You are on page 1of 7

SCIENCE’S COMPASS ● REVIEW

REVIEW: PROTEIN FOLDING

Molecular Chaperones in the Cytosol: from


Nascent Chain to Folded Protein
F. Ulrich Hartl* and Manajit Hayer-Hartl

phobic forces and interchain hydrogen bond-


Efficient folding of many newly synthesized proteins depends on assistance from ing (1, 6). This aggregation process irrevers-
molecular chaperones, which serve to prevent protein misfolding and aggregation in ibly removes proteins from their productive
the crowded environment of the cell. Nascent chain– binding chaperones, including folding pathways, and must be prevented in
trigger factor, Hsp70, and prefoldin, stabilize elongating chains on ribosomes in a vivo by molecular chaperones. A certain lev-
nonaggregated state. Folding in the cytosol is achieved either on controlled chain el of protein aggregation does occur in cells
release from these factors or after transfer of newly synthesized proteins to down- despite the presence of an exclusive chaper-
stream chaperones, such as the chaperonins. These are large, cylindrical complexes one machinery and, in special cases, can lead
that provide a central compartment for a single protein chain to fold unimpaired by to the formation of structured, fibrillar aggre-
aggregation. Understanding how the thousands of different proteins synthesized in a gates, known as amyloid, that are associated
cell use this chaperone machinery has profound implications for biotechnology and with diseases such as Alzheimer’s or Hun-
medicine. tington’s disease (6, 7) (Fig. 1). Compared to
refolding in dilute solution, the tendency of
nonnative states to aggregate in the cell is

T
o become functionally active, newly partially folded intermediates, including mis- expected to be sharply increased as a result of
synthesized protein chains must fold folded states, that tend to aggregate (Fig. 1). the high local concentration of nascent chains
to unique three-dimensional struc- Misfolding originates from interactions be- in polyribosomes and the added effect of
tures. How this is accomplished remains a tween regions of the folding polypeptide macromolecular crowding.
fundamental problem in biology. Although chain that are separate in the native protein Nascent chains. During translation, the fold-
it is firmly established from refolding ex- and that may be stable enough to prevent ing information encoded in the amino acid se-
periments in vitro that the native fold of a folding from proceeding at a biologically rele- quence becomes available in a vectorial fash-
protein is encoded in its amino acid se- ion. The polypeptide exit
quence (1), protein folding inside cells is channel in the large ribo-
not generally a spontaneous process. Evi- somal subunit is 100 Å
dence accumulated over the last decade long, a distance spanned
indicates that many newly synthesized pro- by an extended chain of
teins require a complex cellular machinery ⬃30 amino acid residues
of molecular chaperones and the input of or an ␣ helix of 65 resi-
metabolic energy to reach their native dues (8). The channel is
states efficiently (2–5). The various chap- on average only 15 Å
erone factors protect nonnative protein wide and is expected to
chains from misfolding and aggregation, prohibit folding beyond
but do not contribute conformational infor- helix formation inside the
mation to the folding process. Here we ribosome, unless the tun-
focus on recent advances in our mechanis- nel is conformationally
tic understanding of de novo protein fold- dynamic. Because the
ing in the cytosol and seek to provide a formation of stable ter-
coherent view of the overall flux of newly tiary structure is a coop-
synthesized proteins through the chaperone erative process at the
system. level of protein domains
(50 to 300 amino acid
Protein Aggregation residues), an average do-
Fig. 1. Aggregation of nonnative protein chains as a side-reaction of
Spontaneous refolding in vitro is generally productive folding in the crowded environment of the cell. Enhancement main can complete fold-
efficient for small, single-domain proteins of aggregation and chain compaction by macromolecular crowding (red ing only when its entire
that bury exposed hydrophobic amino acid arrows). U, unfolded protein chain released from ribosome; I, partially sequence has emerged
residues rapidly (within milliseconds) upon folded intermediate; N, native, folded protein. Crowding is predicted to from the ribosome. It
initiation of folding (1). In contrast, larger enhance the formation of amyloid fibrils, but this effect has not yet been takes more than a minute
demonstrated experimentally. [Adapted from (1)]
proteins composed of multiple domains often to synthesize a 300-resi-
refold inefficiently, owing to the formation of due protein in eu-
vant time scale. These nonnative states, karyotes. As a consequence, many nascent
Department of Cellular Biochemistry, Max-Planck-
though compact in shape, often expose hy- chains expose non-native features for a con-
Institut für Biochemie, Am Klopferspitz 18A, D-82152 drophobic amino acid residues and segments siderable length of time and are prone to
Martinsried, Germany. of unstructured polypeptide backbone to the aggregation. This tendency to aggregate is
ⴱTo whom correspondence should be addressed. solvent. They readily self-associate into dis- thought to be greatly increased by the close
E-mail: uhartl@biochem.mpg.de ordered complexes (Fig. 1), driven by hydro- proximity of nascent chains of the same type

1852 8 MARCH 2002 VOL 295 SCIENCE www.sciencemag.org


SCIENCE’S COMPASS
in polyribosome complexes (5), thus leading erone function is combined with an additional including trigger factor and specialized
to the requirement for chaperones to maintain activity, as is the case for certain protein disul- Hsp70 proteins, bind directly to the ribosome
nascent chains in a nonaggregated, folding- fide isomerases and peptidyl-prolyl isomerases, near the polypeptide exit site and are posi-
competent conformation. enzymes that catalyze rate-limiting steps in the tioned to interact generally with nascent
Macromolecular crowding. The excluded folding of some proteins (14). chains (Fig. 2). The majority of small pro-
volume effects resulting
from the highly crowded
nature of the cytosol (300
to 400 g/liter of proteins
and other macromole-
cules in Escherichia coli)
(9) are predicted to en-
hance the aggregation of
nonnative protein chains
substantially by increas-
ing their effective con-
centrations (10) (Fig. 1).
Crowding generally pro-
vides a nonspecific force
for macromolecular com-
paction and association
(11), including the col-
lapse of protein chains
during folding (9) and the
interaction of nonnative
proteins with molecular
chaperones (12).

