ChapterPDF Available

Abstract and Figures

During heat shock conditions a plethora of proteins are found to play a role in maintaining cellular homeostasis. They play diverse roles from folding of non-native proteins to the proteasomal degradation of harmful aggregates. A few out of these heat shock proteins (Hsp) help in the folding of non-native substrate proteins and are termed as molecular chaperones. Various structural and functional adaptations make them work efficiently under both normal and stress conditions. These adaptations involve transitions to oligomeric structures, thermal stability, efficient binding affinity for substrates and co-chaperones, elevated synthesis during shock conditions, switching between ‘holding’ and ‘folding’ functions etc. Their ability to function under various kinds of stress conditions like heat shock, cancers, neurodegenerative diseases, and in burdened cells due to recombinant protein production makes them therapeutically and industrially important biomolecules.
Content may be subject to copyright.
181© Springer International Publishing AG 2018
A. A. A. Asea, P. Kaur (eds.), Regulation of Heat Shock Protein Responses, Heat
Shock Proteins 14, https://doi.org/10.1007/978-3-319-74715-6_8
Chapter 8
Molecular Chaperones: Structure-Function
Relationship andtheir Role inProtein Folding
BhaskarK.Chatterjee, SaritaPuri, AshimaSharma, AshutoshPastor,
andTapanK.Chaudhuri
Abstract During heat shock conditions a plethora of proteins are found to play a
role in maintaining cellular homeostasis. They play diverse roles from folding of
non-native proteins to the proteasomal degradation of harmful aggregates. A few
out of these heat shock proteins (Hsp) help in the folding of non-native substrate
proteins and are termed as molecular chaperones. Various structural and functional
adaptations make them work efciently under both normal and stress conditions.
These adaptations involve transitions to oligomeric structures, thermal stability,
efcient binding afnity for substrates and co-chaperones, elevated synthesis during
shock conditions, switching between ‘holding’ and ‘folding’ functions etc. Their
ability to function under various kinds of stress conditions like heat shock, cancers,
neurodegenerative diseases, and in burdened cells due to recombinant protein pro-
duction makes them therapeutically and industrially important biomolecules.
Keywords Chaperone assisted folding · Heat shock · Molecular chaperones ·
Protein folding · Structure-function of chaperones
Abbreviations
ACD α-crystallin domain
ADP Adenosine di-phosphate
ATP Adenosine tri-phosphate
CCT Chaperonin containing TCP-1
CIRCE Controlling inverted repeat of chaperone expression
B. K. Chatterjee · S. Puri · A. Sharma · A. Pastor · T. K. Chaudhuri (*)
Kusuma School of Biological Sciences, Indian Institute of Technology Delhi,
HauzKhas, New Delhi, India
e-mail: tkchaudhuri@bioschool.iitd.ac.in
Bhaskar K. Chatterjee, Sarita Puri, Ashima Sharma, and Ashutosh Pastor authors are equally
contributed.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
182
CNX Calnexin
CRT Calreticulin
CS Citrate synthase
ER Endoplasmic reticulum
ERAD Endoplasmic reticulum associated degradation
FRET Fluorescence energy resonance transfer
HOP HSP90/HSP70 organizing protein
HSC70 Heat shock cognate
HSEs Heat shock elements
HSFs Heat shock response and specic transcription factors
Hsp Heat shock proteins
HSP Heat shock protein family
HSR Heat shock response
MalZ Maltodextrin glucosidase
NAC Nascent chain associated complex
NEF Nucleotide-exchange factors
NTD n-terminal domain
PBD Peptide binding domain
PPIase Peptidyl-prolyl isomerases
PTP Permeability transition pore complex
RAC Ribosome associated complex
RuBisCO Ribulose-1,5-bisphosphate oxygenase-carboxylase
SHR Steroid hormone receptors
sHsp Small heat shock proteins
sHSP Small heat shock protein family
TF Trigger factor
TPR Tetratricopeptide
TRiC TCP-1 ring complex
UPR Unfolded protein response pathway
8.1 Introduction
Living systems respond to threatening conditions at multiple levels in their quest for
survival. It may in the form of a ght or ight response, which is a result of any
imminent physical threat either to an organism or their inner homeostasis. For
example the temperature, ionic and sugar balance are regulated within a xed range
in our bodies and are probably optimized by evolutionary mechanisms. Similarly,
homeostasis is alsomaintained at the cellular level and maintaining such a balance
is imperative for the survival and efcient functioning of the cell. One of the major
homeostasis mechanisms operating at the cellular level is the protein homeostasis,
commonly referred to as proteostasis (Balch etal. 2008). Starting with maintaining
the structural organization of a cell to catalysing various metabolic reactions; from
the transport of macromolecules within and across cells to various recognition and
B. K. Chatterjee et al.
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
183
immune functions, proteins play vital roles in our bodies and are regarded as the
actual workhorses of the cells. Proteins undergo various post-translational modica-
tions and move through trafcking pathways before they are ready to take up their
function. However, acquiring their specic three-dimensional structure supersedes
all this because only in their specic structural forms can they undertake the func-
tion they are meant to carry out. In this chapter, we shalldiscuss the cellular machin-
ery responsible for maintaining proteins in their functional state during stress
conditions; specically focusing on the Hsp that assist in the folding and refolding
of misfolded and aggregation prone proteins.
8.2 What Is Stress Response?
The stress response can be dened as our involuntary defense reaction to threaten-
ing conditions. This may occur at multiple levels as a response to a different range
of conditions. At a cellular level, the response to any alarming condition, like a
chronic change in the environment away from normalconditions (which may inter-
fere with the physiological functioning of the cells and cause damage to nucleic
acids and proteins) can be identied as a stress response (Kültz 2004). While at the
organism level our adaptive responses are driven by hormonal changes (Charmandari
etal. 2005), the cellular response to damages occurring ata molecular level involves
a cascade of pathways, and molecules that work in cohort to bring the cells back to
their normal functional state. Various kinds of stress at the cellular level include
oxidative, heat, radiation and nutrient deprivation. The major consequences of these
stress are DNA damage, loss of cellular signalling, protein unfolding, misfolding,
aggregation, proteolysis, cellular necrosis and apoptosis. The cellular stress response
may be ofa protective nature where the cell can defend and restore its normal func-
tioning, or of a destructive nature where the conditions are beyond the cell’s ability
to repair. The type, level, and duration of stressful conditions may ultimately deter-
mine the fate a cell. During stress conditions, proteins are misfolded due to changes
in the overall energy landscape. This causes loss of proteinfunction and the accu-
mulation of misfolded proteins in the form of toxic aggregates. The protective stress
response for proteins includes pathways of the heat shock response and the unfolded
protein response (UPR); the destructive response pathways include apoptosis,
necrosis or autophagy (Fulda etal. 2010). The DNA damage response consists of
multiple, complex pathways that restore genomic integrity. Theseinclude the base
excision repair, nucleotide excision repair, and non-homologous end joining
(Kourtis and Tavernarakis 2011). The oxidative stress response helps the cell cope
with the reactive oxygen species, maintain redox homeostasis; and a number of
enzymes like superoxide dismutase and non-enzymatic antioxidants are involved
(Trachootham etal. 2008). In this chapter, we shall mainly focus on the diverse
mechanisms governing heat shock response and the various factors that are involved
in mediating such a response.
8 Molecular Chaperones in Cellular Stress Response
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
184
8.3 Heat Shock Response
The heat shock response was one of the earliest explored stress response mecha-
nisms that wasinitially observed in Drosophila as changes in pufng patterns of
salivary gland chromosomes (Ritossa 1996), followed by changes in gene expres-
sion patterns after heat treatment (Tissiéres etal. 1974; Hightower 1991). The heat
shock response imparts thermo-tolerance to the cells and protects them when they
are stressed due to prolonged exposure to heat. This response is activated by the rise
of a few degreesin temperature from the normal dwelling temperature of the organ-
ism. The regular transcription and translation processes in the cells are halted during
heat shock response, and specic transcription factors (HSFs) which selectively
enhance expression of a set of proteins having protective functions are activated.
However, even during normal conditions these HSFs play important roles in differ-
entiation and development of the organisms (Morimoto etal. 1996). The HSF1
predominantly regulates the heat shock responses, and is itself regulated by its
interactions with heat shock proteins HSP70 and HSP90 (Pirkkala etal. 2001). The
ability to activate transcription and bind toDNA are uncoupled in HSF1 imparting
a higher degree of regulation. The HSF1 exists as a monomer or in complex with
Hspunder normal conditions. During heat shock, HSF1 homotrimerizes and under-
goes hyperphosphorylation which leads to its activation (Cotto etal. 1996). These
HSFsfacilitate the overexpression of Hsp by binding to cis acting sequences on the
genome known as heat shock elements (HSEs). HSP are commonly known as
molecular chaperones for their role in assisting proteins to acquire their native struc-
tures. Hsp prevent heat induceddenaturation and aggregation of proteins, facilitate
the proteins to fold, and assist in the refolding of already denatured proteins
(Lindquist 1986). The Hsp play an active role in facilitating the degradation of
proteins that are unable to fold in order to maintainprotein homeostasis and thus
promote cell survival.
8.4 Cellular Components Providing HS Response
Heat shock response in cells is mediated by concerted actions of heat shock factors
(HSF), heat shock elements (HSE) and Hsp. The heat shock factors as described
above are transcription factors activated during a heat shock. Gram negative bacte-
ria E.coli has aspecic sigma factor 32 (σ32), coded by the rpoH gene, which is a
heat shock promoter specic subunit of RNA polymerase. The σ32 is a positive
regulator and issuppressed by DnaJ during normal conditions (Bukau 1993). The
gram-positive bacteria B. subtilis has a negative regulator HrcA which binds to neg-
atively acting CIRCE elements (controlling inverted repeat of chaperone expression).
Folding of HrcA is mediated by GroE chaperones. During heat shock response,
HrcA doesn’t fold due to the unavailability of free GroE chaperones and hence
the negative regulation of Hsp is switched off (Hietakangas and Sistonen 2006).
B. K. Chatterjee et al.
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
185
A majority of the eukaryotes have eitherone or multiple of the four heat shock fac-
tors HSF1-HSF4; the condition being different in plants. Arabidopsis has 21 genes
encoding various heat shock factors, divided among three classes and 14 different
groups. The plant HSFs are induced and expressed during heat shock (Hietakangas
and Sistonen 2006). Unlike plants, the transcription factors in animals are like
HSF1. They areconstitutively expressedbut functionally repressedduring normal
conditions by HSP70 and HSP90 (Shi etal. 1998). The DNA binding domain of
HSF recognizes the HSE in a major grove of theDNA double helix. HSEs are highly
conserved and contain inverted repeats of nGAAn (e.g. nTTCnnGAAnnTTCn)
(Amin etal. 1988; Akerfelt etal. 2010). There might be multiple HSEs in the pro-
moter region of a heat shock gene and hence multiple HSFs can bind simultane-
ously in a cooperative manner.
Expression of Hsp is invariably upregulated during heat shock irrespective of the
difference in theirmechanism of action,asdiscussed above. Overall, Hsp or molec-
ular chaperones, are one of the seven different classes of proteins to be overex-
pressed during stress response and were the earliest discovered components of the
heat shock response. The second class comprises components of the proteolytic
system which helps inthe clearance of misfolded and aggregated proteins. The pro-
teolytic machinery of the proteasome has a similar structural organization amongst
different organisms differing onlyin their subunit compositions. The protein degra-
dation machinery requires the translocation of misfolded or partially unstruc-
turedintermediatesbetween cytosol, ER and Golgi in eukaryotes and multiple HSP
members coordinating protein folding and degradation in different cellular com-
partments. The third class of proteins help in DNA and RNA repair, like counteract-
ing the heat induced methylation of RNA.The fourth category comprises enzymes
of the metabolic pathways which reorganize the energy supply of the cell. The fth
category includes kinases and transcription factors that further initiate or inhibit
expression cascades to support the stress response. The sixth class of proteins main-
tain the integrity of thecytoskeleton. Transport proteins and membrane-modulating
proteins makeup the seventh class. All these proteins function together to respond
to the heat stress conditions (Richter etal. 2010). However, we will be discussing
the molecular chaperones in detail in the following sections.
The Hsp are divided into multiple families based primarilyon their size. These
different classes of molecular chaperones and their localization in different cellular
compartments provide a great degree of organizational control and distribution of
roles to execute particular functions within the overall cascade of stressresponse.
HSP60, HSP70, HSP90, HSP100, and small HSP are found across all organisms
and are known to have high similarities within their classes among different organ-
isms. One similarity among all chaperones is that they recognize the hydrophobic
surfaces of proteins which areincreasingly exposed among misfolded and unfolded
proteins, and functionto eitherfold the protein (foldases) or mask the hydrophobic
regions to prevent aggregation (holdases). Despite the name suggesting a partic-
ular function, Hsp play signicant roles in multiple stress response pathways in all
organisms, including oxidative stress in all organisms and drought and osmotic
stress in plants. The roles of molecular chaperones have been observedin the correct
8 Molecular Chaperones in Cellular Stress Response
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
186
folding of a newly translated protein, refolding of misfolded proteins, disaggrega-
tion, translocation, and in the degradation of proteins. HSP60, HSP70, and HSP90
are ATP dependent chaperones which actively support protein folding. The small
HSP and other chaperones simply prevent the misfolding of proteins. The heat
shock factors are regulated by molecular chaperones like HSP70 and HSP90,
thusproving their importance in the overall heat shock response of the cell. HSP70
helps in the translocation of proteins to cellular compartments like ER and alsofacil-
itates retrotranslocation (The reverse movement of the protein from ER to the cyto-
sol, where it can be degraded by the proteasomal machinery) during ER associated
degradation. HSP90 clients include many kinases and steroid receptors, which help
in regulating a multitude of functions through signaling pathways. HSP104 is
known to disaggregate proteins and redirect them to correct refolding pathways.
The widespread presence of chaperones in almost all cellular components where
protein folding occurs, proves that they arekey playersin the heat shock response
machinery. Almost all cellular proteins might have to interact with molecular chap-
erones at least once in their lifetimes, be it during synthesis, folding, targeting, or
degradation.
8.5 Role ofChaperones inMediating Cellular Heat Shock
Response
Cells grow optimally within a narrow range of temperature, pH, and other physio-
logical conditions, but adapt to moderate deviation from such conditions. One of the
most well studied cellular adaptations is the heat shock response (HSR) (Guisbert
etal. 2004). During heat shock conditions, many cellular proteins work to either
rescue the cells from dying, or trigger apoptosis when the damage incurred is irre-
versible. These proteins are referred to as heat shock proteins (Hsp) (Herman and
Gross 2000). Few out of several such Hsp protect proteins from undergoing aggre-
gation, unfold aggregated proteins to make them folding compatible and refold
damaged proteins. These proteins are termed as chaperones (Morimoto etal. 1994).
Molecular chaperone is a major class of protein found at all levels of cellular
organizations ranging from bacteria to humans. They have variable organization
and function depending on the cellular location and complexity of the organism.
Bacterial chaperone proteins are found only in the cytosol as they are not
compartmentalized, but in case of higher organisms, these are also localized in
mitochondria, endoplasmic reticulum, and chloroplasts. Structural and functional
organizations of chaperones are evolutionally conserved within the same kingdom
but vary between them. On the basis of gross molecular weight, the major chaper-
one are classied as: HSP60 (GroEL/GroES, Cpn60/Cpn10, HSP60/HSP10, Tric/
CCT, Thermosome), HSP70 (DnaK, DnaJ, GrpE, HSP70), HSP90 (HSP90, TRAP,
HtpG), HSP100 (ClpA, ClpB, ClpP), sHSP and Trigger Factor (Georgopoulous
et al. 1994; Gross 1996). The other important chaperone proteins playing a role
B. K. Chatterjee et al.
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
187
in heat shock are Prefoldin, Calnexin / Calreticulin, GRP94, GRP170, AAA
ATPasesPPIases, PDIases, NAC (Nascent polypeptide Associated Complex), CasA
and HtpX (Rani etal. 2016).
8.5.1 Small Heat Shock Proteins (sHsp)
Most organisms have a well-developed sHSP system, which help in their protection
from the thermal, osmotic and salt stresses (Jakob etal. 1993). sHsp have subunit
molecular masses of 12–43kDa. The common feature of all sHSP is the presence of
a highly conserved stretch of 80–100 amino acids in their C-terminus termed as the
α-crystallin domain” (ACD). It is anked on both side by less conserved N-terminal
domain (NTD) and a C-terminal extension (Kappé etal. 2003; Franck etal. 2004;
Kriehuber etal. 2010). In E.coli major sHsp are IbpA and IbpB.Under normal cel-
lular conditions they help in aggregation prevention and folding of substrate pro-
teins in an ATP independent manner. During stress conditions, along with
anincreased expression, these proteins undergo drastic conformational rearrange-
ments in order to bind to the misfolded proteins and prevent cellular aggregation
(Mani etal. 2016). HSP31 of E.coli functions as a holding chaperone. It cooperates
with the DnaK-DnaJ-GrpE system in managing protein misfolding during stress
conditions (Mujacic et al. 2004). In the pathogenic Halobacterium sp., sHSP1,
HSP-5 and sHSP2 impart protection from thermal stress, solar radiation and high
saltconcentration (Vanghele and Ganea 2010). HSP20 of Mycobacterium tubercu-
losis protects them from macrophage induced stress response and helps in solubili-
zation of heat induced aggregates (Vanghele and Ganea 2010).
In yeast, sHsp HSP26 and HSP42 together, add an additional layer of protection
against a cellular assault like heat shock. Both HSP26 and HSP42 are poorly
expressed during exponential growth, but their expression increases 10-fold under
heat stress suggesting the dominant role they play in a thermally stressed cell
(Haslbeck etal. 2004a). HSP26 exists as a 24-mer under normal conditions, acts
like a holdase for damaged or misfolded proteins and transfers client proteins to the
HSP70 chaperone machinery during heat shock (Haslbeck etal. 2004a, b; 1999).
During heat shock, the 24-mer of HSP26 gets reversibly dissociated into dimers.
This dimeric form then interacts with the unfolded polypeptides and eventually
forms a larger complex, to be presented to chaperones capable of folding the sub-
strate (Stromer etal. 2004). HSP26 also interacts with aggregated proteins, making
them accessible to the HSP104 chaperone (Glover and Lindquist 1998). HSP42
oligomer is a symmetric assembly of dimers organized into two hexameric rings.
