ArticlePDF Available

Roles of Individual Domains and Conserved Motifs of the AAA+ Chaperone ClpB in Oligomerization, ATP Hydrolysis, and Chaperone Activity

Authors:

Abstract

ClpB of Escherichia coli is an ATP-dependent ring-forming chaperone that mediates the resolubilization of aggregated proteins in cooperation with the DnaK chaperone system. ClpB belongs to the Hsp100/Clp subfamily of AAA+ proteins and is composed of an N-terminal domain and two AAA-domains that are separated by a “linker” region. Here we present a detailed structure-function analysis of ClpB, dissecting the individual roles of ClpB domains and conserved motifs in oligomerization, ATP hydrolysis, and chaperone activity. Our results show that ClpB oligomerization is strictly dependent on the presence of the C-terminal domain of the second AAA-domain, while ATP binding to the first AAA-domains stabilized the ClpB oligomer. Analysis of mutants of conserved residues in Walker A and B and sensor 2 motifs revealed that both AAA-domains contribute to the basal ATPase activity of ClpB and communicate in a complex manner. Chaperone activity strictly depends on ClpB oligomerization and the presence of a residual ATPase activity. The N-domain is dispensable for oligomerization and for the disaggregating activity in vitro and in vivo. In contrast the presence of the linker region, although not involved in oligomerization, is essential for ClpB chaperone activity.
Roles of Individual Domains and Conserved Motifs of the AAA
Chaperone ClpB in Oligomerization, ATP Hydrolysis, and
Chaperone Activity*
Received for publication, September 20, 2002, and in revised form, January 23, 2003
Published, JBC Papers in Press, March 6, 2003, DOI 10.1074/jbc.M209686200
Axel Mogk‡§, Christian Schlieker‡, Christine Strub§, Wolfgang Rist‡, Jimena Weibezahn‡,
and Bernd Bukau‡
From the Zentrum fu¨ r Molekulare Biologie Heidelberg, Im Neuenheimer Feld 282, Heidelberg D-69120, Germany and
§Institut fu¨ r Biochemie und Molekularbiologie, Universita¨ t Freiburg, Hermann-Herder-Strasse 7,
79104 Freiburg, Germany
ClpB of Escherichia coli is an ATP-dependent ring-
forming chaperone that mediates the resolubilization of
aggregated proteins in cooperation with the DnaK chap-
erone system. ClpB belongs to the Hsp100/Clp subfamily
of AAAproteins and is composed of an N-terminal do-
main and two AAA-domains that are separated by a
“linker” region. Here we present a detailed structure-
function analysis of ClpB, dissecting the individual roles
of ClpB domains and conserved motifs in oligomeriza-
tion, ATP hydrolysis, and chaperone activity. Our re-
sults show that ClpB oligomerization is strictly depend-
ent on the presence of the C-terminal domain of the
second AAA-domain, while ATP binding to the first AAA-
domains stabilized the ClpB oligomer. Analysis of mu-
tants of conserved residues in Walker A and B and sen-
sor 2 motifs revealed that both AAA-domains contribute
to the basal ATPase activity of ClpB and communicate in
a complex manner. Chaperone activity strictly depends
on ClpB oligomerization and the presence of a residual
ATPase activity. The N-domain is dispensable for oli-
gomerization and for the disaggregating activity in vitro
and in vivo. In contrast the presence of the linker region,
although not involved in oligomerization, is essential for
ClpB chaperone activity.
The Escherichia coli chaperone ClpB belongs to the ring-
forming Clp/Hsp100 proteins. Clp/Hsp100 proteins can be clas-
sified into two distinct subfamilies. Class I proteins (ClpA and
ClpB in E. coli) are composed of two highly conserved nucleo-
tide binding domains (termed ATP-1 and ATP-2), whereas class
II proteins (ClpX and HslU, as representatives of E. coli) con-
tain only a single NBD (homologous to ATP-2) (1). Sequence
analysis of the NBDs revealed a significant sequence homology
between Clp/Hsp100 and AAA proteins (ATPase associated
with a variety of cellular activities), and consequently a new
AAAsuperfamily, representing both protein classes, was pro-
posed (2). The structural basis of this superfamily was con-
firmed by determination of the first Clp/Hsp100 protein struc-
ture, HslU, that showed significant similarity to the AAA
proteins N-ethylmaleimide-sensitive fusion protein and p97 (3,
4). The recently solved crystal structure of the first nucleotide
binding domain of ClpB also demonstrated the close structural
relationship between Clp/Hsp100 and AAA proteins (5). The
conserved AAA-domain (also referred to as AAA module) is
made up of two domains, a core region that forms the nucleo-
tide binding pocket, containing the classical Walker A and B
motifs, and a C-terminal
-helical domain (C-domain). The
ATP binding pocket is located at the interface of neighboring
subunits in the oligomer. The C-domain contacts its own core
ATPase domain and that of adjacent subunits and is involved
in nucleotide binding and hexamerization in HslU (3, 4). Be-
sides sensing the nucleotide status of the core ATPase domain,
C-terminal domains of the second AAA-domain have also been
proposed to mediate substrate interaction and were therefore
termed the sensor and substrate discrimination (SSD)
1
do-
mains (6).
In addition Hsp100/Clp proteins contain variable regions at
their N terminus. ClpA and ClpB have homologous N-domains
of about 150 residues that consist of two sequence repeats and
form an independent structural domain still of unknown func-
tion (7). ClpX possesses a zinc binding domain at the N termi-
nus (8), whereas HslU lacks an N-terminal domain but rather
contains an extra domain (the I-domain) inserted into the AAA-
domain (3, 4). The most striking difference between members
within the class I subfamily is the presence/absence of a region
that is proposed to link the two AAA-domains ATP-1 and
ATP-2. The presence of this variable linker region serves as a
criteria for classification of Hsp100 proteins; the linker is long-
est in ClpB (140 residues) but is absent in ClpA (see Fig. 1)
(1).
ClpB is unique among the Hsp100/Clp proteins because it
does not associate with a proteolytic partner protein. Recently,
an essential docking site of the peptidase ClpP was identified
within ClpX (9). The signature motif (LIV-G-FL) is conserved
in all other ClpP-interacting proteins like ClpA but is missing
in ClpB, thereby explaining why ClpB acts independently of
peptidases. Instead ClpB mediates the resolubilization of ag-
gregated proteins in cooperation with the DnaK chaperone
system (10 –13). The mechanism of the disaggregation reaction
and the basis of ClpB/DnaK cooperation are still not
understood.
Here we report a structure-function analysis of ClpB that is
* This work was supported by Deutsche Forschungsgemeinschaft
Grant Bu617/14-1 and by the Fond der Chemischen Industrie (to B. B.).
The costs of publication of this article were defrayed in part by the
payment of page charges. This article must therefore be hereby marked
advertisement in accordance with 18 U.S.C. Section 1734 solely to
indicate this fact.
To whom correspondence may be addressed. E-mail: a.mogk@
zmbh-uni-heidelberg.de.
To whom correspondence may be addressed. E-mail: bukau@
zmbh.uni-heidelberg.de.
1
The abbreviations used are: SSD, sensor- and substrate discrimina-
tion; DTT, dithiothreitol; ATP
S, adenosine 5-O-(thiotriphosphate);
IPTG, isopropyl-1-thio-
-D-galactopyranoside; MDH, malate dehydro-
genase; WT, wild type; aa, amino acid.
THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 278, No. 20, Issue of May 16, pp. 17615–17624, 2003
© 2003 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A.
This paper is available on line at http://www.jbc.org 17615
by guest on December 27, 2015http://www.jbc.org/Downloaded from
aimed at identifying the roles of individual domains and con-
served motifs in the disaggregation process and the coupling of
the ATPase cycle with the chaperone activity. Constructed
ClpB variants were characterized with respect to their struc-
tural integrity, as determined by oligomerization studies and
partial proteolysis. Additionally the ATPase and chaperone
activities of all constructs were tested.
EXPERIMENTAL PROCEDURES
Strains and PlasmidsE. coli strains used were derivatives of
MC4100 (araD139 (argF-lac)U169 rpsL150 relA1 flbB5301 deoC1
ptsF25 rbsR). clpB mutant allele was introduced by P1 transduction
into MC4100 background to generate strains BB4561 (clpB::kan). The
E. coli clpB gene was cloned by PCR into the pUHE21 expression vector
and verified by DNA sequencing. Mutants and deletion variants were
generated by using standard PCR techniques. ClpB-(410 532) was
constructed by replacing the entire linker region with an SpeI site,
leading to the insertion of two amino acids (Thr and Ser) at the deletion
site. Reinsertion of the linker was obtained by PCR amplification of the
corresponding region (amino acid 410 530) and addition of an SpeI site
5and an NheI site 3to the linker fragment. This fragment was
digested with NheI/SpeI and inserted into the SpeI site of the 410 532
construct, leading to the insertion of two amino acids (Thr and Ser or
Ala and Ser) at each domain boundary (ClpB LBL). Each mutagenesis
was confirmed by DNA sequencing.
ProteinsWild type and mutant ClpB were purified as described
after overproduction in clpB::kan cells (11). Purifications of DnaK,
DnaJ, and GrpE were according to published protocols (14). Pyruvate
kinase and
-casein were purchased from Sigma; malate dehydrogen-
ase (MDH) was from pig heart muscle, and firefly luciferase from Roche
Applied Science. Protein concentrations were determined with the Bio-
Rad Bradford assay using bovine serum albumin as standard. Protein
concentrations refer to the protomer.
Tryptophan FluorescenceMeasurements of the intrinsic trypto-
phan fluorescence of ClpB were performed on a PerkinElmer Life Sci-
ences LS50B spectrofluorimeter. The emission spectra of tryptophan
fluorescence of ClpB (0.5
M) in the absence of nucleotide or the pres-
ence of 2 mMATP/ADP were recorded at 30 °C in buffer A (50 mMTris,
pH 7.5, 150 mMKCl, 20 mMMgCl
2
,2mMDTT) between 300 and 400 nm
at a fixed excitation wavelength of 290 nm.
ATPase Activity AssayATP hydrolysis rates under steady-state
conditions were determined as described (14). Reactions were per-
formed at 30 °C in buffer A (50 mMTris, pH 7.5, 150 mMKCl, 20 mM
MgCl
2
,2mMDTT) containing 0.5
MClpB (wild type or derivatives), 2
mMATP, and [
-
32
P]ATP (0.1
Ci, Amersham Biosciences). ATPase
activities were also determined in the presence of 0.25 mg/ml
-casein
or 100 mMammonium sulfate. Hydrolysis was quantified by using the
program MACBAS version 2.5 (Fuji), and rates of ATP hydrolysis were
determined by using the program GRAFIT version 3.0 (Erithacus
software).
Size Exclusion ChromatographySize exclusion chromatography
was performed at room temperature in buffer A containing 5% (v/v)
glycerol. Nucleotide-dependent oligomerization was followed in the
presence of 2 mMATP or ADP in the running buffer. 10
MClpB was
incubated in buffer A in the absence or presence of nucleotides (2 mM
ATP/ADP) for 5 min at room temperature, followed by injection into the
high pressure liquid chromatography system (PerkinElmer Life Sci-
ences) connected to a SEC 400 column (Bio-Rad). Chromatographic
steps were performed with a flow rate of 0.8 ml/min.