How Chaperones
Prevent Aggregation Fig. 2. Models for the chaperone-assisted folding of newly synthesized polypeptides in the cytosol. (A) Eubacteria. TF,
trigger factor; N, native protein. Nascent chains probably interact generally with TF, and most small proteins (⬃65 to 80%
The cellular chaperone of total) fold rapidly upon synthesis without further assistance. Longer chains (10 to 20% of total) interact subsequently
machinery counteracts with DnaK and DnaJ and fold upon one or several cycles of ATP-dependent binding and release. About 10 to 15% of chains
the aggregation of nonna- transit the chaperonin system—GroEL and GroES—for folding. GroEL does not bind to nascent chains and is thus likely
tive proteins, both during to receive an appreciable fraction of its substrates after their interaction with DnaK. (B) Archaea. PFD, prefoldin; NAC,
de novo folding and un- nascent chain–associated complex. Only some archaeal species contain DnaK/DnaJ. The existence of a ribosome-bound
der conditions of stress, NAC homolog, as well as the interaction of PFD with nascent chains, has not yet been confirmed experimentally. (C)
Eukarya—the example of the mammalian cytosol. Like TF, NAC probably interacts generally with nascent chains. The
such as high temperature, majority of small chains may fold upon ribosome release without further assistance. About 15 to 20% of chains
when some native pro- reach their native states in a reaction assisted by Hsp70 and Hsp40, and a fraction of these must be transferred to
teins unfold. Many chap- Hsp90 for folding. About 10% of chains are co- or posttranslationally passed on to the chaperonin TRiC in a reaction
erones, though constitu- mediated by PFD.
tively expressed, are syn-
thesized at greatly increased levels under stress Protein Flux Through the Chaperone teins are thought to fold rapidly and without
conditions and are classified as stress proteins System further assistance upon completion of synthe-
or heat-shock proteins (Hsps) (3). In general, all Cytosolic chaperones participate in de novo sis and release from this first set of compo-
these chaperones recognize hydrophobic resi- folding mainly through two distinct mecha- nents (Fig. 2A). Longer chains interact
dues and/or unstructured backbone regions in nisms. Chaperones, such as trigger factor and subsequently with members of a second class
their substrates, i.e., structural features typically the Hsp70s, act by holding nascent and newly of nascent chain-binding chaperones, includ-
exposed by nonnative proteins but normally synthesized chains in a state competent for ing the classical Hsp70s and prefoldin, which
buried upon completion of folding. Chaperones folding upon release into the medium. In do not associate directly with the ribosome
that participate broadly in de novo protein fold- contrast, the large, cylindrical chaperonin (20–22). In addition to stabilizing elongating
ing, such as the Hsp70s and the chaperonins, complexes provide physically defined com- chains, these chaperones also assist in co- or
promote the folding process through cycles of partments inside which a complete protein or posttranslational folding, or facilitate chain
substrate binding and release regulated by their a protein domain can fold while being se- transfer to downstream chaperones (Fig. 2, A
adenosine triphosphatase (ATPase) activity and questered from the cytosol. These two classes and C) (17, 20, 21). A subset of slow-folding
by cofactor proteins. Chaperone binding may of chaperone are conserved in all three do- and aggregation-sensitive proteins (10 to
not only block intermolecular aggregation di- mains of life and can cooperate in a topolog- 15% of total) interact with a chaperonin for
rectly by shielding the interactive surfaces of ically and timely ordered manner (15–17) folding in both prokaryotes and eukaryotes
non-native polypeptides, including unas- (Fig. 2, A to C). (22–24). Many eukaryotic kinases and other
sembled protein subunits, but may also prevent Although the essential nature of the chap- signal-transduction proteins use an additional
or reverse intramolecular misfolding. Certain eronins has long been recognized (18, 19), it chaperone pathway from Hsp70 to Hsp90
chaperones of the Hsp100 or Clp family even has proved more difficult to establish the (Fig. 2C), a specialized ATP-dependent
have the ability to unfold proteins or to disrupt essential role of nascent chain-binding chap- chaperone that cooperates with ancillary fac-
small-protein aggregates by an adenosine 5⬘- erones in protein folding, because of consid- tors in protein folding and regulation. [For a
triphosphate (ATP)– dependent mechanism erable functional redundancy among compo- detailed discussion of the Hsp90 system, see
(13). For a growing number of proteins, chap- nents (20, 21). Some of these chaperones, (25, 26).]

