Caenor-
habditis elegans, Drosophila and mouse models of protein-misfolding
disorders, by inducing the expression of heat shock genes. Recently, numer-
ous HSF target genes were identified, such as the classical heat shock genes
and other heat-inducible genes, called nonclassical heat shock genes in this
study. Importance of the expression of the nonclassical heat shock genes
was evidenced by the fact that mouse HSF3 and chicken HSF1 play a sub-
stantial role in the protection of cells from heat shock without inducing
classical heat shock genes. Furthermore, HSF2 and HSF4, as well as
HSF1, shown to have roles in development, were also revealed to be neces-
sary for the expression of certain nonclassical heat shock genes. Thus, the
heat shock response regulated by the HSF family should consist of the

induction of classical as well as of nonclassical heat shock genes, both of
which might be required to maintain protein homeostasis.
Abbreviations
BRG1, brahma-related gene 1; DAF-16, abnormal dauer formation 16; HR, hydrophobic heptad repeat; HSE, heat shock element; HSF, heat
shock factor; HSP, heat shock protein; MEF, mouse embryonic fibroblast; polyQ, polyglutamine.
4112 FEBS Journal 277 (2010) 4112–4125 ª 2010 The Authors Journal compilation ª 2010 FEBS
the HSE with high affinity, and the bound HSF1 rap-
idly induces a robust activation of the classical heat
shock genes [5,6].
There is a single gene encoding HSF in yeast, in
Caenorhabditis elegans and in Drosophila. HSF is
required not only for the heat shock response, but also
for cell growth and differentiation in yeast [7]. In verte-
brates, there are multiple HSF genes, which encode
members of the HSF family (HSF1–4). In mammals,
as in yeast and Drosophila, the HSF1 is required for
the heat shock response, whereas HSF3 is required for
this response in avian species [8,9]. Both mouse HSF1
and chicken HSF3 are necessary for thermotolerance,
at least through the expression of classical heat shock
genes [10,11]. In addition to their role in the heat
shock response, mouse HSFs are critical in develop-
mental processes such as gametogenesis and neuro-
genesis, in the maintenance of sensory and ciliated
tissues, and in immune responses [12–14]. HSF1- and
HSF3-mediated mechanisms of cellular adaptation to
heat shock have been analyzed in detail in chicken
cells, and were considered specific to chicken cells as it
was believed until recently that HSF3 was an avian-
specific factor. In this minireview, we summarize the

evolution of the HSF gene family and HSF-mediated
mechanisms of cellular adaptation to stress in verte-
brates by comparing mammalian and avian cells, and
also review HSF-mediated mechanisms of adaptation
to pathological states related to protein misfolding.
Evolution of the vertebrate HSF gene
family
An HSF protein that binds to the HSEs in the HSP
genes was purified from heat shock-induced Saccharo-
myces cerevisiae, Drosophila and human cells [3]. Anti-
bodies against HSF were used to isolate a single copy
of the S. cerevisiae HSF gene [15–17]. Thereafter, a
single HSF gene was isolated from another budding
yeast Kluyveromyces lactis, and from the fission yeast
Schizosaccharomyces pombe by cross-hybridization
[18,19]. A single copy of the Drosophila HSF was iso-
lated by screening a library with oligonucleotide
probes derived from HSF peptide sequencing. In mam-
mals, human HSF1 and a second HSF gene, HSF2,
were isolated by screening a library with degenerate
oligonucleotide probes [20,21], and the mouse HSF1
and HSF2 genes were isolated by cross-hybridization
with a human HSF1 cDNA probe [22]. In chicken,
HSF1, HSF2 and a third HSF gene, HSF3, were iso-
lated by cross-hybridization with a mouse HSF1
cDNA probe [23]. Furthermore, another HSF gene,
HSF4, was isolated from human and mouse cells by
the screening of human and mouse cDNA libraries
with a chicken HSF3 cDNA probe [24,25], but a mam-
malian orthologue of the chicken HSF3 gene was not

identified. Therefore, HSF3 was considered specific to
avian species, and HSF4 was considered specific to
mammalian species [5,8,9,12].
Although human and mouse genome sequences have
become available [26,27], no HSF3-related sequence
was identified in silico from the genome database [28].
However, analysis of the chicken genome enabled com-
parison of the syntenic regions [29], where the same
genes occur in a similar order along the chromosomes
of different organisms [30]. For example, HSF2 was
flanked by the SERINCI gene in human, mouse and
chicken orthologous segments (Fig. 1) [31]. Likewise,
the chicken HSF3 gene was located between Vsig4 and
HEPH on chromosome 4, and orthologous segments
containing the two genes were found on the human
and mouse X chromosome. Sequencing of a region
between the two genes revealed the mouse HSF3 gene.
Although sequences related to HSF3 were also
observed in an orthologous region of the human gen-
ome, this genomic segment is likely to be an HSF3
pseudogene as no transcript was identified [31]. Fur-
thermore, HSF4 was located in a region between the
TRADD–FBXL8 genes and the NoL3 gene in the
human and mouse genomes, and chicken HSF4 was
identified in an orthologous segment [31].
Comparison of the predicted amino acids of four
members of the vertebrate HSF family revealed that
sequences of the DNA-binding and trimerization
[hydrophobic heptad repeat (HR)-A ⁄ B] domains are
well conserved (Fig. 2) [31]. The identity of the amino

acid sequence in the DNA-binding domain of mouse
HSF1 was much lower for mouse HSF3 (53%) than
for mouse HSF2 or HSF4 (70% and 76%, respec-
tively). Furthermore, the amino acid sequence of the
DNA-binding domain of mouse HSF3 was only 60%
homologous to that of chicken HSF3, whereas the
sequences of mouse HSF1, HSF2 and HSF4 were
much more identical to the corresponding domains of
chicken HSF1, HSF2 and HSF4 (92%, 86% and 71%
identity, respectively). Moreover, a phylogenetic tree,
which was generated from full-length amino acid
sequences of the HSF family, showed the relatedness
of mouse HSF3 with chicken HSF3 to be much
weaker than that of mouse HSF1 with chicken HSF1,
that of mouse HSF2 with chicken HSF2, or even that
of mouse HSF4 with chicken HSF4 (Fig. 3). These
estimations indicate that the nucleic acid sequences of
HSF3 diverged most quickly during evolution, whereas
those of HSF1 and HSF2 were similarly conserved.
The phylogenetic tree also demonstrates the amino
M. Fujimoto and A. Nakai Evolution and function of the HSF family
FEBS Journal 277 (2010) 4112–4125 ª 2010 The Authors Journal compilation ª 2010 FEBS 4113
acid sequence of HSF1 to be most closely related with
that of HSF4 among the HSF family, and the amino
acid sequence of HSF2 to be most closely related with
that of HSF3. These findings are consistent with the
assertion that two rounds of whole-genome duplication
occurred in the vertebrate lineage (Fig. 4) [32–34].
Alignment of the human and chicken HSF genes with
the mouse HSF gene showed that sequences of the ex-

ons are well conserved, whereas those of introns are
not [31], suggesting that four duplicated HSF genes
have been conserved during evolution under selective
pressure, except for human HSF3.
Expression of classical heat shock
genes induced by two heat-responsive
HSFs
After the identification of mammalian HSF1 and
HSF2 [20–22], and of chicken HSF1, HSF2 and HSF3
[23], research was conducted to reveal the factors
responsible for the heat-inducible HSE-binding activ-
ity. In mammalian cells, HSF1 remains an inert mono-
mer in unstressed cells and forms a trimer that binds
to the HSE in response to heat shock [35,36], whereas
the HSE-binding activity of HSF2 is induced during
hemin-induced differentiation of erythroleukemia
cells and is constitutively high during early mouse
development [37–39]. In chicken cells, both an HSF1
monomer and an HSF3 dimer were converted to
homotrimers that bind to the HSE under heat shock
[40]. The disruption of HSF genes in mouse embryonic
fibroblasts (MEFs) clearly demonstrated that mouse
HSF1 is required for the expression of classical heat
shock genes [10], whereas mouse HSF2 is not [41].
Unexpectedly, disruption of chicken HSF3 in chicken
B-lymphocyte DT40 cells resulted in a severe reduction
in the inducible expression of HSP70, and the expres-
sion of HSP110, HSP90a, HSP90b and HSP40 were
not induced at all in chicken HSF3-null cells [11].
These observations imply that duplicated HSF genes