Cross-linking AssaysAll ClpB variants were dialyzed against
buffer B (50 mMHEPES, pH 7.5, 150 mMKCl, 20 mMMgCl
2
,2mM
DTT). 1
MClpB was incubated at 30 °C in the absence or presence of
nucleotides (2 mMATP or ADP) for 5 min. Cross-linking reactions were
started by addition of 0.1% glutaraldehyde and incubated for another
10 min. Reactions were stopped by addition of 1 MTris, pH 7.5, and
cross-linking products were analyzed by SDS-PAGE (4 15%) followed
by silver staining.
Partial Proteolysis and Identification of Cleavage Products1
M
ClpB (wild type or derivative) was incubated at 30 °C in buffer A
without DTT for 5 min in the absence or presence of nucleotide (2 mM
ATP, ADP, and ATP
S). Proteolysis was initiated upon addition of 0.2
g/ml thermolysin or subtilisin, and generated cleavage products were
analyzed by SDS-PAGE (15%) and silver staining. Kinetic analysis of
the degradation reaction revealed the occurrence of stable fragments
after 30 60 min of incubation time. Identity of cleavage products was
determined by N-terminal sequencing (TopLab) and mass spectrome-
try. For mass spectrometry analysis, bands were excised from one-
dimensional Coomassie Silver-stained SDS-polyacrylamide gels and
in-gel-digested with trypsin as described (15). Tryptic peptides were
analyzed by nanoelectrospray tandem mass spectrometry as described
previously (16) using a QSTAR
TM
Pulsar (MDS Sciex, Toronto, Canada)
equipped with a nanoelectrospray ion source (MDS Proteomics, Odense,
Denmark). Sequence searches were performed with the Protein and
Peptide Software Suite (MDS Proteomics).
In Vitro Activity Assays2
MMDH was denatured in buffer A for
30 min at 47 °C. Turbidity of 1
Maggregated MDH was measured
at 30 °C at an excitation and emission wavelength of 550 nm
(PerkinElmer Life Sciences luminescence spectrometer LS50B). De-
crease of light scattering was followed upon addition of 1.5
MClpB
(wild type or derivative), the DnaK chaperone system (1
MDnaK, 0.2
MDnaJ, 0.1
MGrpE), and an ATP-regenerating system (3 mM
phosphoenolpyruvate, 20 ng/ml pyruvate kinase, 2 mMATP). Disag-
gregation rates were derived from the linear decrease of turbidity
between 5 and 30 min.
0.2
Mluciferase was denatured in buffer A for 15 min at 43 °C.
Refolding of 0.1
Maggregated luciferase was initiated by addition of 1
MClpB (wild type or derivative), the DnaK system (0.5
MDnaK, 0.1
MDnaJ, 0.05
MGrpE), and an ATP-regenerating system. Luciferase
activities were determined as described, and refolding rates were cal-
culated from the linear increase of luciferase activities between 15 and
90 min.
In Vivo Activity AssaysSurvival of E. coli cells after exposure to
lethal temperature was determined by calculating the plating effi-
ciency. Cells were grown in LB medium to mid-exponential growth
phase at 30 °C followed by incubation at 50 °C for the indicated time.
Serial dilutions (10
3
to 10
7
) of cells were prepared in LB medium and
spotted onto LB plates. Plates were incubated for 24 h at 30 °C, and
colony numbers were determined afterward. Defects in protein disag-
gregation in clpB::kan mutant cells were followed by isolation of
aggregated proteins as described (17).
RESULTS
Design of ClpB Mutants and Deletions VariantsClpB is
proposed to consist of an N-terminal domain, two AAA-domains
(ATP-1 and ATP-2), which are separated by a linker region, and
the C-terminal SSD domain (as illustrated in Fig. 1). In order
to probe the proposed domain organization of ClpB, several
deletion variants were constructed, based on sequence and
secondary structure analysis of ClpB and its comparison with
AAAproteins of known structure. Due to the presence of an
internal start codon, two versions of E. coli ClpB exist in vivo:
a full-length protein (aa 1857) and an N-terminally trun-
cated variant (aa 149 857). Both versions have been demon-
strated to form mixed oligomers (18). In order to work with
uniform proteins, we decided to mutate the internal start
codon of wild type ClpB, leading to the production of full-
length ClpB only. This ClpB derivative exhibited in vitro and
in vivo chaperone activity, indistinguishable from ClpB ex-
pressed from the wild type gene (data not shown). The cor-
responding construct was used as basis for the construction of
all ClpB derivatives.
The border between the N-domain and the first AAA-domain
is often characterized by the presence of an internal start codon
within the clpB gene (Val-149). Sequence analysis of Hsp100
N-domains and secondary structure prediction of 15 different
ClpB proteins revealed that the ClpB N-domain is built up of
residues 1144 (7) (data not shown). We therefore decided to
use Met-143 as the start site for ClpB N-(143857) version.
The C-terminal SSD was removed in the ClpB SSD variant
(aa 1758). AAA-domains were suggested to be separated by
the insertion of the linker region within ClpB. Recent sequence
and structural analysis of AAAproteins revealed that the
linker region potentially interrupts the first AAA-domain in-
stead of separating both AAA modules (2). Several ClpB dele-
tion variants were constructed to test this possibility: ClpB-
(1409), ClpB-(551857), and ClpB-(1567).
Several conserved motifs and residues are proposed to be
involved in ATP binding and hydrolysis by AAA proteins. The
highly conserved Lys residues of the Walker A motif (Lys-212
Structure-Function Analysis of ClpB17616
by guest on December 27, 2015http://www.jbc.org/Downloaded from
and Lys-611 of ClpB) contact the phosphate groups of the
and
nucleotides in the first and second AAA-domain, respectively
(19). The conserved Glu residues of the Walker B motif (Glu-
279 and Glu-678 of ClpB) are proposed to represent the cata-
lytic base for ATP hydrolysis in the first and second AAA-
domain, respectively. AAAproteins additionally contain a
conserved Arg residue, termed sensor 2 (2). Sensor 2 lies in the
C-terminal domain of each AAA module and contacts the
-phosphate of bound ATP. This Arg residue is part of an
invariant GAR motif (
813
GAR
815
in ClpB) within the SSD do-
main of AAA(Hsp100) proteins. It is proposed that the con-
served arginine can sense the nucleotide status and mediate
conformational changes of the C-terminal domain relative to
the core domain during ATP hydrolysis (19, 20).
The formation of hexameric rings in AAAproteins brings
residues of adjacent protomers into close proximity to ATP,
bound to a neighboring subunit. Such interactions could pro-
vide the basis for an intermolecular catalytic mechanism, re-
sulting in cooperative ATP hydrolysis within the oligomer.
Previous studies on the AAA protein FtsH have demonstrated
that Arg residues (Arg-332 and Arg-756 in ClpB) can poten-
tially act as Arg fingersby contacting the ATP bound to a
neighbored subunit (21). All described motifs, involved in ATP
binding and hydrolysis, were subjected to alanine mutagenesis,
as illustrated in Fig. 1. ClpB mutants and deletion variants
were purified and characterized with respect to oligomeriza-
tion, ATP hydrolysis, and chaperone activity.
Oligomerization of ClpB VariantsClpB proteins form hex-
americ ring structures in the presence of ATP (22). Oligomer-
ization of ClpB mutants and deletion variants was followed by
size exclusion chromatography in the absence or presence of
ATP and ADP. Consistent with published data, ClpB is only
capable of hexamerization in the presence of ATP (Fig. 2A). In
the absence of nucleotide, ClpB eluted predominantly as a
dimer, although ADP addition led to the formation of trimers
and/or tetramers. Analysis of ClpB deletion variants revealed
that isolated AAA-domains (1409, 1567, and 551857) can-
not oligomerize. In agreement with previous data (23), hexa-
merization is strictly dependent on the presence of the SSD
domain, although the N-domain and the linker region are dis-
pensable (Table I). Characterization of ClpB point mutants
revealed that besides the C-terminal SSD domain, the first
ATPase domain also contributes to oligomerization: mutations
in the Walker A motif (K212A) or the Arg finger (R332A) of the
first AAA-domain abolished hexamerization, although corre-
sponding mutations in the second AAA module (K611A and
R756A) domain had no effect (Fig. 2Band Table I). Finally, all
Walker B mutants in the two AAA modules (E279A, E678A,
and E279A/E678A), as well as the sensor 2 mutant (
813
AAA
815
),
did not exhibit oligomerization defects.
In an additional approach, oligomerization characteristics
were studied by glutaraldehyde cross-linking. Fast cross-link-
ing of ClpB monomers to oligomeric species was obtained
within 10 min in the absence of nucleotide, indicating that
ClpB assembly can in principle also occur without nucleotide.
Because such nucleotide-independent hexamerization was not
observed during gel filtration runs, the formed oligomers seem
to be unstable in the absence of ATP.
Kinetic analysis of the cross-linking reaction revealed that a
ladder of cross-linking products preceded the formation of the
fully cross-linked oligomer (data not shown). The presence of
ATP or ADP accelerated the cross-linking reaction and also
slightly changed the size of the fully cross-linked oligomer.
Whereas the ladder of cross-link products could be followed up
to a heptamer in the absence of nucleotide, addition of ATP and
ADP resulted predominantly in cross-linking to the hexameric
and pentameric species, respectively (Fig. 3A). The existence of
heptameric ClpB rings have also been shown by Chung and
co-workers (18) using electron microscopy.
Cross-linking studies of ClpB mutants and deletion variants
confirmed the findings obtained by gel filtration analysis; how-
ever, qualitative differences with respect to the oligomerization
deficiencies were revealed. Isolated AAA-domains (1409 and
551857) stayed as monomers in presence of glutaraldehyde,
whereas in case of the ClpB SSD variant (aa 1758) and a
longer version of the first AAA module (aa 1567) some dimeric
and trimeric species were observed (Fig. 3B). Mutating the
Walker A motif (K212A) or the Arg finger (R332A) of the first
AAA-domain resulted in the formation of mixtures of oligo-
meric species, ranging from monomers to tetramers in case of
K212A and monomers to hexamers for R332A (Fig. 3Aand
Table I). Interestingly, the observed oligomerization defects
were nucleotide-independent, indicating that the mutated res-
idues are also important for subunit interactions within the
oligomer in the absence of ATP. The involvement of these
charged residues in ClpB assembly can also explain the ob-
served salt sensitivity of ClpB and Hsp104 oligomerization (24,
25). All other ClpB variants exhibited cross-linking character-
istics indistinguishable from wild type ClpB with exception of
ClpB-(410 532) that misses the linker region. This variant
exhibited only a ladder of cross-linking products in the absence
of nucleotide and required ATP for full oligomerization. The
observed assembly defect was, however, not primarily caused
by the absence of the linker region but rather by the introduc-
tion of additional amino acids at each boundary of the linker
segment, resulting from the construction strategy of this dele-
tion variant (see Experimental Procedures). Thus a control
construct (termed ClpB LBL), carrying the same additional
amino acids and the reinserted linker region, also exhibited
the same oligomerization defects as ClpB-(410 532) (data
FIG.1.Proposed domain organization of E. coli ClpB and conserved motifs, involved in ATP binding and ATP hydrolysis. ClpB
consists of an N-terminal domain, two nucleotide binding domains (ATP-1 and ATP-2), which are separated by a linker region, and a C-terminal
SSD. Both nucleotide binding domains contain the Walker A (
206
GX
4
GKT
213
and
605
GX
4
GKT
612
) and Walker B (
276
Hy
2
DE
279
and
675
Hy
2
DE
678
)
motifs, where Xany amino (aa) and Hy hydrophobic amino acids. Conserved Arg residues (R) present in both nucleotide binding domains
(Arg-332 and Arg-756) are proposed to serve as Arg fingers, contacting the ATP bound to an adjacent subunit. The invariant sensor 2 motif
(
813
GAR
815
) of the SSD of Hsp100 proteins potentially sense the nucleotide status of ATP-2. Conserved residues of these motifs (marked in
boldface) were subjected to alanine mutagenesis, as indicated by arrows.