www.sciencemag.org SCIENCE VOL 295 8 MARCH 2002 1853


SCIENCE’S COMPASS
Ribosome-Binding Chaperones release from the ribosome (4, 33). However, ATP-dependent manner, extended polypep-
Trigger factor (TF), a eubacterial protein of 48 a direct role for NAC in protein folding re- tide segments that are exposed by proteins in
kD, binds to ribosomes at a 1:1 stochiometry mains to be established. their non-native states.
and interacts with nascent chains as short as 57 Whereas the Hsp70 proteins in bacteria and Structure and reaction cycle. The struc-
residues (27). The nascent chain–TF complex higher eukaryotes act both co- and posttransla- tural and mechanistic aspects of the Hsp70
dissociates, in an ATP-independent manner, af- tionally (see below), yeast and other fungi have system are best understood for the eubacterial
ter chain release from the ribosome (27). Al- cytosolic Hsp70 homologs that are specialized Hsp70, termed DnaK, its Hsp40 cochaper-
though TF exhibits peptidyl-prolyl cis/trans in nascent chain binding. The Ssb1 and Ssb2 one, DnaJ, and the nucleotide exchange fac-
isomerase (PPIase) activity in vitro, recognition proteins in the yeast Saccharomyces cerevisiae tor GrpE. DnaK consists of a ⬃44-kD NH2-
of target polypeptides by TF is independent of interact with the ribosome and with short nas- terminal ATPase domain and a ⬃27-kD
proline residues (28) and is mediated by short cent chains (34). Interestingly, this function of COOH-terminal peptide-binding domain
sequences enriched in hydrophobic (aromatic) the Ssb proteins appears to be mediated by yet (37) (Fig. 3A). The latter is divided into a
amino acids (28). TF has an overlapping chap- another Hsp70, Ssz1, which forms a stable ri- ␤-sandwhich subdomain with a peptide-
erone function with the main bacterial Hsp70 bosome-associated complex (RAC) with zuotin binding cleft and an ␣-helical latchlike seg-
system, DnaK and DnaJ, in stabilizing nascent (35, 36), the Hsp40 partner of Ssb1 and Ssb2 ment (38). Target peptides are ⬃seven resi-
chains in a state competent for subsequent fold- (30). RAC and the Ssb proteins are thought to dues long and are typically hydrophobic in
ing (20, 21). E. coli cells lacking TF (⌬tig) or act in concert in stabilizing nascent chains. their central region, with leucine and isoleu-
DnaK (⌬dnak) exhibit no apparent folding de- cine residues being preferred by DnaK (4, 39)
fects at 37°C; however, deletion of dnaK in a The Hsp70 System (Fig. 3A). These binding sites occur statisti-
⌬tig strain is lethal. In light of this functional The classic, nonribosome-binding members cally every ⬃40 residues in proteins and are
redundancy, the biological significance of the of the Hsp70 family exist in the cytosol of recognized with affinities of 5 nM to 5 ␮M
PPIase activity of TF remained unclear, but a eubacteria, eukarya, and some archaea, as (37). The peptides are bound to DnaK in an
recent study suggests that DnaK has a related well as within eukaryotic organelles, such as extended state through hydrophobic side-
activity in accelerating the cis/trans isomeriza- mitochondria and endoplasmic reticulum. S. chain interactions and hydrogen bonds with
tion of nonprolyl peptide bonds (29). These cerevisiae has four
isomerase activities may allow TF and the nonribosome-bind-
Hsp70s to maintain nascent and newly synthe- ing Hsp70 proteins
sized chains in a flexible state, poised for rapid in the cytosol,
folding upon release. In contrast to DnaK, a role namely, Ssa1 to
of TF in mediating folding posttranslationally Ssa4. The cytosol
has not yet been demonstrated, but would be of higher eu-
consistent with the finding that only half of total karyotes contains
TF is ribosome bound (30) both constitutively
The eukaryotic cytosol lacks TF but con- expressed Hsp70
tains a ribosome-associated heterodimeric homologs (Hsc70)
complex of ␣ (33 kD) and ␤ (22 kD) sub- and stress-inducible
units, termed NAC (nascent chain–associated forms (Hsp70). To-
complex) (Fig. 2C) (31). A homolog of gether with cochap-
␣-NAC appears to be present in some archaea erones of the Hsp40
(32). Although NAC lacks a PPIase domain, (DnaJ) family,
it has properties that suggest a functional these Hsp70s func-
similarity to TF. NAC associates with short tion by binding and
nascent chains and dissociates upon chain releasing, in an

Fig. 3. Structure and function of chaperones with the ability to bind


nascent chains. (A) (Top) Structures of the ATPase domain (40) and the
peptide-binding domain (38) of Hsp70 shown representatively for E.
coli DnaK, generated with the program MOLSCRIPT (87). The ␣-helical
latch of the peptide binding domain is shown in yellow and a ball-and-
stick model of the extended peptide substrate in pink. ATP indicates the
position of the nucleotide binding site. The amino acid sequence of the
peptide is indicated in single-letter code (D, Asp; E, Glu; G, Gly; L, Leu;
N, Asn; R, Arg; T, Thr; and V, Val). (Bottom) The interaction of prokary-
otic and eukaryotic cofactors with Hsp70 is shown schematically.
Residue numbers refer to human Hsp70. Only the Hsp70 proteins of the
eukaryotic cytosol have the COOH-terminal sequence EEVD that is
involved in binding of tetratricopeptide repeat (TPR) cofactors. (B)
Simplified reaction cycle of the DnaK system with DnaK colored as in
(A). J, DnaJ; E, GrpE; S, substrate peptide. GrpE is drawn to reflect the
extended shape of the protein. Not all substrates are presented to DnaK
by DnaJ. The intermediate DnaK-DnaJ-substrate-ATP is probably very
transient, as this is the fastest step of the cycle. (C) (Left) Side view and
dimension of the structure of achaeal PFD with the two ␣ subunits
shown in gold and the four ␤ subunits in gray. (Right) Bottom view of
the PFD complex showing the central space enclosed by the six coiled-
coil segments. Surface representation is shown with hydrophobic patch-
es in yellow and hydrophilic regions in gray [reproduced from (54) with
permission].