Mouse Chr.15
DGAT1
HSF1
BOP1
SCRT1
Chicken Chr.2
HSF1
HSF1
10 kb
Human Chr.8
Mouse Chr.15
DGAT1
HSF1BOP1
SCRT1
HSF1
BOP1
SERINC1
HSF2
HSF2 SERINC1
Human Chr 6
Mouse Chr.10
Chicken Chr.3
HSF2
SERINC1
HSF2
SERINC1
Chicken Chr.4
HSF3
HEPH
Vsig4

10 kb
20 kb
HSF2 SERINC1
Human Chr.
6
HSF3
Human Chr.X
HEPH
Vsig4
Mouse Chr.X
HSF3
Vsig4
100 kb
100 kb
HSF3
HEPH
Chicken Chr.11
TRADD FBXL8
HSF4
NL3
HSF4
Mouse Chr.8
NoL3TRADD FBXL8
HSF4
Human Chr.16
N
o
L3
HSF4
10 kb

Fig. 1. Comparative genomic analysis of
orthologous segments containing vertebrate
HSF genes. The location of each segment is
as follows: human Chr.8 q24.3 and mouse
Chr.15 D3 for HSF1; human Chr.6 q22.31,
mouse Chr.10 B4 and chicken 63.95–
63.98 Mb for HSF2; human Chr. X q12,
mouse Chr. X B4 and chicken 0.252–
0.265 Mb for HSF3; and human Chr.16
q22.1, mouse Chr.8 D3 and chicken 2.44–
2.45 Mb for HSF4. A genomic sequence
corresponding to chicken HSF1 cDNA has
not yet been identified. Arrows indicate the
5¢ to 3¢ orientation of each gene. The
chicken HSF1 gene is located on chromo-
some 2, which is present in three copies in
DT40 cells [42]. The gray box in human
chromosome X is probably an HSF3
pseudogene.
Evolution and function of the HSF family M. Fujimoto and A. Nakai
4114 FEBS Journal 277 (2010) 4112–4125 ª 2010 The Authors Journal compilation ª 2010 FEBS
evolved differently in mammalian and avian species
(Fig. 4).
As the amino acid sequence of HSF1 is highly con-
served in mammalian and chicken cells, the functional
difference was examined in more detail. In fact,
chicken HSF1 is dispensable for the expression of the
classical heat shock genes in DT40 cells [42], and the
ectopic expression of chicken HSF1 in MEF cells defi-
cient in mouse HSF1 does not restore the inducible

expression of the classical heat shock genes [28].
Thus, chicken HSF1 lacks the ability to induce the
expression of classical heat shock genes, whereas
mouse HSF1 is a master regulator of these genes.
Interestingly, the amino-terminal region of chicken
HSF1 containing an alanine-rich sequence and the
DNA-binding domain is sufficient to cause the func-
tional difference between the two orthologues [28]. As
chicken HSF1 can bind to the HSE, its amino-terminal
domain might inhibit exposure of the activation
domain to basal transcriptional machinery. Alterna-
tively, the corresponding domain of mouse HSF1, but
not that of chicken HSF1, could recruit components
required for gene activation.
Recent identification of mouse HSF3 enabled us to
examine the functional difference of HSF3 in mouse
and chicken cells [31]. In cells exposed to heat shock,
mouse HSF3 fused to green fluorescent protein moved
into the nucleus, similarly to chicken HSF3 [40], indi-
cating that both chicken and mouse HSF3 are heat-
responsive factors. Furthermore, overexpression of
DBD HR-A/B HR-
C
DHR
hHSF1
hHSF2
5291 16 120 130 203 384 409
415
446
5361 8 112 119 192 360 385

391
422
649407232225
Human
XY
mHSF1
mHSF2
18 2446 224970
hHSF4
76
100
39
6999
28
79
34
97
24
83
mHSF3
mHSF4
HSF1
16 24153353
30 21313976
58 678592
92
82
Mouse
cHSF1
cHSF2

cHSF3
58 678592
92
82
1342
4647 7
2
76
25 193849 2206
Chicken
cHSF4
30 23
285384
CeHSF1
DmHSF
53
59
58
31
30
hHSFY1
31
ScHSF
CeHSF1
53
31
46
32
hHSFX1
20

hHSF5
37
4921 19 123 130 203 396 427
5251
16
120 130 203 380 405
411
442
5351 8 112 119 192 359 384
390
421
4921 10 114 121 194 355
380
4921 18 122 129 202 395 426
4911 21 125 135 208 346 371
376
408
564
1 22 126 133 206 390 415
421
453
4671 17 121 129 202 364 389
395
426
5101 18 122 130 203 393 424
147 150
691161 232
1 90 194 671210 275
1173
833

276 347
424
1 80 198 401
1 37 155 423
1 14 129 596
Fig. 2. Members of the HSF superfamily. Diagrammatic representation of vertebrate and nonvertebrate HSF family members and of human
HSF-related gene products. The percentage identity between human HSF1 and each HSF was established using the computer program
GENETYX-MAC. The number of amino acids of each HSF is shown at the amino-terminal end. c, chicken; Ce, Caenorhabditis elegans ; DHR,
downstream of HR-C; DBD, DNA-binding domain; DHR, downstream of HR-C; Dm, Drosophila melanogaster; h, human; HR, hydrophobic
heptad repeat; m, mouse; Region X, a region upstream of the HR-C domain; Region Y, a C-terminal region downstream of the HR-C-like
domain; Sc, Saccharomyces cerevisiae. hHSF1 (hHSF1-a) [20]; hHSF2 (hHSF2-a) [21]; hHSF4 (hHSF4b) [24]; mHSF1 (hHSF1-a) and mHSF2
(hHSF2-a) [35]; mHSF3 (mHSF3a) [31]; mHSF4 (mHSF4b) [25]; cHSF1, cHSF2 and cHSF3 [23]; cHSF4 (cHSF4b) [31]; DmHSF [Wu 1990];
CeHSF1 (Swiss-P accession no. Q9XW45); ScHSF [Pelham; Parker 1988]; hHSFY1 (SP accession no. Q96LI6) [106,107]; hHSFX1 ⁄ LW-1 (SP
accession no. Q9UBD0); hHSF5 (SP accession no. Q4G112). The hatched box indicates an HR-C-like domain, in which hydrophobic amino
acids are not well conserved. The DBD domain of HSF family members is conserved with one region in hHSFY1, hHSFX1 and hHSF5 that
may not bind to the HSE.
M. Fujimoto and A. Nakai Evolution and function of the HSF family
FEBS Journal 277 (2010) 4112–4125 ª 2010 The Authors Journal compilation ª 2010 FEBS 4115
chicken HSF3 in HSF1-null MEF cells induced the
constitutive and heat-induced expression of classical
heat shock genes. In marked contrast, overexpression
of mouse HSF3 in the same cells did not affect the
expression of classical heat shock genes at all, even
after heat shock. Therefore, mouse HSF3 lacks the
ability to induce the expression of classical heat shock
genes, whereas chicken HSF3 is a master regulator.
Why does mouse HSF3 fail to induce the expression
of classical heat shock genes? It was revealed, by exam-
ining a DNA-binding transcription factor required for
the activation of the GAL genes in response to galac-

tose (GAL4) site-directed luciferase activity, that mouse
HSF3 has strong potential to induce transcription [31].
Deletion analysis showed that the activation domain of
mouse HSF3 is located in its C-terminal region down-
stream of the HR-C-like domain (region Y) whereas
that of chicken HSF3 is located upstream of the HR-C
domain (region X) (Fig. 2). The amino acid sequence
of the activation domain of mouse HSF3 is not con-
served in chicken HSF3, which is consistent with a
functional divergence of the activation domain during
evolution. Domains of mouse HSF3 were swapped with
the corresponding domains of chicken HSF3, and the
chimeras possessing the chicken HSF3 activation
domain induced the expression of the classical heat
shock genes in response to heat shock. In contrast, the
chimeras possessing only the mouse HSF3 activation
domain did not induce their expression. Furthermore,
the C-terminal activation domain of human HSF1
[43–45] was swapped with the mouse HSF3 activation
domain, and the resultant protein did not induce gene
expression in response to heat shock. These results
indicate that the activation domain of mouse HSF3
does not have the potential to activate the classical
heat shock genes.
Human HSF1 recruits brahma-related gene 1
(BRG1), a component of switch ⁄ sucrose nonferment-
ing (SWI ⁄ SNF) chromatin remodeling complexes, to
the HSP70 promoter through direct interaction [46],
and expression of an HSF1 mutant, which cannot
interact with BRG1, did not restore the induction of