Structure-Function Analysis of ClpB 17617
by guest on December 27, 2015http://www.jbc.org/Downloaded from
not shown), showing that the linker is not essential for
oligomerization.
Finally in a complementary approach we examined the in-
trinsic fluorescence of ClpB as a potential tool to monitor the
oligomeric state. ClpB contains 2 tryptophan residues (Trp-462
and Trp-543). We determined whether tryptophan fluorescence
changes in a nucleotide-dependent fashion. In the absence of
nucleotide, ClpB exhibited a fluorescence emission maximum
at 350.5 nm. Addition of ATP or ADP caused a blue shift in the
fluorescence maximum to 346.5 nm, and a stronger decline of
fluorescence intensity beyond the maximum was observed (Fig.
4A). This change was more pronounced in the presence of ATP
compared with ADP. Several findings suggest that the ob-
served changes in fluorescence are caused by oligomerization of
ClpB. First, the blue shift was also observed when ClpB was
incubated in low salt buffer in the absence of nucleotide (Fig. 4).
Such buffer conditions have been shown for several ClpB ho-
mologues to stabilize the oligomer in the absence of nucleotide
(25, 26). Consistently, the ATP-dependent blue shift was re-
verted under high ionic strength conditions (addition of 100 mM
(NH
4
)
2
SO
4
or 300 mMNH
4
Cl), conditions which have been
shown to inhibit ClpB activity (10). De-oligomerization of ClpB
in the additional presence of 100 mM(NH
4
)SO
4
or 300 mM
NH
4
Cl was also observed in cross-linking experiments. Such
high salt conditions resulted in a strong reduction of ClpB
cross-linking efficiency in the absence or presence of ATP (Fig.
4Band data not shown).
Second, the observed changes in fluorescence also became
nucleotide-independent with increasing ClpB concentrations; a
complete blue-shift was revealed in presence of 4
MClpB (data
not shown). Finally, the analysis of ClpB mutants and deletion
variants revealed a direct correlation between the ability to
form hexamers and the observed changes in tryptophan fluo-
rescence; the ATP-dependent blue shift was not or was only
partially observed for ClpB variants with oligomerization de-
fects (Table I).
In order to dissect the individual contributions of the two
tryptophan residues to the changes in the fluorescence spec-
trum, a ClpB variant (W543F) with only a single tryptophan
residue (Trp-462) was constructed. Although this variant was
not affected in oligomerization and exhibited full chaperone
activity in vitro, nucleotide-dependent changes in ClpB fluores-
cence were no longer observed (data not shown). We conclude
that conformational changes in the close vicinity of Trp-543,
driven by oligomerization, must be responsible for the observed
blue shift in tryptophan fluorescence of ClpB.
Nucleotide-dependent Conformational Changes within ClpB
Revealed by Partial ProteolysisThe ability of Hsp100 mu-
tants to form oligomers is commonly used as a criteria for their
structural integrity. We additionally looked for conformational
changes of wild type ClpB in response to nucleotides, and we
checked whether such structural rearrangements were pre-
served in the constructed ClpB variants. ClpB was subjected to
limited proteolysis by thermolysin or subtilisin in the absence
or presence of nucleotides. In the absence of nucleotides two
highly stable fragments were recovered (Fig. 5A). Fragments
were N-terminally sequenced and identified by mass spectrom-
etry (Fig. 5B). The larger degradation products (aa 3331 and
3351) corresponded to the N-terminal domain of ClpB and the
core domain of the first AAA module. The second fragment (aa
536 756) represented the core domain of the second AAA mod-
ule. Addition of ATP or ADP slowed the proteolysis consider-
ably and resulted in stabilization of full-length ClpB and the
occurrence of two other stable cleavage products. The first
fragment (aa 3331 and 3351) was already obtained by cleav-
age in the absence of nucleotide, although the second fragment
(aa 537857 and 551857) corresponded now to the complete
second AAA module. In the presence of ATP
S cleavage was
further limited, and reduced amounts of the smaller fragments
were obtained (Fig. 5A).
Partial proteolysis of ClpB variants revealed that nucleotide-
dependent stabilization of wild type ClpB can be attributed to
hexamer formation. Mutants with oligomerization deficiencies
(K212A, ClpB SSD) or reduced stability of hexamers (ClpB-
(410 532)) were degraded more quickly in the presence of
ATP compared with WT ClpB and thus did not exhibit stabili-
zation of the full-length version (Fig. 5C). The observed defects
of ClpB-(410 532), missing the linker region, in ATP-depend-
ent stabilization could again be attributed to the presence of
the two additional amino acids at the domain boundary be-
cause the control construct (ClpB LBL) with the reinserted
linker region exhibited the same reduced proteolytic stability.
Isolated AAA-domains (1409 and 551857) did not react to
ATP addition and were processed rapidly to stable cleavage
products (similar to the fragments obtained for WT ClpB in the
FIG.2.Oligomerization of ClpB wild type and mutant deriva-
tives analyzed by gel filtration chromatography. A, elution pro-
files of ClpB wild type (WT) were recorded in the absence ()or
presence of nucleotides (2 mMATP/ADP) in the running buffer (50 mM
Tris, pH 7.5, 20 mMMgCl
2
, 150 mMKCl, 10% (v/v) glycerol). Elution
positions of protein standards (thyroglobulin (669 kDa), ferritin (440
kDa), catalase (232 kDa), and IgG (158 kDa)) are given. B, elution
profiles of the indicated ClpB mutants were recorded in the presence of
ATP (2 mM) in the running buffer.
Structure-Function Analysis of ClpB17618
by guest on December 27, 2015http://www.jbc.org/Downloaded from
absence of nucleotides). In contrast, a longer version of the first
AAA-domain (1567) in the presence of ATP exhibited some
protection from degradation. Further structural analysis of
ClpB point mutants surprisingly revealed that conformational
changes within each AAA-domain can occur independently of
structural deficiencies in the other AAA module. Thus, the
Walker A mutant K212A of the first AAA-domain, although
deficient in oligomerization, still exhibited ATP-dependent con-
formational changes in the second AAA-domain (stabilization
of the second AAA module). On the other hand structural
changes within the second AAA-domain were not or were only
partially observed in the case of K611A and
813
AAA
815
, ClpB
mutants that nevertheless showed stabilization of the full-
length protein (Fig. 5C). Thus oligomerization protects full-
length ClpB even in case of mutants that do not bind nucleotide
tightly at the second AAA module (K611A;
813
AAA
815
). We
conclude that stabilization of full-length ClpB against proteo-
lytic degradation primarily reflects binding of ATP to the first
AAA-domain. These data also demonstrate that the ability to
form oligomers is not per se a sufficient criteria for probing the
structural integrity of the ClpB mutants. Structural integrity
with respect to oligomerization and stability during partial
proteolysis was only completely preserved in Walker B mu-
tants of ClpB, although Walker A and the sensor 2 mutants
exhibited significant structural deficiencies. Interestingly, the
double Walker B mutant (E279A/E678A) was much more re-
TABLE I
Oligomerization of ClpB fragments and mutants
Assembly of the indicated ClpB derivatives was analyzed in the presence of 2 mMATP by gel filtration chromatography, cross-linking, and
tryptophan fluorescence. Size of ClpB oligomers were calculated from a calibration curve with protein standards. Cross-link products formed after
10 min of incubation with glutaraldehyde were analyzed by SDS 4 12% PAGE. Most populated species are underlined. ATP-dependent
oligomerization of ClpB is coupled to a blue shift in the maximum of fluorescence (346.5 nm instead of 350.5 nm in the absence of nucleotide). The
maximum of ClpB fluorescence is indicated. ND, not determined.
Gel filtration
chromatography Cross-linking products Maximum of ClpB
fluorescence
ClpB fragments
1857 (wild type) Hexamer Hexamers 346.5
143857 (N) Hexamer Hexamers 346.5
1758 (SSD) Monomer Monomers-trimers 351
410532 (linker) Hexamer Hexamers 348.5
1409 Monomer Monomer ND
1567 Monomer Monomer ND
551857 Monomer Monomers-dimers ND
ClpB mutants
K212A (Walker A) Dimer Monomers-tetramers 351
K611A (Walker A) Hexamer Hexamers 347
K212A/K611A Dimer Monomers-tetramers 351
E279A (Walker B) Hexamer Hexamers 346.5
E678A (Walker B) Hexamer Hexamers 346.5
E279A/E678A Hexamer Hexamers 346.5
R332A (Arg finger) Dimer Monomer-hexamers 348.5
R756A (Arg finger) Hexamer Hexamers 346.5
813AAA815 (sensor 2) Hexamer Hexamers 346
FIG.3.Assembly of ClpB wild type
and derivatives revealed by cross-
linking. A,1
MClpB wild type (WT) and
the indicated point mutants were incu-
bated for 5 min at 30 °C in the absence of
nucleotides (lane 2) or the presence of 2
mMATP (lane 3) or ADP (lane 4). Cross-
linking reactions were initiated by addi-
tion of glutaraldehyde and proceeded for
10 min. Cross-linked proteins were sepa-
rated on SDS (4 12%)-polyacrylamide
gels, followed by silver staining. Incuba-
tion of ClpB proteins in the absence of
cross-linker (lane 1) served as control. B,
ClpB deletion variants were incubated in
the absence (lane 2) or presence of 2 mM
ATP (lane 3) for 5 min at 30 °C. Cross-
linking was performed as described
above. Incubation of truncated ClpB spe-
cies in the absence of cross-linker (lane 1)
served as control.
Structure-Function Analysis of ClpB 17619
by guest on December 27, 2015http://www.jbc.org/Downloaded from
sistant toward proteolysis than the single Walker B mutants
and ClpB wild type. Because this mutant is deficient in ATP
hydrolysis (see below), the occurrence of the other fragments
(aa 3331 and 3351 and 537857 and 551857) in ClpB wild
type (and various ClpB mutants) can be attributed to ATP
hydrolysis occurring during the digestion reaction. These data
also indicate that ClpB therefore adopts a different conforma-
tion if both AAA-domains have ATP bound, as compared with
the situation where only one AAA-domain has ATP bound and
the other domain has ADP bound.