HSP42 binds with 30% of total yeast cytosolic proteins (Haslbeck etal. 2004a). It
is a more effective chaperone than HSP26, as ahigher HSP26 to substrate ratio is
needed to prevent aggregation (Haslbeck etal. 2004a). HSP12 exhibits low sequence
homology to the sHSP superfamily and is structurally and functionally distinct, as
it exists exclusively as a monomer (Welker etal. 2010). Liketheother sHsp, HSP12
8 Molecular Chaperones in Cellular Stress Response
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
188
is weakly expressed in exponentially growing cells but overexpressed (100-fold)
during heat shock (Welker etal. 2010).
In plants, there are distinct gene families for sHSP found in different organelles
and a total of 6 families have been classied. The HSP17.6 and HSP17.9 reside in
the cytoplasm, HSP21in the chloroplast, HSP22in the ER, HSP23in the mitochon-
dria and HSP22.3in the cell membrane (Wang etal. 2004). They have been sug-
gested to be involvedeither in maintaining the structure of a heat stressed cell
(Lindquist and Craig 1988) or to protect the photosynthesis machinery during heat
shock (Schuster etal. 1988; Chen et al. 1990). Genetically modied plants with
higher thermo-tolerance have been designed by constitutive upregulation of
sHSP.For example, the HSF3 gene of Arabidopsis thaliana was modied to express
it in non-heat shock conditions andwas shown toincrease the basal thermotolerance
of the plant (Prändl et al. 1998). Overexpression of HSP17.6 results in a higher
tolerance to drought and salinity in Arabidopsis thaliana, while the natural elevated
gene expression is observed in case of heat shock (Sun etal. 2001).
In humans, Group I sHsp consists of HSP27, HSP20, HSP22, and αB-crystallin.
They are found in various tissues and are heat inducible. Group II sHsp consists of
HSPB9, HSPB10, HSPB3, HSPB2 and α-crystallin, and they are involved in cell
differentiation and are restricted to certain tissues (Taylor and Benjamin 2005).
HSP27 forms oligomeric species when subjected to higher temperatures and this
results in increased chaperone activity (Bakthisaran etal. 2015). They predomi-
nantly function as ‘holdases’,keeping the substrates in a folding competent globular
state that can later be presented to ‘foldases’ like HSP60/10 or HSP70 to make them
functional (Eyles and Gierasch 2010). Under conditions of heat stress, they also
prevent aggregation by binding to late unfolding intermediates and keep them in a
stable, soluble complex. sHSP may also be involved in the transient/reversible reac-
tivation of early unfolded intermediates and this process may be ATP dependent,
although no ATP hydrolysis has been observed (Rajaraman etal. 2001). ATP bind-
ing is thought to trigger a conformational change that aides Hsp to release their
refolding-competent substrates (Muchowski etal. 1999).
8.5.2 The Chaperonin System (HSP60)
Chaperonins are double ring complexes of 800–900kDa which help in the folding
of many cellular proteins under normal and stress conditions (Spiess etal. 2004;
Hemmingsen etal. 1988; Vabulas etal. 2010). These are further classied as group
I and group II chaperonins (Horwich etal. 2007).
In bacteria and symbiotic organelles like mitochondria and chloroplasts, Group I
chaperonins (cpn60) are found. These chaperonins are termed as GroEL in E.coli,
mtHsp60in mitochondria and Rubisco binding protein in chloroplast (Figueiredo
etal. 2004). These require co-chaperonin GroESor HSP10 to functionin prokary-
otic and eukaryotic cells respectively. Under normal cellular conditions, GroEL/ES
is constitutively expressed in bacterial cytoplasm and helps in the folding of
B. K. Chatterjee et al.
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
189
substrate proteins in an ATP dependent fashion. Under environmental stress condi-
tions, the expression of these proteins increase by 15–20% (Georgopoulos etal.
1994; Gross 1996)and also leads to a few structural and functional modications of
the chaperone system, which enable them to fold or hold the aggregation prone
proteins. One such modication is the phosphorylation of the double ring which
mediates substrate folding in an ATP independent manner (Sherman and Goldberg
1994). GroEL also acts as a holding chamber for substrate proteins during thermal
stress and helps them regain their folding function once normality is restored
(Carrascosa etal. 1998). Mitochondrial Hsp60, with the help of its co-chaperonin
HSP10 helps in thefolding and assembly of imported proteins inside the matrix of
mitochondria. In HSP60 conditional mutants, aggregatesaccumulate andare unable
to assemble into functional complexes. The plastid HSP60 chaperonin found in
chloroplasts consists of two different subunits which make them different from
other members of the HSP60 family (Levy-Rimler etal. 2002). These two distinct
subunits known as CPN60α and CPN60β share about 50% sequence identity
(Boston etal. 1996). The tetradecameric structure consists of α7β7oligomers where
the assembly of α7 is dependent on the β7 homo-oligomer, thus forming two homo-
geneous rings, and the oligomers having a seven fold symmetry (Waldmann etal.
1995). The chloroplast co-chaperonin can bind to both GroEL and ch-CPN60 and
assists in thefolding of proteins in both cases. Its structure however, is markedly
different from GroES and contains two GroES like domains connected by a short
linker region. Some reports have conrmed the presence of 10kDa CPN10 and
20 kDa dimer like CPN20 existing simultaneously in the chloroplasts (Hill and
Hemmingsen 2001; Levy-Rimler etal. 2002).
Group II chaperonins are found in archaea (Thermosome) andthe cytoplasm of
eukaryotes (TCP/CCT) (Ditzel etal. 1998; Leitner etal. 2012). They do not require
the co-factorHSP10 as they have specialized α-helical extensions in their apical
domain that function as a built-in lid (Vabulaset al. 2010). Archaeal Group II chap-
eronins (Thermosome) consists of two stacked octameric rings with two different
kinds of subunits α & β (Klumpp etal. 1997). The mechanism of folding of non-
native protein substrates is similar to GroEL andinvolves binding of the substrate at
the apical domain, followed by ATP hydrolysis and release of thefolded substrate
fromthe cavity (Lopez etal. 2016). During stress conditions, Group II chaperonins
form a large octadecameric β complex which is more efcient insubstrate binding.
It also helps in membrane stabilization of archaeal cells during stress conditions
(Chaston etal. 2016). The group II chaperonins in the eukaryotic cytosol are known
as TRiC (TCP-1 ring complex) or CCT (chaperonin containing TCP-1) and like
their archaeal counterparts have eight or nine rings, each containing eight paralo-
gous subunits (Frydman 2001). The general domain structure of the group II chap-
eronins is akin to GroEL (Ditzel et al. 1998). The closing and opening of these
segments to encapsulate the substrate in the TriC/CCT cavity is ATP dependent
(Meyer etal.2003a). TriCs interact functionally with the co-chaperone prefoldin
(Vainberg etal. 1998; Siegert etal. 2000) and HSP70 (Langer etal. 1992), which
serve to transfer substrates to this chaperonin.
8 Molecular Chaperones in Cellular Stress Response
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
190
8.5.3 HSP70
HSP70 family is involved in a multitude of functions in all organisms and are found
in various cellular compartments. HSP70 facilitates translocation, protein import,
and signal transduction along with assisting in therefolding of substrate proteins
and preventing their aggregation (Frydman 2001; Miemyk 2017). HSP70 consists
of two domains, a highly conserved 44kDa N-terminal ATP binding domain, and a
15kDa C-terminal peptide binding domain (PBD) which consists of a β-sandwich
motif and an α-helical lid segment (Vabulas etal. 2010; Yu etal. 2015). In eukary-
otes, under normal conditions, HSP70 exist as complexes with either HSP90 or
co-chaperones like HSP40 and others, and is known as Heat Shock Cognate 70
(HSC70). The constitutively expressed HSC70 assistin the folding and transloca-
tion of newly synthesized proteins during normal conditions (Hartl etal. 2011).
Under heat stress conditions, the stress-inducible HSP70 isoforms are over-
expressed. This is achieved by itsinteraction with HSF, which transcriptionally
activates several heat shock genes including that of HSP70. HSP70s function is
generally conserved in all organisms (Akerfelt etal. 2010). Nascent polypeptides
form the bulk of its clients and undergo chaperoning via the ATP-dependent reac-
tion cycle of HSP70 which is regulated by HSP40 and nucleotide-exchange factors
(NEF) (Kampinga and Craig 2010; Mayer 2010). The β-sandwich motif of the PBD
recognizes an extended, seven-residue hydrophobic patch of an aggregation prone
protein, especiallywhen they are locally surrounded by positively charged residues
(Rudiger etal. 1997). The binding of the peptide is ATP-dependent and is regulated
by a conformational change in the β-sandwich motif (Mayer 2010). In the ATP-
bound state, the lid adopts an open conformation resulting in a high binding afnity
for the peptide. Hydrolysis of ATP facilitates lid closure and is accelerated by
HSP40, leading to the peptide getting locked inside. Following the hydrolysis of
ATP, NEF binds to the ATPase domain and catalyses ADP-ATP exchange that result
in the opening of the lid and release of the substrate, presumably in their non-native
conformation. These intermediates might then undergo multiple rounds of binding
and release till they acquire their native conformation (Hartl etal. 2011). Under
stress conditions (apart from folding nascent polypeptides) HSP70 prevents aggre-
gation of non-native substrates by transiently shielding exposed hydrophobic seg-
ments and keeping them in a folding competent state, which may subsequently be
transferred to the chaperonin cage for completion of the folding process (Vabulas
etal. 2010). BiP, the HSP70 paralog in the ER, binds to unfolded regions of the
protein manifesting exposed hydrophobic residues and has an ATP dependent
mechanism of substrate binding and release, similar to its cytosolic counterpart. BiP
plays an active role in the Unfolded Protein Response pathway (UPR) and in ER
Associated Degradation (ERAD), both of which occur when misfolded proteins
start accumulating in ER.It does so by interacting with several co-chaperones that
assist in protein folding and quality control. One of them is Erdj3, an HSP40
paralog in the ER, which interacts with partially folded intermediates and presents
them to BiP.Another cohort is BAP, a NEF that plays a similar role as its cytosolic
B. K. Chatterjee et al.
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
191
counterparts, promoting ADP release and facilitating BiPs transition to the open
state (Ma and Hendershot 2004).
In yeast, SSA and SSB are subclasses of cytosolic Hsp70 that are constitutively
expressed, while SSA3 and SSA4 are known to be induced upon exposure toheat
shock (Santoro etal. 1998). Interaction with the chaperone Ydj1 (Hsp40), a homo-
logue of the bacterial DnaJ protein, accelerates the ATPase activity of Hsp70. KAR2
(ER Hsp70) is a yeast homolog of BiP protein which gets induced within 10minutes
of heat shock (Normington etal. 1989). Kar2 is responsible for the translocation of
proteins across the ER membrane through co-translational as well as post-transla-
tional pathways (Brodsky etal. 1995). It is also involved in theretrograde transport
of defective non-functional substrates from the ER to thecytosol for proteosomal
degradation via ERAD, and thereby contributes to ER proteostasis (McCracken and
Brodsky 1996). The promoter region of the Kar2 gene contain unfolded protein
response elements (UPREs), and itgets induced when unfolded polypeptides start
accumulating in the ER lumen (Cox and Walter 1996). As is the case for human
cytosolic Hsp70, Kar 2 interacts with co-chaperonesHsp40/J proteins (Sec63, Scj1,
and Jem1) and nucleotide exchange factors (Sil1) (Nishikawa and Endo 1997;
Sadler etal. 1989; Schlenstedt etal. 1995). Sec63 acts as a Kar2 ATPase activator
(Lyman and Schekman 1995), while Scj1 functions to counter the misfolding of
proteins occurring due to lack of a carbohydrate modication (Silberstein etal. 1998).
SSC1 and SSC 3 are the two mitochondrial Hsp70 chaperones which are involved
in the heat shock response (Wagner etal. 1994). Binding of the SSC1 to the unfolded
proteinis critical for post-stress(Baumann etal. 2000). Mdj1 (J protein),the yeast
mitochondrial HSP70 cofactor, helps in preventing heat induced protein aggrega-
tion in mitochondria (Rowley etal. 1994). It is also involved in protein folding.
Furthermore, in addition to folding, its interaction with SSC1 alsofacilitates the
clearance of misfolded mitochondrial proteins. MGE1 is the only mitochondrial
NEF known to interact with, and therefore assist the Ssc1 chaperone action by pro-
moting the release of bound nucleotide (Wagner etal. 1994).
The HSP70 family in plants has been subdivided into 4 subgroups based on
C-terminal sequence motifs and their cellular localization. The cytosolic HSP70
contains the EEVD motif; the ER HSP70 contains the HDEL motif, the HSP70
molecules found in plastids havethePEGDVIDADFTDSK motifand those in mito-
chondria have aconserved PEAEYEEAKK motif (Guy and Li 1998). These motifs,
known as anchors, are identied by co-chaperones and are involved in substrate
binding through theassistance of co-chaperones (Freeman etal. 1995). Plant cells
are different in that they contain multiple(2-5) HSP70 family members within the
ER (Ray etal. 2016). In Arabidopsis, the totalHSP70 family members are encoded
by about 18 genes, (Lin etal. 2001)and about 12 genes of the HSP70 family have
been identied in the spinach genome (Guy and Li 1998). This gives an idea about
the functional diversity of this chaperone family (Wang etal. 2004). Mechanisms by
which HSP70 mediates its cellular function have been difcult to study in plants
due to the poor survival of deletion mutants and the unavailability of efcient
inhibitors (Sarkar etal. 2013). In prokaryotes, the DnaK/DnaJ system belongs to
HSP70/HSP40 family of heat shock proteins. They function with the prokaryotic
8 Molecular Chaperones in Cellular Stress Response
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
192
NEF GrpE.DnaK is an ATP-dependent chaperone which requires DnaJ and GrpE
for substrate binding through its ATPase cycle. During stress conditions, DnaJ
undergoes certain functional modications which stimulates the ATPase activity of
DnaK (~500 fold). During stress conditions, DnaK/DnaJproteins also interact with
ClpB and reactivate the inactivated proteins (Liberek et al. 1992). DnaK/DnaJ
system is absent in archaeal species except for a few halophiles, hence may not have
any particular role in alleviating archaeal stress. The functional analogs of DnaK/
DnaJ system in archaea are prefoldin and GimC protein (Rani etal. 2016).
8.5.4 HSP90
HSP90 is another major player that functions downstream of HSP70in the confor-
mational maturation and functional activation of several important classes of pro-
teins that actively take part in various cell signaling pathways. While the basal
HSP90 content is about 1% of the total cellular proteincontent, it may increase to
4–6% during the stress response (Young etal. 2001; Wegele etal. 2004). HSP90 is
usually present in the cytosol, but they may also be found in the ER, mitochondria
and plastids (Krishna and Gloor 2001). HSP90 functions as a dimer; the monomer
units covalently joined at their C-terminal domains via the dimerization domain.
The N-terminal domain binds and hydrolyzes ATP and is joined to the C-terminal
domain by a middle domain, which participates in substrate and co-chaperone bind-
ing. In all organisms HSP90 acts as a scaffold for several if not all protein folding
pathways and components. HSP90 dimer undergoes an ATP driven reaction cycle of
substrate binding and release achieved by the transition from an open nucleotide-
free to a closed ATP-bound state ‘committed to hydrolysis. HSP90 works in cohort
with multiple co-chaperones like HSP70,HOP, the J-domain proteins (HSP40) and
p23. These co-chaperones mediate the interactions between HSP90 and the sub-
strates in most of the cases (Li etal. 2012).
In humans, clients of this ~ 90Kda chaperone number over 200, which includes
kinases, nuclear receptors, transcription factors, telomerase and many other proteins
(Zhao etal. 2005; McClellan etal. 2007). Under stress conditions HSP90α is over-
expressed and shows marked increase in interactions with certain clients. HSP90α
has a higher potential for oligomerization and undergoes temperature-dependent
oligomerization above 45°C, and denatured substrates like DHFR bind specically
with these higher order oligomers (4-mer, 6-mer, and 8-mer). Oligomeric forms of
HSP90 display manifold higher afnity for denatured substrates as they themselves
undergo ‘unfolding', wheretheir hydrophobic patches are exposed (Csermely etal.
1998). Like HSP70, HSP90 prevents irreversible denaturation of substrates like
luciferase and can act as ‘holdases’. When cells go back to the resting phase, these
substrates can be captured by ‘foldases’ and reactivated (Minami and Minami
1999). Tumour Necrosis Factor Receptor-Associated protein 1 is the mitochondrial
homologue of Hsp90 (Felts etal. 2000). It is a ~75kDa protein (Chen etal. 1996)
which is structurally similar to its cytosolic counterpart, except the absence of an
B. K. Chatterjee et al.
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
193
MEEVD motif in the C-terminus that binds to co-chaperones like p23, indicating
that the regulation of TRAP-1 might be orchestrated by a different set of co-
chaperones in a yet unknown fashion (Felts etal. 2000). TRAP-1 functionally dif-
fers from its ER counterpart, GRP94in that it adopts a closed conformation upon
ATP binding, but this alone is insufcient for commitment towards ATP hydrolysis
as the propensity to adopt an open conrmation even before ATP hydrolysis is
greater than its ATP hydrolysis rate. This kinetic partitioning essentially reduces the
turnover number of ATP and signies that the dissociation of ATP is favoured to its
ATPase activity (Leskovar etal. 2008). Under heat shock conditions, its ATPase
activity increases by ~ 200 fold (Altieri etal. 2012). Although downstream targets
involved in this process is not fully understood, it is probably the chaperone
Cyclophilin D (a mitochondrial matrix PPIase), a key component of the Permeability
Transition Pore complex (PTP). Opening of the PTP is known to induce mitochon-
drial apoptosis, and by refolding Cyclophilin D in a closed PTP conguration
through its ATPase activity, TRAP-1 is perhaps able to mediate cell survival (Green
and Kroemer 2004; Kang et al. 2007). The proteins classied in HSP90 family
range from 80–94kDa in size and about 70% sequence identity has been observed
between the plant and other eukaryotic counterparts of HSP90 (Lindquist and Craig
1988). The HSP90 family in Arabidopsis has seven members where AtHSP90–1 to
AtHSP90–4 are present in the cytoplasm, AtHSP90–5 is present in the plastids,
AtHSP90–6 in the mitochondria and AtHSP90–7 in the ER (Krishna and Gloor
2001; Wang etal. 2004).It has been observed in Arabidopsis that HSP90 is involved
in mediating the stress response pathways through its interactions with the heat
shock factor (HSF) (Yamada etal. 2007). Overall, studiescarried out in plants sug-
gest important roles of HSP90 in many aspects of plant development and stress
response. HtpG is the Escherichia coli homolog of HSP90. It normally acts as a
holder chaperone that transiently maintains the nascent protein in a conformation
accessible to DnaK. During stress conditions, its expression increases 5–10-fold
and it binds to aggregated proteins with the help of ClpB and re-presents them to the
DnaK/DnaJ system (Thomas and Baneyx 2000). In yeast, HSP90 chaperone is
encoded by the Hsp82 gene, which is overexpressed under heat shock conditions
(Smith etal. 1991). Yeast HSP90 chaperone system is also involved in overcoming
the deleterious impact of heat shock on the cell surface (Imahi & Yahara 2000).