1854 8 MARCH 2002 VOL 295 SCIENCE www.sciencemag.org


SCIENCE’S COMPASS
the peptide backbone (38). Thus, Hsp70 rec- Large proteins ⬎60 kD, which do not fit into tubulin with the eukaryotic chaperonin
ognizes structural features common to most the central cavity of the chaperonin GroEL (17 ). Interestingly, a combined deletion of
nascent chains: exposed hydrophobic amino (see below), constitute an appreciable frac- the Ssb-class Hsp70s and of PFD in yeast
acid side chains, in conjunction with an tion of these substrates, suggesting that DnaK results in a pronounced synthetic growth
accessible polypeptide backbone. facilitates the posttranslational folding of defect (56 ), resembling the synthetically
Rapid peptide binding occurs in the ATP- multidomain proteins through cycles of bind- lethal phenotype of the TF and DnaK dele-
bound state of DnaK in which the ␣-helical ing and release (21). Consistent with this tions in E. coli (20, 21). These findings
latch over the peptide-binding cleft is in an conclusion, depletion of DnaK in TF-deleted underscore the functional redundancy
open conformation (Fig. 3B). Stable holding cells causes the aggregation of many large, among nascent chain– binding chaperones
of peptide involves closing of the latch, a newly synthesized polypeptides (20). Similar and suggest that PFD may have a DnaK or
conformational change that is achieved by to DnaK, mammalian Hsc70 also binds a TF-like role in the archaeal cytosol.
hydrolysis of bound ATP to adenosine 5⬘- wide range of nascent and newly synthesized
diphosphate (ADP). The cycling of DnaK chains (⬎15 to 20% of total) (Fig. 2C), in- The Chaperonins
between these states is regulated by DnaJ (41 cluding many multidomain proteins ⬎50 kD The chaperonins are a conserved class of
kD) and by GrpE, a homodimer of 20-kD (22). large double-ring complexes of ⬃800 kD
subunits (37, 40). The NH2-terminal J do- How do cycles of Hsp70 binding and enclosing a central cavity. They occur in two
main of DnaJ binds to DnaK and accelerates release promote protein folding? Generally, subgroups that are similar in architecture but
hydrolysis of ATP by DnaK, thus facilitating on release from Hsp70, an unfolded chain is distantly related in sequence. Group I chap-
peptide capture (41, 42). The COOH-terminal free to partition to its native state. Rebinding eronins, also known as Hsp60s, are generally
domain of DnaJ (and of other Hsp40s) func- of slow-folding intermediates to Hsp70 fol- found only in eubacteria and in organelles of
tions as a chaperone in recognizing hydro- lows this release and prevents aggregation. endosymbiotic origin—mitochondria and
phobic peptides and can thus recruit DnaK to Assuming that for long protein chains cycling chloroplasts. They cooperate with cofactors
nascent chains (15, 43, 44). GrpE induces the is mediated by multiple Hsp70 molecules at of the GroES or Hsp10 family. Group II
release of ADP from DnaK (40), and upon the level of individual domains, the Hsp70 chaperonins exist in the archaeal and the
rebinding of ATP the DnaK-peptide complex system could promote the folding of multido- eukaryotic cytosol and are GroES indepen-
dissociates, completing the reaction cycle main proteins by preventing (and perhaps dent. The chaperonin mechanism differs fun-
(Fig. 3B). reversing) intramolecular misfolding. The re- damentally from that of the Hsp70 system,
Some eukaryotic Hsp70 homologs, such cently discovered isomerase activity of although in both cases protein binding and
as BiP in the endoplasmic reticulum, cooper- Hsp70 for nonprolyl peptide bonds may sup- release is ATP regulated. Nonnative substrate
ate with J-domain proteins that lack a sepa- port this function (29). Consistent with this protein is first captured through hydrophobic
rate affinity for hydrophobic sequences. model, the Hsp70 system strongly accelerates contacts with multiple chaperonin subunits
These Hsp70s may be able to bind extended the slow, spontaneous refolding of chemical- and is then displaced into the central ring
polypeptide chains more generally, indepen- ly denatured firefly luciferase (⬃60 kD) in cavity where it folds, protected from aggre-
dent of exposed hydrophobic features (45). vitro (48, 49). The enzyme ornithine transcarb- gating with other nonnative proteins.
Whereas all Hsp70s seem to cooperate with J amylase accumulates in a misfolded but soluble Group I chaperonins—structure and re-
proteins, most eukaryotic Hsp70 proteins form in vivo when expressed in Hsp70 (Ssa)- action cycle. E. coli GroEL and its cofactor
may be independent, for their general func- deficient yeast (50). GroES represent the paradigmatic Group I
tion, of a GrpE-like nucleotide exchange fac- Surprisingly, the components of the chaperonin system. In GroEL, two hep-
tor. Such a factor appears to be dispensable Hsp70 system are missing in certain species tameric rings of identical 57-kD subunits are
because the rate-limiting step in the ATPase of archaea (32). How these cells protect nas- stacked back-to-back. Each subunit consists
cycle of eukaryotic Hsp70 is normally not the cent and newly synthesized polypeptides of three domains: The equatorial domain har-
dissociation of bound ADP, but ATP hydro- from aggregating is not yet clear, but a can- bors the ATP binding site and is connected
lysis itself. On the other hand, a small protein, didate chaperone for nascent chains in ar- through an intermediate, hingelike domain to
Bag-1, acts as a nucleotide exchange factor chaea is prefoldin. the apical domain (Fig. 4A). The latter makes
and specific regulator of Hsp70 in the eukary- up the opening of the cylinder and exposes a
otic cytosol (46). The Hsp70-interacting Bag Prefoldin number of hydrophobic residues toward the
domain is structurally unrelated to GrpE (40, Prefoldin (PFD) (51), also known as the Gim ring cavity for substrate binding. GroES is a
47) and in Bag-1 is linked to an NH2-termi- complex (genes involved in microtubule bio- homoheptameric ring of ⬃10-kD subunits
nal ubiquitin-homology domain (see below). genesis) (52), is a ⬃90-kD complex of two ␣ that cycles on and off the ends of the GroEL
Substrates and mechanism of folding. The and four ␤ subunits in the archaeal and eu- cylinder, in a manner regulated by the GroEL
cellular concentration of DnaK (⬃50 ␮M) karyotic cytosol. The eukaryotic ␣ and ␤ ATPase (4, 37, 57) (Fig. 4A).
exceeds that of ribosomes (⬃30 ␮M) (4), subunits are not identical but orthologous The hydrophobic surfaces exposed by the
assuming an even cytosolic distribution. (53). The structure of PFD resembles that of apical domains (Fig. 4A) interact with hydro-
DnaK preferentially associates with elongat- a jellyfish, with six ␣-helical coiled-coil ten- phobic amino acid residues on compact folding
ing polypeptides larger than 20 to 30 kD and tacles emanating from a ␤-barrel body (Fig. intermediates (4, 37, 57). Hydrophobic se-
thus acts on nascent chains subsequent to TF 3C). At the tips these ⬃65 Å long coiled coils quences bind to a flexible groove between two
(21) (Fig. 2A). Upon deletion of TF, the are partially unwound, exposing hydrophobic amphiphilic helices in the apical domain. This
fraction of nascent and newly synthesized amino acid residues for the binding of non- region can accommodate a peptide either as a ␤
polypeptides interacting with DnaK increases native protein (54 ) (Fig. 3C). Substrate hairpin or an amphiphilic ␣-helical conforma-
from ⬃15 to ⬃40% (20, 21). Whereas some binding and release by PFD is ATP inde- tion (58, 59). Stable substrate binding with
chains transit DnaK with half-lives of less pendent, and in vitro, mammalian and ar- nanomolar affinity relies on the interaction of a
than 1 min, consistent with rapid folding chaeal PFD can stabilize nonnative proteins nonnative polypeptide with multiple apical do-
upon completion of synthesis, other newly for subsequent transfer to a chaperonin (51, mains (60). The GroES subunits have mobile
synthesized proteins are released from DnaK 53). PFD binds to nascent chains (55, 56 ) sequence loops that contact the substrate-bind-
slowly, with half-lives of 10 min or more. and cooperates in the folding of actin and ing regions in the apical domains of GroEL and