HSP70 mRNA expression in HSF1-null MEF cells
during heat shock [47,48]. It was revealed that mouse
HSF3 does not bind to BRG1, or recruit BRG1 to the
HSP70 promoter [31], whereas chicken HSF3 does
bind to and recruit BRG1. These observations indicate
that mouse HSF3 does not induce the expression of
the classical heat shock genes, at least in part because
of its inability to interact with BRG1.
HSF-mediated adaptation to thermal
stress
Heat shock induces both apoptotic and necrotic cell
death, but the pathways of cell death and the factors
hHSF5
hHSFY1
hHSFX1
mHSF3
cHSF3
mHSF2
cHSF2
674
991
1000
1000
1000
1000
602
1000
1000
1000
1000

1000
839
908
0.05
684
HSF3
HSF2
HSF4
HSF1
hHSF2
cHSF4
mHSF4
hHSF4
ScHSF
mHSF1
hHSF1
cHSF1
CeHSF1
DmHSF
Fig. 3. The phylogenetic tree generated in CLUSTAL W [108] for members of the HSF family. Gaps were excluded from all phylogenetic
analyses. The numerals represent bootstrap values (1000 bootstrap replicates were performed). The unrooted tree was drawn using the pro-
gram TREEVIEW [109]. The bar represents 0.05 substitutions per site. Amino acid sequences of the HSF family were shown previously
[28,31].
Evolution and function of the HSF family M. Fujimoto and A. Nakai
4116 FEBS Journal 277 (2010) 4112–4125 ª 2010 The Authors Journal compilation ª 2010 FEBS
that are primarily impaired are not clear, as heat
shock causes various types of stress, including proteo-
toxic stress. Cells in different states of metabolism and
in different stages of differentiation may be induced to
die by different mechanisms, and there must be vari-

ous targets of extremely high temperatures that induce
cell death. Therefore, HSPs should not be the only
proteins that protect against cell death. Nevertheless,
HSPs are recognized as major players in the protec-
tion of cells from heat shock, especially from proteo-
toxic stress [1,2]. As the expression of a set of HSPs is
regulated by HSFs, HSFs should be involved in the
protection of cells from heat shock or proteotoxic
stress [49].
It is well established that cells pretreated with suble-
thal heat shock can survive lethal heat shock. This
phenomenon is called induced thermotolerance, and is
regulated by mouse HSF1 and chicken HSF3 through
the activation of the heat shock genes [10,11]. HSPs
prevent the denaturation and aggregation of cellular
proteins, and support their renaturation when the cells
are recovering [1]. At the same time, HSPs inhibit sev-
eral molecules, such as apoptotic peptidase activating
factor 1 (Apaf-1) and cytochrome c, which are
involved in mitochondria-mediated apoptotic pathways
[50].
Both HSF1 and HSF3 complementarily regulate the
constitutive expression of some HSPs in normally
growing chicken DT40 cells [11,42]. In these cells, a
lack of the two factors resulted in increased sensitivity
to a single exposure to high temperature because of
reduced Hsp90a expression, and the cell cycle is
blocked at the G
2
phase [42]. A similar phenotype was

observed in yeast S. cerevisiae harbouring a mutant
HSF [51,52]. Mouse HSF1 also regulates the expres-
sion of some HSPs, including Hsp90, in various mouse
tissues [53–56]. Therefore, HSFs could be involved in
determining a temperature at which cells can survive
by regulating the constitutive expression of HSPs such
as Hsp90.
Curiously, in chicken cells, HSF1 induced only very
low levels of expression of the classical heat shock
genes, but had a significant effect on the protection of
cells from heat shock [28]. This effect was not medi-
ated through the induction of classical heat shock
genes or regulation of the constitutive expression of
heat shock genes such as Hsp90. HSF1-null MEF cells,
which lack induced expression of the classical heat
shock genes, are more sensitive to high temperatures
than wild-type cells. Remarkably, overexpression of
chicken HSF1 in the HSF1-null MEF cells restored
resistance to heat shock [28]. Moreover, mouse HSF3
HSF1
2RWGD
Mouse cell
HSF1
Non-WGD
HSP induction
High expression
HSF2
HSF4
HSF3
Ancestor cell

in the lens
Polyploid
Chicken cell
HSF3
HSF4
HSF4
HSF2
HSF1
HSF1
> 310 Myr ago
High expression
in the lens
HSF3
HSP induction
HSF2
Fig. 4. A model to explain the evolution of HSF genes. Two rounds of whole-genome duplication (WGD) may have occurred in vertebrate
ancestral cells more than 440 million years ago (Ma) [110,111], which resulted in polyploidization. Thereafter, avian and mammalian cells
evolved differently from an ancestral cell 310 Ma. The expression and function of the four HSF genes were conserved or diverged during
evolution [112]. For example, during mammalian evolution, HSF1 retained the ability to induce the expression of heat shock proteins,
whereas it lost this function during avian evolution. Instead, avian HSF3 retained the function. Expression of the HSF4 gene increased in the
lens during both avian and mammalian evolutions (M. Fujimoto and A. Nakai, unpublished). Diamonds, circles and triangles represent regula-
tory regions driving expression in different tissues.
M. Fujimoto and A. Nakai Evolution and function of the HSF family
FEBS Journal 277 (2010) 4112–4125 ª 2010 The Authors Journal compilation ª 2010 FEBS 4117
was able to protect HSF1-null MEF cells from heat
shock without inducing the expression of the classical
heat shock genes [31]. These observations indicate that
chicken HSF1 and mouse HSF3 protect cells from
thermal stress by regulating the expression of heat-
inducible genes other than classical heat shock genes.

Expression of nonclassical heat shock
genes induced by HSFs
Although the heat shock response was originally char-
acterized based on the expression of a limited number
of classical heat shock genes, HSF1 was recently
revealed to regulate the expression of numerous other
genes in the absence or presence of heat shock. Com-
prehensive analyses of HSF-binding regions in the
whole genome revealed that  3% of genes are direct
targets in heat-shocked cells in yeast and Drosophila
[57,58], and expression of the majority of the target
genes is induced during heat shock in unicellular yeast
[57]. Even in mammalian cells, HSF1 (similarly to
yeast and Drosophila HSF) binds to the promoters of
a great number of genes in the whole genome [59,60],
and about half of the target genes are expressed during
heat shock [59]. We now refer to these heat-inducible
genes that are different from the classical heat shock
genes as ‘nonclassical heat shock genes’. As already
discussed, chicken HSF1 and mouse HSF3 play a sub-
stantial role in the protection of cells from heat shock
[28,31], implying significance of the nonclassical heat
shock genes in this process.
To establish whether the expression of nonclassical
heat shock genes was induced by mouse HSF3 or
chicken HSF1, nonclassical heat shock genes were
identified in MEF cells [31]. Induction of one of the
nonclassical heat shock genes, the gene for the PSD-95 ⁄
Dlg-A ⁄ ZO-1 (PDZ) domain-containing protein
PDZK3 ⁄ PDZD2 ⁄ PAPIN (plakophilin-related armadillo