8.5.5 HSP100/Clp
The HSP100 family is primarily known for its unique function of remodeling pro-
tein complexes and disassembling protein aggregates, facilitating either refolding or
degradation of the aggregated proteins (Goloubinoff etal. 1999). The HSP100 are
hexameric ring structures belonging to AAA+ ATPase family (Burton and Baker
2005) with two dened classes based on distinct N and C-domains. The class Ipro-
teins are ClpA, B, C, D, E having two AAA modules, whereas the class II proteins
ClpM, N, X, and Y have only one AAA module (Lee etal. 2007). In E.coli the major
8 Molecular Chaperones in Cellular Stress Response
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
194
HSP100 are ClpA, ClpB, ClpC, ClpX and ClpY.They help in thedisassembly of
oligomers and aggregates during stress (Smith etal. 1999). HSP104 is the yeast
homolog of ClpB. It recognizes misfolded proteins within an aggregate, unfolds
them and ultimately delivers the substrates into various refolding pathways
(Schirmer etal. 1996). Together with HSP70 and HSP40, it resolubilizes and refolds
the substrates. HSP78 (a yeast mitochondrial aggregase), plays a role in the reacti-
vation of proteins damaged due to stress, thus imaprting thermotolerance (Janowsky
etal. 2006). It binds to misfolded polypeptides in the matrix and stabilizes them,
therebypreventing their aggregation (Schmitt etal. 1995). One of the major roles of
HSP78 is resolubilization of the Ssc1 (mtHsp70) chaperone which itself tends to
misfold during stress (Sichting etal. 2005). HEP1 (mtHsp70 escort protein) and
Pim1 (involved in mitochondrial proteolysis) are some of the other proteins in mito-
chondria which are involved in HSR (Sichting et al. 2005; Wagner etal. 1994).
HEP1 plays a complementary role to Hsp78 in maintaining functional SSC1
(Sichting etal. 2005). It assists Ssc1 to maintain its solubility and function during
stress (Sichting etal. 2005). Multiple HSP100 membersexist in Arabidopsis; four
ClpB, two ClpC, and one ClpD.Although they are present at basal levels in cells
under normal conditions, an increased expression is observed during heat stress.
The cytosolic ClpB1 of Arabidopsis, known as AtHSP101 is present in high levels
during heat stress and imparts thermotolerance to plants (Glover and Lindquist
1998; Lee etal. 2007).
8.5.6 Other Chaperones
Many other chaperones also play a role in stress tolerance; e.g., PPIases, AAA
ATPases and so on. PPIase proteins include cyclophilins, FKBPs and parvulins
(Maruyama etal. 2004). They help in cis-trans prolineisomerization in a polypep-
tide chain and help in the fast folding of kinetically trapped proteins. AAA ATPases
are mainly found in archaea and eukaryotes. The major AAA ATPase of archaea is
CDC48 and AMA which function as major proteasomal ATPases. These proteins
regulate proteasomal protein degradation in archaea (Forouzan etal. 2012). In yeast
and humans, most proteins that are translocated into the ER are N-glycosylated with
a branched glucose-3-mannose-9-N-acetylglucosamine-2 (Glc3Man9) glycan chain
(Helenius and Aebi 2004). This oligosaccharide moiety serves as a recognition sig-
nal for lectin-like chaperones calnexin (CNX) and calreticulin (CRT) (Daniels etal.
2003). CRT prevents thermal aggregation and promotes recovery of nonglycosyl-
ated substrates. This happens due to enhanced polypeptide binding property of CRT
under heat stress. CRT also forms oligomers under heat stress via certain conforma-
tional changes that occur in its C-terminal acidic domain. Oligomerization of CRT
and enhanced polypeptide binding are concomitant and explain how CRT acts as a
chaperone. CNX may act to recruit other chaperones like the PDI family member
Erp57 that catalyses disulphide bond formation in a highly oxidized ER lumen or it
may even act as a chaperone, binding to exposed polypeptide stretches of misfolded
AU1
B. K. Chatterjee et al.
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
195
glycoproteins. Both these functions of CNX are enhanced under heat stress condi-
tions. However, these chaperones may also ‘generally’ bind to glycoproteins inde-
pendent of their folding state, suggesting that there is no specic recognition of the
attached polypeptide chain (Buchberger etal. 2010). Other chaperones like GRP94,
Peptidyl-ProlylIsomerases (PPIase) and GRP170 along with the ones mentioned
above form a large ER-localized multi-protein complex that functions as a network
and bind to unfolded proteins in the ER rather than existing as free pools that get
individually assembled onto protein clients. Grp94 most likely acts as a scaffold like
its cytosolic HSP90 counterpart and forms an important part of this multi-chaperone
complex (Ma and Hendershot 2004). Table8.1 lists the major chaperones involved
in attenuating the adverse effects of heat shock and their subcellular localization.
8.6 Detailed Mechanism ofChaperone Assisted
Protein Folding
Out of the many chaperone systems described briey in the previous section, our
group has mainly focussed on a couple of chaperone systems namely GroEL/ES (E.
coli) and HSP90 (human). Based on the various structural and functional studies
carried out by our group since the last decade on the chaperonin GroEL/ES system,
we have better understood the importance of this system in aggregation prevention
of denatured substrates, refolding of substrates and survival of E. coli. While we
have only recently started working on the human HSP90 system, we have obtained
novel information regarding the contribution of HSP90s structure in modulating its
chaperone properties using certain HSP90 inhibitors. The following section
describes in detail the structure and function of GroEL/ES and HSP90 during nor-
mal and under stress conditions.
Table 8.1 Major Chaperones described herein and their subcellular localization (Modied from
(Graner etal. 2014))
Subcellular localization
Major chaperones ER Mitochondria Cytosol Nucleus Cell Surfacea
HSP27 ✓ ✓
HSP60 ✓ ✓
HSP70/HSC70 ✓ ✓
GRP78 (BiP) ✓ ✓
HSP90 ✓ ✓
HSP110 ✓ ✓
GRP94 (gp96) ✓ ✓
CNX/CALRb ✓ ✓
PDIc✓ ✓
aCell surface localization is mostly associated with tumour cell surfaces
bCalnexin/Calreticulin
cProlyl Disulphide Isomerase
8 Molecular Chaperones in Cellular Stress Response
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
t1.1
t1.2
t1.3
t1.4
t1.5
t1.6
t1.7
t1.8
t1.9
t1.10
t1.11
t1.12
t1.13
t1.14
t1.15
t1.16
196
8.7 GroEL/ES Mediated Protein Folding
8.7.1 GroEL/ES Structure
GroEL in its normal state is a porous cylindrical protein made of 14 subunits
arranged in nearly 7- fold rotationally symmetrical rings stacked back to back, and
forms a cage like structure with a central cavity (Braig etal. 1994). Each GroEL
subunit further folds into three domains (Braig etal. 1994; 1995). The apical domain
(residues 190–345) whichis rich in hydrophobic residues and acts as a binding site
for the non-native substrates and co-chaperonin GroES (Fenton etal. 1994). The
intermediate domain (residues 134–190, 377–408) acts like a hinge between the
apical and equatorial domains. This also helps in allosteric communication between
the two domains. The equatorial domain consists of sub-domains E1 (residues
4–133) and E2 (residues 409–523) that form all the intra and inter-subunit interac-
tions required for the proper folding of the monomeric subunit and their assembly
into the tetra-decameric form (Braig etal. 1994, Hayer-Hartl etal. 2016). It also
houses the nucleotide binding pocket since GroEL works in ATP-dependent manner
(Braig etal. 1994). GroES is a single seven membered ring with identical subunits
of 10kDa that binds to one or both ends of the GroEL cylinder in presence of a
nucleotide. Each subunit consists of a β-barrel body with an extended hydrophobic
mobile loop that interacts with the apical domain of GroEL and providesan enclosed
cavity for the folding of substrate proteins (Braig etal. 1994, Fenton et al. 1994;
Hendrick and Hartl 1993). Binding of GroES to the GroEL cylinder doubles the
volume of central cavity to provide sufcient space for the folding of substrate
proteins. Under normal conditions, GroEL interacts with non-native substrates
post- translationally (Fenton etal. 1997), while native proteins under stress are sub-
jected to unfolding that leads to the formation of intermediate ‘aggregation prone’
states that remain bound to GroEL.These can thenbe refolded back to their native
form, presented to other chaperones, or taken up by the proteolysis machinery
(Hartl etal. 2011).
8.7.2 GroEL Mechanism ofSubstrate Folding underNormal
Conditions
GroEL helps in the folding of about 5–15% of total cellular protein under normal
conditions (Ewalt etal. 1997). Depending upon the size of thesubstrate protein, it
assists in folding in two different ways: (I) Cis- mechanism of folding and(II)
Trans- mechanism of folding. The GroEL cavity can encapsulate substrate proteins
ranging between 54–57kDa (Sakikawa etal. 1999) and help in their folding by
providing an Annsen cage inside the cavity. This is termed as the cis-mechanism
of substrate folding. GroEL is not able to encapsulate large proteins (> 60 kDa)inside
its cavity and they can undergo multiple rounds of binding and release at the apical
B. K. Chatterjee et al.
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
197
domain of GroEL before reaching their nal folded or ‘committed to folding’ state.
Both GroES and ATP are required to release the folded/partially folded substrate
from the cis-ring (Fenton and Horwich 1997; Dahiya and Chaudhuri 2014).
8.7.3 Cis- Mechanism ofGroEL Action
The asymmetric complex of GroEL-GroES is the most common form found in the
cellular milieu, along with a few symmetric complexes of GroEL (Weissman etal.
1995). Afew other reports however, show that both types of complexes can exist in
equimolar concentration in cellular milieu and that the ratio of symmetric to the
asymmetric complex is an ADP dependent phenomenon (Yang etal. 2013; Lizuka
and Funatsu 2016). The asymmetric complex has one ring of GroEL occupied by
GroES, while another ring remains ready to accept a non-native substrate protein.
Due to the presence of hydrophobic residues at the apical domain, newly synthe-
sised polypeptides or partially folded substrates bind to the empty GroEL ring. The
binding of substrate protein leads to a conformational change in the same ring which
promotes subsequent binding of GroES and ATP.This leads to the formation of an
isolated cage for the folding of thesubstrate, known as the cis-ring. The conforma-
tional changes that occur during the binding of GroES and ATP make the GroEL
cavity hydrophilic, which in turn providesa proper environment for the folding of
substrate protein. ATP hydrolysis is a slow process with one molecule of ATP
hydrolysed in 8–10seconds. The hydrolysis of ATP lowers the binding afnity of
GroES to the apical domain. This releases GroES and the folded substrate from the
GroEL cis ring; simultaneously ATP binds to the opposite ring along with another
substrate and a new cycle is initiated. (Fenton and Horwich 1997; Weissman etal.
1995). Recent reports show that both rings of GroEL act in a concomitant fashion
during the folding process (Yang etal. 2013). The following schematic explains the
mechanism of cis- folding (Fig.8.1).
8.7.4 Trans Mechanism ofGroEL
Large proteins with molecular weight (>60kDa) are too big to t inside the GroEL
cavity. This does not involve cis-encapsulation, but requires GroES binding to the
trans ring to release either folded or partially folded substrates (Chaudhuri etal.
2001). As thecis-ring binds with the substrate protein, the trans ring acts as a bind-
ing site for GroES, so this mechanism is termed as folding in-trans. There are many
substrate proteins, e.g., 70kDa tail spike protein of phage p22, 86kDa α/β heterodi-
mer, 82kDa mitochondrial aconitase, dimeric citrate synthase, etc. that fold via this
mechanism (Weissman etal. 1995, Chaudhuri etal. 2001). In this process, the non-
native substrate binds to the apical domain of theasymmetric GroEL-GroES com-
plex, and undergoes folding with domain rearrangements or prevents aggregation
8 Molecular Chaperones in Cellular Stress Response
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
198
(Chaudhuri etal. 2001). Theapical domain of GroEL can also bind with the ‘burst
phase intermediate’ state of substrate proteins. To prove this hypothesis, experi-
ments were carried out using slow folding Malate Synthase G (MSG), a 89kDa
multi-domain monomeric protein. Our observations suggest that binding of MSG to
GroEL accelerates the slowest kinetic phase of the spontaneous protein folding
pathway. Due to the large size of thesubstrate, GroES is not able to bind to this ring,
but ATP hydrolysis continues, which helps in the release of the folded substrate
from the ring. Sometimes the release of thesubstrate depends on the binding of both
GroES and ATP to the opposite ring. The binding of GroES and ATP to the opposite
ring also accelerates the release of folded/partially folded substrate from the GroEL
ring (Paul etal. 2007; Dahiya and Chaudhuri 2014). The trans mechanism of sub-
strate folding occurs under normal cellular conditions and also during thermal
stress. It helps in preventing the irreversible aggregation of thermally denatured
proteins. Study of the thermal unfolding pathway of citrate synthase (CS) show that
CS unfolds via an inactive dimeric intermediate. Further unfolding of these interme-
diates led totheir irreversible aggregation. GroEL interacts with this dimeric unfold-
ing intermediate, dissociating them into monomers which stably associate with
GroEL (Grallert etal. 1998). The following schematic explains the mechanism of
Trans-folding (Fig.8.2).
Fig. 8.1 Proposed model for GroEL/ES assisted folding of small proteins via cis- mechanism):
(I) Open ring of GroEL–GroES–ADP complex acts as an acceptor state for the non-native poly-
peptide. (II) Binding of GroES to the GroEL–ATP complex leads to conformational changes in the
apical domain of GroEL; consequently, polypeptide enters in the central cavity. (III) Folding
occurs in the cis cavity before the ATP gets hydrolysed, this weakens the interaction between
GroEL and GroES. (IV) Binding of ATP to the trans ring promotes the release of (N– Native, Ic–
partially folded, U– misfolded) from the cis ring. (V) At the same time, binding of GroES to
GroEL allows GroEL to alternate its rings between binding-active and folding-active states.
(Copyright clearance order number 4177600370026.
B. K. Chatterjee et al.
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
199
8.7.5 Passive Models ofGroEL-Mediated Folding
GroEL passively suppresses protein aggregation (Pelham 1986; Ellis and van der
Vies 1991; Agard 1993) by binding to the exposed hydrophobic regions of aggrega-
tion prone proteins. GroEL specically recognises and binds with the on-pathway
intermediate states, which shifts the overall folding equilibrium away from aggrega-
tion and provides a pool of folding competent monomers that could reach their
native state by further folding or assembly. In a purely passive folding model, high
concentrations of an aggregating molecule cause non-linear increase in the rate of
aggregation (Van den Berg 1999; Ellis 2001). GroEL prevents the formation of
unfavourable intermediates that could lead to aggregation and hence aid in the
proper folding of its substrates.
Fig. 8.2 Proposed model for GroEL/ES assisted folding of large proteins via trans-
mechanism: The burst phase intermediate of MSG is captured by GroEL (orange colored) to form
GroEL-MSG complex. This binding induces minor structural rearrangements to give rise to a more
folding-compatible state. Further addition of GroES/ATP or ATP releases the GroEL-bound form
of MSG, which folds to the native state via formation of a compact intermediate, that is structurally
quite close to the native MSG.GroES (shown in blue) binds in Trans to the folding polypeptide and
doubles the ATP-dependent reactivation rate
8 Molecular Chaperones in Cellular Stress Response
685
686
687
688
689
690
691
692
693
694
695
200
8.8 GroEL inStress
8.8.1 Synthesis ofGroEL during Stress
During heat shock, activation of σ32 transcription factor leads to an increased
production of Hsp inside E.coli. GroELpopulation increases from a basal level of
1–2% to 10–15% (of the total cellular proteincontent). Non-native substrates bind-
ing to GroEL results in a 2-fold increase (30% of the total cytoplasmic protein
content) in the clientele of GroEL, as compared to normal conditions (10–15%)
(Bukau 1993).
8.8.2 Structural andFunctional Modications
duringHeat Shock
Foldase During heat shock, GroEL undergoes phosphorylation allowing it to
function without GroES. Enhanced ATPase activity in the phosphorylated form
ensures relatively fast release of the folded substrate from GroEL.The folding ef-
ciency of phosphorylated GroEL is 50–100 fold higher than its native counterpart.
Phosphorylation is a reversible process and once the cell is relieved of stress, GroEL
undergoes de-phosphorylation and resumes its normal function inside the cell
(Sherman and Goldberg 1994).
Holdase Our studies on monomeric GroELshow that the apical domain is the most
stable region in the GroEL subunit as it requires 4.0M urea and 70°C to undergo
complete unfolding (Golbik etal. 1998; Puri and Chaudhuri 2017). This is a kind of
adaptation in the GroEL structure that can possibly hold aggregating substratesunder
stress conditions. It is also observed that during thermal stress, GroELs protein fold-
ing activity is reduced and itstarts acting as holdase by binding to the aggregation
prone proteins, thusbehaving as a storehouse of proteins. The molecular basis for
such a functional transition is the loss of inter-ring signalling and negative co-
operativity, whichslows down the release of GroES and the unfolded proteins from
the GroEL cavity. This phenomenon is also reversible with GroEL reverting to its
normal function after heat shock (Llorca etal. 1998).
Unfoldase Strong binding of substrate proteins at the GroEL apical domain can be
the main cause of unfolding or ‘stretching’ of bound substrates. The unfolding
activity occurs through inter-domain movement where stretching of the apical
domain helps in ‘opening up’ of the aggregating substrates. The unfolded substrate
remains bound to the GroEL apical domain during stress conditions, but once cel-
lular conditions become normal, GroEL is able to perform its folding function
(Grallert etal. 1998).
B. K. Chatterjee et al.
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
201
8.8.3 ATP Independent Aggregation Prevention
byGroEL Protein
Aggregation prevention tendency of GroEL is an ATP independent process. This
proves useful during stress conditions, because of increasing load of aggregating
proteins and relatively more ATP consumption to maintain cellular homeostasis
(Soini, etal. 2005). In vivo studies in bacteria and yeast demonstrated that depletion
of either GroEL or mitochondrial Hsp60 resulted in aggregation of a large number
of newly translated proteins (Cheng etal. 1989; Horwich etal. 1993). Similarly,
many invitro studies demonstrated that GroEL can rapidly and efciently bind
to non-native states of several proteins and arrest their aggregation e.g., malate
dehydrogenase (MDH; Ranson etal. 1995), ribulose-1,5-bisphosphate oxygenase-
carboxylase (RuBisCO; Goloubinoff etal. 1989), etc. Our current study on aggrega-
tion prevention of Maltodextrin glucosidase (MalZ) demonstrate that GroEL can
prevent aggregation of this protein without the involvement of GroES and ATP (Puri
and Chaudhuri 2017). The passive binding of non-native intermediates to GroEL
can prevent their aggregation by disallowing random hydrophobic interactions.
