40
1 Coordinated ATP - hydrolysis by the Hsp90 dimer Klaus Richter, Paul Muschler, Otmar Hainzl and Johannes Buchner* Institut für Organische Chemie und Biochemie Technische Universität München Lichtenbergstr.4, 85747 Garching * To whom correspondence should be adressed Phone: 0049-89-28913340 Fax: 0049-89-28913345 E-Mail: [email protected] Running Title: Coordinated ATP - hydrolysis of Hsp90 Keywords: Hsp90, ATPase, dimerization, heterodimers, gyrase Copyright 2001 by The American Society for Biochemistry and Molecular Biology, Inc. JBC Papers in Press. Published on July 5, 2001 as Manuscript M103832200 by guest on February 21, 2018 http://www.jbc.org/ Downloaded from

Institut für Organische Chemie und Biochemie

Embed Size (px)

Citation preview

1

Coordinated ATP - hydrolysis by the Hsp90 dimer

Klaus Richter, Paul Muschler, Otmar Hainzl and Johannes Buchner*

Institut für Organische Chemie und Biochemie

Technische Universität München

Lichtenbergstr.4, 85747 Garching

* To whom correspondence should be adressed

Phone: 0049-89-28913340

Fax: 0049-89-28913345

E-Mail: [email protected]

Running Title: Coordinated ATP - hydrolysis of Hsp90

Keywords: Hsp90, ATPase, dimerization, heterodimers, gyrase

Copyright 2001 by The American Society for Biochemistry and Molecular Biology, Inc.

JBC Papers in Press. Published on July 5, 2001 as Manuscript M103832200 by guest on February 21, 2018

http://ww

w.jbc.org/

Dow

nloaded from

2

Abstract

The Hsp90 dimer is a molecular chaperone with an unusual N-terminal ATP binding site. The

structure of the ATP binding site makes it a member of a new class of ATP hydrolysing

enzymes, known as the GHKL-family. While for some of the family members structural data

on conformational changes occuring after ATP binding are available, these are still lacking for

Hsp90. Here we set out to investigate the correlation between dimerization and ATP

hydrolysis by Hsp90. The dimerization constant of wt Hsp90 was determined to be 60nM.

Heterodimers of wt Hsp90 with fragments lacking the ATP binding domain form readily and

exhibit dimerization constants similar to full length Hsp90. However, the ATPase activity of

these heterodimers was significantly lower than that of the wild type protein, indicating

cooperative interactions in the N-terminal part of the protein, that lead to the activation of the

ATPase activity. To further address the contribution of the N-terminal domains to the ATPase

activity, we used an Hsp90 point mutant which is unable to bind ATP. As heterodimers

between the wt protein and this mutant showed wt ATPase activity, this mutant, though

unable to bind ATP, still has the ability to stimulate the activity in its wt partner domain. Thus

contact formation between the N-terminal domains might not depend on ATP bound to both

domains. Together these results suggests a mechanism for coupling the hydrolysis of ATP to

the opening - closing movement of the Hsp90 molecular chaperone.

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

3

Abbreviations

GA, Geldanamycin; HPLC, high pressure liquid chromatography; Hsp90, heat shock protein

90; Hsp90 - 262C, yeast Hsp90 fragment ranging from amino acid 262 to 709; Hsp90 - 527C,

fragment from amino acid 527 to 709 of Hsp90; Hsp90 - D79N, the D79N - point mutant of

Hsp90; Hsp90 - N210, fragment 1 to 210; Hsp90 - N272, fragment 1 to 272; Hsp90 - N529,

fragment 1 to 529; Hsp90 - N599, fragment 1 to 599; IPTG, Isopropyl-thiogalactosid; Kd,

dissociation constant for the Hsp90 dimer; Kd,app apparent dissociation constant; PAGE,

polyacrylamide gel electrophoresis; PCR, polymerase chain reaction; SDS, sodium-

dodecylsulfate; SEC, size exclusion chromatography; wt, wild type.

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

4

Introduction

Hsp90 is an abundant cytosolic protein which belongs to the class of molecular chaperones.

Interaction with non-native and destabilised proteins has been shown in vitro (1, 2, 3, 4, 5). In

vivo an ever increasing number of proteins have been found to be associated with Hsp90 (6,

7). Most of the known in vivo substrates are involved in signal transduction pathways, like

tyrosin and serine/threonine kinases, steroid hormone receptors, helix-loop-helix transcription

factors and tumour supressor proteins (6, 7, 8). Other proteins, like reverse transcriptase (9) or

telomerase (10) were found to be dependent on Hsp90 action as well. Consistent with the

importance of the substrate proteins, Hsp90 was found to be an essential protein in yeast (11).

Although the mechanism of Hsp90 is still far from being understood, Hsp90 is thought to

maintain an otherwise unstable conformation of the substrate (7). ATP binding and hydrolysis

were found to be essential for the function of Hsp90 (12, 13) and competitive inhibitors for

ATP-binding, like geldanamycin were shown to be potent anti-proliferative agents (14, 15).

First indications that Hsp90 conformation is influenced by ATP were reported by Csermely et

al. (16) and Grenert et al. (17). Experiments using spin-labeled ATP showed that Hsp90 binds

ATP weakly (18). The nucleotide binding site was identified in the crystal structure of the N-

terminal domain of yeast Hsp90 in the presence of ADP (19, 20). This binding site exhibits a

new type of fold similar to that of DNA gyrase B and MutL (21). The nucleotide is bound in

an unusual kinked conformation with the adenosine base and the ribose buried inside the

protein in a cleft. The β-phosphate and probably even more the γ-phosphate of ATP are

solvent accessible in the crystal structure. Interestingly, the crystal structures of the domain in

the presence or absence of nucleotide were almost identical (19). The kinetic analysis of the

ATPase cycle of Hsp90 revealed that after ATP binding, a conformational change occurs in

Hsp90 which traps the ATP molecule (22). This trapped ATP molecule proved to be

committed to hydrolysis, as it was impossible to exchange it for unbound ATP. This

conformational change requires parts of the protein that are C-terminal to the binding site. A

conformational change was also observed in the crystal structure of the homologous protein

MutL (23, 24). MutL traps ATP by contacts between the γ-phosphate of ATP and a domain

further C-terminal from the ATP binding site. Conformational changes of Hsp90 upon ATP

binding were also reported by Prodromou et al. (25). Here, cross-linking data indicate that the

N-terminal domains associate in the presence of AMP-PNP, which had been suggested

previously based on electron microscopic data (26). This seems to be a prerequisite for the

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

5

association of Hsp90 with the co-chaperone p23, which is known to occur after ATP-binding

(27, 28, 29). In the studies of Weikl et al. (22) and Prodromou et al. (25) fragments of Hsp90

lacking C-terminal domains were found to be considerably less active, than wild-type Hsp90.

