author's personal copy · author's personal copy thermodynamics of radicicol binding to...

12
This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and education use, including for instruction at the authors institution and sharing with colleagues. Other uses, including reproduction and distribution, or selling or licensing copies, or posting to personal, institutional or third party websites are prohibited. In most cases authors are permitted to post their version of the article (e.g. in Word or Tex form) to their personal website or institutional repository. Authors requiring further information regarding Elsevier’s archiving and manuscript policies are encouraged to visit: http://www.elsevier.com/copyright

Upload: others

Post on 19-Oct-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Author's personal copy · Author's personal copy Thermodynamics of radicicol binding to human Hsp90 alpha and beta isoforms Asta Zubrien a, Malgorzata Gutkowskab,c, Jurgita Matulien

This article appeared in a journal published by Elsevier. The attachedcopy is furnished to the author for internal non-commercial researchand education use, including for instruction at the authors institution

and sharing with colleagues.

Other uses, including reproduction and distribution, or selling orlicensing copies, or posting to personal, institutional or third party

websites are prohibited.

In most cases authors are permitted to post their version of thearticle (e.g. in Word or Tex form) to their personal website orinstitutional repository. Authors requiring further information

regarding Elsevier’s archiving and manuscript policies areencouraged to visit:

http://www.elsevier.com/copyright

Page 2: Author's personal copy · Author's personal copy Thermodynamics of radicicol binding to human Hsp90 alpha and beta isoforms Asta Zubrien a, Malgorzata Gutkowskab,c, Jurgita Matulien

Author's personal copy

Thermodynamics of radicicol binding to human Hsp90 alpha and beta isoforms

Asta Zubrienė a, Malgorzata Gutkowska b,c, Jurgita Matulienė a, Romanas Chaleckis a, Vilma Michailovienė a,Aliona Voroncova a, Česlovas Venclovas d, Alicja Zylicz b, Maciej Zylicz b, Daumantas Matulis a,⁎a Laboratory of Biothermodynamics and Drug Design, Institute of Biotechnology, LT-02241 Vilnius, Lithuaniab International Institute of Molecular and Cell Biology in Warsaw, 02-109 Warsaw, Polandc The Nencki Institute of Experimental Biology, 02-093 Warsaw, Polandd Laboratory of Bioinformatics, Institute of Biotechnology, LT-02241 Vilnius, Lithuania

a b s t r a c ta r t i c l e i n f o

Article history:Received 7 July 2010Received in revised form 14 September 2010Accepted 15 September 2010Available online 12 October 2010

Keywords:Isothermal titration calorimetryThermal shift assayThermoFluorHeat shock protein 90Enthalpy

Radicicol is a natural antibiotic that specifically inhibits chaperone Hsp90 activity and binds to its active sitewith nanomolar affinity. Radicicol has been widely used as a lead compound to generate synthetic analogswith reduced toxicity and increased stability that could be employed clinically. Here we present a detailedthermodynamic description of radicicol binding to human Hsp90 and yeast Hsc82 studied by isothermaltitration calorimetry and thermal shift assay. Titrations as a function of pH showed a linked protonation eventupon radicicol binding. The intrinsic binding constant and the thermodynamic parameters (including theenthalpy, entropy, and heat capacity) were determined for yeast Hsc82, and human alpha and beta Hsp90.Recent experimental evidence in literature shows that yeast Hsc82 has significant differences from humanHsp90 isozymes. Here we support this by demonstrating differences in radicicol binding thermodynamics tothese proteins. The intrinsic enthalpy of radicicol binding to Hsc82 was −46.7 kJ/mol, to Hsp90alpha−70.7 kJ/mol, and to Hsp90beta was −66.8 kJ/mol. The enthalpies of binding were significantly different,while the intrinsic dissociation constants were quite similar, equal to 0.25, 0.04, and 0.15 nM, respectively.The structural features responsible for such large difference in binding enthalpy but small difference in theintrinsic binding Gibbs free energy are discussed.