Once captured proteins reach the folding competent state, they must be released
back into thefree solution, to undergo completion of the nal steps of folding or
oligomer assembly without ATP (Lin and Rye 2006).
8.9 HSP90 Mediated Protein Folding
Structure
HSP90 functions as a exible dimer inside the cell; the dimers consist of two mono-
mers joined at their C-terminus via the dimerization domain. The N-terminus consists
of a conserved Bergerat ATP/ADP binding fold that also binds to Geldanamycin and
Radicicol, competitive inhibitors of ATPbinding (Pearl 2016; Bergerat etal. 1997;
Stebbins etal. 1997; Roe etal. 1999). An N-terminus ‘lid’ segment that responds to
ATP binding has been found, consisting of two highly conserved glycine clusters
(Prodromou etal. 2000). The N-terminus is connected to the Middle (M) domain by
a exible linker that has been shown to convey allosteric modulations from the
M-domain and the C-terminus to the N-terminus, resulting in global conforma-
tional changes. The M-domain binds to co-chaperones that “present” client pro-
teins to HSP90. The C-terminus contains a unique, conserved MEEVD domain that
binds to tetratricopeptide (TPR) domain containing co-chaperones. The nucleotide
binding site lies in a deep pocket on the helical face of the N-domain. The adenine
base, sugar, and the α- phosphate group make extensive contacts within this pocket,
whereas the β- and the γ- phosphate groups display weak and no contacts respec-
tively. A hydrogen bond connects the adenine base with Asp79, while all other
contacts observed are polar in nature with water molecules interacting with the
8 Molecular Chaperones in Cellular Stress Response
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
202
ribose sugar moiety. The α- and the β- phosphate groups are bound to an octahedrally
co-ordinated Mg2+ ion, making an indirect coupling with the protein(Pearl 2016;
Panaretou etal. 1998).
ATPase Cycle
HSP90s function as a chaperone is dependent on its inherent ATPase activity. The
molecule undergoes a series of conformational changes upon ATP and client protein
bindingthat culminatesin ATP hydrolysis and release of the partially folded client
protein(Li etal. 2012). Understanding the mechanism of this ATPase coupled con-
formational cycle has progressively advanced with several research groups using
sophisticated biophysical techniques like FRET and Analytical Ultracentrifugation
to identify the kinetic states in this cycle and the switch that occurs between these
states (Hessling etal. 2009; Mickler etal. 2009). The dening mechanism is the
‘molecular clamp mechanism’, (Ali et al. 2006) where ATP binding to the open
V-shaped constitutive HSP90 dimer (apo-state) induces structural changes via inter-
mediates, the formation of each of which is regulated by co-chaperones that have
specic roles to play in the cycle. The rate limiting step is the formation of a closed
‘tensed’ state, where the N-domains are dimerized and associated with the
M-domains. This closed conformation is committed to ATP hydrolysis withsubse-
quent release of the substrate and ADP occurring concomitantlyas the N-domains
dissociate and HSP90 returns to its open conformation, ready for another round of
theATPase cycle (Pearl 2016; Li etal. 2012). The rate limiting step involves major
structural changes occurring in the N-domain in two distinct ‘switch’ regions: (a)
the β-strand in the N-domain of one monomer hydrogen bonds with the main β-sheet
in the N-domain of the other monomer. Simultaneous movement of the α-helix
exposes a large hydrophobic patch that dimerizes with the equivalent patch of the
other monomer; and (b) the ‘lid’ segment, which ips over ~180° from its apo con-
formation to fold over the bound nucleotide in the pocket and ‘cradle’ the γ-phosphate
of ATP in a series of main-chain hydrogen bonds(Pearl 2016). Additionally, the
exible loop from the M-domain associates with the lid segment of the N-domain
via hydrophobic residues. This intra-molecular docking of the N and M-domains
facilitates the interaction between the γ-phosphate and theR380 residue, resulting
in the assembly of the two- halves of a split active site for ATP hydrolysis(Pearl
2016; Meyer etal. 2003a, b; Cunningham etal. 2012). The stable HSP90C-terminal
dimer model was challenged by Ratzke and his team (Ratzke et al. 2010) who
observed through FRET experiments that apart from the transient dimerization
observed at the N-domains, the C-terminus can also open and close with fast kinet-
ics, even when ATP/ADP is bound to the N-domain. They proposed a unique mech-
anism where HSP90 may undergo multiple transitions from being a dimer at the
C-domain and open at the N-domains, to being open at the C-terminus while
N-domains are dimerized, thereby associating a higher degree of dynamic exibility
that may inuence substrate release (Ratzke etal. 2010). The following schematic
depicts the conformational change from the open to the closed state of HSP90 via
several transition intermediates (Fig.8.3).
B. K. Chatterjee et al.
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
203
HSP90 co-Chaperones
HSP90 can only keep itssubstrates in a folding competent state and prevent them
from undergoing aggregation, but cannot refold any of its clients by itself (Freeman
and Morimoto 1996). It requires the assistance of other proteins to carry out its
biological function. These proteins are also known as co-chaperones. Around 20
co-chaperones have been identied in eukaryotes and while some of them have been
well-characterized, the mechanism by which some of the other co-chaperones func-
tion is unclear. These co-chaperones mainly regulate the ATPase activity and bind-
ing of client proteins to HSP90 (Prodromou etal. 1999; 2002; Richter etal. 2004;
Roe etal. 2004; Chen and Smith 1998). They associate and dissociate dynamically
and help inthe transition from one intermediate conformation to another, or stabi-
lize a certain conformation in the protein folding cycle. Some of these co- chaperones
like HSP90/HSP70 Organizing Protein (HOP) (Johnson et al. 1998), protein
phosphatase PP5 (Silverstein et al. 1997) and PPIase family members FKBP51
Fig. 8.3 An overview of the conformational cycle of HSP90: Several co-chaperones and ATP
mediate the transition between the early, intermediate and late stages of substrate-bound HSP90
(Copyright clearance order number- 4177600710200)
8 Molecular Chaperones in Cellular Stress Response
812
813
814
815
816
817
818
819
820
821
822
823
824
825
204
(Nair etal. 1997) and FKBP52 (Cox etal. 2007; Johnson and Toft 1994) have a TPR
domain that binds to the MEEVD sequence located in the C-terminus of the
chaperone (Scheuer etal. 2000; Das etal. 1998). Table8.2 (Modied from Li etal.
2012) lists some of the well-studied co-chaperones that have been identied and
their role in the ATPase cycle of Hsp90.
HSP90 co-chaperone Cycle
The most recent understanding of the chaperone cycle is that there are three differ-
ent complexes formed in a chronological fashion with different co-chaperone com-
position (Smith 1993). The rst complex, called the early complex consists of
HSP70, HSP40 and a client protein (Smith etal. 1992; Patricia Hernández etal.
2002; Cintron and Toft 2006), whichpresumably isa misfolded or nascent polypep-
tide. This complex docks to HSP90 via HOP which acts as a scaffold. One TPR
domain of HOP binds to one MEEVD motif of the HSP90 dimer while another
HOP domain binds to the early complex (Li etal. 2011). This HOP bound HSP90
complex can be called the intermediate complex. Addition of ATP and a PPIase
results in the formation of an asymmetric complex, where the TPR domain of the
PPIase binds with the unoccupied MEEVD motif of the other monomer, and
Table 8.2 List of some well-studied co-chaperones of HSP90 and their role in the ATPase cycle
(Modied from (Li etal. 2012))
Protein Name Function
TPR containing
co-chaperones
HOP Stabilizes the open conformation of HSP90; inhibits ATP hydrolysis;
simultaneously binds HSP90 and HSP70 and aids in transfer of nascent
polypeptides recognized by HSP70.
FKBP51/52 Maturation of steroid hormone receptors (SHR); chaperone.
CYP40 Maturation of Estrogen receptors specically; chaperone (Chen etal. 1998;
Pirkl and Buchner 2001).
PP5 Post translational modication of HSP90; dephosphorylates HSP90 and
CDC37; plays a role in processing of client proteins.
TPR2 Recognizes both HSP90 and HSP70 through its TPR domain; may act in the
client transfer from HSP70 to HSP90 (Brychzy etal. 2003).
Non-TPR
co-chaperones
AHA1 Stimulates ATPase activity; induces conformational change (Retzlaff etal.
2010; Sato etal. 2000).
P23 Binds and stabilizes closed client bound HSP90 heterocomplex; maturation
of client proteins; inhibits ATP hydrolysis; chaperone (Johnson and Toft
1994; Obermann etal. 1998; Bose etal. 1996; Freeman etal. 1996).
CDC37 Binds specically to client kinases and ‘presents’ them to HSP90 by binding
to HSP90s N-terminal domain via its C-terminal; inhibits ATP binding and
ATPase activity of HSP90; chaperone (MacLean and Picard 2003; Gaiser
etal. 2010; Siligardi etal. 2002; Ali etal. 2006).
B. K. Chatterjee et al.
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
t2.1
t2.2
t2.3
t2.4
t2.5
t2.6
t2.7
t2.8
t2.9
t2.10
t2.11
t2.12
t2.13
t2.14
t2.15
t2.16
t2.17
t2.18
t2.19
t2.20
t2.21
t2.22
t2.23
t2.24
t2.25
t2.26
205
concomitantlyATP binds to the N-terminus (Smith 1993). The chaperone is still in
its open conrmation. HSP90 adopts the closed ‘committed to ATP hydrolysis’
conformation after p23 binds to the intermediate complex (Johnson et al. 1994;
Johnson and Toft 1995; McLaughlin etal. 2006; Freeman etal. 2000). This is called
the late complex where HOP and HSP70/40 dissociate from HSP90 as their binding
afnity is weakened due to the conformational change. After ATP hydrolysis, p23
and PPIase is released along with the partially folded client protein. Not much is
known about the ADP bound conformation that re-positions the relative orientations
of the N-domains just prior to the release of the client (Pearl 2016). Dynamic X-ray
scattering data has revealed that HSP90 can exist in a highly exible conformational
ensemble (Rice etal. 2008; Zhang etal. 2004), even in the nucleotide free state and
that the ADP bound state is a partially closed one with the N-domains close to each
other, but different to the ATP bound closed complex where the N-domains dimer-
ize (Scherrer etal. 1990).
Expression and Regulation of HSP90s Function under Thermal Stress
The upregulation of chaperones under heat shock conditions is mediated by the
Heat Shock Transcription Factor1 (HSF-1) (Fiorenza et al. 1995). HSF-1 under
normal conditions remains in its inactive monomeric state bound to HSP90. Upon
heat shock, HSP90 dissociates from the complex and HSF-1 undergoes trimeriza-
tion (Baler etal. 1993; Sarge etal. 1993). This trimer can bind to DNA regulatory
elementscalled the Heat Shock Elements (HSEs) and upregulate the transcription
of HSP genes like HSP90 and HSP70 (Akerfelt etal. 2010). Elevated levels of HSP
lead to inactivation of HSF-1; HSP90 and its cohorts FKBP2 and p23 bind tothe
trimeric state and attenuates its DNA binding afnity (Zou etal. 1998; Bharadwaj
etal. 1999; Guo etal. 2001) while HSP70/HSP40 bind to HSF-1 and prevent its
transactivation (Shi etal. 1998; Abravaya etal. 1992; Baler etal. 1992). This nega-
tive feedback loop regulates the chaperones at the transcription level, and thereby
modulates the levels of misfolded nascent polypeptides inside the cell. Additionally,
certain post translational modications affect the chaperone functions of HSP90
(Scroggins and Neckers 2007). Under normal circumstances, HSP90 remains exten-
sively phosphorylated, with as many as four phosphorylatedresidues in each iso-
form (Sefton etal. 1978; Kelley and Schlesinger 1982). Under heat shock conditions,
the general understanding is that HSP90 is rapidly dephosphorylated, leading to the
loss of HSP90s ability to stimulate the activity of its client proteins (Lees-Miller and
Anderson 1989; Morange and Bensaude 1991). Dephosphorylation is mediated by
Protein Phosphatase 1 (PP1), while phosphorylation is mediated by several kinases
like CKII, DNA-PK, and Akt (Wandinger etal. 2006; Dougherty etal. 1987; Walker
etal. 1985; Lees-Miller and Anderson 1989; Chalovich and Eisenberg 2012). After
clients that are bound to HSP90 under normal temperature get released due to
dephosphorylation, one of HSP90s client proteins, heme-regulated Inhibitor Kinase
(HRI), is activated by the chaperone upon rapid phosphorylation and this in-turn
down regulates protein synthesis by inactivating eukaryotic initiation Factor-2α
subunit (Scroggins and Neckers 2007). This cycle of dephosphorylation and phos-
phorylation of HSP90 reduces the load of misfolded proteins inside the cell.
8 Molecular Chaperones in Cellular Stress Response
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
206
8.10 Combinatorial Assistance ofVarious Chaperones
There exists various checkpoints to ensure correct folding from the beginning of
protein synthesis tillit attains its native, biologically active conformation. Molecular
chaperones act as a crucial buffer providing conditions conducivefor the partially
unfolded intermediate to fold. The chaperone machinery helps newly synthesized
protein to navigate the complex energy landscape in order to achieve their native
form. The following paragraph describes the presence of various chaperones at
different spatial and temporal points, coordinating with each other to regulate
protein folding.
The chaperones are present at different levels: The rst chaperonic tier consists
of ribosome associated chaperones. These chaperones stabilize the nascent poly-
peptides which are being synthesized on the ribosome and initiate their folding
process. The second tier of components act subsequently and aid in the complete
folding of the proteins. Both systems cooperate to form one single folding pathway.
Chaperones involved in the rst tier include prokaryotic chaperones Trigger Factor
(TF) and eukaryotic Ribosome Associated Complex (RAC) (Kramer et al. 2009).
These are present on the large ribosome near the exit tunnel of a polypeptide chain.
This RACcomplex comprises HSP70 homologs Ssb1, Ssb2 and Ssz1, zuotin (in
yeast) and the corresponding homologs in higher eukaryotes (Hundley etal. 2005;
Otto et al. 2005). These chaperones primarily bind to the exposed hydrophobic
regions of the proteins. TF works in an ATP independent manner to fold a newly
synthesized polypeptide. The bound polypeptide tries to bury its hydrophobic
regions which facilitate the foldingprocess. Interestingly most of the nascent chains
(~70%) seek the assistance of TF and are successfully folded without any further
assistance. Small proteins also fold spontaneously after their synthesis without any
further assistance (Ferbitz etal. 2004; Merz etal. 2008).
Another class of proteins (~ 20%) possess strong hydrophobic elements (Teter
etal. 1999; Thulasiraman et al. 1999) and thus TF and other ribosome associated
chaperones do not provide enough assistance for their folding. It has been reported
that TF dissociates from the ribosome while bound to the polypeptide chain (Agashe
etal. 2004), thereby presenting the unfolded polypeptides to the downstream chap-
erones like DnaJ/DnaK (Martinez-Hackert and Hendrickson 2009). The DnaJ/
DnaK system further interacts with the longer polypeptide chains and helps in their
folding in an ATP dependent manner. In eukaryotes the second tier also consists of
NAC (Nascent chain associated complex) and like TF, it interacts with the newly
synthesized polypeptides and helps them attain their folded conformation with the
assistance of RAC, HSP70, and other cofactors.
HSP70s along with other cofactors like HSP40, J-proteins and various NEFs aid
in the folding of substrates in an ATP dependent manner (Zhu etal. 1996; Mayer
etal. 2000; Pellecchia etal. 2000). HSP40 also binds with the misfolded polypep-
tides and recruits HSP70 to assist in the proper folding of substrates (Young etal.
2003). In eukaryotes, subsequent to the HSP70 system, HSP90 with the help of
numerous regulators and co-chaperones acts as a nisher, helping the polypeptide
B. K. Chatterjee et al.
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
207
attain its biologically active structure (Pearl and Prodromou 2006; Zhao and Houry
2007; Scheuer etal. 2000). The remaining 10% of substrates that remain in a non-
functional, yet non-aggregated state, are shifted to the chaperonin cage for their
folding (Hartl 1996; Horwich etal. 2007). The environment in the nano cage of
chaperonin facilitates the misfolded proteins to attain a native-like structure
(Gromiha and Selvaraj 2004). In bacteria, GroEL/ES help in thefolding of substrate
proteins. In eukaryotes, substrate proteins are presented to the chaperonin TRiC.The
reaction is mediated by HSP70 and Prefoldin which interact directly with the TRiC
system resulting in the release of folded, functional proteins (Hartl and Hayer-Hartl
2002).
8.11 Conclusions
Compilation of studies both past and recent, on chaperone mediated protein folding
unequivocally show that chaperones play a basal role in maintaining protein homeo-
stasis inside the cell. While some of them function as ‘foldases’, hydrolysing ATP
to carry out the folding and release of various substrates, a few others function
independent of ATP, primarily as ‘holdases’. Under stress conditions however,
chaperones undergo multiple levels of modication; from the enhanced rates of ATP
hydrolysis andmore efcient substrate binding to various post-translational modi-
cations that aid in the transcriptional upregulation of several hsp genes. Chaperones
can also prevent aggregation of misfolded proteins that tend to accumulate under
heat shock conditions. Chaperones play a dual role inside the cell; either rescuing
misfolded or unfolded polypeptides and preventing aggregation or triggering degra-
dation of substrates when cell damage is irreversible, both of which needs to be
tightly regulated to maintain proteostasis. Such properties of chaperones are cur-
rently being utilized in the industry as well as in designing therapy for certain debil-
itating diseases like cancer and neurodegenerative disorders. Overexpression of
certain proteins can be a major problem, especially when they are being expressed
in a different host. Several studies including some of our own have shown that
achaperone or a combination of chaperones have been proven effective in increas-
ing the yield as well as the active fraction of total protein expressed. Some of those
proteins are considered therapeutically important; hence, large-scale productions of
such proteins are regularly uptaken by the pharmaceutical industry. Chaperones
thus ablate a signicant clog in this giant wheel of industrial-scale protein produc-
tion and such strategies have come to the rescue over the years. HSP90 has been
used as a target to design inhibitors that have been successful in the amelioration of
tumor progression and development, and is considered a hot prospect for drug-
based therapy for the treatment of certain types of cancer. Although a promising
approach, only a few compounds have reached the clinical trials and none of them
have made it to the market. On the other end of the spectrum, the ability to prevent
aggregation of misfolded proteins and even refold partially folded ‘aggregation-
prone’ intermediates can be used to treat neurodegenerative disorders. We are
8 Molecular Chaperones in Cellular Stress Response
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
208
currently working on one such compound that stimulates the chaperone functions of
HSP90 in-vitro, and plan to carry out several studies to investigate its potential in
preventing certain neurodegenerative conditions. Overall, chaperones make an
interesting topic of study, not only because they play such important roles in regu-
lating protein function, stability and degradation but also because they possess tre-
mendous value both industrially and therapeutically.