These fragments are thought to be monomeric, in contrast to full length Hsp90, which is a

dimer. The dimerization site was shown to reside in the very C-terminal domain of Hsp90 (30,

31).

To gain further insight into the ATPase mechanism of Hsp90 we investigated the importance

of dimer formation for the ATPase activity. To this end we used a set of Hsp90 deletion

mutants and analysed their quaternary structure as well as their ATPase activities. Hsp90 as

well as C-terminal fragments were shown to be dimeric with dissociation constants of

0.06µM. Analysis of the ATPase activity of these fragments demonstrated that the presence of

two N-terminal domains is required to stimulate the ATPase activity, as heterodimers formed

between C-terminal fragments and the wt protein had diminished ATPase activities.

Surprisingly, the stimulation of hydrolysis does not require ATP bound on both sides of the

dimer. The presence of two N-terminal domains is sufficient to activate ATP-hydrolysis.

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

6

Material and Methods

Materials

Geldanamycin (GA) was a kind gift of the NIH, NCI Experimental Drug Division, Bethesda,

Maryland, USA.

Hsp90 constructs

Fragments of yeast Hsp90 were constructed using the plasmid pET28-Hsp82, containing the

full length Hsp82 gene of S. cerevisiae with an N-terminal His-Tag as a template. All the

PCR-fragments were cloned into the Qiagen vectors pQE30, resulting in the constructs

pQE30-N529, pQE30-N599 and pQE30-527C. In addition, fragments constructed by Scheibel

et al. (1998) were used. The mutagenesis leading to the D79N mutation in full length Hsp90

was done by overlap extention PCR in N210 using two primers conatining the mutation (5’-

CCAATACCAGAGTTTCTGATTTCC-3’ and 5’-GGAAATCAGAAACTCTGGTATT GG-

3’) and subsequent insertion of this fragment into the full-length gene, using an N-terminal

BamHI - restriction site and the internal XbaI site. The identity of all constructs was

confirmed by DNA sequencing.

Protein expression and purification

His-Hsp90 was expressed in the strain BL21 (DE3) cop+ (Stratagene, La Jolla, USA) in LBKan

and induced with 1mM IPTG. For all other constructs the strain M15 prep and 2mM IPTG

were used. Cells were lysed using a Cell Disruption System (Constant Systems, Warwick,

UK). The fragments and the His-tagged full-length protein were purified on a Chelating

Sepharose Column (Amersham Pharmacia Biotech, Uppsala, Sweden) preloaded with 100mM

NiSO4. Cell lysis and loading of the proteins was done in loading buffer (40mM KPO4, pH

8.0, 400mM KCl, 6mM imidazole). The column was washed with washing buffer (40mM

KPO4, pH 8.0, 400mM KCl, 20mM imidazole) before elution was performed in a step

gradient with washing buffer containing 300mM imidazole. A Resource Q column

(Amersham Pharmacia Biotech, Uppsala, Sweden) was used to further purify the proteins. The

protein was loaded in 50mM Tris, pH 8.0, 20mM KCl and eluted with a gradient from 20mM

to 1000mM KCl. As a final purification step, a Superdex 200 HiLoad (Amersham Pharmacia

Biotech, Uppsala, Sweden) or Superdex 75 HiLoad (Amersham Pharmacia Biotech, Uppsala,

Sweden), depending on protein size was run in 40mM HEPES, pH 7.5, 300mM KCl, 5%

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

7

glycerol. Proteins were stored in 40mM HEPES, pH 7.5, 20mM KCl, 5% glycerol at

concentrations of 1.5mg/ml to 6mg/ml at -70°C.

Protein structure and stability

Far-UV CD-Spectroscopy was used to confirm the secondary - structure of the fragments.

Measurements were performed in 40mM potassium phosphate, pH 7.0 at protein

concentrations of 200µg/ml and 20°C. CD-spectra were collected between 195nM and 250nM

on a J-715 spetropolarimeter (Jasco, Groß-Umstadt, Germany).

Urea transitions were performed with 20µg/ml protein and urea concentrations ranging from

0M to 7.5M in 40mM HEPES, pH 7.5, 20mM KCl. Changes in tertiary structure were

detected by fluorescence measurements on a FluoroMax-2 (Spex, Edison, USA) at 25°C. The

excitation wavelength was set to 280nM, while emission spectra were collected from 300nM

to 400nM. The midpoint of the urea transition was obtained from a plot of the fluorescence

signal at 328nM against the urea concentration.

ATPase activity

ATPase activities were measured using the EnzCheck phosphate detection kit (Molecular

Probes, Leiden, Netherlands). The assays were performed in 120µl cuvettes and the

production of phosphate was detected by the increase of adsorbance at 360nm, using a

Pharmacia 40/60 spectrophotometer. The temperature was set to 37°C. Assays were

performed in 40mM HEPES, pH 7.5, 150mM KCl, 5mM MgCl2 and 2mM ATP. Typical

protein concentrations were 2.5µM for Hsp90 and up to 50µM for the less active N-terminal

fragments. To subtract Hsp90 contaminating ATPase activities that usually co-purify with

Hsp90, GA was added at concentrations of 20µM for Hsp90 and 100µM for the N-terminal

fragments. The remaining ATPase activity was Hsp90 independent.