ATPase Activities of ClpB Mutants and Deletion Variants
ATPase activities of ClpB variants were determined in the
absence or presence of the artificial substrate casein, which is
known to stimulate ATP hydrolysis by ClpB (27). Wild type
ClpB exhibited an ATPase rate of 0.021/s, and the ATPase
activity was increased by 34-fold in the presence of saturating
concentrations of
-casein (20-fold excess over ClpB mono-
mers). ClpB N had a slightly increased ATPase activity but
was less stimulable by casein, in agreement with published
data (23, 28). Removal of the linker region (ClpB-(410 532))
strongly decreased the basal ATPase activity by 4-fold, al-
though stimulation by casein was not affected. Similar results
were obtained with the control construct ClpB LBL, bearing a
reinserted linker, indicating that the reduced basal ATPase
activity is primarily caused by the observed oligomerization
deficiencies. All other deletion variants (aa 1409, 1567, 551
857, and 1758) with even stronger defects in oligomerization
did not exhibit any ATPase activity, even in presence of casein,
indicating that ATP binding and/or hydrolysis is strictly linked
to the formation of hexamers. Consistent with this hypothesis,
the basal ATPase activities of mutants with oligomerization
defects were sensitive to the buffer conditions; ATP activities of
K212A and Arg-332 were enhanced 23-fold in low salt buffer,
conditions that favor oligomerization (data not shown). In
agreement with these findings, addition of 100 mM(NH
4
)
2
SO
4
or 300 mMNH
4
Cl diminished the ATPase rate of full-length
ClpB or ClpB N 3- and 5-fold, respectively (Table II, data not
shown). Further analysis of Walker A and Walker B single
mutants, bearing only one functional AAA-domain, revealed
that especially the first AAA-domain was sensitive toward high
FIG.4.A, nucleotide-dependent blue shift in the wavelength of max-
imum ClpB fluorescence. Emission spectra of tryptophan fluorescence
of ClpB (0.5
M) in the absence of nucleotide (black line) or the presence
of2m
MATP (gray line) or ADP (thin black line) in buffer A (50 mMTris,
pH 7.5, 150 mMKCl, 20 mMMgCl
2
,2mMDTT) are shown. Additionally,
ClpB fluorescence was recorded in low salt buffer (buffer A without 150
mMKCl) (black dotted line) or in the presence of 2 mMATP in high salt
buffer (buffer A plus 100 mM(NH
4
)
2
SO
4
)(gray dotted line). The maxi-
mum values in the wavelength of tryptophan fluorescence are indi-
cated. B, high salt concentrations inhibit cross-linking of ClpB to oligo-
meric species. Cross-linking of ClpB (1
M) by glutaraldehyde was
performed in the absence or presence of ATP (2 mM) and/or (NH
4
)
2
SO
4
(100 mM). Cross-linked proteins were separated on SDS (4 12%)-poly-
acrylamide gels, followed by silver staining.
FIG.5. Nucleotide-dependent structural changes in ClpB re-
vealed by limited proteolysis. A, partial proteolysis of ClpB wild
type by thermolysin and subtilisin was performed in the absence of
nucleotide or in the presence of 2 mMATP/ADP/ATP
S for 60 min.
Cleavage products were separated on SDS (15%)-polyacrylamide gels,
followed by silver staining. Cleavage sites were identified by N-terminal
sequencing and mass spectrometry. B, proposed domain organization of
ClpB. Domains were defined from sequence analysis, secondary struc-
ture prediction, and results obtained from partial proteolysis. ClpB
domains are as follows: N-domain (1144), ATP-1 (144 340), C-1 (341
409 and 525550), the linker region (410 525), ATP-2 (550755), and
C-2 (755857). ATP-1/2 corresponds to the core domains of the AAA
modules. C-1/2 represents the C-terminal helical domains of the AAA
modules. Sites of cleavage by thermolysin and subtilisin are indicated
(open arrows, cleavage site in absence of ATP; filled arrows, cleavage
sites in presence of nucleotides). C, partial proteolysis of ClpB mutants
and fragments by thermolysin. Proteins were incubated in the absence
or presence of ATP for 60 min. Cleavage products were separated on
SDS (15%)-polyacrylamide gels, followed by silver staining.
Structure-Function Analysis of ClpB17620
by guest on December 27, 2015http://www.jbc.org/Downloaded from
ionic strength conditions. Addition of 100 mM(NH
4
)
2
SO
4
strongly reduced the ATPase activity of mutants, in which the
second AAA-domain was inactivated (K611A and E678A),
whereas variants with only an active second AAA module
(E279A and K211A) were only partially affected (Table II),
thereby underlining the functional importance of the first AAA-
domain for ClpB assembly. Higher concentrations of ammo-
nium sulfate (200 mM) completely inhibited ATP hydrolysis by
ClpB (data not shown).
Single Walker A or Walker B mutations resulted in a com-
plete loss of ATPase activity in the corresponding AAA-domain,
because double Walker A or Walker B mutants were deficient
in ATP hydrolysis, even in presence of casein. Because signif-
icant ATPase activities were measurable for all Walker A and
Walker B single mutants, both ATPase domains seemed to
contribute to the basal ATPase activity of ClpB. Additionally,
both AAA modules were stimulated to similar degrees by ca-
sein. Variations in the basal ATPase activities of ClpB point
mutants compared with wild type ClpB indicate a communica-
tion between both AAA-domains. The basal ATPase rate of
ClpB E678A was increased 4-fold compared with wild type.
However, analysis of other ClpB mutants revealed that no easy
conclusions with respect to the directionality of the communi-
cation between the AAA-domains can be drawn. Mutating the
Arg finger of the second AAA module (R756A) resulted in a
nearly complete loss of ATPase activity, although the overall
structural integrity was preserved as for the single Walker B
mutant (E678A). The signaling between both AAA-domains
seemed to be rather complex, and subtle conformational
changes within ClpB mutants can obviously have completely
different consequences on the other AAA module.
The mechanism of ATPase stimulation by casein is un-
known. Because ATP hydrolysis by ClpB is sensitive to its
oligomeric status, we tested the possibility that interaction of
ClpB with casein stabilizes the hexameric state, thereby in-
creasing the ATPase activity. We compared the ATPase activ-
ity of ClpB concentration at different protein concentrations
both in the absence and presence of casein. In the absence of
substrate, low ClpB concentrations (0.10.2
M) had no activity
or a strongly reduced specific ATPase activity (Fig. 6). Increas-
ing ClpB concentrations activated ATP hydrolysis in a cooper-
ative manner as revealed by the sigmoidal shape of the curve.
In contrast, in the presence of casein high ATPase activities
were already determined at low ClpB concentrations (0.1
M),
implying that stabilization of ClpB hexamers by substrates
causes induction of the ATPase activity. Such stabilization by
casein was indeed observed by gel filtration analysis for the
ClpB Walker B mutant E279A/E678A, which is deficient in
ATP hydrolysis.
2
Besides stabilization of the oligomeric state,
conformational changes within ClpB, triggered by substrate
binding, may also contribute to the stimulation of ATP hydrol-
ysis, because saturating ClpB concentrations (1
M) still
showed a lower specific ATPase activity in the absence of casein
(Fig. 6).
Chaperone Activities of ClpB DerivativesWe tested the
chaperone activities of the various ClpB variants in vitro by
following the ClpB/DnaK-dependent reactivation of heat-ag-
gregated MDH and firefly luciferase. Resolubilization of MDH
aggregates by ClpB/DnaK was directly followed by measuring
the decrease of aggregate turbidity in light scattering experi-
ments. Disaggregation of aggregated luciferase was followed by
determining the refolding rate of luciferase in the presence of
the bi-chaperone system. All disaggregation reactions were
performed in the presence of non-saturating ClpB concentra-
tions which allowed for sensitive detection of potential activity
defects. Results are summarized in Table III.
All ClpB deletion variants (aa 1409, 1567, 551857, and
1758) with severe oligomerization defects were inactive in
MDH and luciferase disaggregation. Deletion of the linker re-
gion (ClpB-(410 532)) also caused a complete loss of the dis-
aggregation activity of ClpB. Importantly, this inactivation was
not caused by the observed oligomerization deficiency of ClpB-
(410 532), because the control construct ClpB LBL with the
reinserted linker region exhibited significant disaggregation
activity (60% of ClpB wild type; data not shown). ClpB N had
the same chaperone activity as full-length ClpB. Full disaggre-
gation activity of ClpB N was also observed when the ClpB
concentrations in the activity assays were further reduced,
thereby increasing the dependence of the disaggregation on
ClpB (data not shown).
ClpB variants with point mutations in conserved motifs ex-
hibited none or only partial chaperone activity (Table III). In
general a disaggregation activity was only observed for mu-
tants that are still able to form oligomers. Thus the Walker A
(K212A) and Arg finger (R332A) mutations of the first AAA-
domain, which are deficient in hexamer formation, were inac-
2
J. Weibezahn, C. Schlieker, B. Bukau, and A. Mogk, manuscript
in preparation.
TABLE II
ATPase activities of ClpB fragments and mutants
ATPase activities of the indicated ClpB derivatives were determined
in buffer A. The factors of ATPase stimulation in the presence of 0.25
mg/ml
-casein and the factors of ATPase inhibition in the presence of
100 mMammonium sulfate are given. ND, not determined.
ATPase
activity
(1/s)
Factor of
ATPase
stimulation
by
-casein
Factor of
ATPase
inhibition
by sulfate
ClpB fragments
1857 (wild type) 0.021 3.3 3
143857 (N) 0.037 1.9 4.9
1758 (SSD) 0 0 ND
410532 (linker) 0.005 6 ND
1409 0 0 ND
1567 0 0 ND
551857 0 0 ND
ClpB mutants
K212A (Walker A) 0.006 4.4 1.4
K611A (Walker A) 0.019 5.3 4.2
K212A/K611A 0 0 ND
E279A (Walker B) 0.023 3.8 1.7
E678A (Walker B) 0.089 2.7 5.3
E279A/E678A 0 0 ND
R332A (Arg finger) 0.019 2.4 ND
R756A (Arg finger) 0.003 0 ND
813AAA815 (SSD motif) 0.036 2.8 ND
FIG.6. The ATPase activity of ClpB depends on protein con-
centration and protein substrates. The specific ATPase activity of
ClpB was measured at different protein concentrations in the absence
or presence of
-casein (0.25 mg/ml).
Structure-Function Analysis of ClpB 17621
by guest on December 27, 2015http://www.jbc.org/Downloaded from
tive. In contrast, the corresponding mutations in the second
AAA-domain (K611A and R756A) exhibited low chaperone ac-
tivities to varying degrees (below 10% of the WT control).
Partial activities were also obtained in case of single Walker B
mutants, although the double mutant (E279A/E678A) did not
have any disaggregation activity. These findings indicate that
whereas ATP hydrolysis is strictly required for chaperone ac-
tivity, a lack of ATP hydrolysis in one AAA-domain does not
completely inhibit the disaggregation activity.
Disaggregation activities of all ClpB variants were also
tested in vivo by complementation of the phenotypes of clpB
mutant cells by IPTG induction of plasmid-encoded clpB mu-
tant alleles. Resolubilization of protein aggregates, formed dur-
ing severe heat stress to 45 °CinE. coli cells, is severely
affected in clpB mutants. This deficiency is directly linked to
a strongly reduced survival rate of clpB mutant cells at lethal
(50 °C) temperatures compared with wild type cells (thermo-
tolerance). We therefore tested the ClpB variants for their
ability to re-establish protein disaggregation and thermotoler-
ance in clpB mutant cells (Fig. 7). Complementation studies
were performed in the presence of 25
MIPTG, leading to
34-fold increased ClpB levels as compared with heat-shocked
wild type cells lacking any plasmid (data not shown). In sum-
mary the in vivo disaggregation activities reflected those ob-
served in vitro. Only wild type ClpB and the ClpB N could
efficiently mediate the resolubilization of protein aggregates
and the development of thermotolerance (Fig. 7 and Table III).
Constructs with partial chaperone activities in vitro, such as
E279A, also exhibited some activity in vivo.