www.sciencemag.org SCIENCE VOL 295 8 MARCH 2002 1855


SCIENCE’S COMPASS
mediate substrate dissoci- teome of Ureaplasma
ation (37, 57, 61, 62). urealyticum (68), the first
GroEL is functionally eubacterium that lacks
asymmetrical; the two GroEL and GroES, may
rings are coupled by neg- now offer an opportunity
ative allostery and do not to test these hypotheses.
occur in the same nucle- About 50 proteins in-
otide-bound state. The teracting with GroEL in
chaperonin reaction be- the E. coli cytosol have
gins by the binding of been identified, and many
substrate polypeptide to of them contain two or
the free end (i.e., the more domains with ␣/␤
trans ring) of a GroEL- folds (24). Proteins with
GroES complex (Fig. such complex topologies
4B). This step is closely often fold slowly and are
followed by the binding aggregation prone, owing
of seven ATP molecules to the exposure of exten-
and GroES, resulting in sive hydrophobic surfaces
the displacement of sub- in their non-native states.
strate into a GroES- Stringent model substrates
capped cavity and caus- of GroEL, such as bacteri-
ing the dissociation of al RuBisCo (ribulose-1,5-
the seven ADP mole- bisphosphate carboxylase-
cules and GroES from oxygenase), share this do-
the former cis complex. main topology and fold ef-
Upon binding to GroES, ficiently only when in the
the apical domains un- GroEL-GroES cage (67).
dergo a massive rotation In addition to prevent-
and upward movement ing aggregation during
(61, 63), resulting in an folding, encapsulation of
enlargement of the cavity nonnative RuBisCo (50
and a shift in its surface kD) in the hydrophilic
properties from hydro- cage speeds up the fold-
phobic to hydrophilic ing reaction substantially
(Fig. 4A). Non-native (67). Confinement in the
proteins up to ⬃60 kD Fig. 4. The GroEL-GroES chaperonin system. (A) (Left) View of the asymmetric GroEL- cage may smooth the en-
can be encapsulated and GroES-(ADP)7 complex generated with the coordinates 1AON (61) and program Weblab ergy landscape of folding
are free to fold in the re- ViewLite 4.0 (Molecular Simulations). The equatorial, intermediate, and apical domains of for some larger proteins,
one subunit each in the cis and trans ring of GroEL are colored in pink, yellow, and dark
sulting GroEL-GroES blue, respectively, and one subunit of GroES is colored red. (Right) The accessible surface either by preventing the
cage (also termed “An- of the central cavity of the GroEL-GroES complex. Polar and charged side-chain atoms, formation of certain ki-
finsen cage”) (64–67). blue; hydrophobic side-chain atoms, yellow; backbone atoms, white; and solvent-excluded netically trapped interme-
Folding is allowed to surfaces at subunit interfaces, gray. [Reprinted from (61) with permission] (B) Simplified diates or by facilitating
proceed for ⬃10 s, timed reaction of protein folding in the GroEL-GroES cage. I, folding intermediate bound by the their progression toward
by the hydrolysis of the apical domains of GroEL; N, native protein folded inside the cage. For a typical GroEL the compact, native state
substrate, multiple rounds of chaperonin action are required for folding; both I and N
seven ATP molecules in accumulate after a single reaction cycle and exit the cage upon GroES dissociation. I is then (Fig. 4C). This accelera-
the cis ring. Upon com- rapidly re-bound by GroEL. (C) Mechanisms of accelerated folding. Simple energy diagrams tion of RuBisCo folding
pletion of hydrolysis, are shown for a protein that forms a kinetically trapped intermediate during spontaneous had previously been at-
binding of seven ATP folding (left). In the iterative annealing model, this intermediate is thought to be actively tributed to a mechanism
molecules to the trans unfolded by GroEL/GroES (69) and allowed to repartition (middle), whereas confinement of “iterative annealing”
ring triggers the opening of nonnative protein in the narrow, hydrophilic environment of the GroEL-GroES cage is (57, 69), rather than to an
suggested to result in a smoothing of the energy landscape (right), such that formation of
of the cage. Both folded certain trapped intermediates is avoided (67). Both proposed mechanisms would result in effect of confinement. In
and nonnative protein accelerated folding. this alternative model the
exit at this point (Fig. chaperonin is suggested
4B), but folding interme- to facilitate folding by
diates that still expose extensive hydrophobic (⬃30 ␮M). Most of these proteins are between cycles of unfolding kinetically trapped states,
surfaces are rapidly recaptured and folding 20 and 60 kD in size and leave GroEL with followed by repartitioning of the unfolded
cycles are repeated until the protein reaches half-lives between 15 s and several minutes protein between productive and nonproduc-
its native state. Oligomeric assembly occurs (23). The essential nature of GroEL and GroES tive folding pathways (57) (Fig. 4C). Active
in solution after subunit folding inside the (18) may be explained in principle by the exis- unfolding of RuBisCo was suggested to result
cage. tence of at least one essential E. coli protein with from GroES-induced movements of the api-
Substrates and folding mechanisms. About an absolute chaperonin dependence. Alterna- cal GroEL domains, exerting a stretching
10% of newly synthesized polypeptides normal- tively, loss of chaperonin could be lethal be- force on the bound polypeptide (69), but this
ly transit GroEL posttranslationally (23, 24), cause it results in a reduced efficiency, rather effect has not yet been confirmed with any
consistent with the cytosolic concentration of than a complete loss, of folding for many pro- other GroEL-dependent protein (69, 70).
GroEL (⬃3 ␮M) relative to that of ribosomes teins. Analysis of the highly streamlined pro- As shown recently, GroEL also interacts