ATPase - Competition experiments

For the ATPase assays Hsp90 concentrations were 2.5µM, while the concentrations of

fragments were varied between 1µM and 16µM. To allow equilibration between homodimers

and heterodimers, the samples were incubated for 10 minutes at 37°C, which was sufficient to

reach equilibrium before ATP was added. Again, GA was added later on to ensure, that the

detected ATPase activity is due to ATP hydrolysis by Hsp90. The data analysis of the titration

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

8

relied on the assumption that the binding of fragments to each other is as effective as the

binding of fragments to Hsp90 and the binding of Hsp90 to itself. To correct for potential

mistakes in the concentration of one of the proteins, a stoichiometry factor n was incorporated

that should, in case that all the concentrations are correct, be close to one. n-values smaller

than one could also indicate that the fragments bind with a lower affinity to Hsp90 than Hsp90

to itself. The resulting equation was

[ ][ ] [ ]

[ ][ ] [ ]Activity Act

Hsp

Hsp n FragAct

Hsp

Hsp n Fraghet= ⋅

+ ⋅+ ⋅ −

+ ⋅

homo

90

901

90

90 (equation 1)

Acthomo is the activity of the homodimer and Acthet is the activity of the heterodimer.

Cross-linking

Cross-linking was performed with glutaraldehyde (Merck, Darmstadt, Germany). 20µl of a

2.4µM Hsp90 solution containing different amounts of Hsp90 - 527C were incubated for 10

minutes at room temperature. 1µl of a 1:10 solution of 25% glutaraldeyde in reaction buffer

(40mM HEPES, pH 7.5, 150mM KCl, 5mM MgCl2) was added for two minutes. 2mM AMP-

PNP was included optional in the reaction buffer. Then the reaction was stopped by adding

5µl of 1M Tris, pH 8.0. After addition of 5µl of 5x Laemmli-buffer, the proteins were

separated on precast 4-12% Polyacrylamide gels (Invitrogen, Groningen, Netherlands) and

stained with Coomassie Blue (Sigma-Aldrich, St.Louis, USA). Evaluation of the stained gels

was performed with the programs LabScan and ImageMaster (Amersham Pharmacia Biotech).

Size -exclusion HPLC and data analysis

Gelfiltration HPLC was performed on a PU-1580 system (Jasco, Groß-Umstadt, Germany)

with a Superdex 200HR - column (Amersham Pharmacia Biotech, Uppsala, Sweden) in

40mM HEPES, pH 7.5, 150mM KCl. The Hsp90 concentrations applied to the column were

varied between 12µM and 9nM. Protein fluorescence was detected using a FP 1520-S

fluorescence detector (Jasco, Victoria, Canada) with excitation at 280nm and emission at

328nm. The gain of the photomultiplier was varied to be able to record the signal over a

concentration range of three magnitudes. No interaction of monomeric and dimeric species

with the column matrix was observed (with the exception of 527C) as the peak shapes were

similar for the dimeric and the monomeric species. The analysis of the dissociation curves was

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

9

based on the assumption that at any time monomeric and dimeric species were in equilibrium.

Data were fitted using the Scientist program (MicroMath, Salt Lake City, USA) and equation

2:

ET ET ET ETotein

otein Kmono mono erd app

= − − ⋅+

( )[Pr ]

[Pr ]dim,

(equation 2)

where ET is the elution time, ETmono the elution time of the monomeric species and ETdimer the

elution time of the dimeric species. The obtained dissociation konstant Kd,app is only an

aproximation, as the protein is diluted severalfold during the course of the experiment. To

determine the dilution factor the peak area was analysed. At the time of injection, the width of

the peak was 0.2 minutes, which corresponds to 0.1ml injected volume. Dividing the peak

area by 0.2 minutes gave a signal, that is about seven times as high as the maximum of the

originally observed peak. This factor was used as the dilution factor for the calculation of the

”real” Kd by dividing Kd,app with the dilution factor. As the dilution factors were found to be

similar for the different Hsp90 fragments they allowed to directly compare their association

behaviours. All values in the following text are already corrected for dilution.

Analytical ultracentrifugation

Data for sedimentation equilibrium runs were collected on a Beckman XL-A analytical

ultracentrifuge using a Ti60 rotor. All runs were performed at 4°C for 48 hours with the

addition of protease inhibitors (complete, Roche Molecular Biochemicals, Mannheim,

Germany) to avoid degradation. The protein gradient was detected at 280nm. Data analysis

was done with the programme Origin (Beckman, Fullerton, USA).

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

10

Results

N-terminal fragments of yeast Hsp90 show reduced ATPase activity

Earlier studies showed that truncation mutants of Hsp90 which contain the N-terminal ATP-

binding site, but lack C-terminal regions are considerably less active in ATP-hydrolysis than

the wt-protein (5, 25, 22). We were interested in further defining the requirements for full

ATPase activity of Hsp90. For this purpose, we designed a set of Hsp90 fragments based on

proteolytic digests (5), sequence alignments with homologous proteins and hydropathy plots

(Figure 1). The fragments, varing in size from 210 amino acids (1 - 210) to 599 amino acids (1

- 599), were purified and their structures and stabilities were analysed using CD-spectroscopy

and urea transitions. The midpoint of the unfolding transition were between 3M and 5M urea

for all fragments tested. This compares to the stability of the wt protein. All domains used

were thus shown to be folded and stable (data not shown). As reported previously, Hsp90 -

N210 and Hsp90 - N272 showed extremely weak ATPase activities (0.05µmole ATP per

µmole protein per minute), while Hsp90 - N529 and, more pronounced, Hsp90 - N599

exhibited increased ATPase activities (0.12 and 0.18 µmole ATP per µmole protein per

minute, respectively). However, these activities are still by a factor of six lower than the

ATPase activity of wild-type Hsp90, which was 1.1 µmole ATP per µmole protein per minute

(Figure 2A).