© 2010 Elsevier B.V. All rights reserved.

1. Introduction

Heat shock protein 90 (Hsp90) is a component of the cellularchaperonemachinery [1,2]. There are a number of recent developmentsin the understanding of the interesting and complex mechanism ofHsp90 action [3–7]. Hsp90 is over-expressed in cancer cells and Hsp90inhibitors have shown selectivity for cancer cells. Therefore, small-molecule inhibitors are being developed as anticancer therapeutics[8–12].

Two groups of natural product inhibitors of Hsp90, based ongeldanamycin (GM) and radicicol (RD) have been discovered thatbind to the N-terminal domain at the active site ATP-binding pocket.Bothnatural compoundshavebeenused as leads todevelopcompoundswith desired pharmaceutical properties such as increased potency andreduced toxicity [1,9]. RD is themore potent binder of the two. Radicicol,also known as monorden, was originally discovered as an antifungalsubstance of fungal origin in 1953 [13] and is a specific Hsp90 inhibitor[14]. The co-crystal structures of radicicol bound to the Hsp90 N-terminal domains from various species have been determined, and theyshow numerous hydrogen bondswith the ATP-binding pocket of the N-terminal domain [15,16]. As determined using isothermal titration

calorimetry (ITC) full-length yeast Hsc82 binds radicicol with a Kd of19 nM, and theN-terminal domain of yeast Hsc82 binds radicicol with aKd of 2.7 nM [15].

Here, we determine the thermodynamics of RD binding to the N-terminal domain of human Hsp90 using a combination of ITC andthermal shift assay (TSA). Isothermal titration calorimetry fullycharacterizes the thermodynamics of the binding reaction, includingnot only the Kd (Gibbs free energy), but also the enthalpy, entropy, andheat capacity of binding [17–20]. The full thermodynamic descriptionof binding is important for structure-based drug development [18,20].The thermal shift assay [21], also known as differential scanningfluorimetry [22] andThermoFluor® [23], is a high-throughput screeningmethod for hit selection and the determination of the protein-ligandbinding constants used in the pharmaceutical industry [24]. Thisbiophysical technique can be applied to anyprotein-ligandnon-covalentbinding reaction, independent of whether the ligand stabilizes ordestabilizes the protein upon binding [25]. In cases when there is asingle binding event, the ITC and TSA results complement each other forincreased precision of themeasurements [26]. In addition, themethod isuseful for the characterization of protein stability in the presence ofvarious excipients [27] and the optimization of conditions for proteincrystallization [28].

We have redetermined radicicol binding thermodynamics to yeastHsc82 and determined binding to human Hsp90 alpha and beta

Biophysical Chemistry 152 (2010) 153–163

⁎ Corresponding author. Tel.: +370 5269 1884; fax: +370 5 260 2116.E-mail address: [email protected] (D. Matulis).

0301-4622/$ – see front matter © 2010 Elsevier B.V. All rights reserved.doi:10.1016/j.bpc.2010.09.003

Contents lists available at ScienceDirect

Biophysical Chemistry

j ourna l homepage: ht tp : / /www.e lsev ie r.com/ locate /b iophyschem

Page 3: Author's personal copy · Author's personal copy Thermodynamics of radicicol binding to human Hsp90 alpha and beta isoforms Asta Zubrien a, Malgorzata Gutkowskab,c, Jurgita Matulien

Author's personal copy

isozymes using ITC and TSAmethods to estimate the binding affinity ofvery tight interactions. Furthermore, linked protonation effects weredetermined and dissected from intrinsic binding thermodynamicsshowing large differences in intrinsic enthalpies but small differencesin the intrinsic Gibbs free energies of binding. Our findings supportrecent demonstrations [29–33] which show that Hsc82, used as amodel and archetypical Hsp90machine, showsdifferences in themodeof action compared to other family members.