1856 8 MARCH 2002 VOL 295 SCIENCE www.sciencemag.org


SCIENCE’S COMPASS
with, and assists in the folding of, certain The mechanistic principles underlying this ist in which different chaperones function in a
proteins too large to be encapsulated by functional cooperation are not yet well under- processive manner to minimize the exposure
GroES (23, 24, 71). Mitochondrial aconitase stood, but plausible models have been devel- of nonnative proteins to the bulk cytosol.
(82 kD), for example, can fold through ATP- oped based on a combination of in vitro and Whereas in both prokaryotes and eukaryotes
regulated cycles of GroEL binding and re- in vivo studies. specific chaperones are recruited to nascent
lease of nonnative states, with protein release Cotranslational domain folding. It has chains by virtue of their affinity for the ribo-
being triggered by the binding of GroES to been suggested that translation itself can have some (Fig. 2), the existence of processive chap-
the opposite (trans) ring of GroEL (71). a “chaperone-like” role in the folding of larg- erone pathways has so far been demonstrated
Group II chaperonins. The Group II chap- er proteins composed of multiple domains only in the eukaryotic system. As shown in
eronin of the eukaryotic cytosol, TRiC (77). Attempts to refold such proteins in vitro yeast and mammalian cells, folding intermedi-
(TCP-1 ring complex, also called CCT for often result in intramolecular misfolding and ates generated during biosynthesis are not free-
chaperonin-containing TCP-1), contains aggregation (5). Cotranslational and sequen- ly exposed to the bulk cytosol, but rather are
eight orthologous subunits per ring that differ tial domain folding, i.e., the folding of one functionally compartmentalized (17, 22). In
primarily in their apical domains. The sim- domain well before another is synthesized, these experiments, certain nascent and newly
pler archaeal chaperonin, referred to as ther- avoids this problem, as demonstrated with synthesized protein chains do not bind to a
mosome, consists of up to three different artificial two-domain fusion proteins combin- heterologously expressed, noncycling mutant of
subunits, which are arranged in eight- or ing the ⬃20-kD proteins H-ras and dihydro- GroEL (Trap-GroEL), but instead interact pro-
nine-membered rings. The backbone trace of folate reductase (DHFR) (77). ductively with the endogenous eukaryotic chap-
the chaperonin II apical domain is virtually Domain folding during translation occurs erones. In the specific case of actin, an obliga-
identical to that of GroEL, with the exception in the prokaryotic and eukaryotic cytosol tory substrate of TRiC, protection from expo-
of an ␣-helical insertion that protrudes from (78), but a difference in the efficiencies of sure to the bulk cytosol during folding is medi-
the ring opening (72, 73) and, in the absence folding was noted for certain multidomain ated by PFD. The speed and efficiency of actin
of a separate GroES-like cofactor, is thought proteins when comparing both systems (77). folding is markedly reduced in PFD-deficient
to function as a built-in lid of the central The bacterial two-domain protein OmpR and yeast, with nonnative chains being released into
cavity. the ras-DHFR fusion proteins fold cotransla- the cytosol (17). PFD may deliver substrate
The mechanism of Group II chaperonins tionally in a mammalian cell lysate or in proteins to TRiC by binding both to nascent
is not yet well understood, and the nature and intact cells but posttranslationally upon syn- chains (55, 56) and to TRiC itself (51). In
exact location of the substrate binding site(s) thesis in the E. coli system. Although the addition, PFD and TRiC seem to cooperate
on the apical domains are still undefined. The individual domains fold efficiently in E. coli, functionally in actin folding, such that nonna-
most abundant substrates of TRiC are the bacterial expression of ras-DHFR does not tive chains are not released into the cytosol
cytoskeletal proteins actin and tubulin. Strik- result in an active protein. Eukaryotic cells during folding cycles (17) (Fig. 2C).
ingly, folding of these proteins cannot be contain a much greater number of proteins Another example of chaperone coupling
mediated by GroEL and GroES, suggesting a with multiple domains than bacteria (77). in the eukaryotic cytosol is the cooperation
more specific role for TRiC in folding be- Although the sizes of protein domains are between Hsc70 and Hsp90 in the folding of
yond prevention of aggregation. Actin binds uniformly distributed in all domains of life, in signal-transduction proteins (26). Substrate
to TRiC through at least two distinct regions E. coli only 13% of all 4300 proteins exceed transfer from Hsc70 to Hsp90 is mediated by
and interacts with specific TRiC subunits (74, a length of 500 residues (⬃55 kD), compared Hop (Hsp organizing protein; also known as
75). ATP binding induces encapsulation of with 38% of the 5800 proteins in S. cerevi- p60), an adaptor protein that physically links
the protein by the apical-domain protrusions siae. Thus, the eukaryotic translation and both chaperones. Hop contains two tetratri-
and initiates folding (75, 76). Through its folding machineries may have been opti- copeptide repeat (TPR) domains, which bind
built-in lid mechanism, TRiC may act co- mized in evolution to facilitate cotransla- the extended COOH-terminal sequences of
translationally in the folding of discrete do- tional domain folding. This optimization may Hsc70 and Hsp90, respectively (80) (Fig.
mains of proteins that are too large to be be reflected in the 5- to 10-times slower 3A). As shown recently, similar mechanisms
encapsulated as a whole (16) (Fig. 2C). speed of translation in eukaryotes compared are involved in regulating the transfer of non-
The subunit heterogeneity of TRiC sug- with bacteria and in a functional adaptation of native or irreversibly misfolded proteins from
gested that the cytosolic chaperonin may be the eukaryotic chaperone machinery. Eukary- these chaperones to the ubiquitin-proteasome
adapted to assisting the folding of a small set otic TRiC, for example, may mediate cotrans- machinery. The protein CHIP associates with
of specific proteins, including actins and tu- lational domain folding for some proteins Hsp90 through an NH2-terminal TPR domain
bulins. However, as determined by a recent (16, 56, 79), whereas folding in the bacterial and targets certain Hsp90 substrates for deg-
pulse-chase analysis in mammalian cells, GroEL-GroES cage is strictly posttransla- radation through a COOH-terminal ubiquitin
TRiC interacts transiently with a wider range tional (23) (Fig. 2). ligase domain (81, 82). CHIP cooperates
of newly synthesized proteins of 30 to 120 Processivity of chaperone action. The no- functionally with Bag-1 (see above), which
kD in size, constituting ⬃12% of total syn- tion of cooperation between mechanistically binds to Hsc70 and to the proteasome (83).
thesized chains (22). The list of model sub- distinct chaperones in protein folding is now These findings provide the first insight into
strates includes, among others, firefly lucif- firmly established (15, 16, 20, 21, 26), but the mechanisms that integrate chaperone-as-
erase, ␣-transducin, and the von Hippel– how the different components of the folding sisted folding and proteolytic degradation,
Lindau tumor suppressor protein (5). machinery are functionally integrated is not the two main components of protein quality
yet well understood. In principle, substrate control in the cytosol.
Coordination of Translation and transfer between chaperones could be accom-
Chaperone Activities plished by free partitioning of nonnative Perspectives
To ensure an efficient use of the cytosolic states through the solution. However, consid- Recent years have seen major advances in our
folding machinery, protein synthesis on ribo- ering the highly crowded nature of the cy- understanding of the basic mechanisms of
somes must be coordinated with the activities tosol, it is difficult to envisage how aggrega- chaperone-assisted protein folding. Future ef-
of the various chaperone systems in stabiliz- tion is avoided in this model. Alternatively, forts will define more comprehensively the
ing nascent chains and in promoting folding. ordered pathways of cellular folding may ex- rules for how the thousands of different pro-