We were unable to detect any concentration dependence of the ATPase activity for Hsp90 -

N210 and for the full-length protein. However, Hsp90 - N529 and Hsp90 - N599 showed a

reproducable increase of activity with protein concentration (Figure 2B).

C-terminal fragments and wt - Hsp90 share similar dimerization properties

To determine the dissociation constant (Kd) for Hsp90 we decided to employ size-exclusion

HPLC (SEC - HPLC) with fluorescence detection at different protein concentrations. Using

Hsp90 concentrations of 2nM to 1µM, only one peak was observed at every concentration

tested. However, the elution times differed depending on the protein concentration used

(Figure 3A). A shift could be observed starting from an elution time of 20 minutes (high

protein concentration) to an elution time of 23 minutes (low protein concentration). The

presence of one peak at every concentration is consistent with fast equilibration compared to

the time scale of the experiment. Using a model which assumes a monomer-dimer equilibrium

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

11

at every time point of the experiment, we fitted the data points and obtained a Kd of 60nM +/-

12nM (Figure 3B).

Having calculated the Kd for wt - Hsp90, we were interested to investigate whether C-terminal

fragments containing the dimerization site show similar dimerization characteristics. Two

fragments were constructed, one ranging from amino acids 262 to 708 (Hsp90 - 262C) and the

other containing amino acids 527 to 709 (Hsp90 - 527C). Hsp90 - 262C gave a transition

curve similar to Hsp90 (Figure 3C). The Kd obtained for Hsp90 - 262C was 45nM +/- 12nM.

For Hsp90 - 527C no titration curve was obtained, as this protein interacted with the

gelfiltration matrix, resulting in significant peak broadening.

These data clearly show, that Hsp90 and its C-terminal fragments are dimeric under

conditions used for the ATPase assay, where protein concentrations usually are by at least one

order of magnitude higher than the dimerization constant of Hsp90.

N-terminal Hsp90 fragments are monomeric

To analyse the quaternary structure of N-terminal fragments the same experimental setup was

used. For Hsp90 - N529, no changes in the elution time were observed at concentrations

ranging from 0.04µM to 5µM. This, together with equilibrium sedimentation data from

analytical ultracentrifugation collected for Hsp90 - N529 (data not shown) led us to conclude,

that Hsp90 - N529 is monomeric within the concentration range tested. Similarly SEC - HPLC

experiments were performed with Hsp90 - N599. Here no change in quaternary structure was

observed within the concentration range tested (0.04µM to 2µM) as well (Figure 3C).

Next we investigated, whether the oligomerization properties of the N-terminal fragments

change in the presence of ATP. Hsp90 - N529 as well as Hsp90 - N599 showed a slight

concentration dependence in SEC - HPLC experiments (data not shown). In agreement with

the ATPase assays (Fig. 2B) which were performed at concentrations between 2 and 10µM for

Hsp90 - N599, these data indicate that this could be the beginning of a monomer - dimer

transition curve, but still be far away from the actual dimerization constant.

Taken together, these data suggest that the fragments lacking the C-terminal domain are

monomeric and that ATP does not alter the oligomerization behaviour of these proteins in the

concentration range tested. At higher concentrations, as used in the ATPase assays (Figure

2B), they dimerize with dimerization constants in the range of 10µM for Hsp90 - N599 and

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

12

about 70µM for Hsp90 - N530, as obtained by analysis of the concentration dependent

ATPase assays.

Heterodimers can be formed between Hsp90 and C-terminal Hsp90 fragments

Having established that the dimerization properties for Hsp90 and its C-terminal fragment

262C are similar, we were interested whether heterodimers can be obtained by mixing Hsp90

with C-terminal fragments. To detect dimers we used glutaraldehyde which is known to

efficiently crosslink dimeric Hsp90 (32). Crosslinking is not complete since about 20%

monomers can be detected. (Figure 4). Uncomplete glutaraldehyde - crosslinking of Hsp90

has been observed previously (32). Hsp90 - 527C was added to wt - Hsp90. In the presence of

Hsp90 - 527C, a decrease in the intensity of the dimeric wt - Hsp90 band was visible on SDS-

PAGE, while an additional band appeared, consisting of one molecule Hsp90 - 527C and one

molecule Hsp90 (Figure 4A). Thus, it is possible to form heterodimers between Hsp90 and its

C-terminal fragments in a concentration - dependent manner. Even though the concentration

of the C-terminal fragment was much higher than the concentration of wt - Hsp90 in these

experiments, still the sample seems to contain a fraction of homodimers of wt - Hsp90. A

quantitative densitometric analysis of the scanned SDS-gels showed that the heterodimer band

increases with increased concentrations of Hsp90 – 527C. At the highest concentration of

Hsp90 - 527C the cross-linked sample contains about equal concentrations of monomeric

Hsp90, dimeric Hsp90 and heterodimeric Hsp90 (Figure 4B). This ratio does not necessarily

reflect the amount present under equilibrium conditions, especially as it is known from the

dissociation constant, that dimerization of Hsp90 is complete at the concentrations used. Thus,

this result might reflect a higher crosslinking efficiency of the homodimer compared to the

heterodimer and the C-terminal fragment. Crosslinks made in the absence and presence of

AMP-PNP did not show detectable differences in the intensity of the heterodimer band (data

not shown). Thus the N-terminal association induced by AMP-PNP does not contribute

significantly to the ability of 527C - Hsp90 to disrupt Hsp90 homodimers.

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

13

Heterodimers of Hsp90 and C-terminal Hsp90 fragments show reduced ATPase activity

Next, ATPase assays were performed with Hsp90 and Hsp90 - 527C. Hsp90 - 527C itself was

found to have no ATPase activity. When the concentration of Hsp90 - 527C was increased in

the presence of a constant amount of Hsp90, a decrease in ATPase activity was observed (Fig.

5A). This result suggests that the heterodimers have reduced ATPase activity. A statistical

model which assumes equal probabilities for the formation of homodimers and heterodimers

of Hsp90 and Hsp90 - 527C gave an excellent fit for the data points (Figure 5A).