Because the binding affinity is a combined function of the bindingenthalpy and the binding entropy, improved affinity could resultwhen any or both terms are designed to contribute more favorably tobinding [34]. The enthalpy and heat capacity of binding correlate withstructural parameters such as hydrogen bond formation and hydro-phobic contacts more closely than the Gibbs free energy. Despitestructural determination of radicicol binding to Hsc82, the thermody-namic characterization of the linked protonation reactions [35] andcomparison with human Hsp90α and β isoforms have been insuffi-ciently detailed to understand the binding energetics.

2. Results

2.1. Isothermal titration calorimetry of radicicol binding to Hsp90

The energetics of radicicol binding to various constructs of humanHsp90α and Hsp90β, and yeast Hsc82wasmeasured using ITC. Fig. 1Ashows a representative raw data titration of Hsp90αN with radicicol.The binding reaction is strongly exothermic. Thefirst 12peaks represent

the heats of binding and heats of dilution whereas the small peaks afterthe 12th injection represent the heats of dilution. The binding ofradicicol to Hsp90 was found to be stoichiometric under all conditionsexamined. Variation of the time between injections did not affect theobserved enthalpy indicating that the binding measurements weredone under reversible equilibrium binding conditions.

Fig. 1B shows the integrated ITC curves of radicicol binding toHsp90αN in two buffers, 50 mM sodium phosphate or 50 mM Tris,both at pH 7.5, and at 25 °C. Similarly, Fig. 1C shows the integrated ITCcurves of radicicol binding to yeast Hsc82F in the same two buffers.The enthalpies of binding are different in buffers of different ionizationenthalpies. Therefore, one or more proton binding/release events arelinked to radicicol binding for both human and yeast proteins. Todetermine the intrinsic thermodynamic parameters of binding, theproton binding energetics has to be dissected from the energetics ofradicicol binding.

Fig. 1D compares the integrated ITC curves of radicicol binding tothe N-terminal domains of α and β Hsp90 isozymes with the full-length proteins. It was important to determine whether there is adifference in the binding thermodynamics to α and β Hsp90 isoforms,as both isozymes are present in the human body. The binding to the N-terminal ATP-binding domain of the protein is often studied using theHsp90 protein construct lacking the middle and C-terminal domains.Therefore, it was also important to determine whether the N-terminaldomain of both isozymes bind radicicol with the same energetics asthe full-length native protein. Differences between radicicol bindingto the N-terminal and full-length protein were negligible and within

1,40

1,45

1,50

1,55

1,60

1,65

1,70

0 1000 2000 3000 4000 5000

Pow

er, µ

cal/s

ec

Time, sec

A

D

-90-80-70-60-50-40-30-20-10

010

0.0 0.5 1.0 1.5 2.0

[Rdc]/[Hsp90], mol/mol

δ bH

obs,

kJ/m

ol

Hsp90αFHsp90αN

Hsp90βN

Hsp90βF

C

-70

-60

-50

-40

-30

-20

-10

0

10

0.0 0.5 1.0 1.5 2.0

[Rdc]/[Hsc82], mol/mol

Δ bH

obs,

kJ/m

ol

Phosphate

Tris

-80

-70

-60

-50

-40

-30

-20

-10

0

10

0.0 0.5 1.0 1.5 2.0

[Rdc]/[Hsp90αN], mol/mol

Δ Δ bH

obs,

kJ/m

ol

Phosphate

Tris

B

Fig. 1. Representative isothermal titration calorimetry curves of radicicol binding to various human and yeast Hsp90 family member protein constructs. Panel A. The raw data curvesof radicicol binding to Hsp90αN in 50 mM sodium phosphate, pH 7.5, at 25 °C. Insert shows the chemical structure of radicicol. Panel B. The integrated ITC data curves of radicicolbinding to Hsp90αN in two buffers: ■ — in sodium phosphate, and Δ — in Tris chloride buffer, both at pH 7.5 and 25 °C. Note that the observed enthalpy of binding is significantlymore exothermic in phosphate than in Tris buffer. Panel C. The integrated ITC data curves of radicicol binding to yeast Hsc82 in the same conditions. Note that the observedenthalpies are significantly different from Hsp90αN. Panel D. Integrated ITC data curves of radicicol binding to various constructs of Hsp90: ▲ — Hsp90αN, Δ — Hsp90αF, ■ —