www.sciencemag.org SCIENCE VOL 295 8 MARCH 2002 1857


SCIENCE’S COMPASS
teins in a cell use the chaperone machinery. 16. J. Frydman, E. Nimmesgern, K. Ohtsuka, F. U. Hartl, 52. S. Geissler, K. Siegers, E. Schiebel, EMBO J. 17, 952
Nature 370, 111 (1994). (1998).
Global analyses of chaperone usage and fold- 17. K. Siegers et al., EMBO J. 18, 75 (1999). 53. M. R. Leroux et al., EMBO J. 18, 6730 (1999).
ing properties in eukaryotes and prokaryotes 18. O. Fayet, T. Ziegelhoffer, C. Georgopoulos, J. Bacte- 54. R. Siegert, M. R. Leroux, C. Scheufler, F. U. Hartl, I.
will address these questions through a com- riol. 171, 1379 (1989). Moarefi, Cell 103, 621 (2000).
bination of proteomics and high-throughput 19. A. L. Horwich, K. B. Low, W. A. Fenton, I. N. Hirshfield, 55. W. J. Hansen, N. J. Cowan, W. J. Welch, J. Cell Biol.
K. Furtak, Cell 74, 909 (1993). 145, 265 (1999).
protein expression (84). These studies may 20. E. Deuerling, A. Schulze-Specking, T. Tomoyasu, A. 56. K. Siegers, F. U. Hartl., personal communication.
eventually offer a rational basis to optimize Mogk, B. Bukau, Nature 400, 693 (1999). 57. P. B. Sigler et al., Annu. Rev. Biochem. 67, 581 (1998).
recombinant protein production in organisms 21. S. A. Teter et al., Cell 97, 755 (1999). 58. J. Chatellier, A. M. Buckle, A. R. Fersht, J. Mol. Biol.
22. V. Thulasiraman, C. F. Yang, J. Frydman, EMBO J. 18, 292, 163 (1999).
that have been genetically modified to pro- 85 (1999). 59. L. L. Chen, P. B. Sigler, Cell 99, 757 (1999).
vide the appropriate folding machineries. 23. K. L. Ewalt, J. P. Hendrick, W. A. Houry, F. U. Hartl, Cell 60. G. W. Farr et al., Cell 100, 561 (2000).
In addition to its biotechnological interest, 90, 491 (1997). 61. Z. H. Xu, A. L. Horwich, P. B. Sigler, Nature 388, 741
24. W. A. Houry, D. Frishman, C. Eckerskorn, F. Lottspeich, (1997).
understanding the complex functions of the F. U. Hartl, Nature 402, 147 (1999). 62. A. Richardson, S. J. Landry, C. Georgopoulos, Trends
chaperone arsenal will likely prove useful in 25. J. Buchner, Trends Biochem. Sci. 24, 136 (1999). Biochem. Sci. 23, 138 (1998).
dissecting the mechanisms by which protein 26. J. C. Young, I. Moarefi, F. U. Hartl, J. Cell Biol. 154, 63. A. M. Roseman, S. X. Chen, H. White, K. Braig, H. R.
267 (2001). Saibil, Cell 87, 241 (1996).
misfolding and aggregation cause disease. Are 27. T. Hesterkamp, S. Hauser, H. Lutcke, B. Bukau, Proc. 64. R. J. Ellis, Curr. Biol. 11, R1038 (2001).
chaperones capable of preventing the deposi- Natl. Acad. Sci. U.S.A. 93, 4437 (1996). 65. M. Mayhew et al., Nature 379, 420 (1996).
tion of amyloid aggregates, and if so, why do 28. H. Patzelt et al., Proc. Natl. Acad. Sci. U.S.A. 98, 66. J. S. Weissman, H. S. Rye, W. A. Fenton, J. M.
these defense mechanisms fail in the millions of 14244 (2001). Beechem, A. L. Horwich, Cell 84, 481 (1996).
29. C. Schiene-Fischer, J. Habazettl, G. Fischer, person- 67. A. Brinker et al., Cell 107, 223 (2001).
patients suffering from neurodegenerative mal- al communication. 68. J. I. Glass et al., Nature 407, 757 (2000).
adies such as Alzheimer’s or Huntington’s dis- 30. B. Bukau, E. Deuerling, C. Pfund, E. A. Craig, Cell 101, 69. M. Shtilerman, G. H. Lorimer, S. W. Englander, Sci-
ease? Recent reports that an up-regulation of 119 (2000). ence. 284, 822 (1999).
31. B. Wiedmann, H. Sakai, T. A. Davis, M. Wiedmann, 70. J. W. Chen, S. Walter, A. L. Horwich, D. L. Smith,
the Hsp70 system can suppress the neurotoxic- Nature 370, 434 (1994). Nature Struct. Biol. 8, 721 (2001).
ity of certain amyloidogenic proteins (85, 86) 32. M. R. Leroux, Adv. Appl. Microbiol. 50, 219 (2001). 71. T. K. Chaudhuri, G. W. Farr, W. A. Fenton, S. Rospert,