Addition of longer C-terminal fragments containing the middle domain in addition to the C-

terminal domain (amino acid 262 - 709, cf. Fig.1) also led to decreased ATPase activities

(Figure 5B). The inhibition curves were very similar to that of Hsp90 - 527C and again the

statistical model gave a good fit to the data points. The model used to fit the data points also

provides information on the ATPase activity of the heterodimer. The calculated activity of the

heterodimer is one third of the wt activity. This indicates that the ATPase is not completely

inhibited in the heterodimer, at least not the same extend as in the isolated N-terminal

domains. The fact that further increase in the concentrations of N - terminal truncated mutants

did not result in additional decrease of the ATPase activity suggests that the endpoint of the

titration is reached as expected based on the concentrations of the two proteins. This suggests

that beside the N-terminal dimerization, C-terminal regions seem to be important for efficacy

of the ATPase cycle. Also, fast dissociation and reassociation of dimers could result in

formation of homodimers, which might be sufficient to stimulate the ATPase activity and thus

prevent complete inhibition.

Taken together, the data suggest that the presence of two N-terminal domains is required to

activate the ATPase activity in wt Hsp90. Additionally, they confirm that the dimerization

properties of Hsp90 - 262C and Hsp90 - 527C are very similar to that of wild-type Hsp90,

indicating that interactions in the N-terminal part of the protein do not contribute significantly

to the dissociation constant of the wild-type protein.

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

14

Hsp90 - D79N does not inhibit the ATPase activity of Hsp90

Next we wanted to know, whether ATP binding is required in both N-terminal domains of the

homodimer to achieve wild-type activity. To test this, we used a point mutant of Hsp90,

Hsp90 - D79N, which does not bind nor hydrolyse ATP (12, 13).

Competition experiments performed with increasing concentrations of Hsp90 - D79N and a

fixed amount of Hsp90 showed that Hsp90 - D79N did not affect the ATPase activity of

Hsp90, even if an eightfold excess of the Hsp90 mutant was used (Figure 6A). This implies

that the activity of an individual wt-monomer in the dimer context is the same, independent of

whether the N-terminal partner domain is active or inactive. Thus a defect in ATP binding

does not compromise stimulation of ATP hydrolysis in the partner domain.

As we could not detect any change in ATPase activity, we wanted to rule out the possibility,

that formation of heterodimers did not occur in the competition experiment. Hsp90 is known

to be very sensitive to C-terminal proteolysis, and the loss of even a small fragment from the

C-terminal end affects the dimerization properties significantly (data not shown). Thus, we

determined the dimerization constant for Hsp90 - D79N by SEC-HPLC. Here we obtained a

dimerization constant of 80nM +/- 20nM (Figure 6B). This value indicates, that the

dimerization of Hsp90 - D79N does not differ significantly from that of wt - Hsp90.

Dimerization of Hsp90 is a dynamic process

We used the inhibiton of the ATPase activity of Hsp90 by formation of heterodimers to obtain

insight into the dynamics of the dimerization reaction. To this end, we added Hsp90 - 262C to

the ATPase reaction while monitoring the progress of ATP hydrolysis. As expected, the

formation of heterodimers resulted in a pronounced decrease in ATP hydrolysis immediately

after mixing (Figure 7A).

Additionally, we monitored the kinetics of formation of active dimers by adding Hsp90 -

D79N to an ATPase-suppressed heterodimeric form of Hsp90 and Hsp90 - 262C. This

experiment gave the expected increase in activity (Figure 7B), as the addition of Hsp90 -

D79N leads to the formation of Hsp90/Hsp90 - D79N - heterodimers, in which the intact N-

terminal domain exhibits wild type ATPase activity (cf. Figure 6A). Again, the increase in

activity was obtained within the first seconds of the experiment.

These results indicate, that the Hsp90 dimer is a highly dynamic structure.

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

15

Discussion

The X-ray structure of an N-terminal fragment of Hsp90 in complex with ADP showed a

nucleotide binding site which buries the ribose backbone and the adenine base inside a cleft

but leaves the β-phosphate pointing towards the surface of the molecule (19). It is obvious

from this structure, that for ATP the γ-phosphate would be completely solvent-exposed. More

importantly, no conformational changes were observed in the crystal structure of the N-

terminal domain upon binding of ADP or GA (19, 20). Subsequent studies showed that

conformational changes occur in other parts of the protein, which trap the ATP molecule and

commit it to hydrolysis (22). In addition, the N-terminal domains come in close contact (25).

The functional consequences of these changes are still largely unclear. One of the few

functional correlations of these rearrangements is that binding of the cofactor p23 requires

ATP bound to Hsp90 (28) and the dimeric form of the N-terminal domains (22, 29).

In this study, the Hsp90 dimer was shown to exhibit a dissociation constant of about 60nM.

The major dimerization site resides in the C-terminal region, as an Hsp90 fragment consisting

of amino acids 527 - 709 formed heterodimers with a affinity comparable to the native dimer.

This is in agreement with previous qualitative studies, using truncated fragments or deletion

mutants of Hsp90. For human Hsp90, Jibard et al. (31) mapped sites important for

dimerization to amino-acids 548-567, 661-677 and 679-728. These sites can be found in

fragment 527C of yeast Hsp90 indicating that this domain is responsible for dimerization. In

addition to this, our results suggest that much weaker (Kd > 70µM) interaction sites are

present within amino acid 1 - 529.

Our approach to obtain the dimerization constant of Hsp90 differs from other methods, in that

it does not require labelling. The quantitative analysis of gelfiltration allows a wide

concentration range to be covered using sensitive fluorescence detection. Previously

dimerization constants for homodimers were obtained for enzymes like lactate

dehydrogenases using enzymatic activity, as the probe for the native tetrameric state (33, 34).

The value obtained is 1nM. A dissociation constant of 60nM for Hsp90 may reflect the

requirement for a dynamic monomer/dimer equilibrium. Most interestingly, the dissociation

constant for the dimer of CheA, a histidine kinase with an ATP-binding site homologous to

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

16

Hsp90 was determined by enzymatic assays and found to be in the range of 200 to 400nM

(35). Dissociation and association in the Hsp90 dimer occurs fast, as we were unable to

measure the rates of subunit-exchange in experiments with a deadtime of about 20 seconds.