DISCUSSION
Here we present a detailed structure-function analysis of the
AAAchaperone ClpB, which mediates resolubilization of pro-
tein aggregates in cooperation with the DnaK chaperone sys-
tem. Oligomerization of ClpB was dependent on both AAA-
domains; however, each domain seemed to play a different role
in ClpB assembly. Deletion of the
-helical C-domain of the
second AAA-domain (ClpB SSD) resulted in a severe defect in
oligomerization defect. ClpB SSD stayed predominantly mon-
omeric even in cross-linking experiments. On the other hand
we could demonstrate that ATP binding to the first AAA-
domain was also necessary for stabilizing the ClpB hexamer.
Several observations support this model. First, the Walker A
(K212A) and Arg finger (R332A) mutants of the first AAA
module did not form stable hexamers in the presence of ATP,
whereas corresponding mutations in the second nucleotide
binding domain (K611A and R756A) had no influence on ClpB
oligomerization, although these mutants exhibited severe
structural deficiencies. Second, the introduction of amino acids
into the C-domain of the first AAA module (ClpB-(410 532)
and ClpB LBL) also resulted in destabilization of ClpB oli-
gomers, underlining the importance of this domain in hexa-
merization. We also could show that high salt conditions (ad-
dition of 100 mM(NH
4
)
2
SO
4
) cause dissociation of ClpB
oligomers by interfering predominantly with subunit contacts
between the first AAA-domains. Interestingly, the recently de-
termined structure of monomeric ClpA suggests that electro-
static interactions between the first AAA-domains of ClpA
could play a much more important role than the second AAA
modules in protein oligomerization (29). These findings might
also explain the observed salt sensitivity of ClpB oligomeriza-
tion and underline the functional importance of the first AAA-
domain in this process. Finally, the observed changes in fluo-
rescence of Trp-543, located in the C-domain of the first AAA
module, revealed a conformational rearrangement of this re-
gion in response to nucleotides. The observed blue shift (4 nm)
TABLE III
Chaperone activities of ClpB mutants and fragments
Chaperone activity (disaggregation rate of aggregated MDH; refold-
ing rate of aggregated luciferase) of ClpB wild type was set as 100%. In
vivo activity reflects the ability of ClpB derivatives to restore thermo-
tolerance and protein disaggregation in E. coli clpB mutant cells (,
full activity; (), residual activity; , no activity). ND, not determined.
Disaggregation
of aggregated
MDH
Refolding of
aggregated
luciferase
In vivo
activity
% activity
ClpB fragments
1857 (wild type) 100 100
143857 (N) 98 99
1758 (SSD) 0 0
410532 (linker) 0 0
1409 0 0 ND
1567 0 0 ND
551857 0 0 ND
ClpB mutants
K212A (Walker A) 0 0
K611A (Walker A) 45810 ()
K212A/K611A 0 0
E279A (Walker B) 79810 ()
E678A (Walker B) 2345()
E279A/E678A 0 0
R332A (Arg finger) 0 0
R756A (Arg finger) 0.51 0.51
813AAA815 (SSD motif) 0 0
FIG.7. In vivo activity of ClpB derivatives. A,E. coli wild type
(MC4100), clpB mutant cells, and clpB mutants, bearing plasmid-
encoded clpB alleles under the control of an IPTG-regulatable promo-
tor, were grown in LB medium at 30 °C in the presence of 25
MIPTG
to mid-exponential phase. Subsequently strains were heat-shocked to
50 °C and incubated for the indicated time. Various dilutions of the
stressed cells were spotted on LB plates and incubated at 30 °C. After
24 h, colony numbers were counted, and survival rates were calculated
in relation to unstressed cells. B, cultures of clpB mutant strains,
bearing plasmid-encoded clpB alleles under the control of an IPTG-
regulatable promotor, were grown in LB medium in the presence of 25
MIPTG at 30 °C to logarithmic phase. Cells were then shifted to 45 °C
for 30 min, followed by a recovery phase at 30 °C for 60 min. Aggregated
proteins were isolated before (0 min) and after the recovery phase (60
min) and analyzed by SDS-PAGE followed by staining with Coomassie
Brilliant Blue.
Structure-Function Analysis of ClpB17622
by guest on December 27, 2015http://www.jbc.org/Downloaded from
in the wavelength for maximum fluorescence could be attrib-
uted to ClpB oligomerization. We suggest that changes in Trp-
543 fluorescence reflect interactions of the first C-domain not
only with itself but also adjacent ATPase domains, thereby
leading to increased shielding of Trp-543 and stabilization of
the oligomer. Nucleotide-dependent conformational changes of
the C-terminal
-helical domain was also demonstrated by
limited proteolysis. Whereas the C-domain was rapidly de-
graded in the absence of nucleotides, it became largely resist-
ant to proteases upon nucleotide addition (Fig. 5B). Similarly
the C-domain of the second AAA module was also protected by
addition of nucleotides. This protection is likely to be caused by
the interaction of C-domains with their own AAA-domain and
that of adjacent subunits, thereby becoming less accessible to
proteases. Consistently, ClpB mutants with defects in nucleo-
tide binding (K212A and K611A;
813
AAA
815
) did not exhibit
stabilization of the corresponding C-domains.
A functional importance of the first ATPase domain for
Hsp100 oligomerization has also been reported for E. coli ClpA
(30, 31) and ClpB from Thermus thermophilus (32). It is in-
triguing that the contributions of the individual AAA-domains
differ in the ClpB homologues Hsp104 and Hsp78 from Saccha-
romyces cerevisiae. Here mutations of the Lys residue in the
Walker A motif of the second AAA-domain resulted in a loss of
ability to form hexamers (3335).
Hexamerization of ClpB is a prerequisite for its chaperone
activity. Because ATP is bound at the interface of two neigh-
boring subunits in the hexamer, oligomerization additionally
influences the ATPase activity of ClpB. Consequently high salt
conditions inhibit ATP hydrolysis by ClpB. Furthermore, ClpB
variants with defects in oligomerization exhibited little or no
ATPase activity. On the other hand, the stimulation of ATP
hydrolysis by substrates, such as
-casein, is most likely be-
cause of the stabilization of ClpB oligomers.
Mutations of conserved Lys and Glu residues in the Walker
A and Walker B motifs of both AAA modules completely abol-
ished ATP hydrolysis by the corresponding AAA-domain. Be-
cause single mutants still retained significant ATPase activity,
both AAA-domains seem to contribute to the basal rate of ATP
hydrolysis by ClpB. Interestingly, ClpB homologues from dif-
ferent organisms differ significantly in their ATPase cycle.
Although both AAA-domains of ClpB from T. thermophilus (26,
32) also contribute to the basal ATPase activity, ATP hydroly-
sis by the yeast homologue Hsp104 is dominated by the first
AAA module (25). Such differences may potentially explain the
observed species specificity in the cooperation of ClpB/Hsp104
proteins with the corresponding Hsp70 partners (36). Varia-
tions in the basal ATPase activity of ClpB mutants indicate
communication between both AAA-domains. However, differ-
ent mutations in the same AAA-domain (E678A and R756A)
exhibited different influences on the ATPase activity of the
other AAA module. The signaling between both AAA-domains
appears to be rather complex, and subtle conformational
changes within ClpB mutants can have completely different
consequences on the other AAA module.
The function of the N-domain still remains enigmatic. Iso-
lated N-domains from ClpA and ClpB have been shown to form
stable, monomeric domains, which are probably separated from
the AAA-domains by a flexible linker (7, 37, 38). Consistent
with this model, N-domains are not essential for oligomeriza-
tion of ClpA or ClpB (23, 37, and this work). The connector is
apparently sterically inaccessible to proteases because partial
proteolysis of ClpB did not release significant amounts of iso-
lated N-domains but rather produced a fragment comprising
the N-domain and the core domain of the first AAA module (aa
1353). Similar findings have been reported for ClpA (37). The
reported consequences of N-terminal deletions of ClpB on its
chaperone activity are contradictory. Zolkiewski and co-work-
ers (23) showed that a ClpB variant, starting from the internal
start site (aa 149 857), was completely inactive in refolding of
aggregated luciferase. In contrast, a ClpB N-(143857) vari-
ant showed the same chaperone activity as wild type ClpB both
in vitro and in vivo (resolubilization of protein aggregates and
development of thermotolerance). In agreement with these
data, an N-terminal truncated ClpB derivative of Synechococ-
cus sp. PCC7942 conferred the same degree of thermotolerance
in vivo as full-length ClpB; likewise, the N-terminal truncated
version of T. thermophilus ClpB was also shown to be active in
protein disaggregation in vitro (39, 40). The existence of ClpB
homologues completely lacking an N-domain in Mycoplasma
sp. also argues against an essential function of the N-domains
in the disaggregating activities of ClpB (41, 42).
What might be the function of the N-domain with respect to
these conflicting results? N-domains have been proposed to
mediate substrate binding; however, the reported activities of
N-terminally truncated ClpB variants clearly rule out an es-
sential function in this process. Interestingly, defects in sub-
strate binding were also not the basis of the inactivation of
ClpB-(149 857), because this deletion variant interacted with
casein and unfolded luciferase indistinguishable from ClpB
wild type (38). N-domains may therefore be involved in a mech-
anism for coupling the ATPase cycle of ClpB with its proposed
unfolding activity. Because a longer deletion of the N-domain,
including parts of the flexible linker to the first AAA-domain,
has been reported to be inactive, this linker could potentially
contribute to induced structural changes in bound substrates.
Alternatively, N-domains may be involved in other ClpB activ-
ities, which are so far unknown and are not related to protein
disaggregation. Such new activities have been described for a
second ClpB homologue in Synechococcus, which is essential for
cell viability but is not involved in thermotolerance (43). Inter-
estingly, the N-domain of this ClpB variant seems to be crucial
for its unknown activity and may serve as binding sites for
special substrates or, alternatively, for specific adaptor pro-
teins. Binding of adaptor proteins to N-domains has been dem-
onstrated for ClpA (44) and the AAA protein p97 (45, 46).
The linker region of ClpB was originally suggested to sepa-
rate both AAA-domains (1). We propose that the linker instead
interrupts the C-domain of the first AAA-domain as was sug-
gested recently (47) for the yeast homologue Hsp104. A com-
parison of different ClpB fragments (aa 1409 and 1567) with
respect to their oligomerization and their resistance to prote-
olysis supports this model. First, cross-linking studies revealed
that ClpB-(1567) can form dimeric species in contrast to the
shorter variant (aa 1409), which remains monomeric under
all conditions. Additionally, partial proteolysis revealed a more
pronounced protection of the full-length version (aa 1567),
which was not observed for ClpB-(1409). Formation of dimeric
species and the observed partial stabilization is likely to be
caused by the interaction of the helical C-domain, which can
only be formed in ClpB-(1567), with its own and an adjacent
AAA-domain. Interestingly, a very similar domain organiza-
tion has been proposed for ClpA (37). In ClpA, which is missing
the linker region, a short basic loop (KRKK) is also inserted
into the C-domain of the first AAA module.
The linker region of ClpB, in contrast to the N-domain, is
essential for chaperone activity. It is proposed to form a 4 times
repeated coiled-coil (48), which is very likely to play an impor-
tant role in ClpB function. Conformational changes in response
to nucleotides within the linker region were shown by its in-
creased proteolytic stability in the context of full-length ClpB.
We assume that the linker is not an integral part of the ClpB
Structure-Function Analysis of ClpB 17623
by guest on December 27, 2015http://www.jbc.org/Downloaded from
hexamer, because oligomerization was still possible in case of a
ClpB variant missing the linker region (ClpB-(410 532)).