Acknowledgments The authors acknowledge the nancial assistance from IIT Delhi and infra-
structural facility from IIT Delhi, India. AS and AP acknowledge nancial assistance from CSIR,
Government of India for providing fellowships in their doctoral course programme. SP acknowl-
edge nancial assistance from UGC, Government of India for providing fellowships in their doc-
toral course programme. BKC acknowledges IIT Delhi for providing fellowship in the doctoral
course program.
References
Abravaya, K., etal. (1992). The human heat shock protein hsp70 interacts with HSF, the tran-
scription factor that regulates heat shock gene expression. Genes and Development, 6(7),
1153–1164.
Agard, D. (1993). To fold or not to fold. Science, 260(5116), 1903–1904.
Agashe, V. R., et al. (2004). Function of trigger factor and DnaK in multidomain protein folding.
Cell, 117(2), 199–209.
Akerfelt, M., Morimoto, R.I., & Sistonen, L. (2010). Heat shock factors: Integrators of cell stress,
development and lifespan. Nature Reviews. Molecular Cell Biology, 11(8), 545–555.
Ali, M.M. U., etal. (2006). Crystal structure of an Hsp90–nucleotide–p23/Sba1 closed chaperone
complex. Nature, 440(7087), 1013–1017.
Altieri, D.C., et al. (2012). TRAP-1, the mitochondrial Hsp90. Biochimica et Biophysica Acta-
Molecular Cell Research, 1823(3), 767–773.
Amin, J., Ananthan, J., & Voellmy, R. (1988). Key features of heat shock regulatory elements.
Molecular and Cellular Biology, 8(9), 3761–3769.
Bakthisaran, R., Tangirala, R., & Rao, C.M. (2015). Small heat shock proteins: Role in cellular
functions and pathology. Biochimica et Biophysica Acta - Proteins & Proteomics, 1854(4),
291–319.
Balch, W. E., et al. (2008). Adapting proteostasis for disease intervention. Science, 319(5865),
916–919.
Baler, R., Welch, W.J., & Voellmy, R. (1992). Heat shock gene regulation by nascent polypeptides
and denatured proteins: hsp70 as a potential autoregulatory factor. The Journal of Cell Biology,
117(6), 1151–1159.
Baler, R., Dahl, G., & Voellmy, R. (1993). Activation of human heat shock genes is accompanied
by oligomerization, modication, and rapid translocation of heat shock transcription factor
HSF1. Molecular and Cellular Biology, 13(4), 2486–2496.
Baumann, F., Milisav, I., Neupert, W., & Herrmann, J.M. (2000). Ecm10, a novel hsp70 homolog
in the mitochondrial matrix of the yeast Saccharomyces Cerevisiae. FEBS Letters, 487(2),
307–312.
van den Berg, B. (1999). Effects of macromolecular crowding on protein folding and aggregation.
The EMBO Journal, 18(24), 6927–6933.
Bergerat, A., etal. (1997). An atypical topoisomerase II from archaea with implications for meiotic
recombination. Nature, 386, 414–417.
B. K. Chatterjee et al.
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
1001
1002
1003
1004
1005
1006
1007
1008
1009
1010
1011
1012
1013
1014
1015
209
Bharadwaj, S., Ali, A., & Ovsenek, N. (1999). Multiple components of the HSP90 chaperone
complex function in regulation of heat shock factor 1 invivo. Molecular and Cellular Biology,
19(12), 8033–8041.
Bose, S., et al. (1996). Chaperone function of Hsp90-associated proteins. Science, 274(5293),
1715–1717.
Boston, R.S., Viitanen, P.V., & Vierling, E. (1996). Molecular chaperones and protein folding in
plants. Plant Molecular Biology, 32(1), 191–222.
Braig, K., Otwinowski, Z., Hegde, R., Boisvert, D.C., Joachimiak, A., Horwich, A.L., & and-
Sigler, P. B. (1994). The crystal structure of bacterial chaperonin GroEL at 2.8Ao. Nature,
371(6498), 578–586.
Braig, K., Adams, P.D., & Brunger, A.T. (1995). Conformational variability in the rened struc-
ture of the chaperonin GroEL at 2.8Ao resolution. Nature Structural and Molecular Biology,
2(12), 1083–1094.
Brodsky, J.L., Goeckeler, J., & Schekman, R. (1995). BiP and Sec63p are required for both co-
and posttranslational protein translocation into the yeast endoplasmic reticulum. Proceedings
of the National Academy of Sciences, 92(21), 9643–9646.
Brychzy, A., etal. (2003). Cofactor Tpr2 combines two TPR domains and a Jdomain to regulate
the Hsp70/Hsp90 chaperone system. The EMBO Journal, 22(14), 3613–3623.
Buchberger, A., Bukau, B., & Sommer, T. (2010). Protein quality control in the cytosol and the
endoplasmic reticulum: Brothers in arms. Molecular Cell, 40(2), 238–252.
Bukau, B. (1993). Regulation of the Escherichia Coli heat-shock response. Molecular Microbiology,
9(4), 671–680.
Burton, B.M., & Baker, T.A. (2005). Remodeling protein complexes: Insights from the AAA+
unfoldase ClpX and mu transposase. Protein Science, 14(8), 1945–1954.
Carrascosa, L., Muga, A., & Valpuesta, M. (1998). GroEL under Heat-Shock. The Journal of
Biological Chemistry, 273(49), 32587–32594.
Chalovich, J.M., & Eisenberg, E. (2012). A proteomic screen identied stress-induced chaperone
proteins as targets of Akt phosphorylation in Mesangial cells. Biophysical Chemistry, 257(5),
2432–2437.
Charmandari, E., Tsigos, C., & Chrousos, G. (2005). Endocrinology of the stress response. Annual
Review of Physiology, 67, 259–284.
Chaston, J.J., etal. (2016). Structural and functional insights into the evolution and stress adapta-
tion of type II Chaperonins. Structure, 24(3), 364–374.
Chaudhuri, T. K., Farr, G. W., Fenton, W. A., Rospert, S., & Horwich, A. L. (2001). GroEL/GroES-
mediated folding of a protein too large to be encapsulated. Cell, 107(2), 235–246.
Chen, S., & Smith, D.F. (1998). Hop as an adaptor in the heat shock protein 70 (Hsp70) and Hsp90
chaperone machinery. The Journal of Biological Chemistry, 273(52), 35194–35200.
Chen, Q., etal. (1990). Accumulation, stability, and localization of a major chloroplast heat-shock
protein. The Journal of Cell Biology, 110(6), 1873–1883.
Chen, C.F., etal. (1996). A new member of the hsp90 family of molecular chaperones interacts
with the retinoblastoma protein during mitosis and after heat shock. Molecular and Cellular
Biology, 16(9), 4691–4699.
Chen, S., et al. (1998). Differential interactions of p23 and the TPR-containing proteins hop,
Cyp40, FKBP52 and FKBP51 with Hsp90 mutants. Cell Stress & Chaperones, 3(2), 118–129.
Cheng, M. Y., et al. (1989). Mitochondrial heat-shock protein hsp60 is essential for assembly of
proteins imported into yeast mitochondria. Nature, 337(6208), 620–625.
Cintron, N.S., & Toft, D. (2006). Dening the requirements for Hsp40 and Hsp70in the Hsp90
chaperone pathway. The Journal of Biological Chemistry, 281(36), 26235–26244.
Cotto, J., Kline, M., & Morimoto, R.I. (1996). Activation of heat shock factor 1 DNA binding
precedes stress- induced serine phosphorylation. The Journal of Biological Chemistry, 271(7),
3355–3358.
Cox, J.S., & Walter, P. (1996). A novel mechanism for regulating activity of a transcription factor
that controls the unfolded protein response. Cell, 87(3), 391–404.
8 Molecular Chaperones in Cellular Stress Response
1016
1017
1018
1019
1020
1021
1022
1023
1024
1025
1026
1027
1028
1029
1030
1031
1032
1033
1034
1035
1036
1037
1038
1039
1040
1041
1042
1043
1044
1045
1046
1047
1048
1049
1050
1051
1052
1053
1054
1055
1056
1057
1058
1059
1060
1061
1062
1063
1064
1065
1066
1067
1068
210
Cox, M.B., etal. (2007). FK506-binding protein 52 phosphorylation: A potential mechanism for
regulating steroid hormone receptor activity. Molecular Endocrinology, 21(12), 2956–2967.
Csermely, P., etal. (1998). The 90-kDa molecular chaperone family: Structure, function, and clini-
cal applications. A comprehensive review. Pharmacology & Therapeutics, 79(2), 129–168.
Cunningham, C.N., etal. (2012). The conserved arginine 380 of Hsp90 is not a catalytic resi-
due, but stabilizes the closed conformation required for ATP hydrolysis. Protein Science, 21,
1162–1171.
Dahiya, V., & Chaudhuri, T.K. (2014). GroEL/ES accelerates the refolding of a multi-domain
protein through modulating on pathway intermediates. The Journal of Biological Chemistry,
289(1), 286–298.
Daniels, R., etal. (2003). N-linked glycans direct the cotranslational folding pathway of inuenza
hemagglutinin. Molecular Cell, 11(1), 79–90.
Das, A.K., et al. (1998). The structure of the tetratricopeptide repeats of protein phosphatase
5: Implications for TPR-mediated protein-protein interactions. The EMBO Journal, 17(5),
1192–1199.
Ditzel, L., Lowe, J., Stock, D., Stetter, K.O., Huber, H., Huber, R., & Steinbacher, S. (1998).
Crystal structure of the thermosome, the archaeal chaperonin and homolog of CCT. Cell, 93(1),
125–138.
Dougherty, J., etal. (1987). Identication of the 90Kda substrate of rat liver type II casein kinase
with the heat shock protein which binds steroid receptors. Biochimica et Biophysica Acta, 927,
74–80.
Ellis, R. J. (2001). Macromolecular crowding: Obvious but underappreciated. Trends in
Biochemical Sciences, 26(10), 597–604.
Ellis, R. J., & van der Vies, S. M. (1991). Molecular Chaperones. Annual Review of Biochemistry,
60(1), 321–347.
Ewalt, K. L., Hendrick, J. P., Houry, W. A., & Hartl, F. U. (1997). In Vivo observation of polypep-
tide ux through the bacterial chaperonin system. Cell, 90(3), 491–500.
Eyles, S.J., & Gierasch, L.M. (2010). Nature’s molecular sponges: Small heat shock proteins
grow into their chaperone roles. Proceedings of the National Academy of Sciences, 107(7),
2727–2728.
Felts, S. J., et al. (2000). Protein structure and folding: The hsp90-related protein TRAP1 is a
mitochondrial protein with distinct functional properties the hsp90-related protein TRAP1 is a
mitochondrial protein with distinct functional properties. The Journal of Biological Chemistry,
275(5), 3305–3312.
Ferbitz, L., et al. (2004). Trigger factor in complex with the ribosome forms a molecular cradle for
nascent proteins. Nature, 431(7008), 590–596.
Fenton, W. A., & Horwich, A. L. (1997). Review, GroEL mediated protein folding. Protein
Science, 6, 743–760.
Fenton, W. A., Kashi, Y., Furtak, K., & Horwich, A. L. (1994). Residues in chaperonin GroEL
required for polypeptide binding and release. Nature, 371(6498), 614–619.
Figueiredo, L., Klunker, D., Ang, D., Naylor, D.J., Kerner, M.J., Georgopoulos, C., Hartl, F.U., &
Hayer-Hartl, M. (2004). Functional characterization of an archaeal GroEL/GroES chaperonin
system: Signicance of substrate encapsulation. The Journal of Biological Chemistry, 279(2),
1090–1099.
Fiorenza, M. T., etal. (1995). Complex expression of murine heat shock transcription factors.
Nucleic Acids Research, 23(3), 467–474.
Forouzan, D., Ammelburg, M., Hobel, C.F., Ströh, L.J., Sessler, N., Martin, J., & Lupas, A.N.
(2012). The archaeal proteasome is regulated by a network of AAA ATPases. The Journal of
Biological Chemistry, 287(46), 39254–39262.
Franck, E., etal. (2004). Evolutionary diversity of vertebrate small heat shock proteins. Journal of
Molecular Evolution, 59(6), 792–805.
Freeman, B.C., & Morimoto, R. I. (1996). The human cytosolic molecular chaperones hsp90,
hsp70 (hsc70) and hdj-1 have distinct roles in recognition of a non-native protein and protein
refolding. The EMBO Journal, 15(12), 2969–2979.
B. K. Chatterjee et al.
1069
1070
1071
1072
1073
1074
1075
1076
1077
1078
1079
1080
1081
1082
1083
1084
1085
1086
1087
1088
1089
1090
1091
1092
1093
1094
1095
1096
1097
1098
1099
1100
1101
1102
1103
1104
1105
1106
1107
1108
1109
1110
1111
1112
1113
1114
1115
1116
1117
1118
1119
1120
1121
1122
211
Freeman, B.C., etal. (1995). Identication of a regulatory motif in Hsp70 that affects ATPase
activity, substrate binding and interaction with HDJ-1. The EMBO Journal, 14(10), 2281–2292.
Freeman, B.C., Toft, D.O., & Morimoto, R.I. (1996). Molecular chaperone machines: Chaperone
activities of the cyclophilin Cyp-40 and the steroid aporeceptor-associated protein p23. Science,
274, 1718–1720.
Freeman, B.C., etal. (2000). The p23 molecular chaperones act at a late step in intracellular recep-
tor action to differentially affect ligand efcacies. Genes and Development, 14(4), 422–434.
Frydman, J.(2001). Folding of newly translated proteins invivo: The role of molecular chaper-
ones. Annual Review of Biochemistry, 70, 603–647.
Fulda, S., et al. (2010). Cellular stress responses: Cell survival and cell death. Int JCell Biol,
2010(2010), 1–23.
Gaiser, A.M., Kretzschmar, A., & Richter, K. (2010). Cdc37-Hsp90 complexes are responsive to
nucleotide-induced conformational changes and binding of further cofactors. The Journal of
Biological Chemistry, 285(52), 40921–40932.
Georgopoulos, C., Liberek, K., Zylicz, M., & Ang, D. (1994). Properties of the heat shock proteins
of Escherichia coli and the autoregulation of the heat shock response. Biol Heat Shock Prot
Mol Chaperones, 209–249.
Glover, J. R., & Lindquist, S. (1998). Hsp104, Hsp70, and Hsp40: A novel chaperone system that
rescues previously aggregated proteins. Cell, 94(1), 73–82.
Golbik, R., Zahn, R., Harding, S.E., & Alan, F.R. (1998). Thermodynamic stability and folding of
GroEL minichaperones. Journal of Molecular Biology, 276, 505–515.
Goloubinoff, P., Gatenby, A. A., & Lorimer, G. H. (1989). GroE heat-shock proteins promote
assembly of foreign prokaryotic ribulose bisphosphate carboxylase oligomers in Escherichia
coli. Nature, 337(6202), 44–47.
Goloubinoff, P., etal. (1999). Sequential mechanism of solubilization and refolding of stable pro-
tein aggregates by a bichaperone network. Proceedings of the National Academy of Sciences,
96(24), 13732–13737.
Grallert, H., Rutkat, K., & Buchner, J.(1998). GroEL traps dimeric and monomeric unfolding
intermediates of citrate synthase. The Journal of Biological Chemistry, 273(50), 33305–33310.
Graner, M.W., Lillehei, K.O., & Katsanis, E. (2014). Endoplasmic reticulum chaperones and their
roles in the immunogenicity of cancer vaccines. Frontiers in Oncology, 4(379), 1–12.
Green, D.R., & Kroemer, G. (2004). The pathophysiology of mitochondrial cell death. Science,
305(5684), 626–629.
Gromiha, M. M., & Selvaraj, S. (2004). Inter-residue interactions in protein folding and stability.
Progress in Biophysics and Molecular Biology, 86(2), 235–277.
Gross, C.A. (1996). Function and regulation of the heat shock proteins in Escherichia coli and
Salmonella cellular and molecular biology. American Society for Microbiology, 1382–1399.
Guisbert, E., Herman, C., Lu, C.Z., & Gross, C.A. (2004). A chaperone network controls the heat
shock response in E.Coli. Genes & Development, 18(22), 2812–2821.
Guo, Y., et al. (2001). Evidence for a mechanism of repression of heat shock factor 1 transcrip-
tional activity by a multichaperone complex. The Journal of Biological Chemistry, 276(49),
45791–45799.
Guy, C.L., & Li, Q.-B. (1998). The organization and evolution of the spinach stress 70 molecular
chaperone gene family. Plant Cell, 10(4), 539–556.
Hartl, F. U. (1996). Molecular chaperones in cellular protein folding. Nature, 381(6583), 571–580.
Hartl, F.U., Bracher, A., & Hartl, M.H. (2011). Molecular chaperones in protein folding and pro-
teostasis. Nature, 475(7356), 325–332.
Hartl, F. U., & Hayer-Hartl, M. (2002). Molecular chaperones in the cytosol: From nascent chain
to folded protein. Science, 295(5561), 1852–1858.
Haslbeck, M., Walke, S., Stromer, T., Ehrnsperger, M., White, H. E., Chen, S., Saibil, H.R., &
Buchner, J.(1999). Hsp26: A temperature-regulated chaperone. The EMBO Journal, 18(23),
6744–6751.
8 Molecular Chaperones in Cellular Stress Response
1123
1124
1125
1126
1127
1128
1129
1130
1131
1132
1133
1134
1135
1136
1137
1138
1139
1140
1141
1142
1143
1144
1145
1146
1147
1148
1149
1150
1151
1152
1153
1154
1155
1156
1157
1158
1159
1160
1161
1162
1163
1164
1165
1166
1167
1168
1169
1170
1171
1172
1173
1174
212
Haslbeck, M., Braun, N., Stromer, T., Richter, B., Model, N., Weinkauf, S., & Buchner, J. (2004a).