This shows, that the dynamic of the dimerization of Hsp90 is at least in the same range as

ATP hydrolysis, leading to the interesting possibility, that Hsp90 might be able to dissociate

during the ATPase cycle.

The data of this study, combined with earlier data (22, 25) allow us to define the key steps of

the Hsp90 ATPase as summarized below (Fig. 8).

In the first step of our model, ATP binding occurs independently at both N-terminal domains

of the dimer in a fast reaction (Fig. 8, step 1). Cooperativity does not seem to be involved in

the binding reaction, since the binding constant for AMP-PNP to wt - Hsp90 is identical to

that of Hsp90 - N529 or Hsp90 - N210 (data not shown).

Following the binding of ATP, a conformational change occurs in Hsp90 (Fig. 8, step 2),

which requires amino-acids 1 - 451 and was suggested to use Lys-342 as the acceptor for the

γ-phosphate of the ATP (22). These movements seem to be a prerequisite for a functional N-

terminal dimerization reaction (25), that we consider to be the next step (Fig. 8, step 3).

This dimerization reaction was shown to be possible for Hsp90 - N529, but much more

efficient for Hsp90 - N599, indicating that major interaction sites exist between amino acids

529 and 599. Our results show that the N-terminal dimerization is the prerequisite for an

efficient ATP hydrolysis reaction. Formation of heterodimers with fragments that compete for

the dimerization site on the wt-protein leads to a significant inhibition of the Hsp90 ATPase

activity. A heterodimer comprising one wild - type protein and one fragment lacking the first

261 amino acids shows reduced ATPase - activity suggesting that the two N - terminal

fragments hydrolyse their ATP-molecules in a cooperative manner (Fig. 8, step 4 and 5). This

is reminescent of the ATP-dependent chaperone GroEL, which is known to hydrolyse its

ATP-molecules in a highly regulated way (36, 37) and where the hydrolysis of ATP is

preceeded by conformational changes, that trap the ATP molecule and commit it to hydrolysis

(36).

In the case of Hsp90, surprisingly, addition of Hsp90 - D79N did not change the activity of the

wild-type protein, indicating that the wt - monomer in complex with Hsp90 - D79N is as

active as the wt - monomer in the homodimer. This implies that Hsp90 - D79N fulfills all the

requirements needed to activate the ATPase activity of wt - protein, although its part is only a

passive one. Thus, in principle, Hsp90 is able to work even with only one ATP bound to the

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

17

dimer, although this case might be unphysiological given the high ATP concentrations inside

the cell and the fact that ATP binding is known to be very fast compared to the following

steps in the hydrolysis cycle (22). After hydrolysis, Hsp90 presumably opens and releases the

ADP molecule (Fig. 8, step 6), which brings Hsp90 back to the conformation competent to

bind ATP.

Additional evidence for this model comes from studies of homologous proteins, known to

share a structurally related ATP-binding site with Hsp90. N-terminal fragments of gyrase B

show a concentration-dependent ATPase-activity, that closely resembles that observed for

Hsp90. Most striking is the crystal structure of an N-terminal fragment of gyrase B (38). Here,

the structural organization seems to be very similar to that of Hsp90. The model proposed for

the ATPase mechanism of gyrase involves the transient association of the N-terminal domains

in their ATP-bound state (39). This may lead to changes in the active center of the enzyme,

which are the basis for the cooperative hydrolysis of ATP (40). A cooperative mechanism of

this kind would guarantee, that ATP hydrolysis is closely coupled to the coordinated

movement of the two N-terminal domains. It might be envisioned, that Hsp90 this way

couples the energy of ATP hydrolysis to the coordinated movement of domains.

It remains to be seen, how these movements are influenced by the partner proteins of Hsp90

and how they affect the conformational processing of client proteins.

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

18

Literature

1. Wiech, H., Buchner, J., Zimmermann, R. & Jakob, U. (1992) Nature 358, 169-170.

2. Miyata, Y. & Yahara, I. (1992) J. Biol. Chem. 267, 7042-7047.

3. Freeman, B. C. & Morimoto, R. I. (1996) EMBO J. 15, 2969-2979.

4. Jakob, U., Lilie, H., Meyer, I. & Buchner, J. (1995) J. Biol. Chem. 270, 7288-7294.

5. Scheibel, T., Weikl, T. & Buchner, J. (1998), Proc. Nat. Acad. Sci. U.S.A. 95, 1495-

1499.

6. Pratt, W. B. (1998) Proc. Soc. Exp. Biol. Med. 217, 420-434.

7. Buchner, J. (1999) Trends. Biochem. Sci. 24, 136-141.

8. Richter, K. & Buchner, J. (2001) J. Cell. Physiol., in press.

9. Hu, J. & Seeger, C. (1996) Proc. Natl. Acad. Sci. U.S.A. 93, 1060-1064.

10. Holt, S. E., Aisner, D. L., Baur, J., Tesmer, V. M., Dy, M., Ouellette, M., Trager, J. B.,

Morin, G. B., Toft, D. O., Shay, J. W., Wright, W. E. & White, M. A. (1999) Genes

Dev. 13, 817-826.

11. Nathan, D. F. & Lindquist, S. (1995) Mol. Cell. Biol. 15, 3917-3925.

12. Obermann, W. M. J., Sondermann, H., Russo, A. A., Pavlevitch, N. P. & Hartl, F. U.

(1998) J. Cell. Biol. 143, 901-910.

13. Panaretou, B., Prodromou, C., Roe, S. M., O’Brien, R., Ladbury, J. E., Piper, P. W. &

Pearl, L. H. (1998) EMBO J. 17, 4829-4836.

14. Whitesell, L., Mimnaugh, E. G., DeCosta, B., Myers, C. E. & Neckers, L. M. (1994)

Proc. Natl. Acad. Sci. U.S.A. 91, 8324-8328.