point toward molecular chaperones as promis- 33. B. Beatrix, H. Sakai, M. Wiedmann, J. Biol. Chem. 275, A. L. Horwich, Cell 107, 235 (2001).
37838 (2000). 72. M. Klumpp, W. Baumeister, L. O. Essen, Cell 91, 263
ing targets in the quest for treatment of protein- 34. C. Pfund et al., EMBO J. 17, 3981 (1998). (1997).
misfolding diseases. 35. M. Gautschi et al., Proc. Natl. Acad. Sci. U.S.A. 98, 73. L. Ditzel et al., Cell 93, 125 (1998).
3762 (2001). 74. O. Llorca et al., EMBO J. 19, 5971 (2000).
36. M. Gautschi, A. Mun, S. Ross, S. Rospert, Proc. Natl. 75. O. Llorca et al., EMBO J. 20, 4065 (2001).
Acad. Sci. U.S.A., in press. 76. I. Gutsche, J. Holzinger, N. Rauh, W. Baumeister, R. P.
References and Notes
1. C. Dobson, M. Karplus, Curr. Opin. Struct. Biol. 9, 92 37. B. Bukau, A. L. Horwich, Cell 92, 351 (1998). May, J. Struct. Biol. 135, 139 (2001).
(1999). 38. X. T. Zhu et al., Science 272, 1606 (1996). 77. W. J. Netzer, F. U. Hartl, Nature 388, 343 (1997).
39. S. Rüdiger, L. Germeroth, J. Schneider-Mergener, B. 78. A. V. Nicola, W. Chen, A. Helenius, Nature Cell Biol. 1,
2. R. J. Ellis, S. M. Hemmingsen, Trends Biochem. Sci. 14,
Bukau, EMBO J. 16, 1501 (1997). 341 (1999).
339 (1989).
40. C. J. Harrison, M. Hayer-Hartl, M. Di Liberto, F. U. 79. C. D. McCallum, H. Do, A. E. Johnson, J. Frydman,
3. M.-J. Gething, J. Sambrook, Nature 355, 33 (1992).
Hartl, J. Kuriyan, Science 276, 431 (1997). J. Cell Biol. 149, 591 (2000).
4. F. U. Hartl, Nature 381, 571 (1996).
41. M. P. Mayer et al., Nature Struct. Biol. 7, 586 (2000). 80. C. Scheufler et al., Cell 101, 199 (2000).
5. J. Frydman, Annu. Rev. Biochem. 70, 603 (2001). 42. M. Pellecchia et al., Nature Struct. Biol. 7, 298 (2000). 81. J. Demand, S. Alberti, C. Patterson, J. Höhfeld, Curr.
6. S. E. Radford, Trends Biochem. Sci. 25, 611 (2000). 43. S. Rüdiger, J. Schneider-Mergener, B. Bukau, EMBO J. Biol. 11, 1569 (2001).
7. C. M. Dobson, Trends Biochem. Sci. 24, 329 (1999). 20, 1042 (2001). 82. P. Connell et al., Nature Cell Biol. 3, 93 (2001).
8. P. Nissen, J. Hansen, N. Ban, P. B. Moore, T. A. Steitz, 44. B. D. Sha, S. Lee, D. M. Cyr, Struct. Fold. Des. 8, 799 83. J. Luders, J. Demand, J. Höhfeld, J. Biol. Chem. 275,
Science 289, 920 (2000). (2000). 4613 (2000).
9. R. J. Ellis, Curr. Opin. Struct. Biol. 11, 114 (2001). 45. B. Misselwitz, O. Staeck, T. A. Rapoport, Mol. Cell 2, 84. S. A. Lesley, Protein Expr. Purif. 22, 159 (2001).
10. B. van den Berg, R. J. Ellis, C. M. Dobson, EMBO J. 18, 593 (1998). 85. P. Kazemi-Esfarjani, S. Benzer, Science 287, 1837
6927 (1999). 46. J. Höhfeld, S. Jentsch, EMBO J. 16, 6209 (1997). (2000).
11. A. P. Minton, Curr. Opin. Struct. Biol. 10, 34 (2000). 47. H. Sondermann et al., Science 291, 1553 (2001). 86. C. J. Cummings et al., Hum. Mol. Genet. 10, 1511
12. J. Martin, F. U. Hartl, Proc. Natl. Acad. Sci. U.S.A. 94, 48. A. Szabo et al., Proc. Natl. Acad. Sci. U.S.A. 91, 10345 (2001).
1107 (1997). (1994). 87. P. J. Kraulis, J. Appl. Crystallogr. 24, 946 (1991).
13. A. P. Ben-Zvi, P. Goloubinoff, J. Struct. Biol. 135, 84 49. R. Herbst, K. Gast, R. Seckler, Biochemistry 37, 6586 88. We thank our colleagues B. Bukau, E. Deuerling, R. J.
(2001). (1998). Ellis, G. Fischer, A. Horwich, S. Rospert, and K. Siegers
14. C. Schiene, G. Fischer, Curr. Opin. Struct. Biol. 10, 40 50. S. Kim, B. Schilke, E. A. Craig, A. L. Horwich, Proc. for providing access to unpublished information, and
(2000). Natl. Acad. Sci. U.S.A. 95, 12860 (1998). R. J. Ellis and members of the Hartl group for critically
15. T. Langer et al., Nature 356, 683 (1992). 51. I. E. Vainberg et al., Cell 93, 863 (1998). reading the manuscript.

1858 8 MARCH 2002 VOL 295 SCIENCE www.sciencemag.org

You might also like