Similarly, the I-domain of HslU, although inserted into the
AAA-domain, forms an independent structural domain and is
exposed at the surface of the oligomer (3, 4). The function of the
linker region is still unknown. The postulated coiled-coil struc-
ture might be involved in protein-protein interaction. Because
the ATPase activity of ClpB-(410 532) was still strongly
stimulable by casein, the linker region cannot serve as a pri-
mary substrate-binding site, at least for this type of substrate.
Alternatively, the linker region is necessary for coupling ATP
hydrolysis and substrate unfolding, as proposed recently by
Lindquist and co-workers (47).
AcknowledgmentsWe thank D. Dougan and K. Turgay for discus-
sions and critical reading of the manuscript. We also thank A. Schulze-
Specking for technical assistance.
REFERENCES
1. Schirmer, E. C., Glover, J. R., Singer, M. A., and Lindquist, S. (1996) Trends
Biochem. Sci. 21, 289 296
2. Neuwald, A. F., Aravind, L., Spouge, J. L., and Koonin, E. V. (1999) Genome
Res. 9, 2743
3. Bochtler, M., Hartmann, C., Song, H. K., Bourenkov, G. P., Bartunik, H. D.,
and Huber, R. (2000) Nature 403, 800 805
4. Sousa, M. C., Trame, C. B., Tsuruta, H., Wilbanks, S. M., Reddy, V. S., and
McKay, D. B. (2000) Cell 103, 633643
5. Li, J., and Sha, B. (2002) J. Mol. Biol. 318, 11271137
6. Smith, C. K., Baker, T. A., and Sauer, R. T. (1999) Proc. Natl. Acad. Sci.
U. S. A. 96, 6678 6682
7. Lo, J. H., Baker, T. A., and Sauer, R. T. (2001) Protein Sci 10, 551559
8. Banecki, B., Wawrzynow, A., Puzewicz, J., Georgopoulos, C., and Zylicz, M.
(2001) J. Biol. Chem. 276, 1884318848
9. Kim, Y. I., Levchenko, I., Fraczkowska, K., Woodruff, R. V., Sauer, R. T., and
Baker, T. A. (2001) Nat. Struct. Biol. 8, 230 233
10. Goloubinoff, P., Mogk, A., Peres Ben Zvi, A., Tomoyasu, T., and Bukau, B.
(1999) Proc. Natl. Acad. Sci. U. S. A. 96, 1373213737
11. Mogk, A., Tomoyasu, T., Goloubinoff, P., Ru¨diger, S., Ro¨der, D., Langen, H.,
and Bukau, B. (1999) EMBO J. 18, 6934 6949
12. Motohashi, K., Watanabe, Y., Yohda, M., and Yoshida, M. (1999) Proc. Natl.
Acad. Sci. U. S. A. 96, 71847189
13. Zolkiewski, M. (1999) J. Biol. Chem. 274, 2808328086
14. Laufen, T., Mayer, M. P., Beisel, C., Klostermeier, D., Reinstein, J., and
Bukau, B. (1999) Proc. Natl. Acad. Sci. U. S. A. 96, 54525457
15. Shevchenko, A., Wilm, M., Vorm, O., and Mann, M. (1996) Anal. Chem. 68,
850 858
16. Wilm, M., Shevchenko, A., Houthaeve, T., Breit, S., Schweigerer, L., Fotsis, T.,
and Mann, M. (1996) Nature 379, 466 469
17. Tomoyasu, T., Mogk, A., Langen, H., Goloubinoff, P., and Bukau, B. (2001)
Mol. Microbiol. 40, 397413
18. Kim, K. I., Cheong, G. W., Park, S. C., Ha, J. S., Woo, K. M., Choi, S. J., and
Chung, C. H. (2000) J. Mol. Biol. 303, 655666
19. Ogura, T., and Wilkinson, A. J. (2001) Genes Cells 6, 575597
20. Nicholls, A., Sharp, K. A., and Honig, B. J. (1991) Proteins 11, 281296
21. Karata, K., Inagawa, T., Wilkinson, A. J., Tatsuta, T., and Ogura, T. (1999)
J. Biol. Chem. 274, 2622526232
22. Zolkiewski, M., Kessel, M., Ginsburg, A., and Maurizi, M. R. (1999) Protein Sci.
8, 1899 1903
23. Barnett, M. E., Zolkiewska, A., and Zolkiewski, M. (2000) J. Biol. Chem. 275,
3756537571
24. Schmid, F. X. (1991) Curr. Biol. 1, 36 41
25. Hattendorf, D. A., and Lindquist, S. L. (2002) EMBO J. 21, 1221
26. Schlee, S., Groemping, Y., Herde, P., Seidel, R., and Reinstein, J. (2001) J. Mol.
Biol. 306, 889 899
27. Woo, K. M., Kim, K. I., Goldberg, A. L., Ha, D. B., and Chung, C. H. (1992)
J. Biol. Chem. 267, 20429 20434
28. Park, J. H., Lee, Y. S., Chung, C. H., and Goldberg, A. L. (1988) J. Bacteriol.
170, 921926
29. Guo, F., Maurizi, M. R., Esser, L., and Xia, D. (2002) J. Biol. Chem. 277,
4674346752
30. Singh, S. K., and Maurizi, M. R. (1994) J. Biol. Chem. 269, 2953729545
31. Seol, J. H., Baek, S. H., Kang, M.-S., Ha, D. B., and Chung, C. H. (1995) J. Biol.
Chem. 270, 80878092
32. Watanabe, Y. H., Motohashi, K., and Yoshida, M. (2002) J. Biol. Chem. 277,
5804 5809
33. Parsell, D. A., Kowal, A. S., and Lindquist, S. (1994) J. Biol. Chem. 269,
4480 4487
34. Schirmer, E. C., Queitsch, C., Kowal, A. S., Parsell, D. A., and Lindquist, S.
(1998) J. Biol. Chem. 273, 15546 15552
35. Krzewska, J., Konopa, G., and Liberek, K. (2001) J. Mol. Biol. 314, 901910
36. Krzewska, J., Langer, T., and Liberek, K. (2001) FEBS Lett. 489, 9296
37. Singh, S. K., Rozycki, J., Ortega, J., Ishikawa, T., Lo, J., Steven, A. C., and
Maurizi, M. R. (2001) J. Biol. Chem. 276, 29420 29429
38. Tek, V., and Zolkiewski, M. (2002) Protein Sci. 11, 11921198
39. Clarke, A. K., and Eriksson, M. J. (2000) J. Bacteriol. 182, 70927096
40. Beinker, P., Schlee, S., Groemping, Y., Seidel, R., and Reinstein, J. (2002)
J. Biol. Chem. 277, 47160 47166
41. Fraser, C. M., Gocayne, J. D., White, O., Adams, M. D., Clayton, R. A.,
Fleischmann, R. D., Bult, C. J., Kerlavage, A. R., Sutton, G., Kelley, J. M.,
et al. (1995) Science 270, 397403
42. Himmelreich, R., Hilbert, H., Plagens, H., Pirkl, E., Li, B. C., and Herrmann,
R. (1996) Nucleic Acids Res. 24, 4420 4449
43. Eriksson, M. J., Schelin, J., Miskiewicz, E., and Clarke, A. K. (2001) J.
Bacteriol. 183, 73927396
44. Dougan, D. A., Reid, B. G., Horwich, A. L., and Bukau, B. (2002) Mol. Cell 9,
673683
45. Rouiller, I., Butel, V. M., Latterich, M., Milligan, R. A., and Wilson-Kubalek,
E. M. (2000) Mol. Cell 6, 14851490
46. Meyer, H. H., Shorter, J. G., Seemann, J., Pappin, D., and Warren, G. (2000)
EMBO J. 19, 21812192
47. Cashikar, A. G., Schirmer, E. C., Hattendorf, D. A., Glover, J. R., Ramakrishnan,
M. S., Ware, D. M., and Lindquist, S. L. (2002) Mol. Cell 9, 751760
48. Celerin, M., Gilpin, A. A., Schisler, N. J., Ivanov, A. G., Miskiewicz, E., Krol,
M., and Laudenbach, D. E. (1998) J. Bacteriol. 180, 51735182
Structure-Function Analysis of ClpB17624
by guest on December 27, 2015http://www.jbc.org/Downloaded from
Bernd Bukau
Strub, Wolfgang Rist, Jimena Weibezahn and
Axel Mogk, Christian Schlieker, Christine
and Chaperone Activity
ClpB in Oligomerization, ATP Hydrolysis,
Conserved Motifs of the AAA+ Chaperone
Roles of Individual Domains and
PROTEIN STRUCTURE AND FOLDING:
doi: 10.1074/jbc.M209686200 originally published online March 6, 2003
2003, 278:17615-17624.J. Biol. Chem.
10.1074/jbc.M209686200Access the most updated version of this article at doi:
.JBC Affinity SitesFind articles, minireviews, Reflections and Classics on similar topics on the
Alerts:
When a correction for this article is posted When this article is cited
to choose from all of JBC's e-mail alertsClick here
http://www.jbc.org/content/278/20/17615.full.html#ref-list-1
This article cites 48 references, 27 of which can be accessed free at
by guest on December 27, 2015http://www.jbc.org/Downloaded from
... As shown by the docking studies, these inhibitors mainly interact with the NBD1 domain of Mtb ClpB; some of the interacting residues being parts of Walker A and B motifs. Lys 212 of the Walker A motif has been earlier shown to be crucial for the ATPase activity, whereas the other residues are crucial for oligomerization of ClpB [31]. Gentamicin and ribostamycin do show some interaction through hydrogen bonding with Lys212. ...
... However, the ATPase activity of Mtb ClpB was not found to be significantly affected in the presence of four drugs investigated in the current study. It has been shown earlier that both NBD1 and 2 are responsible for the basal level of ATPase activity of ClpB [31]. In the current study, the drugs appear to be inhibiting the oligomerization of Mtb ClpB resulting in the inhibition of its chaperonic activity. ...
Article
Mycobacterium tuberculosis (Mtb) caseinolytic protease B (ClpB) is a chaperone possessing a unique ability to resolubilize the aggregated proteins in vivo. ClpB has been shown to be important for the survival of Mtb within the host. Thus, it appears to be a promising target to develop new therapeutic molecules against tuberculosis. In this study, we have screened FDA approved compounds in silico to identify inhibitors against Mtb ClpB. In our screen, several compounds interacted with ClpB. The top four compounds, namely framycetin, gentamicin, ribostamycin and tobramycin showing the highest binding energy were selected for further investigation. MD simulations and tryptophan-based quenching of ClpB-drug complexes established that the selected inhibitors stably interacted with the target protein. The inhibitor and protein complexes were found to be stabilized by hydrogen bonding, and hydrophobic interactions. Although, the compounds did not affect the ATPase activity of ClpB significantly, the protein resolubilization activity of ClpB was remarkably reduced in their presence. All four compounds potently inhibited the growth of Mtb H37Ra. The antimycobacterial activity of the compounds appears to be due the inhibition of functional ClpB oligomer formation, in turn affecting its chaperonic activity.
... To directly demonstrate this function, we fused the ClpL NTD to ΔN-ClpB-K476C (ΔN-ClpB*) generating L N -ClpB* ( Figure 3F). The ClpB NTD mediates binding to soluble unfolded proteins (Rosenzweig et al., 2015), however, it is dispensable for disaggregation activity (Iljina et al., 2021;Mogk et al., 2003). The K476C mutation is located in the ClpB MD, abrogating ClpB repression and allowing for high ATPase activity in the absence of DnaK. ...