repeat protein-interacting protein) [61], decreased, but
was still observed in HSF1-null MEF cells during heat
shock [31]. Overexpression of mouse HSF3 or chicken
HSF1 in the HSF1-null MEF cells restored the marked
induction of expression of PDZK3, whereas knock-
down of mouse HSF3 completely abolished the induc-
tion. Induction of another nonclassical heat shock
gene, that for a membrane glycoprotein, prominin-2
(PROM2) [62], was abolished in HSF1-null MEF cells,
but was restored when mouse HSF3 or chicken HSF1
was overexpressed [31]. These observations clearly
demonstrated that evolutionally conserved mouse
HSF3 and chicken HSF1 uniquely regulate only non-
classical genes, suggesting importance of the regulation
of the nonclassical heat shock genes.
It is worth noting that HSF4 also regulates nonclas-
sical heat shock genes in lens cells although it is not a
heat-responsive factor. A set of HSF4-binding regions
was identified in lens cells, and the expression of genes
located in and near these regions was examined [63].
Interestingly, a great number of the genes (33%) were
expressed in response to heat shock, and, unexpect-
edly, the expression of these genes was not induced in
HSF1-null lens cells. Surprisingly, HSF4 was required
for the expression of half of the genes, in part by facil-
itating the binding of HSF1 to the promoters [63].
Moreover, the expression of satellite III repeat
sequences during heat shock was extensively studied
[64,65]. HSF1 is required for expression of the satellite
III gene during heat shock, but HSF2 also greatly

affects its expression, possibly by interacting with
HSF1 [66]. Taken together, all members of the verte-
brate HSF family are involved in the regulation of
gene expression during heat shock (Fig. 5).
HSF is essential in yeast and human
cancer cells
In the budding yeast S. cerevisiae, HSF is essential for
survival under normal conditions [15,16], consistent
with the notion that S. cerevisiae HSF is constitutively
Classical heat shock genes Nonclassical heat shock genes
HSF2
HSF1
HSF1
HSF4
HSF3
HSE
HSE
HSPs
PDZK3
PROM2
Sat III
etc
Protein homeostasis
?
Adaptation to proteotoxic stress
Fig. 5. Adaptation to proteotoxic stress by the HSF family in mice.
HSF1 remains mostly as an inert monomer in unstressed cells, and
is converted to an active trimer that binds to the HSE located in
the proximal promoter region of a limited number of classical heat
shock genes during heat shock, which results in induction of the

expression of HSPs. HSF2 may modulate this process by interact-
ing with HSF1 [6,113]. Members of the HSF family coordinately
bind to the less-conserved HSE located on and near numerous non-
classical heat shock genes, and greatly affect heat-induced expres-
sion of the genes including PDZK3, PROM2 and satellite III
[31,63,66]. HSF3 is not expressed in human cells.
Evolution and function of the HSF family M. Fujimoto and A. Nakai
4118 FEBS Journal 277 (2010) 4112–4125 ª 2010 The Authors Journal compilation ª 2010 FEBS
a trimer that binds to the HSE. However, the binding
of HSF to the promoter of a heat shock gene markedly
increased during heat shock in S. cerevisiae in vivo [67].
Furthermore, HSF was essential even in the fission
yeast S. pombe, in which HSF forms a trimer only
under stress, similar that observed for vertebrate HSF1
[19]. These observations implied that a balance of the
monomer and trimer HSF affects the amount of HSF
bound to the HSE in vivo, but even a small amount of
the trimer could regulate the gene expression, which is
required for survival under normal growth conditions
in unicellular yeasts. In fact, HSF binds to many target
genes in vivo, and their products have a broad range of
biological functions, including protein folding and deg-
radation, energy generation and protein trafficking
[57]. Human HSF2, but not HSF1, forms a trimer and
functionally complements the viability defect of yeast
cells lacking HSF, and both human HSF1 and HSF2
partially rescue the induction of heat-inducible genes,
which is associated with acquired thermotolerance [68].
These observations suggest that the roles of HSF
under normal growth conditions can be distinguished

from those under stress.
Among multicellular organisms, HSF-null Drosophila
was the first to be generated, although a single HSF
was required for oogenesis and early larval develop-
ment, indicating that HSF is dispensable for cell
growth and survival under normal conditions [69].
Subsequently, HSF1-, HSF2- and HSF4-deficient mice
were generated, indicating that these HSFs are also
dispensable [41,70–73]. Detailed analysis of these mice
demonstrated that meiosis was impaired in the absence
of HSF1 and HSF2 in female [53,74] and male [71,75–
77] germ cells. Neuronal differentiation and migration
were affected in the HSF2-null cerebral cortex [78].
Furthermore, the differentiation and maintenance of
sensory placodes required HSF1 and HSF4
[13,54,72,73]. Thus, members of the HSF family exert
essential activities in the absence of stress and are
required for the differentiation of many types of cells
during development [14]. However, they are not neces-
sary for cell proliferation and survival under normal
conditions in multicellular organisms.
Cancer cells proliferate and survive in different ways
from normal cells, and many of the signalling path-
ways and transcription factors display a striking
dependence on the chaperone machinery [79,80]. Fur-
thermore, HSF1 expression is elevated in human can-
cer cells [81,82], suggesting that cancer cells may be
dependent on the heat shock response. Therefore, the
effects of loss of HSF1 function on cancer cell prolifer-
ation and survival were examined. First, human cervi-

cal epithelial HeLa cells stably expressing short hairpin
RNA for HSF1 were generated (these cells show >
95% reduction in the HSF1 level), and were highly
sensitive to combined treatment with both elevated
temperature and anticancer reagents [83]. Then,
decreased lymphomagenesis in a p53-deficient mouse
model was shown in the absence of HSF1 [84]. Unex-
pectedly, in addition to being required for carcinogene-
sis in mice, HSF1 is required for proliferation and
survival in various human cancer cell lines, including
HeLa cells, but not in normal or immortalized cells
[85]. This observation suggests the possibility of com-
mon HSF-mediated mechanisms for cell proliferation
and survival in cancer cells and in yeast cells.
Is the HSF1 in cancer cells activated? As HSF1
expression, which is correlated with HSE-binding
activity [35], is elevated in human cancer cells [81,82],
the HSE-binding activity of HSF1 might be higher in
cancer cells than in normal cells. Even so, the HSE-
binding activity is robustly induced in response to heat
shock in cancer cells, such as HeLa cells, compared
with normal cells [35,36] or in fission yeast cells [19].
HSF1 is involved in regulating translation, ribosome
biogenesis and glucose metabolism in cancer cells [85],
and is also required for the expression of numerous
genes, including those for inflammatory cytokines,
chemokine-related genes and interferon-related genes,
even in normally growing primary cultures of MEF
cells [86]. Furthermore, the ability of HSF1 to form a
trimer is required for the gene expression [87]. Taken

together, only a little amount of HSF1 trimer regulates
the gene expression in normal cells, which is not
required for cell growth and survival. However, the tri-
meric HSF1 is increased in cancer cells and may regu-
late the expression of genes, which is indispensable for
cell growth and survival, as in fission yeast cells for
example [19].
Adaptation to misfolding-related
pathological conditions
An imbalance of protein homeostasis is associated with
aging and age-related pathological conditions such as
neurodegenerative disorders, including Alzheimer’s dis-
ease, Parkinson’s disease, amyotrophic lateral sclerosis,
prion disease and polyglutamine diseases. These diseases
are characterized by conformational changes in disease-
causing proteins that results in misfolding and aggrega-
tion, and are therefore termed protein-misfolding
disorders or protein conformational disorders [49,88].
Polyglutamine (polyQ) diseases are caused by an expan-
sion of CAG repeats, coding for glutamine, in their
respective proteins. Misfolding and aggregation of
aggregation-prone polyQ proteins results in cellular
M. Fujimoto and A. Nakai Evolution and function of the HSF family
FEBS Journal 277 (2010) 4112–4125 ª 2010 The Authors Journal compilation ª 2010 FEBS 4119
toxicity. Gain of HSF1 function significantly inhibits
the aggregation of polyQ protein and prolongs life span
in C. elegans models of polyQ diseases, whereas loss of
HSF1 function accelerates the aggregation of polyQ and
shortens life span [89,90]. It is considered that HSF1
function is mediated through the expression of HSPs