15. An, W. G., Schnur, R. C., Neckers, L. & Blgosklonny, M. V. (1997) Cancer Chemother.

Pharmacol. 40, 60-64.

16. Csermely, P., Kajtar, J., Hollosi, M., Jalsovszky, G., Holly, S., Kahn, C. R., Gergely, P.,

Soti, C., Mihaly, K. & Somogyi, J. (1993) J Biol Chem. 268, 1901-1907.

17. Grenert, J. P., Sullivan, W. P., Fadden, P, Haystead, T. A., Clark, J., Mimnaugh, E.,

Krutzsch, H., Ochel, H. J., Schulte, T. W., Sausville, E., Neckers, L. M. & Toft, D.O.

(1997) J. Biol. Chem. 272, 23832-23850.

18. Scheibel, T., Neuhofen, S., Weikl, T., Mayr, C., Reinstein, J., Vogel, P. D. & Buchner,

J. (1997) J. Biol. Chem. 272, 18608-18613.

19. Prodromou, C., Roe, S. M., O’Brien, R., Ladbury, J. E., Piper, P. W. & Pearl, L. H.

(1997) Cell 90, 65-75.

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

19

20. Stebbins, C. E., Russo, A. A., Schneider, C., Rosen, N., Hartl, F. U. & Pavletich, N. P.

(1997) Cell 89, 239-250.

21. Dutta, R. & Inouye, M. (2000) Trends. Biochem. Sci. 25, 24-28.

22. Weikl, T., Muschler, P., Richter, K., Veit, T., Reinstein, J. & Buchner, J. (2000) J. Mol.

Biol. 303, 583-592.

23. Ban, C. & Yang, W. (1998) Cell 95, 541-552.

24. Ban, C., Junop, M. & Yang, W. (1999) Cell 97, 85-97.

25. Prodromou, C., Panaretou, B., Chohan, S., Siligardi, G., O’Brien, R., Ladbury, J. E.,

Roe, S. M., Piper, P. W. & Pearl, L. H. (2000) EMBO J. 16, 4383-4392.

26. Maruya M., Sameshima M., Nemoto T. & Yahara I. (1999) J. Mol. Biol. 285, 903-907.

27. Johnson, J. L. & Toft, D. O. (1994) J. Biol. Chem. 269, 24989-24993.

28. Johnson, J. L., Corbisier, R., Stensgard, B. & Toft, D. O. (1996) J. Steroid. Biochem.

Mol. Biol. 56, 31-37.

29. Chadli, A., Bouhouche, I., Sullivan, W., Stensgard, B., McMahon, N., Catelli, M. G. &

Toft, D. O. (2000) Proc. Natl. Acad. Sci. USA 97, 12524-12529.

30. Nemoto, T., Ohara-Nemoto, Y., Ota, M., Takagi, T. & Yokoyama, K. (1995) Eur. J.

Biochem. 233, 1-8.

31. Jibard, N., Meng, X., Leclerc, P., Rajkowski, K., Fortin, D., Schweizer-Groyer, G.,

Catelli, M. G., Baulieu, E. E. & Cadepond, F. (1999) Exp. Cell. Res. 247, 461-474.

32. Jakob, U., Meyer, I., Bügl, H., Andre, S., Bardwell, J. C. & Buchner, J. (1995) J. Biol.

Chem. 270, 14412-14419.

33. Bartholmes, P., Durchschlag, H. & Jaenicke, R. (1973) Eur. J. Biochem. 39, 101-108.

34. Berr, K., Wassenberg, D., Lilie, H., Behlke, J. & Jaenicke, R. (2000) Eur. J. Biochem.

267, 5413-5420.

35. Surette, M. G., Levit, M., Liu, Y., Lukat, G., Ninfa, E. G., Ninfa, A. & Stock, J. B.

(1996) J. Biol. Chem. 271, 939-945.

36. Todd, M. J., Viitanen, P. V. & Lorimer, G. H. (1994) Science 265, 659-666.

37. Weissman, J. S., Rye, H. S., Fenton, W. A., Beechem, J. M. & Horwich, A. L. (1996)

Cell 84, 481-490.

38. Wigley, D. B., Davies, G. J., Dodson, E. J., Maxwell, A. & Dodson, G. (1991) Nature

351, 624-629.

39. Kampranis, S. C. & Maxwell, A. (1998) J. Biol. Chem. 273, 26305-26309.

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

20

40. Brino, L., Urzhumtsev, A., Mousli, M., Bronner, C., Mitschler, A., Oudet, P. & Moras,

D. (2000) J. Biol. Chem. 275, 9468-9475.

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

21

Acknowledgements

The authors would like to thank Dr. Stefan Walter for stimulating discussions and Alex Frenzl

and Martin Haslbeck for help with the artwork. J.B. was supported by grants from the

Deutsche Forschungs Gemeinschaft and the Fonds der chemische Industrie.

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

22

Figure Legends

Figure 1.

Fragments of Hsp90 used in this study. The ATP binding site is located in the N - terminal

domain (■ ), while the dimerization site is in the C - terminal domain (■ ). The middle part of

Hsp90 (❑ ) may consist of several domains. The hydropathyblot was calculated using the

Expasy-Tool ProtScale software.

Figure 2. ATPase activity of Hsp90 fragments

ATPase assays were performed at 37°C in a buffer containing 40mM HEPES, 150mM KCl,

5mM MgCl2, 2mM ATP, pH 7.5.

A) ATPase activity of N-terminal Hsp90 fragments. All values were corrected for background

activity using geldanamycin. Protein concentrations were 20µM for Hsp90 – N210 and Hsp90

– N272, 5µM for Hsp90 – N530 and 3µM for Hsp90 – N599 and wt-Hsp90.