Hsp42 is the general small heat shock protein in the cytosol of Saccharomyces cerevisiae. The
EMBO Journal, 23(3), 638–649.
Haslbeck, M., Ignatiou, A., Saibil, H., Helmich, S., Frenzl, E., Stromer, T., & Buchner, J.(2004b).
A domain in the N-terminal part of Hsp26 is essential for chaperone function and oligomeriza-
tion. Journal of Molecular Biology, 343(2), 445–455.
Hayer-Hartl, M., Bracher, A., & Hartl, F. U. (2016). The GroEL-GroES chaperonin machine: A
nano-cage for protein folding. Trends in Biochemical Sciences, 41(1), 62–76.
Helenius, A., & Aebi, M. (2004). Roles of N-linked glycans in the endoplasmic reticulum. Annual
Review of Biochemistry, 73, 1019–1049.
Hemmingsen, S.M., etal. (1988). Homologous plant and bacterial protein chaperone oligomeric
protein assembly. Nature, 333, 330–334.
Hendrick, J. P., & Hartl, F. U. (1993). Molecular chaperone functions of heat-shock proteins.
Annual Review of Biochemistry, 62(1), 349–384.
Herman, C., & Gross, C.A. (2000). Heat stress. In J.Lederberg (Ed.), Encyclopedia of microbiol-
ogy (pp.598–606). Academic Press.
Hessling, M., Richter, K., & Buchner, J.(2009). Dissection of the ATP-induced conformational
cycle of the molecular chaperone Hsp90. Nature Structural & Molecular Biology, 16(3),
287–293.
Hietakangas, V., & Sistonen, L. (2006). Regulation of the heat shock response by heat shock tran-
scription factors. Chaperones, 1–34.
Hightower, L.E. (1991). Heat shock, stress proteins, chaperones, and Proteotoxicity. Cell, 66(2),
191–197.
Hill, J.E., & Hemmingsen, S. M. (2001). Arabidopsis Thaliana type I and II chaperonins. Cell
Stress and Chaperones, 6(3), 190–200.
Horwich, A.L., Fenton, W.A., Chapman, E., & Farr, G.W. (2007). Two families of chaperonin:
Physiology and mechanism. Annual Review of Cell and Developmental Biology, 23, 115–145.
Horwich, A. L., Low, K. B., Fenton, W. A., Hirsheld, I. N., & Furtak, K. (1993). Folding in vivo
of bacterial cytoplasmic proteins: Role of GroEL. Cell, 74(5), 909–917.
Hundley, H. A. (2005). Human Mpp11 J protein: Ribosome-tethered molecular chaperones are
ubiquitous. Science, 308(5724), 1032–1034.
Imai, J., & Yahara, I. (2000). Role of HSP90in salt stress tolerance via stabilization and regulation
of calcineurin. Molecular and Cellular Biology, 20(24), 9262–9270.
Jakob, U., Gaestel, M., Engel, K., & Buchner, J.(1993). Small heat shock proteins are molecular
chaperones. The Journal of Biological Chemistry, 268(3), 1517–1520.
Janowsky, B., Major, T., Knapp, K., & Voos, W. (2006). The disaggregation activity of the mito-
chondrial ClpB homolog Hsp78 maintains Hsp70 function during heat stress. Journal of
Molecular Biology, 357(3), 793–807.
Johnson, J.L., & Toft, D.O. (1994). A novel chaperone complex for steroid receptors involving
heat shock proteins, immunophilins, and p23. The Journal of Biological Chemistry, 269(40),
24989–24993.
Johnson, J.L., & Toft, D.O. (1995). Binding of p23 and hsp90 during assembly with the proges-
terone receptor. Molecular Endocrinology, 9(6), 670–678.
Johnson, J.L., etal. (1994). Characterization of a novel 23-kilodalton protein of unactive proges-
terone receptor complexes. Molecular and Cellular Biology, 14(3), 1956–1963.
Johnson, B.D., et al. (1998). Hop modulates hsp70 / hsp90 interactions in protein folding. The
Journal of Biological Chemistry, 273(6), 3679–3686.
Kampinga, H.H., & Craig, E.A. (2010). The HSP70 chaperone machinery: Jproteins as drivers of
functional specicity. Nature Reviews. Molecular Cell Biology, 11(8), 579–592.
Kang, B.H., et al. (2007). Regulation of tumor cell mitochondrial homeostasis by an organelle-
specic Hsp90 chaperone network. Cell, 131(2), 257–270.
Kappé, G., etal. (2003). The human genome encodes 10 alpha-crystallin-related small heat shock
proteins: HspB1-10. Cell Stress and Chaperones, 8(1), 53–61.
B. K. Chatterjee et al.
1175
1176
1177
1178
1179
1180
1181
1182
1183
1184
1185
1186
1187
1188
1189
1190
1191
1192
1193
1194
1195
1196
1197
1198
1199
1200
1201
1202
1203
1204
1205
1206
1207
1208
1209
1210
1211
1212
1213
1214
1215
1216
1217
1218
1219
1220
1221
1222
1223
1224
1225
1226
1227
213
Kelley, P.M., & Schlesinger, M.J. (1982). Antibodies to two major chicken heat shock proteins
cross-react with similar proteins in widely divergent species. Molecular and Cellular Biology,
2(3), 267–274.
Klumpp, M., Baumeister, W., & Essen, L.O. (1997). Structure of the substrate binding domain of
the thermosome, an archaeal group II chaperonin. Cell, 91(2), 263–270.
Kourtis, N., & Tavernarakis, N. (2011). Cellular stress response pathways and ageing: Intricate
molecular relationships. The EMBO Journal, 30(13), 2520–2531.
Kramer, G., Boehringer, D., Ban, N., & Bukau, B. (2009). The ribosome as a platform for co-
translational processing, folding and targeting of newly synthesized proteins. Nature Structural
& Molecular Biology, 16(6), 589–597.
Kriehuber, T., etal. (2010). Independent evolution of the core domain and its anking sequences
in small heat shock proteins. The FASEB Journal, 24(10), 3633–3642.
Krishna, P., & Gloor, G. (2001). The Hsp90 family of proteins in Arabidopsis Thaliana. Cell Stress
and Chaperones, 6(3), 238–246.
Kültz, D. (2004). Molecular and evolutionary basis of the cellular stress response. Annual Review
of Physiology, 67(1), 225–257.
Langer, T., et al. (1992). Successive action of DnaK, DnaJ and GroEL along the pathway of
chaperone- mediated protein folding. Nature, 356(6371), 683–689.
Lee, U., etal. (2007). The Arabidopsis ClpB/Hsp100 family of proteins: Chaperones for stress and
chloroplast development. The Plant Journal, 49(1), 115–127.
Lees-Miller, S.P., & Anderson, C.W. (1989). Two human 90-kDa heat shock proteins are phos-
phorylated invivo at conserved serines that are phosphorylated invitro by casein kinase II. The
Journal of Biological Chemistry, 264(5), 2431–2437.
Leitner, A., Joachimiak, L.A., Bracher, A., Mönkemeyer, L., Walzthoeni, T., Chen, B., Pechmann,
S., Holmes, S., Cong, Y., & Ma, B. (2012). The molecular architecture of the eukaryotic chap-
eronin TRiC/CCT. Structure, 20(5), 814–825.
Leskovar, A., etal. (2008). The ATPase cycle of the mitochondrial Hsp90 analog trap1. The Journal
of Biological Chemistry, 283(17), 11677–11688.
Levy-Rimler, G., etal. (2002). Type I chaperonins: Not all are created equal. FEBS Letters, 529(1),
1–5.
Li, J., Richter, K., & Buchner, J.(2011). Mixed Hsp90-cochaperone complexes are important for
the progression of the reaction cycle. Nature Structural and Molecular Biology, 18(1), 61–66.
Li, J., Soroka, J., & Buchner, J. (2012). The Hsp90 chaperone machinery: Conformational
dynamics and regulation by co-chaperones. Biochimica et Biophysica Acta – Molecular Cell
Research, 1823(3), 624–635.
Liberek, K., Galitski, T.P., Zylicz, M., & Georgopoulos, C. (1992). The DnaK chaperone modu-
lates the heat shock response of Escherichia coli by binding to the σ32 transcription factor.
Proceedings of the National Academy of Sciences, 89, 3516–3520.
Lin, Z., & Rye, H.S. (2006). GroEL mediated protein folding: Making the impossible, possible.
Critical Reviews in Biochemistry and Molecular Biology, 41(4), 211–239.
Lin, B.L., etal. (2001). Genomic analysis of the Hsp70 superfamily in Arabidopsis Thaliana. Cell
Stress and Chaperones, 6(3), 201–208.
Lindquist, S. (1986). The heat-shock response. Annual Review of Biochemistry, 55(1), 1151–1191.
Lindquist, S., & Craig, E.A. (1988). The heat-shock proteins. Annual Review of Genetics, 22(1),
631–677.
Lizuka, R., & Funatsu, T. (2016). Chaperonin GroEL uses asymmetric and symmetric reac-
tion cycles in response to the concentration of non-native substrate proteins. Biophy and
Physicobiol, 13, 63–69.
Llorca, O., Galán, A., Carrascosa, J. L., Muga, A., & Valpuesta, J. M. (1998). GroEL under heat-
shock. Journal of Biological Chemistry, 273(49), 32587–32594.
Lopez, T., Dalton, K., & Frydman, J.(2016). The mechanism and function of group II chaperonins.
Journal of Molecular Biology, 427(18), 2919–2930.
8 Molecular Chaperones in Cellular Stress Response
1228
1229
1230
1231
1232
1233
1234
1235
1236
1237
1238
1239
1240
1241
1242
1243
1244
1245
1246
1247
1248
1249
1250
1251
1252
1253
1254
1255
1256
1257
1258
1259
1260
1261
1262
1263
1264
1265
1266
1267
1268
1269
1270
1271
1272
1273
1274
1275
1276
1277
1278
1279
214
Lyman, S.K., & Schekman, R. (1995). Interaction between BiP and Sec63p is required for the
completion of protein translocation into the ER of Saccharomyces Cerevisiae. The Journal of
Cell Biology, 131(5), 1163–1171.
Ma, Y., & Hendershot, L.M. (2004). ER chaperone functions during normal and stress conditions.
Journal of Chemical Neuroanatomy, 28(1–2), 51–65.
MacLean, M., & Picard, D. (2003). Cdc37 goes beyond Hsp90 and kinases. Cell Stress &
Chaperones, 8(2), 114–119.
Mani, N., et al. (2016). Multiple oligomeric structures of a bacterial small heat shock protein.
Scientic Reports, 6, 1–12.
Martinez-Hackert, E., & Hendrickson, W. A. (2009). Promiscuous substrate recognition in folding
and assembly activities of the trigger factor chaperone. Cell, 138(5), 923–934.
Maruyama, T., Suzuki, R., & Furutani, M. (2004). Archaeal peptidylprolyl cis-trans isomerases
(PPIases) update. Frontiers in Bioscience, 9, 1680–1700.
Mayer, M.P. (2010). Gymnastics of molecular chaperones. Molecular Cell, 39(3), 321–331.
Mayer, M. P., Schröder, H., Rüdiger, S., Paal, K., Laufen, T., & Bukau, B. (2000). Multistep mech-
anism of substrate binding determines chaperone activity of Hsp70. Nature Structural Biology,
7, 586–593.
McClellan, A.J., etal. (2007). Diverse cellular functions of the Hsp90 molecular chaperone uncov-
ered using systems approaches. Cell, 131(1), 121–135.
McCracken, A. A., & Brodsky, J.L. (1996). Assembly of ER-associated protein degradation
in vitro: Dependence on cytosol, calnexin, and ATP. The Journal of Cell Biology, 132(3),
291–298.
McLaughlin, S.H., et al. (2006). The co-chaperone p23 arrests the Hsp90 ATPase cycle to trap
client proteins. Journal of Molecular Biology, 356(3), 746–758.
Merz, F., et al. (2008). Molecular mechanism and structure of trigger factor bound to the translat-
ing ribosome. The EMBO Journal, 27(11), 1622–1632.
Meyer, A. S., et al. (2003a). Closing the folding chamber of the eukaryotic chaperonin requires the
transition state of ATP hydrolysis. Cell, 113(3), 369–381.
Meyer, P., et al. (2003b). Structural and functional analysis of the middle segment of Hsp90:
Implications for ATP hydrolysis and client protein and Cochaperone interactions. Molecular
Cell, 11, 647–658.
Mickler, M., etal. (2009). The large conformational changes of Hsp90 are only weakly coupled to
ATP hydrolysis. Nature Structural and Molecular Biology, 16(3), 281–286.
Miemyk, J.(2017). The 70 kDa stress-related proteins as molecular chaperones. Trends in Plant
Science, 2(5), 180–187.
Minami, Y., & Minami, M. (1999). Hsc70/Hsp40 chaperone system mediates the Hsp90-dependent
refolding of rey luciferase. Genes to Cells, 4(12), 721–729.
Morange, M., & Bensaude, O. (1991). Heat shock increases turnover of 90 groups in HeLa cells.
Intl JBiol Stress, l(2), 359–362.
Morimoto, R.I., Tissieres, A., & Georgopoulos, C. (1994). Progress and perspectives on the biol-
ogy of heat shock proteins and molecular chaperones. Biol Heat Shock Prot Mol Chaperones,
1–30.
Morimoto, R.I., Kroeger, P.E., & Cotto, J.J. (1996). The transcriptional regulation of heat shock
genes: A plethora of heat shock factors and regulatory conditions. In U.Feige et al. (Eds.),
Stress-Inducible Cellular Responses (Vol. 77, pp.139–163). Basel: Birkhäuser Basel EXS.
Muchowski, P. J., et al. (1999). ATP and the core “alpha-Crystallin” domain of the small heat-
shock protein alphaB-crystallin. The Journal of Biological Chemistry, 274(42), 30190–30195.
Mujacic, M., Bader, M.W., & Baneyx, F. (2004). Escherichia coli Hsp31 functions as a holding
chaperone that cooperates with the DnaK-DnaJ-GrpE system in the management of protein
misfolding under severe stress conditions. Molecular Microbiology, 51, 849–859.
Nair, S.C., etal. (1997). Molecular cloning of human FKBP51 and comparisons of immunophilin
interactions with Hsp90 and progesterone receptor. Molecular and Cellular Biology, 17(2),
594–603.
B. K. Chatterjee et al.
1280
1281
1282
1283
1284
1285
1286
1287
1288
1289
1290
1291
1292
1293
1294
1295
1296
1297
1298
1299
1300
1301
1302
1303
1304
1305
1306
1307
1308
1309
1310
1311
1312
1313
1314
1315
1316
1317
1318
1319
1320
1321
1322
1323
1324
1325
1326
1327
1328
1329
1330
1331
1332
215
Nishikawa, S., & Endo, T. (1997). The yeast Jem1p is a DnaJ-like protein of the endoplasmic retic-
ulum membrane required for nuclear fusion. The Journal of Biological Chemistry, 272(20),
12889–12892.
Normington, K., Kohno, K., Kozutsumi, Y., Gething, M.J., & Sambrook, J.(1989). S.Cerevisiae
encodes an essential protein homologous in sequence and function to mammalian BiP. Cell,
57(7), 1223–1236.
Obermann, W.M. J., etal. (1998). In vivo function of Hsp90 is dependent on ATP binding and ATP
hydrolysis. The Journal of Cell Biology, 143(4), 901–910.
Otto, H., et al. (2005). The chaperones MPP11 and Hsp70L1 form the mammalian ribosome-
associated complex. Proceedings of the National Academy of Sciences, 102(29), 10064–10069.
Panaretou, B., etal. (1998). ATP binding and hydrolysis are essential to the function of the Hsp90
molecular chaperone invivo. The EMBO Journal, 17(16), 4829–4836.
Panaretou, B., etal. (2002). Activation of the ATPase activity of Hsp90 by the stress-regulated
cochaperone Aha1. Molecular Cell, 10(6), 1307–1318.
Patricia Hernández, M., Chadli, A., & Toft, D. O. (2002). HSP40 binding is the rst step in
the HSP90 chaperoning pathway for the progesterone receptor. The Journal of Biological
Chemistry, 277(14), 11873–11881.
Paul, S., Singh, C., Mishra, S., & Chaudhuri, T. K. (2007). The 69-kDa Escherichia Coli
Maltodextrin Glucosidase does not get encapsulated underneath GroES and folds through trans
mechanism during GroEL/GroES assisted folding. The FASEB Journal, 21(11), 2874–2885.
Pearl, L.H. (2016). Review: The HSP90 molecular chaperone- an enigmatic ATPase. Biopolymers,
105(8), 594–607.
Pearl, L. H., & Prodromou, C. (2006). Structure and mechanism of the Hsp90 molecular chaperone
machinery. Annual Review of Biochemistry, 75(1), 271–294.
Pelham, H. R. B. (1986). Speculations on the functions of the major heat shock and glucose-
regulated proteins. Cell, 46(7), 959–961.
Pellecchia, M., Montgomery, D. L., Stevens, S. Y., Vander Kooi, C. W., Feng, H., Gierasch, L.
M., & Zuiderweg, E. R. P. (2000). Structural insights into substrate binding by the molecular
chaperone DnaK. Nature Structural Biology, 7, 298–303.
Pirkkala, L., Nykänen, P., & Sistonen, L. (2001). Roles of the heat shock transcription factors in
regulation of the heat shock response and beyond. The FASEB Journal, 15(7), 1118–1131.
Pirkl, F., & Buchner, J.(2001). Functional analysis of the Hsp90-associated human peptidylprolyl
cis/trans isomerases FKBP51, FKBP52 and Cyp40. Journal of Molecular Biology, 308(4),
795–806.
Prändl, R., etal. (1998). HSF3, a new heat shock factor from Arabidopsis Thaliana, derepresses
the heat shock response and confers thermotolerance when overexpressed in transgenic plants.
Molecular & General Genetics, 258(3), 269–278.
Prodromou, C., et al. (1999). Regulation of Hsp90 ATPase activity by tetratricopeptide repeat
(TPR)-domain co-chaperones. The EMBO Journal, 18(3), 754–762.
Prodromou, C., etal. (2000). The ATPase cycle of Hsp90 drives a molecular ` clamp via transient
dimerization of the N-terminal domains. The EMBO Journal, 19(16), 4383–4392.
Puri, S., & Chaudhuri, T. K. (2017). Folding and unfolding pathway of chaperonin GroEL
monomer and elucidation of thermodynamic parameters. International Journal of Biological
Macromolecules, 96, 713–726.
Rajaraman, K., etal. (2001). Interaction of human recombinant alphaA- and alphaB-crystallins
with early and late unfolding intermediates of citrate synthase on its thermal denaturation.