[49,88]. Interestingly, a forkhead box (FOXO) family
transcription factor, DAF-16 (abnormal dauer forma-
tion 16), which is a component of the insulin-like signal-
ling pathway that regulates life span, also inhibits polyQ
aggregates, indicating that aging and protein homeosta-
sis are highly related [89,90]. In a C. elegans model of
Alzheimer’s disease, HSF1 inhibited the formation of
toxic aggregates of an aggregation-prone peptide Ab
(1–42) whereas DAF-16 promoted the formation of less-
toxic high-molecular-weight aggregates [91]. Thus,
HSF1 and DAF-16 regulate distinct pathways that
reduce the toxicity of aggregation-prone proteins.
Among mouse polyQ models, the R6 ⁄ 2 polyQ model
has been extensively studied as it is transgenic only for
the 5¢ end of the human huntingtin gene carrying 115–
150 CAG repeat expansions [92]. The formation of
polyQ aggregates is observed not only in the brain but
also in nonneuronal tissues, including the skeletal mus-
cle, heart, liver and pancreas, in mice [93]. Ubiquitous
overexpression of HSP70 in the R6 ⁄ 2 Huntington’s
model had no effect on the life span or neuronal phe-
notypes of the mice and delayed aggregation only
slightly [94,95]. There is only one HSF1 transgenic
mouse model, in which an actively mutated HSF1 is
expressed in tissues such as the testis, skeletal muscle,
heart and stomach, but not in the brain [75]. Remark-
ably, overexpression of an active HSF1 in nonneuronal
tissues in R6 ⁄ 2 mice crossed with HSF1 transgenic
mice suppressed polyQ aggregates, at least in skeletal
muscle, and markedly extended the life span [96].

Inversely, HSF1 deficiency dramatically shortened the
life span of the prion disease model mice, in which
scrapie prions were inoculated [97], and even resulted
in impaired protein homeostasis of the untreated neu-
ronal cells in some genetic backgrounds [98]. These
observations imply significant beneficial effects of the
overexpression of an active HSF1 on the progression
of protein-misfolding disorders in mice. Interestingly,
mouse HSF3 and chicken HSF1 suppressed the forma-
tion of aggregates in a cellular polyQ model [28,31],
suggesting that the nonclassical heat shock genes as
well as classical heat shock genes play roles in amelio-
rating disease progression.
One therapeutic strategy for protein-misfolding dis-
orders such as polyQ disease would be to elevate the
levels of HSPs that assist normal protein folding and
prevent abnormal folding and aggregation [99]. It was
shown that treatment with arimoclomol, a co-inducer
of HSPs through activating HSF1, delays disease pro-
gression in the amyotrophic lateral sclerosis mouse
model, which overexpresses human mutant Cu ⁄ Zn
superoxide dismutase-1 [100]. HSF1-activating reagents,
17-allylamino-17-demethoxygeldanamycin (17-AAG) and
geranylgeranylacetone (GGA), also ameliorated disease
progression in a Drosophila model of spinocerebellar
ataxias [101] or in a mouse model of spinal and bulbar
muscular atrophy [102,103]. Furthermore, novel small
molecules that activate HSF1 were identified and
shown to ameliorate protein misfolding and toxicity of
polyQ aggregates [104,105]. Thus, activation of HSF1

is actually a promising therapeutic approach for pro-
tein-misfolding disorders.
Conclusions and perspectives
We have learned from the identification and character-
ization of HSF1 and HSF3 in chicken and mouse cells
that the function of these two heat-responsive factors
has diverged greatly during the evolution of vertebrates,
even though the nucleic acid sequence of each has been
well conserved. Importantly, chicken HSF1 and mouse
HSF3 are involved in protecting cells from heat shock
and in maintaining protein homeostasis without induc-
ing the expression of the classical heat shock genes. It
was revealed recently that there are numerous nonclas-
sical heat shock genes, whose expression is induced dur-
ing heat shock in various organisms. Remarkably,
chicken HSF1 and mouse HSF3, as well as mammalian
HSF2 and HSF4, play a role in inducing the expression
of only nonclassical heat shock genes. These observa-
tions suggest the importance of the regulation and func-
tion of the nonclassical heat shock genes. Analysis of
these new findings will help us to understand why the
activation of HSF1 suppresses the progression of pro-
tein-misfolding disorders more than HSPs and should
be beneficial in identifying pathways involved in adap-
tation to proteotoxic stress. Furthermore, these analy-
ses would develop our understanding of the biological
significance of the heat shock response.
Acknowledgements
We thank members of our laboratory for discussions
and Naoki Hayashida for comments on the manu-

script. This work was supported in part by Grants-in-
Aid for Scientific Research and on Priority Area-a
Nuclear System of DECODE, from the Ministry of
Education, Culture, Sports, Science and Technology,
Japan, and by the Yamaguchi University Research
Project on STRESS.
Evolution and function of the HSF family M. Fujimoto and A. Nakai
4120 FEBS Journal 277 (2010) 4112–4125 ª 2010 The Authors Journal compilation ª 2010 FEBS
References
1 Lindquist S (1986) The heat-shock response. Ann Rev
Biochem 55, 1151–1191.
2 Gross CA, Strauss DB, Erickson JW & Yura T (1990)
The function and regulation of heat shock proteins in
Escherichia coli. In Stress Proteins in Biology and Medi-
cine (Morimoro RI, Tissieres A & Georgopoulos C eds),
pp. 167–189. Cold Spring Harbor Laboratory Press,
Cold Spring Harbor, NY.
3 Wu C (1995) Heat shock transcription factors: struc-
ture and regulation. Annu Rev Cell Biol 11, 441–469.
4 Fernandes M, O’Breien T & Lis JT (1994) Structure
and regulation of heat shock gene promoters. In The
Biology of Heat Shock Proteins and Molecular Chaper-
ones (Morimoro RI, Tissieres A & Georgopoulos C
eds), pp. 375–393. Cold Spring Harbor Laboratory
Press, Cold Spring Harbor, NY.
5 Morimoto RI (1998) Regulation of the heat shock
transcriptional response: cross talk between a family of
heat shock factors, molecular chaperones, and negative
regulators. Genes Dev 12, 3788–3796.
6 Bjo

¨
rk JK & Sistonen L (2010) Regulation of the mem-
bers of the mammalian heat shock factor family. FEBS
J 277, 4126–4139.
7 Morano KA & Thiele DJ (1999) Heat shock factor
function and regulation in response to cellular stress,
growth, and differentiation signals. Gene Expr 7, 271–
282.
8 Nakai A (1999) New aspects in the vertebrate heat
shock factor system: Hsf3 and Hsf4. Cell Stress Chap-
erones 4, 86–93.
9 Pirkkala L, Nykanen P & Sistonen L (2001) Roles of the
heat shock transcription factors in regulation of the heat
shock response and beyond. FASEB J 15, 1118–1131.
10 McMillan DR, Xiao X, Shao L, Graves K &
Benjamin IJ (1998) Targeted disruption of heat shock
transcription factor 1 abolishes thermotolerance and
protection against heat-inducible apoptosis. J Biol
Chem 273, 7523–7528.
11 Tanabe M, Kawazoe Y, Takeda S, Morimoto RI,
Nagata K & Nakai A (1998) Disruption of the HSF3
gene results in the severe reduction of heat shock gene
expression and loss of thermotolerance. EMBO J 17,
1750–1758.
12 Akerfelt M, Trouillet D, Mezger V & Sistonen L
(2007) Heat shock factors at a crossroad between stress
and development. Ann NY Acad Sci 1113, 15–27.
13 Nakai A (2009) Heat shock transcription factors and
sensory placode development. BMB Rep 42, 631–635.
14 Abane R & Mezger V (2010) Roles of heat shock fac-