Article
Full-text available
Heat stress can cause cell death by triggering the aggregation of essential proteins. In bacteria, aggregated proteins are rescued by the canonical Hsp70/AAA+ (ClpB) bi-chaperone disaggregase. Man-made, severe stress conditions applied during, e.g., food processing represent a novel threat for bacteria by exceeding the capacity of the Hsp70/ClpB system. Here, we report on the potent autonomous AAA+ disaggregase ClpL from Listeria monocytogenes that provides enhanced heat resistance to the food-borne pathogen enabling persistence in adverse environments. ClpL shows increased thermal stability and enhanced disaggregation power compared to Hsp70/ClpB, enabling it to withstand severe heat stress and to solubilize tight aggregates. ClpL binds to protein aggregates via aromatic residues present in its N-terminal domain (NTD) that adopts a partially folded and dynamic conformation. Target specificity is achieved by simultaneous interactions of multiple NTDs with the aggregate surface. ClpL shows remarkable structural plasticity by forming diverse higher assembly states through interacting ClpL rings. NTDs become largely sequestered upon ClpL ring interactions. Stabilizing ring assemblies by engineered disulfide bonds strongly reduces disaggregation activity, suggesting that they represent storage states.
... Strain DJPM01 has genomic evidence for adaptation to cold environments (Figure 4; Supplementary Table S2). The genome contains genes homologous to those encoding protein chaperones, including two copies of ClpB (Mogk et al., 2003) and DnaJ (Qiu et al., 2006), three copies of DnaK (Thompson et al., 2012), and one copy each of GroEL (Lenz and Ron, 2014), GroES (Takei et al., 2012), and GrpE (Wu et al., 1996). There was also evidence for genes encoding cold shock proteins involved in repairing misfolded mRNA, such as ATP-dependent RNA helicases DpbA (Tsu et al., 2001), DeaD (Prud'homme-Généreux et al., 2004), and three copies of RhlE (Bizebard et al., 2004). ...
Article
Full-text available
The McMurdo Dry Valleys of Antarctica experience a range of selective pressures, including extreme seasonal variation in temperature, water and nutrient availability, and UV radiation. Microbial mats in this ecosystem harbor dense concentrations of biomass in an otherwise desolate environment. Microbial inhabitants must mitigate these selective pressures via specialized enzymes, changes to the cellular envelope, and the production of secondary metabolites, such as pigments and osmoprotectants. Here, we describe the isolation and characterization of a Gram-negative, rod-shaped, motile, red-pigmented bacterium, strain DJPM01, from a microbial mat within the Don Juan Pond Basin of Wright Valley. Analysis of strain DJMP01’s genome indicates it can be classified as a member of the Massilia frigida species. The genome contains several genes associated with cold and salt tolerance, including multiple RNA helicases, protein chaperones, and cation/proton antiporters. In addition, we identified 17 putative secondary metabolite gene clusters, including a number of nonribosomal peptides and ribosomally synthesized and post-translationally modified peptides (RiPPs), among others, and the biosynthesis pathway for the antimicrobial pigment prodigiosin. When cultivated on complex agar, multiple prodiginines, including the antibiotic prodigiosin, 2-methyl-3-propyl-prodiginine, 2-methyl-3-butyl-prodiginine, 2-methyl-3-heptyl-prodiginine, and cycloprodigiosin, were detected by LC–MS. Genome analyses of sequenced members of the Massilia genus indicates prodigiosin production is unique to Antarctic strains. UV-A radiation, an ecological stressor in the Antarctic, was found to significantly decrease the abundance of prodiginines produced by strain DJPM01. Genomic and phenotypic evidence indicates strain DJPM01 can respond to the ecological conditions of the DJP microbial mat, with prodiginines produced under a range of conditions, including extreme UV radiation.
... The C-terminal domain of E. coli ClpB (Supplementary Fig. S1A) was shown to primarily support protein self-association and thus hexamer formation, which is required for ATP binding and chaperone activity. E. coli ClpB lacking the C-terminal domain could not form hexamers and had no chaperone activity (Barnett et al., 2000;Barnett and Zolkiewski, 2002;Mogk et al., 2003). Although the truncation in the clpb3-1 mutant removes only a small portion of the conserved part in the C-terminal domain ( Supplementary Fig. S1A), the missing sequences appear to be quite important for stability and functionality of CLPB3, as judged by the following observations: (i) the truncation obviously leads to a reduced accumulation of the protein (Fig. 2); (ii) truncated CLPB3 appears to localize in aggregates already under ambient conditions to which complementing wild-type CLPB3 is attracted (Fig. 5); (iii) truncated CLPB3 massively accumulates in aggregates during heat stress (Fig. 6); (iv) the clpb3-1 mutant is much more impaired in its ability to resolve heat-induced protein aggregates than the clpb3-2 mutant, albeit both accumulate similar levels of residual CLPB3 (Figs 2, 6); and (v) the clpb3-1 mutant is more thermosensitive than the clpb3-2 mutant (Fig. 7). ...
Article
Full-text available
In the cytosol of plant cells, heat-induced protein aggregates are resolved by ClpB/Hsp100 family member HSP101, which is essential for thermotolerance. For chloroplast family member CLPB3 this is less clear with controversial reports on its role in conferring thermotolerance. To shed light onto this issue, we have characterized two clpb3 mutants in Chlamydomonas reinhardtii. We show that chloroplast CLPB3 is required for resolving heat-induced protein aggregates containing stromal TIG1 and the small heat shock proteins HSP22E/F in vivo and for conferring thermotolerance under heat stress. Although CLPB3 accumulates to similarly high levels as stromal HSP70B under ambient conditions, we observed no prominent constitutive phenotypes. However, we found decreased accumulation of the ribosomal subunit PRPL1 and increased accumulation of the stromal protease DEG1C in the clpb3 mutants, suggesting that reduction in chloroplast protein synthesis capacity and increase in proteolytic capacity may compensate for loss of CLPB3 function. Under ambient conditions, CLPB3 was distributed throughout the chloroplast but reorganized into stromal foci upon heat stress, which mostly disappeared during recovery. CLPB3 foci were localized next to HSP22E/F, which accumulated largely to the thylakoid membrane occupied area. This suggests a possible role for CLPB3 in disentangling protein aggregates from the thylakoid membrane system.
... Regarding other groups of proteins over-synthesized in S. oneidensis at 30 • C, we found GroEL (Spot 420), which increases the solubilization of aggregates (Niwa et al., 2012), the ClpB protein (Spot 152), and the DnaK (Spot 244), correlating the immunodetection results observed by WB. Protein GroEL belongs to the Hsp100/Clp AAA + ATPase family and cooperates with the chaperone system DnaK/DnaJ/GrpE in efficient solubilization and reactivation of aberrant protein aggregates formed as a consequence of cellular heat stress (Kedzierska et al., 2003;Mogk et al., 2003;Lee et al., 2004;Doyle et al., 2007). This chaperone restores the activity of other membrane proteins that also appeared over-synthesized in S. oneidensis at 30 • C, the porin (spot 856), which forms hydrophilic channels in the membrane and is refolded by GroEL (Goulhen et al., 2004). ...
Preprint
Heat stress can cause cell death by triggering the aggregation of essential proteins. In bacteria, aggregated proteins are rescued by the canonical Hsp70/AAA+ (ClpB) bi-chaperone disaggregase. Man-made, severe stress conditions applied during e.g. food-processing represent a novel threat for bacteria by exceeding the capacity of the Hsp70/ClpB system. Here, we report on the potent autonomous AAA+ disaggregase ClpL from Listeria monocytogenes that provides enhanced heat resistance to the food-borne pathogen enabling persistence in adverse environments. ClpL shows increased thermal stability and enhanced disaggregation power compared to Hsp70/ClpB, enabling it to withstand severe heat stress and to solubilize tight aggregates. ClpL binds to protein aggregates via aromatic residues present in its N-terminal domain (NTD) that adopts a partially folded and dynamic conformation. Target specificity is achieved by simultaneous interactions of multiple NTDs with the aggregate surface. ClpL shows remarkable structural plasticity by forming diverse higher assembly states through interacting ClpL rings. NTDs become largely sequestered upon ClpL ring interactions. Stabilizing ring assemblies by engineered disulfide bonds strongly reduces disaggregation activity, suggesting that they represent storage states.
Article
Severe heat stress causes massive loss of essential proteins by aggregation, necessitating a cellular activity that rescues aggregated proteins. This activity is executed by ATP-dependent, ring-forming, hexameric AAA+ disaggregases. Little is known about the recognition principles of stress-induced protein aggregates. How can disaggregases specifically target aggregated proteins, while avoiding binding to soluble non-native proteins? Here, we determined by NMR spectroscopy the core structure of the aggregate-targeting N1 domain of the bacterial AAA+ disaggregase ClpG, which confers extreme heat resistance to bacteria. N1 harbors a Zn²⁺-coordination site that is crucial for structural integrity and disaggregase functionality. We found that conserved hydrophobic N1 residues located on a β-strand are crucial for aggregate targeting and disaggregation activity. Analysis of mixed hexamers consisting of full-length and N1-truncated subunits revealed that a minimal number of four N1 domains must be present in a AAA+ ring for high-disaggregation activity. We suggest that multiple N1 domains increase substrate affinity through avidity effects. These findings define the recognition principle of a protein aggregate by a disaggregase, involving simultaneous contacts with multiple hydrophobic substrate patches located in close vicinity on an aggregate surface. This binding mode ensures selectivity for aggregated proteins while sparing soluble, non-native protein structures from disaggregase activity.
Preprint
Full-text available
AAA+ proteins (ATPases associated with various cellular activities) comprise a family of powerful ring-shaped ATP-dependent translocases that carry out numerous vital substrate-remodeling functions. ClpB is a AAA+ protein disaggregation machine that forms a two-tiered hexameric ring, with flexible pore loops protruding into its center and binding to substrate-proteins. It remains unknown whether these pore loops contribute only passively to substrate-protein threading or have a more active role. Recently, we have applied single-molecule FRET (smFRET) spectroscopy to directly measure the dynamics of substrate-binding pore loops in ClpB. We have reported that the three pore loops of ClpB (PL1-3) undergo large-scale fluctuations on the microsecond timescale that are likely to be mechanistically important for disaggregation. Here, using smFRET, we study the allosteric coupling between the pore loops and the two nucleotide binding domains of ClpB (NBD1-2). By mutating the conserved Walker B motifs within the NBDs to abolish ATP hydrolysis, we demonstrate how the nucleotide state of each NBD tunes pore loop dynamics. This effect is surprisingly long-ranged; in particular, PL2 and PL3 respond differentially to a Walker B mutation in either NBD1 or NBD2, as well as to mutations in both. We characterize the conformational dynamics of pore loops and the allosteric paths connecting NBDs to pore loops by molecular dynamics simulations and find that both principal motions and allosteric paths can be altered by changing the ATPase state of ClpB. Remarkably, PL3, which is highly conserved in AAA+ machines, is found to favor an upward conformation when only NBD1 undergoes ATP hydrolysis, but a downward conformation when NBD2 is active. These results explicitly demonstrate a significant long-range allosteric effect of ATP hydrolysis sites on pore-loop dynamics. Pore loops are therefore established as active participants that undergo ATP-dependent conformational changes to translocate substrate proteins through the central pores of AAA+ machines. Statement of Significance Molecular machines function by coupling the energy of ATP hydrolysis to mechanical motion. How this coupling occurs and what timescales are involved remains an open question. In this study, we use a powerful single-molecule FRET technique to measure the real-time dynamics of pore loops, which are essential protein-translocating elements of the ATP-dependent disaggregation machine ClpB. Using a series of mutations of the ATP-hydrolysis motifs of ClpB, we find that, although the motions of these pore loops take place on the microsecond time scale, they are markedly affected by the much slower changes in the nucleotide state of the machine. Generally, this study shows that protein machines, such as ClpB, are wired to harness ATP binding and hydrolysis to allosterically affect distal events, such as the function-related mechanics of pore-loops.