FEBS Letters, 497(2–3), 118–123.
Rani, S., Srivastava, A., Kumar, M., & Goel, M. (2016). CrAgDb- a database of annotated chaper-
one repertoire in archaeal genomes. FEMS Microbiology Letters, 363(6), 1–6.
Ranson, N. A., Dunster, N. J., Burston, S. G., & Clarke, A. R. (1995). Chaperonins can catalyse
the reversal of early aggregation steps when a protein misfolds. Journal of Molecular Biology,
250(5), 581–586.
8 Molecular Chaperones in Cellular Stress Response
1333
1334
1335
1336
1337
1338
1339
1340
1341
1342
1343
1344
1345
1346
1347
1348
1349
1350
1351
1352
1353
1354
1355
1356
1357
1358
1359
1360
1361
1362
1363
1364
1365
1366
1367
1368
1369
1370
1371
1372
1373
1374
1375
1376
1377
1378
1379
1380
1381
1382
1383
1384
216
Ratzke, C., etal. (2010). Dynamics of heat shock protein 90 C-terminal dimerization is an impor-
tant part of its conformational cycle. Proceedings of the National Academy of Sciences,
107(37), 16101–16106.
Ray, D., etal. (2016). Plant stress response: Hsp70in the spotlight. In A.A. A.Asea, P.Kaur, &
S.K. Calderwood (Eds.), Heat shock Prot plants, Heat shock proteins (Vol. 10, pp.123–147).
Cham: Springer.
Retzlaff, M., etal. (2010). Asymmetric activation of the Hsp90 dimer by its Cochaperone Aha1.
Molecular Cell, 37(3), 344–354.
Rice, L.M., etal. (2008). Article multiple conformations of E .coli Hsp90in solution : Insights into
the conformational dynamics of Hsp90. Structure, 16(5), 755–765.
Richter, K., Walter, S., & Buchner, J.(2004). The co-chaperone Sba1 connects the ATPase reac-
tion of Hsp90 to the progression of the chaperone cycle. Journal of Molecular Biology, 342(5),
1403–1413.
Richter, K., Haslbeck, M., & Buchner, J.(2010). The heat shock response: Life on the verge of
death. Molecular Cell, 40(2), 253–266.
Ritossa, F. (1996). Discovery of the heat shock response. Cell Stress and Chaperones, 1(2), 97–98.
Roe, S.M., etal. (1999). Structural basis for inhibition of the Hsp90 molecular chaperone by the
antitumor antibiotics Radicicol and Geldanamycin. Journal of Medicinal Chemistry, 42(2),
260–266.
Roe, S. M., et al. (2004). The mechanism of Hsp90 regulation by the protein kinase-specic
Cochaperone p50cdc37. Cell, 116(1), 87–98.
Rowley, N., Prip-Buus, C., Westermann, B., Brown, C., Schwarz, E., Barreil, B., & Neupertt, W.
(1994). Mdj1p, a novel chaperone of the DnaJ family, is involved in mitochondrial biogenesis
and protein folding. Cell, 77(2), 249–259.
Rudiger, S., Buchberger, A., & Bukau, B. (1997). Interaction of Hsp70 chaperones with substrates.
Nature, 4(5), 686–689.
Sadler, I., Chiang, A., Kurihara, T., Rothblatt, J., Way, J., & Silver, P. (1989). A yeast gene impor-
tant for protein assembly into the endoplasmic reticulum and the nucleus has homology to
DnaJ, an Escherichia coli heat shock protein. The Journal of Cell Biology, 109(6), 2665–2675.
Sakikawa, C., Taguchi, H., Makino, Y., & Yoshida, M. (1999). On the maximum size of proteins
to stay and fold in the cavity of GroEL underneath GroES. Journal of Biological Chemistry,
274(30), 21251–21256.
Santoro, N., Johansson, N., & Thiele, D.J. (1998). Heat shock element architecture is an important
determinant in the temperature and transactivation domain requirements for heats hock tran-
scription factor. Molecular and Cellular Biology, 18(11), 6340–6352.
Sarge, K.D., Murphy, S.P., & Morimoto, R.I. (1993). Activation of heat shock gene transcrip-
tion by heat shock factor 1 involves oligomerization, acquisition of DNA-binding activity, and
nuclear localization and can occur in the absence of stress. Molecular and Cellular Biology,
13(3), 1392–1407.
Sarkar, N.K., Kundnani, P. and Grover, A. (2013). Functional analysis of Hsp70 superfamily
proteins of rice (Oryza Sativa). Cell Stress Chaperones. Dordrecht: Springer Netherlands 18
(4), pp.427–437.
Sato, S., Fujita, N., & Tsuruo, T. (2000). Modulation of Akt kinase activity by binding to Hsp90.
Proceedings of the National Academy of Sciences, 97(20), 10832–10837.
Scherrer, C., etal. (1990). Structural and functional reconstitution of the glucorcorticoid receptor-
Hsp90 complex. The Journal of Biological Chemistry, 265(35), 21397–21400.
Scheuer, C., etal. (2000). Structure of TPR domain-peptide complexes: Critical elements in the
assembly of the Hsp70-Hsp90 multichaperone machine. Cell, 101(2), 199–210.
Schirmer, E.C., Glover, J.R., Singer, M.A., & Lindquist, S. (1996). HSP100/Clp proteins: A com-
mon mechanism explains diverse functions. Trends in Biochemical Sciences, 21(8), 289–296.
Schlenstedt, G., Harris, S., Risse, B., Lill, R., & Silver, P.A. (1995). A yeast DnaJ homologue,
Scj1p, can function in the endoplasmic reticulum with BiP/ Kar2p via a conserved domain that
species interactions with Hsp70s. The Journal of Cell Biology, 129(4), 979–988.
B. K. Chatterjee et al.
1385
1386
1387
1388
1389
1390
1391
1392
1393
1394
1395
1396
1397
1398
1399
1400
1401
1402
1403
1404
1405
1406
1407
1408
1409
1410
1411
1412
1413
1414
1415
1416
1417
1418
1419
1420
1421
1422
1423
1424
1425
1426
1427
1428
1429
1430
1431
1432
1433
1434
1435
1436
1437
217
Schmitt, M., Neupert, W., & Langer, T. (1995). Hsp78, a Clp homologue within mitochondria,
can substitute for chaperone functions of mt-hsp70. The EMBO Journal, 14(14), 3434–3444.
Schuster, G., etal. (1988). Evidence for protection by heat-shock proteins against photoinhibition
during heat-shock. The EMBO Journal, 7(1), 1–6.
Scroggins, B.T., & Neckers, L. (2007). Post-translational modication of heat-shock protein 90:
Impact on chaperone function. Expert Opinion on Drug Discovery, 2(10), 1403–1414.
Sefton, B.M., Beemon, K., & Hunter, T. (1978). Comparison of the expression of the src gene of
Rous sarcoma virus invitro and invivo. Journal of Virology, 28(3), 957–971.
Sherman, M., & Goldberg, A.L. (1994). Heat shock induced phosphorylation of GroEL alters its
binding and dissociation from unfolded protein. The Journal of Biological Chemistry, 269(50),
31479–31483.
Shi, Y., Mosser, D.D., & Morimoto, R.I. (1998). Molecular chaperones as HSF1-specic tran-
scriptional repressors. Genes and Development, 12(5), 654–666.
Sichting, M., Mokranjac, D., Azem, A., Neupert, W., & Hell, K. (2005). Maintenance of structure
and function of mitochondrial Hsp70 chaperones requires the chaperone Hep1. The EMBO
Journal, 24(5), 1046–1056.
Siegert, R., etal. (2000). Structure of the molecular chaperone Prefoldin. Cell, 103(4), 621–632.
Silberstein, S., Schlenstedt, G., Silver, P.A., & Gilmore, R. (1998). A role for the DnaJ homo-
logue Scj1p in protein folding in the yeast endoplasmic reticulum. The Journal of Cell Biology,
143(4), 921–933.
Siligardi, G., etal. (2002). Regulation of Hsp90 ATPase activity by the co-chaperone Cdc37p/p50
cdc37. The Journal of Biological Chemistry, 277(23), 20151–20159.
Silverstein, A.M., et al. (1997). Protein phosphatase 5 is a major component of glucocorticoid
receptor°hsp90 complexes with properties of an FK506-binding immunophilin. The Journal of
Biological Chemistry, 272(26), 16224–16230.
Smith, D.F. (1993). Dynamics of heat shock protein 90-progesterone receptor binding and the
disactivation loop model for steroid receptor complexes. Molecular Endocrinology, 7(11),
1418–1429.
Smith, etal. (1999). Lon and Clp family proteases and chaperones share homologous substrate-
recognition domains. Proceedings of the National Academy of Sciences, 96, 6678–6682.
Smith, B.J., & Yaffe, M.P. (1991). A mutation in the yeast heat-shock factor gene causes tempera-
ture sensitive defects in both mitochondrial protein import and the cell cycle. Molecular and
Cellular Biology, 11(5), 2647–2655.
Smith, D.F., etal. (1992). Assembly of progesterone receptor with heat shock proteins and receptor
activation are ATP mediated events. The Journal of Biological Chemistry, 267(2), 1350–1356.
Spiess, C., etal. (2004). Mechanism of the eukaryotic chaperonin: Protein folding in the chamber
of secrets. Trends in Cell Biology, 14(11), 598–604.
Stebbins, C.E., etal. (1997). Crystal structure of an Hsp90– Geldanamycin complex: Targeting of
a protein chaperone by an antitumor agent. Cell, 89(2), 239–250.
Stromer, T., Fischer, E., Richter, K., Haslbeck, M., & Buchner, J.(2004). Analysis of the regula-
tion of the molecular chaperone Hsp26 by temperature- induced dissociation: The N-terminal
domain is important for oligomer assembly and the binding of unfolding proteins. The Journal
of Biological Chemistry, 279(12), 11222–11228.
Sun, W., et al. (2001). At-HSP17.6A, encoding a small heat-shock protein in Arabidopsis, can
enhance osmotolerance upon overexpression. The Plant Journal, 27(5), 407–415.
Taylor, R.P., & Benjamin, I.J. (2005). Small heat shock proteins: A new classication scheme in
mammals. Journal of Molecular and Cellular Cardiology, 38(3), 433–444.
Teter, S. A., et al. (1999). Polypeptide ux through bacterial Hsp70. Cell, 97(6), 755–765.
Thomas, J.G., & Baneyx, F. (2000). ClpB and HtpG facilitate de novo protein folding in stressed
Escherichia Coli cells. Molecular Microbiology, 36(6), 1360–1370.
Thulasiraman, V., Yang, C., & Frydman, J. (1999). In vivo newly translated polypeptides are
sequestered in a protected folding environment. The EMBO Journal, 18(1), 85–95.
Tissiéres, A., Mitchell, H. K., & Tracy, U.M. (1974). Protein synthesis in salivary glands of
Drosophila Melanogaster: Relation to chromosome puffs. Journal of Molecular Biology, 84(3),
389–398.
8 Molecular Chaperones in Cellular Stress Response
1438
1439
1440
1441
1442
1443
1444
1445
1446
1447
1448
1449
1450
1451
1452
1453
1454
1455
1456
1457
1458
1459
1460
1461
1462
1463
1464
1465
1466
1467
1468
1469
1470
1471
1472
1473
1474
1475
1476
1477
1478
1479
1480
1481
1482
1483
1484
1485
1486
1487
1488
1489
1490
1491
1492
218
Trachootham, D., et al. (2008). Redox regulation of cell survival. Antioxidants and Redox
Signaling, 10(8), 1343–1374.
Vabulas, R.M., et al. (2010). Protein folding in the cytoplasm and the heat shock response. Cold
Spring Harbor Perspectives in Biology, 1–18.
Vainberg, I. E., etal. (1998). Prefoldin, a chaperone that delivers unfolded proteins to cytosolic
chaperonin. Cell, 93(5), 863–873.
Vanghele, M., & Ganea, E. (2010). The role of bacterial molecular chaperones in pathogen survival
within the host. Rom. Journal of Biochemistry, 47(1), 87–100.
Wagner, I., Arlt, H., van Dyck, L., Langer, T., & Neupert, W. (1994). Molecular chaperones coop-
erate with PIM1 protease in the degradation of mis- folded proteins in mitochondria. The
EMBO Journal, 13(21), 5135–5145.
Waldmann, T., etal. (1995). The Thermosome of Thermoplasma acidophilum and its relationship
to the eukaryotic Chaperonin TRiC. European Journal of Biochemistry, 227(3), 848–856.
Walker, A.I., etal. (1985). Double-stranded DNA induces the phosphorylation of several proteins
including the 90000 mol. Wt. heat-shock protein in animal cell extracts. The EMBO Journal,
4(1), 139–145.
Wandinger, S. K., et al. (2006). The phosphatase Ppt1 is a dedicated regulator of the molecular
chaperone Hsp90. The EMBO Journal, 25(2), 367–376.
Wang, W., etal. (2004). Role of plant heat-shock proteins and molecular chaperones in the abiotic
stress response. Trends in Plant Science, 9(5), 244–252.
Wegele, H., Müller, L. and Buchner, J.(2004). Hsp70 and Hsp90—a relay team for protein folding.
Reviews of Physiology, Biochemistry and Pharmacology, Springer, Berlin, Heidelberg 151,
pp.1–44.
Weissman, J. S., et al. (1995). Mechanism of GroEL action: Productive release of polypeptide from
a sequestered position under GroES. Cell, 83(4), 577–587.
Welker, S., Rudolph, B., Frenzel, E., Hagn, F., Liebisch, G., Schmitz, G., Scheuring, J., Kerth,
A., Blume, A., Weinkauf, S., Haslbeck, M., Kessler, H., & Buchner, J.(2010). Hsp12 is an
intrinsically unstructured stress protein that folds upon membrane association and modulates
membrane function. Molecular Cell, 39(4), 507–520.
Yamada, K., etal. (2007). Cytosolic HSP90 regulates the heat shock response that is responsible
for heat acclimation in Arabidopsis Thaliana. The Journal of Biological Chemistry, 282(52),
37794–37804.
Yang, D., Yea, X., & Lorimer, G.H. (2013). Symmetric GroEL: GroES2 complexes are the protein-
folding functional form of the chaperonin nanomachine. Proceedings of the National Academy
of Sciences, 110(46), 4298–4305.
Young, J. C., Moare, I., & Hartl, F.U. (2001). Hsp90. The Journal of Cell Biology, 154(2),
267–274.
Young, J. C., Barral, J. M., & Hartl, F. U. (2003). More than folding: Localized functions of cyto-
solic chaperones. Trends in Biochemical Sciences, 28(10), 541–547.
Yu, A., etal. (2015). Roles of Hsp70s in stress responses of microorganisms, plants, and animals.
BioMed Res Intl, 2015(2015), 1–8.
Zhang, W., et al. (2004). Biochemical and structural studies of the interaction of Cdc37 with
Hsp90. Journal of Molecular Biology, 340(4), 891–907.
Zhao, R., etal. (2005). Navigating the chaperone network: An integrative map of physical and
genetic interactions mediated by the hsp90 chaperone. Cell, 120(5), 715–727.
Zhao, R., & Houry, W. A. (2007). Molecular interaction network of the Hsp90 chaperone system,
BT - molecular aspects of the stress response: Chaperones, membranes and networks. In P.
Csermely & L. Vígh (Eds.), (pp. 27–36). NY: Springer.
Zhu, X., et al. (1996). Structural analysis of substrate binding by the molecular chaperone DnaK.
Science, 272(5268), 1606–1614.
Zou, J., etal. (1998). Repression of heat shock transcription factor HSF1 activation by HSP90
(HSP90 complex) that forms a stress-sensitive complex with HSF1. Cell, 94(4), 471–480.
B. K. Chatterjee et al.
1493
1494
1495
1496
1497
1498
1499
1500
1501
1502
1503
1504
1505
1506
1507
1508
1509
1510
1511
1512
1513
1514
1515
1516
1517
1518
1519
1520
1521
1522
1523
1524
1525
1526
1527
1528
1529
1530
1531
1532
1533
1534
1535
1536
1537
1538
1539
1540
1541
1542
1543
1544
Author Query
Chapter No.: 8 0003407372
Query Details Required Author’s Response
AU1 Please check sentence starting “Although they are present at...
for completeness.
... The abovementioned studies attempted to initiate molecular processes that might inhibit the onset of apoptotic pathways during cryopreservation. The mild sub-lethal stress can cause an increased expression of particular proteins from the chaperone family that is essential for cellular activities like fatty acid and energy metabolism, protein equilibration, redox control, and defective protein elimination (Chatterjee et al. 2018;Verghese et al. 2012;Zhang et al. 2008). The existence of heat shock proteins (molecular chaperones) on the surface of the sperm is widely recognized, which may maintain sperm motility and fertility during stress conditions Pribenszky et al. 2011). ...
... The hypothesis of sub-lethal stress proved useful as it stimulates phosphorylation leading to increased heat shock proteins (HSPs) production (Pribenszky et al. 2011(Pribenszky et al. , 2010. These HSPs aid to stabilize and repairing sperm DNA and enhance sperm resistance during stressful situations that ultimately help to reinstate sperm homeostasis (Chatterjee et al. 2018;Pribenszky et al. 2010). ...
Article
Full-text available
A novel strategy, focused on the induction of sub-lethal oxidative stress to optimize sperm cryosurvival, has been used before cryopreservation. The present study compared the effect of preconditioning with various concentrations of nitric oxide-donor (sodium nitroprusside, SNP) and peroxynitrite-generator (3-morpholinosydnonimine, SIN-1) on in vitro sperm functions and lipid peroxidation status (LPO) of cryopreserved Karan-Fries (KF) crossbred bull semen. To optimize the concentration of additives, spermatozoa obtained from 36 ejaculates were supplemented with different concentrations of SNP (0.01, 0.05, 0.1 μM) and SIN-1 (80, 160, 200 μM) versus control in the extender. The post-freezing sperm motility and viability were greater (p < 0.05) in 0.1 μM SNP and 80 μM SIN-1 in comparison to other concentrations used. Furthermore, the spermatozoa obtained from 48 ejaculates were supplemented with 0.1 μM SNP and 80 μM SIN-1 in the extender. A significant increase (p < 0.05) was observed in progressive motility, viability and membrane integrity in SNP and SIN-1 treated extender at 24 h, 15 days, and 2-month post-cryopreservation (PC) periods. There was no significant difference in sperm abnormality in the extended groups and the control group. The seminal plasma of SNP-treated extender had less (p < 0.05) lipid peroxidation as compared to SIN-1 treated and control groups. In post-thaw semen, both SNP and SIN-1 showed a higher (p < 0.05) proportion of acrosome intact (FITC-PNA) sperm with a greater decrease (p < 0.05) in membrane scrambling and lipid peroxidation. SNP and SIN-1 improved (p < 0.05) the proportion of sperm with higher mitochondrial membrane potential (Δψm) as compared to the control. In conclusion, it seems that the preconditioning of SNP and SIN-1 at lower doses may have beneficial effects on post-thawed crossbred bull sperm quality.