tors in gametogenesis and development. FEBS J 277,
4150–4172.
15 Sorger PK & Pelham HRB (1988) Yeast heat shock
factor is an essential DNA-binding protein that
exhibits temperature-dependent phosphorylation. Cell
54, 855–864.
16 Wiederrecht G, Seto D & Parker C (1988) Isolation of
the gene encoding the S. cerevisiae heat shock tran-
scription factor. Cell 54, 841–853.
17 Clos J, Westwood JT, Becker PB, Wilson S & Wu C
(1990) Molecular cloning and expression of a hexamer-
ic Drosophila heat shock factor subject to a negative
regulation. Cell 63, 1085–1097.
18 Jakobsen BK & Pelham HRB (1991) A conserved
heptapeptide restrains the activity of the yeast heat
shock transcription factor. EMBO J 10, 369–375.
19 Gallo GJ, Prentice H & Kingston RE (1993) Heat
shock factor is required for growth at normal tempera-
ture in the fission yeast Schizosacccharomyces pombe.
Mol Cell Biol 13, 749–761.
20 Rabindran SK, Giorgi G, Clos J & Wu C (1991)
Molecular cloning and expression of a human heat
shock factor, HSF1. Proc Natl Acad Sci USA 88,
6906–6910.
21 Schuetz TJ, Gallo GJ, Sheldon L, Tempst P & Kings-
ton RE (1991) Isolation of a cDNA for HSF2: evi-
dence for two heat shock factor genes in humans. Proc
Natl Acad Sci USA 88, 6911–6915.
22 Sarge KD, Zimarino V, Holm K, Wu C & Morimoto
RI (1991) Cloning and characterization of two mouse

heat shock factors with distinct inducible and constitu-
tive DNA-binding ability. Genes Dev 5, 1902–1911.
23 Nakai A & Morimoto RI (1993) Characterization of a
novel chicken heat shock transcription factor, heat
shock factor 3, suggests a new regulatory pathway.
Mol Cell Biol 13, 1983–1997.
24 Nakai A, Tanabe M, Kawazoe Y, Inazawa J, Morimoto
RI & Nagata K (1997) HSF4, a new member of the
human heat shock factor family which lacks properties
of a transcriptional activator. Mol Cell Biol 17, 469–481.
25 Tanabe M, Sasai N, Nagata K, Liu XD, Liu PC, Thi-
ele DJ & Nakai A (1999) The mammalian HSF4 gene
generates both an activator and a repressor of heat
shock genes by alternative splicing. J Biol Chem 274,
27845–27856.
26 International Human Genome Sequencing Consortium
(2001) Initial sequencing and analysis of the human
genome. Nature 409, 860–921.
27 Mouse Genome Sequencing Consortium (2002) Initial
sequencing and comparative analysis of the mouse gen-
ome. Nature 420, 520–562.
28 Inouye S, Katsuki K, Izu H, Fujimoto M, Sugahara
K, Yamada S, Shinkai Y, Oka Y, Katoh Y & Nakai A
(2003) Activation of heat shock genes is not necessary
for protection by heat shock transcription factor 1
against cell death due to a single exposure to high tem-
peratures. Mol Cell Biol 23, 5882–5895.
29 International Chicken Genome Sequencing Consortium
(2004) Sequence and comparative analysis of the
M. Fujimoto and A. Nakai Evolution and function of the HSF family

FEBS Journal 277 (2010) 4112–4125 ª 2010 The Authors Journal compilation ª 2010 FEBS 4121
chicken genome provide unique perspectives on verte-
brate evolution. Nature 432, 695–716.
30 Koonin EV, Aravind L & Kondrashov AS (2000) The
impact of comparative genomics on our understanding
of evolution. Cell 101, 573–576.
31 Fujimoto M, Hayashida N, Katoh T, Oshima K,
Shinkawa T, Prakasam R, Tan K, Inouye S, Takii R
& Nakai A (2010) A novel mouse HSF3 has the
potential to activate non-classical heat shock genes
during heat shock. Mol Biol Cell 20, 106–116.
32 Ohno S (1970) Evolution by Gene Duplication. Springer
Verlag, New York.
33 Holland PW, Garcia-Ferna
`
ndez J, Williams NA &
Sidow A (1994) Gene duplications and the origins of
vertebrate development. Development (Suppl.), 125–
133.
34 Putnam NH, Butts T, Ferrier DE, Furlong RE,
Hellsten U, Kawashima T, Robinson-Rechavi M,
Shoguchi E, Terry A, Yu JK et al. (2008) The amphi-
oxus genome and the evolution of the chordate karyo-
type. Nature 453, 1064–1071.
35 Sarge KD, Murphy SP & Morimoto RI (1993)
Activation of heat shock gene transcription by heat
shock factor 1 involves oligomerization, acquisition of
DNA-binding activity, and nuclear localization and
can occur in the absence of stress. Mol Cell Biol 13,
1392–1407.

36 Baler R, Dahl G & Voellmy R (1993) Activation of
human heat shock genes is accompanied by oligomeri-
zation, modification, and rapid translocation of heat
shock transcription factor HSF1. Mol Cell Biol 13,
2486–2496.
37 Sistonen L, Sarge KD, Phillips B, Abravaya K &
Morimoto RI (1992) Activation of heat shock factor 2
during hemin-induced differentiation of human
erythroleukemia cells. Mol Cell Biol 12, 4104–4111.
38 Sistonen L, Sarge KD & Morimoto RI (1994) Human
heat shock factors 1 and 2 are differentially activated
and can synergistically induce hsp70 gene transcription.
Mol Cell Biol 14, 2087–2099.
39 Murphy SP, Gorzowski JJ, Sarge KD & Phillips B
(1994) Characterization of constitutive HSF2 DNA-
binding activity in mouse embryonal carcinoma cells.
Mol Cell Biol 8, 5309–5317.
40 Nakai A, Kawazoe Y, Tanabe M, Nagata K &
Morimoto RI (1995) The DNA-binding properties of
two heat shock factors, HSF1 and HSF3 are induced
in the avian erythroblast cell line HD6. Mol Cell Biol
15, 5168–5178.
41 McMillan DR, Christians E, Forster M, Xiao X,
Connell P, Plumier JC, Zuo X, Richardson J, Morgan
S & Benjamin IJ (2002) Heat shock transcription factor
2 is not essential for embryonic development, fertility,
or adult cognitive and psychomotor function in mice.
Mol Cell Biol 22, 8005–8014.
42 Nakai A & Ishikawa T (2001) Cell cycle transition
under stress conditions controlled by vertebrate heat

shock factors. EMBO J 20, 2885–2895.
43 Green M, Schuetz T, Sullivan K & Kingston RE
(1995) A heat shock-responsive domain of human
HSF1 that regulates transcription activation domain
function. Mol Cell Biol, 15, 3354–3362.
44 Shi Y, Kroeger PE & Morimoto RI (1995) The car-
boxyl-terminal transactivation domain of heat shock
factor 1 is negatively regulated and stress responsive.
Mol Cell Biol 15, 4309–4318.
45 Zuo J, Rungger D & Voellmy R (1995) Multiple layers
of regulation of human heat shock transcription
factor 1. Mol Cell Biol 15, 4319–4330.
46 Sullivan EK, Weirich CS, Guyon JR, Sif S & Kingston
RE (2001) Transcriptional activation domains of
human heat shock factor 1 recruit human SWI ⁄ SNF.
Mol Cell Biol 21, 5826–5837.
47 Corey LL, Weirich CS, Benjamin IJ & Kingston RE
(2003) Localized recruitment of a chromatin-remodel-
ing activity by an activator in vivo drives transcrip-
tional elongation. Genes Dev 17, 1392–1401.
48 Sakurai H & Enoki Y (2010) Novel aspects of heat
shock factors: regulation of the mammalian HSF fam-
ily members. FEBS J 277, 4140–4149.
49 Morimoto RI (2008) Proteotoxic stress and inducible
chaperone networks in neurodegenerative disease and
aging. Genes Dev 22, 1427–1438.
50 Beere HM (2001) Stressed to death: regulation of
apoptotic signaling pathways by the heat shock pro-
teins. Sci STKE 2001, re1.
51 Zarzov P, Boucherie H & Mann C (1997) A yeast heat

shock transcription factor (Hsf1) mutant is defective in
both Hsc82 ⁄ Hsp82 synthesis and spindle pole body
duplication. J Cell Sci 110, 1879–1891.
52 Morano KA, Santoro N, Koch KA & Thiele DJ
(1999) A trans-activation domain in yeast heat shock
transcription factor is essential for cell cycle progres-
sion during stress. Mol Cell Biol 19, 402–411.
53 Metchat A, Akerfelt M, Bierkamp C, Delsinne V,
Sistonen L, Alexandre H & Christians ES (2009)
Mammalian heat shock factor 1 is essential for oocyte
meiosis and directly regulates Hsp90alpha expression.
J Biol Chem 284, 9521–9528.
54 Takaki E, Fujimoto M, Sugahara K, Nakahari T,
Yonemura S, Tanaka Y, Hayashida N, Inouye S,
Takemoto T, Yamashita H et al. (2006) Maintenance
of olfactory neurogenesis requires HSF1, a major heat
shock transcription factor in mice. J Biol Chem 281,
4931–4937.
55 Yan LJ, Christians ES, Liu L, Xiao X, Sohal RS &
Benjamin IJ (2002) Mouse heat shock transcription
factor 1 deficiency alters cardiac redox homeostasis
and increases mitochondrial oxidative damage. EMBO
J 21, 5164–5172.
Evolution and function of the HSF family M. Fujimoto and A. Nakai
4122 FEBS Journal 277 (2010) 4112–4125 ª 2010 The Authors Journal compilation ª 2010 FEBS
56 Takaki E, Fujimoto M, Nakahari T, Yonemura S,
Miyata Y, Hayashida N, Yamamoto K, Vallee RB,
Mikuriya T, Sugahara K et al. (2007) Heat shock tran-
scription factor 1 is required for maintenance of ciliary
beating in mice. J Biol Chem 282, 37285–37292.