B) The concentration dependence of the ATPase activity of N-terminal Hsp90 fragments. (▼)

Hsp90 - N599, (❍ ) Hsp90 - N529, (● ) Hsp90 - N210. Lines represent the fits of the data to the

estimated end point of the wt - Hsp90 activity of 1.2µM phosphate / min*µM Hsp90

Figure 3. Analysis of Hsp90 dimerization by SEC - HPLC

A) Size-exclusion HPLC of Hsp90. The running buffer contained 40mM HEPES, 150mM

KCl, pH 7.5 at 20°C. Protein was detected by fluorescence with an excitation wavelength of

280nm and an emission wavelength of 328nm. The shift of the peaks from 20min to 24min

elution time represents the dissociation of Hsp90. ( ____ ) 1µM Hsp90, (.......) 0.03µM Hsp90,

(------) 5nM Hsp90.

B) Dissociation curve of the Hsp90 dimer. The data points of the SEC analysis were fitted

according to equation 1 (see materials and methods section). The resulting dissociation

constant is 60nM +/- 12nM.

C) Dissociation curve of the C-terminal fragment Hsp90 - 262C (● ) and Hsp90 - N529 (❍ )

and Hsp90 – N599 (❑ ) The dissociation constant for Hsp90 - 262C is 45nM +/- 12nM. The

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

23

elution time of Hsp90 - N529, which is suggested to be monomeric does not change over a

wide range of concentrations, as does the elution time of Hsp90 – N599.

Figure 4. Demonstration of heterodimer formation between Hsp90 and Hsp90 - 527C by

crosslinks

A) SDS-PAGE of crosslinked Hsp90 complexes in a buffer containing 40mM HEPES, pH

7.5, 150mM KCl, 5mM MgCl2 and 2mM AMP-PNP. Hsp90 - 527C was added at a

stoichiometry of 8:1, 4:1, 2:1, 1:1, 1:2 and 1:4 (lanes 4-9) to Hsp90. Crosslinking was

performed for 2 minutes at 30°C with glutaraldehyde. Lane 1: high molecular weight standard,

lane 2: Hsp90 without cross-linking, lane 3 and lane 10: crosslinking of Hsp90 - 527C and wt

- Hsp90 respectively. D: Dimer, M: Monomer, HD: Heterodimer, D2: Dimer of Hsp90 -

527C.

B) Densitometric analysis of the SDS-PAGE. (▼) wt - Hsp90 dimer, (● ) wt - Hsp90

monomer, (❍ ) heterodimer of wt - Hsp90 and Hsp90 - 527C.

Figure 5. Influence of heterodimer formation on the ATPase activity of Hsp90

To form heterodimers during the ATPase assay the concentration of the C - terminal fragment

was varied between 1µM and 20µM and the wt - Hsp90 concentration was kept constant at

2.5µM. ATPase assays were performed as described in Material and Methods.

A) Addition of Hsp90 - 527C to Hsp90. The decrease can be fitted to a statistical model. The

n-value was obtained as 0.8, suggesting similar association constants for the homodimer and

the heterodimer.

B) Addition of Hsp90 - 262C to Hsp90. The specific activity of 2.5µM wt-Hsp90 decreases

with increasing concentrations of Hsp90 - 262C. The n-value obtained from fitting the data

points to a statistical model was 0.7.

Figure 6. Influence of an inactive N - terminal domain on the ATPase activity of Hsp90

heterodimers

Hsp90 - D79N was added with increasing concentrations to 2.5µM wt - Hsp90. ATPase

assays was performed as described in Material and Methods.

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

24

A) The specific activity of wt - Hsp90 is not influenced by the presence of increasing amounts

of Hsp90 - D79N, which is unable to bind ATP (● ). These values are compared to the

inhibition obtained with Hsp90 - 262C (❍ ) (cf. Fig. 5B).

B) The dissociation curve of Hsp90 - D79N shows its ability to form dimers. The experiment

was performed as described in Figure 3. The dissociation constant is 85nM +/- 20nM.

Figure 7. Kinetics of subunit exchange in the Hsp90 dimer

A) ATPase inhibition of Hsp90 by heterodimer formation with Hsp90 - 262C. The ATPase

activity of Hsp90 was analysed as described. Starting with Hsp90 (❍ ), Hsp90 - 262C was

added to the ATPase assay after 2 minutes (see arrow) (▲). After roughly 20 seconds of

mixing, the ATPase assay was continued. The ATPase activity levelled off after about 30

seconds, indicating that subunit exchange occurs very efficient.

B) Restoration of ATPase activity by adding Hsp90 - D79N to a solution containing

Hsp90/Hsp90 - 262C heterodimers. The specific activity increases after addition of Hsp90 -

D79N (see arrow). Again subunit exchange is fast compared to the deadtime of the

measurement. (▲) ATPase assay without addition of Hsp90 - D79N, (❑ ) ATPase assay after

addition of Hsp90 - D79N.

Figure 8. Model of the ATPase cycle of Hsp90

The kinetic values are from Weikl et al. (22). These measurements were performed at 25°C

under buffer conditions similar to the ATPase - assays in this study. The model does not

reflect data from Maruya et al., (26), that indicate an antiparallel organization of the Hsp90

dimer in its ATP free state. The N-terminal interaction, as demonstrated in this study,

stimulates the ATP hydrolysis reaction in both domains, leading to the activated species

painted in brackets. After ATP hydrolysis, the phosphate and the ADP are likely to be released

after opening, as the ATP had been trapped inside the protein before hydrolysis occured.

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

25

Figure 1

amino acid number

0 200 400 600

hy

dro

ph

ob

icit

y s

co

re

-4

-3

-2

-1

0

1

2

His6 -

His6 -

His6 -

His6 -

wt - Hsp90

Hsp90 - N210

Hsp90 - N272

Hsp90 - N529

Hsp90 - N599

Hsp90 - 262C

Hsp90 - 527C

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from

Klaus Richter, Paul Muschler, Otmar Hainzl and Johannes BuchnerCoordinated ATP hydrolysis by the Hsp90 dimer

published online July 5, 2001J. Biol. Chem. 

  10.1074/jbc.M103832200Access the most updated version of this article at doi:

 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

by guest on February 21, 2018http://w

ww

.jbc.org/D

ownloaded from