... The mechanism for the recovery of proteins from aggregation often requires the assistance of another ATP-dependent chaperone system. The HSP70 family solubilizes the aggregated protein and extracts it in a process that can be repeated with the aid of a specific HSP family of genes [13]. This review focuses on recent discoveries of molecular and cellular mechanisms of HSP70 that govern the tolerance of plants in unfavorable environmental conditions. ...
... They may be regularly expressed in plants constitutively but their expression is regulated by using various environmental conditions which include heat and salt [47]. Research have verified their presence in cytosol, mitochondria and chloroplast and play vital function in remodeling machines that participate in maintaining the integrity of the mobile proteome via facilitating protein reworking, disaggregation, reactivation or degradation of misfolded and inactive protein [13]. ...
Chapter
Full-text available
Heat stress is considered to induce a wide range of physiological and biochemical changes that cause severe damage to plant cell membrane, disrupt protein synthesis, and affect the efficiency of photosynthetic system by reducing the transpiration due to stomata closure. A brief and mild heat shock is known to induce acquired thermo tolerance in plants that is associated with concomitant production of heat shock proteins’ (HSPs) gene family including HSP70. The findings from different studies by use of technologies have thrown light on the importance of HSP70 to heat, other abiotic stresses and environmental challenges in desserts. There is clear evidence that under heat stress, HSP70 gene stabilized the membrane structure, chlorophyll and water breakdown. It was also found that under heat stress, HSP70 decreased the malondialdehyde (MDA) content and increased the production of superoxide dismutase (SOD) and peroxidase (POD) in transgenic plants as compared to non-transgenic plants. Some reactive oxygen species (ROS) such as superoxide, hydrogen peroxide and hydroxyl radical are also synthesized and accumulated when plants are stressed by heat. Hence HSP70 can confidently be used for transforming a number of heat tolerant crop species.
... As shown in Table 4, majorly 6 genes showed direct homology with HSPs. HSPs act as chaperones regulating the folding, accumulation, localization, and degradation of normal proteins and combating the damaging effects of heat and other stress on plants [63][64][65]. Mishra et al. [66] have also reported 6 genes (viz., HSFA6e, HSP90, HSP17, MAPK, CDPK, and SOD) associated to heat stress in wheat. Among which, HSP17 has shown manifold change in their expression in heat stress as compared to the control condition [66]. ...
Article
Full-text available
Climate change has become a major source of concern, particularly in agriculture, because it has a significant impact on the production of economically important crops such as wheat, rice, and maize. In the present study, an attempt has been made to identify differentially expressed heat stress-responsive long non-coding RNAs (lncRNAs) in the wheat genome using publicly available wheat transcriptome data (24 SRAs) representing two conditions, namely, control and heat-stressed. A total of 10,965 lncRNAs have been identified and, among them, 153, 143, and 211 differentially expressed transcripts have been found under 0 DAT, 1 DAT, and 4 DAT heat-stress conditions, respectively. Target prediction analysis revealed that 4098 lncRNAs were targeted by 119 different miRNA responses to a plethora of environmental stresses, including heat stress. A total of 171 hub genes had 204 SSRs (simple sequence repeats), and a set of target sequences had SNP potential as well. Furthermore, gene ontology analysis revealed that the majority of the discovered lncRNAs are engaged in a variety of cellular and biological processes related to heat stress responses. Furthermore, the modeled three-dimensional (3D) structures of hub genes encoding proteins, which had an appropriate range of similarity with solved structures, provided information on their structural roles. The current study reveals many elements of gene expression regulation in wheat under heat stress, paving the way for the development of improved climate-resilient wheat cultivars.
... Temperature stress causes enzymatic responses, e.g., protease production has been observed in thermophilic and thermotolerant organisms (Ordaz-Hernández et al., 2016;Walker and White, 2017). Temperature stress is also related to protein glycosylation that stabilizes protease attacks and protein secretion (Chatterjee et al., 2018). Therefore, in this work, we determined the changes in the production of laccase and protease by Fomes sp. ...
... Klebsiella pneumoniae is a multidrug resistant (MDR) bacterium that causes septicemia, lung infections, liver abscesses and urinary tract infection, which is considered as an opportunistic pathogen in both hospital and community-acquired diseases [1]. The adhesion properties of K. pneumoniae are generally mediated by type-1 and type-3 fimbria, which are consists of globular proteins that enables K. pneumoniae attach to the host cells as the first step in infectious process [2]. Type-1 fimbrial genes are essential for colonization, invasion and persistence of K. pneumoniae in the UTI. ...
Article
Full-text available
Klebsiella pneumoniae is a member of coliform bacteria that causes wide ranges of infections including circulatory, respiratory system, urinary tract infections (UTIs), and wounds infections. This study aimed to find the correlation between type 1 and 3 fimbrial genes expression with multidrug resistance (MDR) K. pneumoniae isolates towards antibiotics. Sixty clinical isolates of K. pneumoniae were collected from three main types of samples including blood, wound and burn swabs, and urine samples. The diagnosis was confirmed by VITEK-2 system and 16s rRNA housekeeping gene. The antibiotic sensitivity profile included 16 antimicrobial agents, with extended-spectrum beta-lactamase production. PCR technique was applied to detect four genes of type-1 fimbrial genes: (usher-1, chaperon-1L, chaperon-1S, and fim-H1), beside type-3 fimbrial genes: (MrkA, MrkB, MrkC, MrkD, and MrkF). The results showed that K. pneumoniae isolates were hundred percent (100%) resistant towards ampicillin, no resistance (0%) was recorded towards tigecycline and ertapenem, while the percentages of resistance for ceftazidem, cefepime, amikacine, and amipenem were 15%, 20%, 51.7%, and 50% respectively, and the isolates showed about (13-71%) resistance to the rest antimicrobials agents. The production of extended-spectrum beta-lactamase was in 40 (66.67%) of total the 60 isolates. There was no relationship according to the statistical analysis between the types of specimen with the antibiotic resistance rates. For fimbriae type 1 genes, the largest occurrence (90%) was reported in Chaperon-1S gene and the lowest one was in Usher-1 gene (56.6%), while it was above 70% in Chaperon-1L gene and fim-H1 gene of the total K. pneumoniae isolates. The percentages of type 3 genes MrkA, MrkB, MrkC, MrkD, and MrkF were: 28.3, 76.6, 85, 51.6, and 63.3% respectively. The type-1 fimbrial genes had no significant correlation among them; however, the type-3 fimbrial genes had significance in their presence at 0.01 and 0.05 levels, as they are located on the same Mrk operon. Finally, the correlation between type 1 and 3 fimbrial genes with the type of specimen and antibiotic resistance was not significant at all.
Article
Full-text available
Background Heat stress (HS) poses significant threats to the sustainability of livestock production. Genetically improving heat tolerance could enhance animal welfare and minimize production losses during HS events. Measuring phenotypic indicators of HS response and understanding their genetic background are crucial steps to optimize breeding schemes for improved climatic resilience. The identification of genomic regions and candidate genes influencing the traits of interest, including variants with pleiotropic effects, enables the refinement of genotyping panels used to perform genomic prediction of breeding values and contributes to unraveling the biological mechanisms influencing heat stress response. Therefore, the main objectives of this study were to identify genomic regions, candidate genes, and potential pleiotropic variants significantly associated with indicators of HS response in lactating sows using imputed whole-genome sequence (WGS) data. Phenotypic records for 18 traits and genomic information from 1,645 lactating sows were available for the study. The genotypes from the PorcineSNP50K panel containing 50,703 single nucleotide polymorphisms (SNPs) were imputed to WGS and after quality control, 1,622 animals and 7,065,922 SNPs were included in the analyses. Results A total of 1,388 unique SNPs located on sixteen chromosomes were found to be associated with 11 traits. Twenty gene ontology terms and 11 biological pathways were shown to be associated with variability in ear skin temperature, shoulder skin temperature, rump skin temperature, tail skin temperature, respiration rate, panting score, vaginal temperature automatically measured every 10 min, vaginal temperature measured at 0800 h, hair density score, body condition score, and ear area. Seven, five, six, two, seven, 15, and 14 genes with potential pleiotropic effects were identified for indicators of skin temperature, vaginal temperature, animal temperature, respiration rate, thermoregulatory traits, anatomical traits, and all traits, respectively. Conclusions Physiological and anatomical indicators of HS response in lactating sows are heritable but highly polygenic. The candidate genes found are associated with important gene ontology terms and biological pathways related to heat shock protein activities, immune response, and cellular oxidative stress. Many of the candidate genes with pleiotropic effects are involved in catalytic activities to reduce cell damage from oxidative stress and cellular mechanisms related to immune response.
Chapter
Full-text available
Molecular chaperones are a group of heat shock proteins, essential for the maintenance of cellular proteostasis. Genetic and acquired defects in their structure and function due to mutations, truncations, oxidation, altered expression and post-translational modifications are responsible for various protein conformational diseases collectively known as chaperonopathies. Major chaperonopathies include neurodegenerative diseases, myopathies and diabetes. In addition, the chaperone families Hsp60, Hsp70 and Hsp90 stabilize several client proteins implicated in many types of cancers. The complexities of the cellular protein network make therapeutic interventions particularly challenging in the treatment of these often-deadly diseases. The development of therapeutic strategies for various chaperonopathies is a pressing need and has been a subject of active research. While some strategies target pathogenic molecular chaperones with the help of small molecule inhibitors, others make use of chaperones to mitigate proteinopathies. Development of specific inhibitors against chaperones, use of pharmacological chaperones, and over-expression of different chaperone combinations including Hsp70/Hsp40/Hsp90, chaperonins, and sHsp are promising approaches to overcome these diseases. In this article, we focus on the various defects in molecular chaperones, their associated diseases, and potential therapeutic approaches that may help alleviate these chaperonopathies.
Chapter
Molecular chaperones are a group of heat shock proteins, essential for the maintenance of cellular proteostasis. Genetic and acquired defects in their structure and function due to mutations, truncations, oxidation, altered expression and post-translational modifications are responsible for various protein conformational diseases collectively known as chaperonopathies. Major chaperonopathies include neurodegenerative diseases, myopathies and diabetes. In addition, the chaperone families Hsp60, Hsp70 and Hsp90 stabilize several client proteins implicated in many types of cancers. The complexities of the cellular protein network make therapeutic interventions particularly challenging in the treatment of these often-deadly diseases. The development of therapeutic strategies for various chaperonopathies is a pressing need and has been a subject of active research. While some strategies target pathogenic molecular chaperones with the help of small molecule inhibitors, others make use of chaperones to mitigate proteinopathies. Development of specific inhibitors against chaperones, use of pharmacological chaperones, and over-expression of different chaperone combinations including Hsp70/Hsp40/Hsp90, chaperonins, and sHsp are promising approaches to overcome these diseases. In this article, we focus on the various defects in molecular chaperones, their associated diseases, and potential therapeutic approaches that may help alleviate these chaperonopathies.
Article
Full-text available
Heat shock proteins (HSPs) are a large molecular chaperone family classified by their molecular weights, including HSP27, HSP40, HSP60, HSP70, HSP90, and HSP110. HSPs are likely to have antiapoptotic properties and participate actively in various processes such as tumor cell proliferation, invasion, metastases, and death. In this review, we discuss comprehensively the functions of HSPs associated with the progression of colorectal cancer (CRC) and metastasis and resistance to cancer therapy. Taken together, HSPs have numerous clinical applications as biomarkers for cancer diagnosis and prognosis and potential therapeutic targets for CRC and its related metastases.
Article
Small heat shock proteins (sHSPs) constitute a class of molecular chaperones, which are evolutionarily conserved yet diverse group of molecules, rapidly produced in response to stress. In this study, we sought to identify plant sHSPs, especially chaperonin 10 (Cpn10) family members in major evolutionary lineages, and determine their biological significance. Multiple sequence alignment of Cpn10 domains revealed divergent amino acids as well as conserved sites. Phylogenetic tree depicted the diversification and expansion of Cpn10 gene family. During the process of evolution, the Ka/Ks ratio of orthologous and paralogous pairs was <1, suggesting their evolutionary convergence and biological relevance. Functional annotations demonstrated that Cpn10 are involved in protein folding, regulation of metabolic processes and abiotic stress responses. Furthermore, subcellular localization prediction revealed that Cpn10 proteins are localized in multiple compartments, indicating a critical cell-coordinated defense. In-silico gene expression analysis exhibited their expression in most tissues examined, implying functional redundancy. Interactome analysis illustrated their interaction with chloroplast and mitochondrial genes, which are majorly involved in protein folding and assembly. The transcriptional regulation revealed their stress-responsive and distinct physiological roles. Our findings would contribute to new insights on the evolutionary history of Cpn10 gene family and the distinct biological roles.
Article
Full-text available
To better understand assembly mechanisms of progesterone receptor (PR) complexes, we have developed a cell-free system for studying PR interactions with the 90- and 70-kDa heat shock proteins (hsp90 and hsp70), and we have used this system to examine requirements for hsp90 binding to PR. Purified chick PR, free of hsp90 and immobilized on an antibody affinity resin, will rebind hsp90 in rabbit reticulocyte lysate when several conditions are met. These include: 1) absence of progesterone, 2) elevated temperature (30 degrees C), 3) presence of ATP, and 4) presence of Mg2+. We have obtained maximal hsp90 binding to receptor when lysate is supplemented with 3 mM MgCl2 and an ATP-regenerating system. ATP depletion of lysate by dialysis or by enzymatic means blocks hsp90 binding to PR; likewise, addition of EDTA to lysate blocks hsp90 binding, but binding is restored by the addition of excess Mg2+. Addition to lysate of monoclonal antibody against hsp70 inhibits hsp90 binding to PR and destabilizes preformed complexes. Stabilization of hsp90-receptor complexes also requires ATP, indicating that ATP and hsp70 are needed to form and to maintain hsp90 complexes. Hormone-dependent activation of reconstituted receptor complexes was also examined. The addition of progesterone to the reticulocyte lysate promotes dissociation of hsp90 and hsp70 from the receptor. This also appears to require ATP and dissociation is most efficient in the presence of an ATP-regenerating system. In conclusion, these studies indicate that PR-hsp90 complexes do not self-assemble; instead, assembly is probably a multistep process requiring ATP and other cellular factors.
Article
Full-text available
Heat stress results in misfolding and aggregation of cellular proteins. Heat shock proteins (Hsp) enable the cells to maintain proper folding of proteins, both in un-stressed as well as stressed conditions. Hsp70 genes encode for a group of highly conserved chaperone proteins across the living systems encompassing bacteria, plants, and animals. In the cellular chaperone network, Hsp70 family proteins interconnect other chaperones and play a dominant role in various cell processes. To assess the functionality of rice Hsp70 genes, rice genome database was analyzed. Rice genome contains 32 Hsp70 genes. Rice Hsp70 super-family genes are represented by 24 Hsp70 family and 8 Hsp110 family members. Promoter and transcript expression analysis divulges that Hsp70 superfamily genes plays important role in heat stress. Ssc1 (mitochondrial Hsp70 protein in yeast) deleted yeast show compromised growth at 37°C. Three mitochondrial rice Hsp70 sequences (i.e., mtHsp70-1, mtHsp70-2, and mtHsp70-3) complemented the Ssc1 mutation of yeast to differential extents. The information presented in this study provides detailed understanding of the Hsp70 protein family of rice, the crop species that is the major food for the world population.
Article
Full-text available
Small heat shock proteins are ubiquitous molecular chaperones that form the first line of defence against the detrimental effects of cellular stress. Under conditions of stress they undergo drastic conformational rearrangements in order to bind to misfolded substrate proteins and prevent cellular protein aggregation. Owing to the dynamic nature of small heat shock protein oligomers, elucidating the structural basis of chaperone action and oligomerization still remains a challenge. In order to understand the organization of sHSP oligomers, we have determined crystal structures of a small heat shock protein from Salmonella typhimurium in a dimeric form and two higher oligomeric forms: an 18-mer and a 24-mer. Though the core dimer structure is conserved in all the forms, structural heterogeneity arises due to variation in the terminal regions.
Article
Full-text available
The HSP90 molecular chaperone is involved in the activation and cellular stabilization of a range of 'client' proteins, of which oncogenic protein kinases and nuclear steroid hormone receptors are of particular biomedical significance. Work over the last two decades has revealed a conformational cycle critical to the biological function of HSP90, coupled to an inherent ATPase activity that is regulated and manipulated by many of the co-chaperones proteins with which it collaborates. Pharmacological inhibition of HSP90 ATPase activity results in degradation of client proteins in vivo, and is a promising target for development of new cancer therapeutics. Despite this, the actual function that HSP90s conformationally-coupled ATPase activity provides in its biological role as a molecular chaperone remains obscure. This article is protected by copyright. All rights reserved.
Chapter
Heat Shock Protein 70 (Hsp70) is an evolutionarily conserved family of proteins which carry out multiple cellular functions such as protein biogenesis, protection during stress, prevention of formation of protein aggregates, assistance in protein translocation and many others. Hsp70, being the major cytoprotective molecular chaperone, plays a crucial role in protecting against a stunning array of stresses and in the re-establishment of cellular homeostasis. This book chapter gives an overview of the multifaceted Hsp70s in plants, with special emphasis on their association with plant response to various stress conditions and eventually, stress acclimation. The contribution of plant stress-responsive proteomics studies towards putting Hsp in the spotlight has also been brought forth. The road ahead is to decipher the underlying mechanisms of Hsp70-mediated multiple cross tolerance, that is likely to lead to new strategies to enhance crop tolerance to environmental stress.
Article
The Escherichia coli chaperonin GroEL is an essential molecular chaperone that mediates protein folding in association with its cofactor, GroES. It is widely accepted that GroEL alternates the GroES-sealed folding-active rings during the reaction cycle. In other words, an asymmetric GroEL–GroES complex is formed during the cycle, whereas a symmetric GroEL–(GroES)2 complex is not formed. However, this conventional view has been challenged by the recent reports indicating that such symmetric complexes can be formed in the GroEL–GroES reaction cycle. In this review, we discuss the studies of the symmetric GroEL–(GroES)2 complex, focusing on the molecular mechanism underlying its formation. We also suggest that GroEL can be involved in two types of reaction cycles (asymmetric or symmetric) and the type of cycle used depends on the concentration of non-native substrate proteins.