57 Hahn JS, Hu Z, Thiele DJ & Iyer VR (2004) Genome-
wide analysis of the biology of stress responses through
heat shock transcription factor. Mol Cell Biol 24,
5249–5256.
58 Brich-Machin I, Gao S, Huen D, McGirr R, White
RA & Russell S (2005) Genomic analysis of heat-shock
factor targets in Drosophila. Genome Biol 6, R63.
59 Trinklein ND, Murray JI, Hartman SJ, Botstein D &
Myers RM (2004) The role of heat shock transcription
Factor 1 in the genome-wide regulation of the mam-
malian heat shock response. Mol Biol Cell 15, 1254–
1261.
60 Page TJ, Sikder D, Yang L, Pluta L, Wolfinger RD,
Kodadek T & Thomas RS (2006) Genome-wide analy-
sis of human HSF1 signaling reveals a transcriptional
program linked to cellular adaptation and survival.
Mol Biosyst 2, 627–639.
61 Deguchi M, Iizuka T, Hata Y, Nishimura W, Hirao K,
Yao I, Kawabe H & Takai Y (2000) PAPIN. A novel
multiple PSD-95 ⁄ Dlg-A ⁄ ZO-1 protein interacting with
neural plakophilin-related armadillo repeat pro-
tein ⁄ delta-catenin and p0071. J Biol Chem 275, 29875–
29880.
62 Fargeas CA, Florek M, Huttner WB & Corbeil D
(2003) Characterization of prominin-2, a new member
of the prominin family of pentaspan membrane glyco-
proteins. J Biol Chem 278, 8586–8596.
63 Fujimoto M, Oshima I, Shinkawa T, Wang BB, Inouye
S, Hayashida N, Takii R & Nakai A (2008) Analysis
of HSF4 binding regions reveals its necessity for gene

regulation during development and heat shock
response in mouse lenses. J Biol Chem 283, 29961–
29970.
64 Jolly C, Metz A, Govin J, Vigneron M, Turner BM,
Khochbin S & Vourc’h C (2004) Stress-induced tran-
scription of satellite III repeats. J Cell Biol 164, 25–33.
65 Rizzi N, Denegri M, Chiodi I, Corioni M, Valgards-
dottir R, Cobianchi F, Riva S & Biamonti G (2004)
Transcriptional activation of a constitutive heterochro-
matic domain of the human genome in response to
heat shock. Mol Cell Biol 15, 543–551.
66 Sandqvist A, Bjo
¨
rk JK, Akerfelt M, Chitikova Z,
Grichine A, Vourc’h C, Jolly C, Salminen TA,
Nymalm Y & Sistonen L (2009) Heterotrimerization of
heat-shock factors 1 and 2 provides a transcriptional
switch in response to distinct stimuli. Mol Biol Cell 20,
1340–1347.
67 Giardina C & Lis JT (1995) Dynamic protein-DNA
architecture of a yeast heat shock promoter. Mol Cell
Biol 15, 2737–2744.
68 Liu XD, Liu PC, Santoro N & Thiele DJ (1997) Con-
servation of a stress response: human heat shock tran-
scription factors functionally substitute for yeast HSF.
EMBO J 16, 6466–6477.
69 Jedlicka P, Mortin MA & Wu C (1997) Multiple func-
tions of Drosophila heat shock transcription factor in
vivo. EMBO J 16, 2452–2462.
70 Xiao X, Zuo X, Davis AA, McMillan DR, Curry BB,

Richardson JA & Benjamin IJ (1999) HSF1 is required
for extra-embryonic development, postnatal growth
and protection during inflammatory responses in mice.
EMBO J 18, 5943–5952.
71 Kallio M, Chang Y, Manuel M, Alastalo TP, Rallu M,
Gitton Y, Pirkkala L, Loones MT, Paslaru L, Larney
S et al. (2002) Brain abnormalities, defective meiotic
chromosome synapsis and female subfertility in HSF2
null mice.
EMBO J 21, 2591–2601.
72 Fujimoto M, Izu H, Seki K, Fukuda K, Nishida T,
Yamada S, Kato K, Yonemura S, Inouye S & Nakai
A (2004) HSF4 is required for normal cell growth and
differentiation during mouse lens development. EMBO
J 23, 4297–4306.
73 Min JN, Zhang Y, Moskophidis D & Mivechi NF
(2004) Unique contribution of heat shock transcription
factor 4 in ocular lens development and fiber cell dif-
ferentiation. Genesis 40, 205–217.
74 Christians E, Davis AA, Thomas SD & Benjamin IJ
(2000) Maternal effect of Hsf1 on reproductive success.
Nature 407, 693–694.
75 Nakai A, Suzuki M & Tanabe M (2000) Arrest of
spermatogenesis in mice expressing an active heat
shock transcription factor 1. EMBO J 19, 1545–1554.
76 Izu H, Inouye S, Fujimoto M, Shiraishi K, Naito K &
Nakai A (2004) Heat-shock transcription factor 1 is
involved in quality control mechanisms in male germ
cells. Biol Reprod 70, 18–24.
77 Wang G, Ying Z, Jin X, Tu N, Zhang Y, Phillips M,

Moskophidis D & Mivechi NF (2004) Essential
requirement for both hsf1 and hsf2 transcriptional
activity in spermatogenesis and male fertility. Genesis
38, 66–80.
78 Chang Y, Ostling P, Akerfelt M, Trouillet D, Rallu M,
Gitton Y, Fatimy EI, Fardeau V, Le Crom S, Mor-
ange M et al. (2006) Role of heat-shock factor 2 in
cerebral cortex formation and as a regulator of p35
expression. Genes Dev 20, 836–847.
79 Whitesell L & Lindquist SL (2005) HSP90 and the
chaperoning of cancer. Nat Rev Cancer 5, 761–772.
80 Calderwood SK, Khaleque MA, Sawyer DB &
Ciocca DR (2006) Heat shock proteins in cancer:
chaperones of tumorigenesis. Trends Biochem Sci 31,
164–172.
81 Hoang AT, Huang J, Rudra-Ganguly N, Zheng J,
Powell WC, Rabindran SK, Wu C & Roy-Burman P
(2000) A novel association between the human heat
M. Fujimoto and A. Nakai Evolution and function of the HSF family
FEBS Journal 277 (2010) 4112–4125 ª 2010 The Authors Journal compilation ª 2010 FEBS 4123
shock transcription factor 1 (HSF1) and prostate ade-
nocarcinoma. Am J Pathol 156, 857–864.
82 Tang D, Khaleque MA, Jones EL, Theriault JR, Li C,
Wong WH, Stevenson MA & Calderwood SK (2005)
Expression of heat shock proteins and heat shock pro-
tein messenger ribonucleic acid in human prostate car-
cinoma in vitro and in tumors in vivo. Cell Stress
Chaperones 10, 46–58.
83 Rossi A, Ciafre
`