Article
Hsp100 chaperones, also known as Clp proteins, constitute a family of ring-forming ATPases that differ in 3D structure and cellular function from other stress-inducible molecular chaperones. While the vast majority of ATP-dependent molecular chaperones promote the folding of either the nascent chain or a newly imported polypeptide to reach its native conformation, Hsp100 chaperones harness metabolic energy to perform the reverse and facilitate the unfolding of a misfolded polypeptide or protein aggregate. It is now known that inside cells and organelles, different Hsp100 members are involved in rescuing stress-damaged proteins from a previously aggregated state or in recycling polypeptides marked for degradation. Protein degradation is mediated by a barrel-shaped peptidase that physically associates with the Hsp100 hexamer to form a two-component system. Notable examples include the ClpA:ClpP (ClpAP) and ClpX:ClpP (ClpXP) proteases that resemble the ring-forming FtsH and Lon proteases, which unlike ClpAP and ClpXP, feature the ATP-binding and proteolytic domains in a single polypeptide chain. Recent advances in electron cryomicroscopy (cryoEM) together with single-molecule biophysical studies have now provided new mechanistic insight into the structure and function of this remarkable group of macromolecular machines.
Article
Full-text available
The ClpX heat shock protein of Escherichia coli is a member of the universally conserved Hsp100 family of proteins, and possesses a putative zinc finger motif of the C(4) type. The ClpX is an ATPase which functions both as a substrate specificity component of the ClpXP protease and as a molecular chaperone. Using an improved purification procedure we show that the ClpX protein is a metalloprotein complexed with Zn(II) cations. Contrary to other Hsp100 family members, ClpXZn(II) exists in an oligomeric form even in the absence of ATP. We show that the single ATP-binding site of ClpX is required for a variety of tasks, namely, the stabilization of the ClpXZn(II) oligomeric structure, binding to ClpP, and the ClpXP-dependent proteolysis of the lambdaO replication protein. Release of Zn(II) from ClpX protein affects the ability of ClpX to bind ATP. ClpX, free of Zn(II), cannot oligomerize, bind to ClpP, or participate in ClpXP-dependent proteolysis. We also show that ClpXDeltaCys, a mutant protein whose four cysteine residues at the putative zinc finger motif have been replaced by serine, behaves in similar fashion as wild type ClpX protein whose Zn(II) has been released either by denaturation and renaturation, or chemically by p-hydroxymercuriphenylsulfonic acid.
Article
Full-text available
Escherichia coli FtsH is an ATP-dependent protease that belongs to the AAA protein family. The second region of homology (SRH) is a highly conserved motif among AAA family members and distinguishes these proteins in part from the wider family of Walker-type ATPases. Despite its conservation across the AAA family of proteins, very little is known concerning the function of the SRH. To address this question, we introduced point mutations systematically into the SRH of FtsH and studied the activities of the mutant proteins. Highly conserved amino acid residues within the SRH were found to be critical for the function of FtsH, with mutations at these positions leading to decreased or abolished ATPase activity. The effects of the mutations on the protease activity of FtsH correlated strikingly with their effects on the ATPase activity. The ATPase-deficient SRH mutants underwent an ATP-induced conformational change similar to wild type FtsH, suggesting an important role for the SRH in ATP hydrolysis but not ATP binding. Analysis of the data in the light of the crystal structure of the hexamerization domain ofN-ethylmaleimide-sensitive fusion protein suggests a plausible mechanism of ATP hydrolysis by the AAA ATPases, which invokes an intermolecular catalytic role for the SRH.
Article
Full-text available
The clpB gene in Escherichia coli encodes a heat-shock protein that is a close homolog of the clpA gene product. The latter is the ATPase subunit of the multimeric ATP-dependent protease Ti (Clp) in E. coli, which also contains the 21-kDa proteolytic subunit (ClpP). The clpB gene product has been purified to near homogeneity by DEAE-Sepharose and heparin-agarose column chromatographies. The purified ClpB consists of a major 93-kDa protein and a minor 79-kDa polypeptide as analyzed by polyacrylamide gel electrophoresis in the presence of sodium dodecyl sulfate. Upon gel filtration on a Superose-6 column, it behaves as a 350-kDa protein. Thus, ClpB appears to be a tetrameric complex of the 93-kDa subunit. The purified ClpB has ATPase activity which is stimulated 5-10-fold by casein. It is also activated by insulin, but not by other proteins, including globin and denatured bovine serum albumin. ClpB cleaves adenosine 5'-(alpha,beta-methylene)-triphosphate as rapidly as ATP, but not adenosine 5'-(beta,gamma-methylene)-triphosphate. GTP, CTP, and UTP are hydrolyzed 15-25% as well as ATP. ADP strongly inhibits ATP hydrolysis with a Ki of 34 microM. ClpB has a Km for ATP of 1.1 mM, and casein increases its Vmax for ATP without affecting its Km. A Mg2+ concentration of 3 mM is necessary for half-maximal ATP hydrolysis. Mn2+ supports ATPase activity as well as Mg2+, and Ca2+ has about 20% their activity. Anti-ClpB antiserum does not cross-react with ClpA nor does anti-ClpA antiserum react with ClpB. In addition, ClpB cannot replace ClpA in supporting the casein-degrading activity of ClpP. Thus, ClpB is distinct from ClpA in its structural and biochemical properties despite the similarities in their sequences.
Article
Full-text available
Protease Re, a new cytoplasmic endoprotease in Escherichia coli, was purified to homogeneity by conventional procedures, using [3H]casein as the substrate. The enzyme consists of a single polypeptide of 82,000 molecular weight. It is maximally active between pH 7 and 8.5 and is independent of ATP. It has a pI of 6.8 and a Km of 10.8 microM for casein. Since diisopropyl fluorophosphate and phenylmethylsulfonyl fluoride inhibited this enzyme, it appears to be a serine protease. Protease Re was sensitive to inhibition by L-1-tosylamido-2-phenylethylchloromethylketone but not to that by 1-chloro-3-tosylamido-7-aminoheptanone, thiol-blocking reagents, chelating agents, or various peptide aldehydes. Re also degraded [125I]globin, [125I]glucagon, and 125I-labeled denatured bovine serum albumin to acid-soluble products (generally oligopeptides of greater than 1,500 daltons), but it showed no activity against serum albumin, growth hormone, insulin, or a variety of fluorometric peptide substrates. It also hydrolyzed oxidatively inactivated glutamine synthetase (generated by ascorbate, oxygen, and iron) four- to fivefold more rapidly than the native protein. Protease Re appears to be identical to the proteolytic enzyme isolated by Roseman and Levine (J. Biol. Chem. 262:2101-2110, 1987) by its ability to degrade selectively oxidatively damaged glutamine synthetase in vivo. Its role in intracellular protein breakdown is uncertain.
Article
Full-text available
The complete nucleotide sequence (580,070 base pairs) of the Mycoplasma genitalium genome, the smallest known genome of any free-living organism, has been determined by whole-genome random sequencing and assembly. A total of only 470 predicted coding regions were identified that include genes required for DNA replication, transcription and translation, DNA repair, cellular transport, and energy metabolism. Comparison of this genome to that of Haemophilus influenzae suggests that differences in genome content are reflected as profound differences in physiology and metabolic capacity between these two organisms.
Article
Evidence is rapidly accumulating that, in addition to enzymes that catalyze slow steps in protein folding, molecular ‘chaperones’ bind rapidly to newly formed polypeptide chains. This interaction ensures correct folding and prevents side reactions, such as aggregation.
Article
Functional chaperone cooperation between Hsp70 (DnaK) and Hsp104 (ClpB) was demonstrated in vitro. In a eubacterium Thermus thermophilus, DnaK and DnaJ exist as a stable trigonal ring complex (TDnaK\cdot J complex) and the dnaK gene cluster contains a clpB gene. When substrate proteins were heated at high temperature, none of the chaperones protected them from heat inactivation, but the TDnaK\cdot J complex could suppress the aggregation of proteins in an ATP- and TGrpE-dependent manner. Subsequent incubation of these heated preparations at moderate temperature after addition of TClpB resulted in the efficient reactivation of the proteins. Reactivation was also observed, even though the yield was low, if the substrate protein alone was heated and incubated at moderate temperature with the TDnaK\cdot J complex, TGrpE, TClpB, and ATP. Thus, all these components were necessary for the reactivation. Further, we found that TGroEL/ES could not substitute TClpB.
Article
Self-association of ClpB (a mixture of 95- and 80-kDa subunits) has been studied with gel filtration chromatography, analytical ultracentrifugation, and electron microscopy. Monomeric ClpB predominates at low protein concentration (0.07 mg0mL), while an oligomeric form is highly populated at >4 mg/mL. The oligomer formation is enhanced in the presence of 2 mM ATP or adenosine 5′-O-thiotriphosphate (ATPS). In contrast, 2 mMADP inhibits full oligomerization of ClpB. The apparent size of the ATP- or ATPS-induced oligomer, as determined by gel filtration, sedimentation velocity and electron microscopy image averaging, and the molecular weight, as determined by sedimentation equilibrium, are consistent with those of a ClpB hexamer. These results indicate that the oligomerization reactions of ClpB are similar to those of other Hsp100 proteins.
Article
AAA ATPases play central roles in cellular activities. The ATPase p97, a prototype of this superfamily, participates in organelle membrane fusion. Cryoelectron microscopy and single-particle analysis revealed that a major conformational change of p97 during the ATPase cycle occurred upon nucleotide binding and not during hydrolysis as previously hypothesized. Furthermore, our study indicates that six p47 adaptor molecules bind to the periphery of the ring-shaped p97 hexamer. Taken together, these results provide a revised model of how this and possibly other AAA ATPases can translate nucleotide binding into conformational changes of associated binding partners.
Article
We demonstrate in this work that the surface tension, water-organic solvent, transfer-free energies and the thermodynamics of melting of linear alkanes provide fundamental insights into the nonpolar driving forces for protein folding and protein binding reactions. We first develop a model for the curvature dependence of the hydrophobic effect and find that the macroscopic concept of interfacial free energy is applicable at the molecular level. Application of a well-known relationship involving surface tension and adhesion energies reveals that dispersion forces play little or no net role in hydrophobic interactions; rather, the standard model of disruption of water structure (entropically driven at 25 degrees C) is correct. The hydrophobic interaction is found, in agreement with the classical picture, to provide a major driving force for protein folding. Analysis of the melting behavior of hydrocarbons reveals that close packing of the protein interior makes only a small free energy contribution to folding because the enthalpic gain resulting from increased dispersion interactions (relative to the liquid) is countered by the freezing of side chain motion. The identical effect should occur in association reactions, which may provide an enormous simplification in the evaluation of binding energies. Protein binding reactions, even between nearly planar or concave/convex interfaces, are found to have effective hydrophobicities considerably smaller than the prediction based on macroscopic surface tension. This is due to the formation of a concave collar region that usually accompanies complex formation. This effect may preclude the formation of complexes between convex surfaces.