S, Balsamo M, Pierimarchi P &
Santoro MG (2006) Targeting the heat shock factor 1
by RNA interference: a potent tool to enhance hyper-
thermochemotherapy efficacy in cervical cancer. Cancer
Res 66, 7678–7685.
84 Min JN, Huang L, Zimonjic DB, Moskophidis D &
Mivechi NF (2007) Selective suppression of lymphomas
by functional loss of Hsf1 in a p53-deficient mouse
model for spontaneous tumors. Oncogene 26, 5086–
5097.
85 Dai C, Whitesell L, Rogers AB & Lindquist S (2007)
Heat shock factor 1 is a powerful multifaceted modifier
of carcinogenesis. Cell 130, 1005–1018.
86 Inouye S, Izu H, Takaki E, Suzuki H, Shirai M,
Yokota Y, Ichikawa H, Fujimoto M & Nakai A
(2004) Impaired IgG production in mice deficient for
heat shock transcription factor 1. J Biol Chem 279,
38701–38709.
87 Inouye S, Fujimoto M, Nakamura T, Takaki E,
Hayashida N, Hai T & Nakai A (2007) Heat shock
transcription factor 1 opens chromatin structure of
interleukin-6 promoter to facilitate binding of an
activator or a repressor. J Biol Chem 282, 33210–
33217.
88 Balch WE, Morimoto RI, Dillin A & Kelly JW (2008)
Adapting proteostasis for disease intervention. Science
319, 916–919.
89 Hsu AL, Murphy CT & Kenyon C (2003) Regulation
of aging and age-related disease by DAF-16 and heat-
shock factor. Science 300, 1142–1145.

90 Morley JF & Morimoto RI (2004) Regulation of lon-
gevity in Caenorhabditis elegans by heat shock factor
and molecular chaperones. Mol Biol Cell 15, 657–664.
91 Cohen E, Bieschke J, Perciavalle RM, Kelly JW &
Dillin A (2006) Opposing activities protect against
age-onset proteotoxicity. Science 313, 1604–1610.
92 Mangiarini L, Sathasivam K, Seller M, Cozens B,
Harper A, Hetherington C, Lawton M, Trottier Y,
Lehrach H, Davies SW et al. (1996) Exon 1 of the HD
gene with an expanded CAG repeat is sufficient to
cause a progressive neurological phenotype in trans-
genic mice. Cell 87, 493–506.
93 Sathasivam K, Hobbs C, Turmaine M, Mangiarini L,
Mahal A, Bertaux F, Wanker EE, Doherty P, Davies
SW & Bates GP (1999) Formation of polyglutamine
inclusions in non-CNS tissue. Hum Mol Genet 8, 813–
822.
94 Hansson O, Nylandsted J, Castilho RF, Leist M,
Ja
¨
a
¨
ttela
¨
M & Brundin P (2003) Overexpression of heat
shock protein 70 in R6 ⁄ 2 Huntington’s disease mice
has only modest effects on disease progression. Brain
Res 970, 47–57.
95 Hay DG, Sathasivam K, Tobaben S, Stahl B,
Marber M, Mestril R, Mahal A, Smith DL,

Woodman B & Bates GP (2004) Progressive decrease
in chaperone protein levels in a mouse model of
Huntington’s disease and induction of stress proteins
as a therapeutic approach. Hum Mol Genet 13, 1389–
1405.
96 Fujimoto M, Takaki E, Hayashi T, Kitaura Y,
Tanaka Y, Inouye S & Nakai A (2005) Active HSF1
significantly suppresses polyglutamine aggregate forma-
tion in cellular and mouse models. J Biol Chem 280,
34908–34916.
97 Steele AD, Hutter G, Jackson WS, Heppner FL,
Borkowski AW, King OD, Raymond GJ, Aguzzi A &
Lindquist S (2008) Heat shock factor 1 regulates life-
span as distinct from disease onset in prion disease.
Proc Natl Acad Sci USA 105, 13626–13631.
98 Homma S, Jin X, Wang G, Tu N, Min J, Yanasak N
& Mivechi NF (2007) Demyelination, astrogliosis, and
accumulation of ubiquitinated proteins, hallmarks of
CNS disease in hsf1-deficient mice. J Neurosci 27,
7974–7986.
99 Powers ET, Morimoto RI, Dillin A, Kelly JW &
Balch WE (2009) Biological and chemical approaches
to diseases of proteostasis deficiency. Annu Rev
Biochem 78, 959–991.
100 Kieran D, Kalmar B, Dick JR, Riddoch-Contreras J,
Burnstock G & Greensmith L (2004) Treatment with
arimoclomol, a coinducer of heat shock proteins,
delays disease progression in ALS mice. Nat Med 10,
402–405.
101 Fujikake N, Nagai Y, Popiel HA, Okamoto Y,

Yamaguchi M & Toda T (2008) Heat shock transcrip-
tion factor 1-activating compounds suppress polygluta-
mine-induced neurodegeneration through induction of
multiple molecular chaperones. J Biol Chem 283,
26188–26197.
102 Waza M, Adachi H, Katsuno M, Minamiyama M,
Sang C, Tanaka F, Inukai A, Doyu M & Sobue G
(2005) 17-AAG, an Hsp90 inhibitor, ameliorates
polyglutamine-mediated motor neuron degeneration.
Nat Med 11, 1088–1095.
103 Katsuno M, Sang C, Adachi H, Minamiyama M,
Waza M, Tanaka F, Doyu M & Sobue G (2005)
Pharmacological induction of heat-shock proteins
alleviates polyglutamine-mediated motor neuron
disease. Proc Natl Acad Sci USA 102, 16801–16806.
104 Westerheide SD, Bosman JD, Mbadugha BN,
Kawahara TL, Matsumoto G, Kim S, Gu W,
Devlin JP, Silverman RB & Morimoto RI (2004)
Evolution and function of the HSF family M. Fujimoto and A. Nakai
4124 FEBS Journal 277 (2010) 4112–4125 ª 2010 The Authors Journal compilation ª 2010 FEBS
Celastrols as inducers of the heat shock response and
cytoprotection. J Biol Chem 279, 56053–56060.
105 Neef DW, Turski ML & Thiele DJ (2010) Modulation
of heat shock transcription factor 1 as a therapeutic
target for small molecule intervention in neurodegener-
ative disease. PLoS Biol 8, e1000291.
106 Shinka T, Sato Y, Chen G, Naroda T, Kinoshita K,
Unemi Y, Tsuji K, Toida K, Iwamoto T &
Nakahori Y (2004) Molecular characterization of heat
shock-like factor encoded on the human Y

chromosome, and implications for male infertility.
Biol Reprod 71 , 297–306.
107 Tessari A, Salata E, Ferlin A, Bartoloni L, Slongo ML
& Foresta C (2004) Characterization of HSFY, a novel
AZFb gene on the Y chromosome with a possible role
in human spermatogenesis. Mol Hum Reprod 10, 253–
258.
108 Thompson JD, Higgins DG & Gibson TJ (1994)
CLUSTAL W: improving the sensitivity of progressive
multiple sequence alignment through sequence
weighting, position-specific gap penalties and weight
matrix choice. Nucleic Acids Res 22, 4673–4680.
109 Page RD (1996) TreeView: an application to display
phylogenetic trees on personal computers. Comput Appl
Biosci 12, 357–358.
110 Sidow A (1996) Gen(om)e duplications in the evolu-
tion of early vertebrates. Curr Opin Genet Dev 6,
715–722.
111 Meyer A & Schartl M (1999) Gene and genome dupli-
cations in vertebrates: the one-to-four (-to-eight in fish)
rule and the evolution of novel gene functions. Curr
Opin Cell Biol 11, 699–704.
112 Se
´
mon M & Wolfe KH (2007) Consequences of
genome duplication. Curr Opin Genet Dev 17, 505–512.
113 Ostling P, Bjo
¨
rk JK, Roos-Mattjus P, Mezger V &
Sistonen L (2007) Heat shock factor 2 (HSF2) con-

tributes to inducible expression of hsp genes through
interplay with HSF1. J Biol Chem 282, 7077–7086.
M. Fujimoto and A. Nakai Evolution and function of the HSF family
FEBS Journal 277 (2010) 4112–4125 ª 2010 The Authors Journal compilation ª 2010 FEBS 4125

×