Hsp90βN, and □ — Hsp90βF. Solid lines show curve fits for the N-terminal Hsp90 while the dashed lines show curve fits for the full-length Hsp90. Differences between the N-terminal and full-length protein binding were within the experimental error. All of the tested constructs bound radicicol stoichiometrically. The intrinsic parameters of radicicolbinding to various Hsp90 constructs and isozymes are summarized in Table 2.

154 A. Zubrienė et al. / Biophysical Chemistry 152 (2010) 153–163

Page 4: Author's personal copy · Author's personal copy Thermodynamics of radicicol binding to human Hsp90 alpha and beta isoforms Asta Zubrien a, Malgorzata Gutkowskab,c, Jurgita Matulien

Author's personal copy

the experimental error of ITC measurements. Furthermore, the ITCexperiments were performed as a function of buffer, pH, temperature,and the type of Hsp90/Hsc82 construct to dissect the thermodynamicparameters of binding from the linked protonation values. All of theHsp90 and Hsc82 constructs tested exhibited the protonation eventlinked to radicicol binding. The observed enthalpies of binding showeda marked difference between the values obtained in the Tris andphosphate buffers in the alkaline pH region. Human and yeast proteinsbound radicicol with significantly different observed enthalpies andKds. Table 1 shows the observed binding parameters of selected ITCruns. The enthalpies of buffer ionization were taken from reference[36].

A theoretical treatment demonstrating the use of ITCmeasurementsto dissect proton linkage from ligand bindingwas given byMurphy andcoworkers [37–39]. If a protein has a single ligand-binding site and asingle proton uptake is linked to the binding process (i.e., a proton istaken from the buffer solution), then there are four linked processesdescribed by the thermodynamic parameters labeled F in Fig. 2, Panel A.

When ligand binding affects and shifts the pKa of any ionizablegroup on a protein molecule, the binding and protonation events arelinked. All thermodynamic parameters, such as the Gibbs free energies(or binding constants), enthalpies, entropies, and the heat capacities,are additive. As it takes energy to shift the pKa, the binding constant ofa ligand would be diminished at the pH at which the proton needs tobe taken or given to the buffer. If ligand binding is linked to the

binding of a single proton, then the observed binding constant (Kobs)and the intrinsic binding constant to the protonated protein form(Kintr) are related by:

Kobs = Kintr1 + 10pH−pKb

a

1 + 10pH−pKfa

ð1Þ

Kab and Ka

f are the proton dissociation constants from the liganded andunliganded protein, respectively. The change in the number of protonsbound by the protein upon binding of the ligand (n) is the differencebetween the fractional saturation of protons in the liganded and freeprotein:

n = f bp−f fp =10pKb

a−pH

1 + 10pKba−pH

− 10pKfa−pH

1 + 10pKfa−pH

ð2Þ

The value of n can be determined by ITC because it contributes tothe observed binding enthalpy (ΔbHobs):

ΔbHobs = ΔbH + nΔbHbuffer ð3Þ

ΔbH is the enthalpy that would be measured in a buffer that has anionization enthalpy ΔbHbuffer equal to zero. However, it is not equal tothe intrinsic binding enthalpy. A series of observed enthalpiesmeasuredat constant pH in buffers with different ionization enthalpies (Fig. 2B)plottedas a functionof buffer ionization enthalpywill yield a slopeequalto n and an intercept of ΔbH (Fig. 2C). The n values (Eq. (2)) plotted as afunction of the pH are shown in Fig. 2D. The data points are obtainedfrom the slopes as shown in Fig. 2C, whereas the line is fitted to Eq. (2)and yields a pKaf equal to 6.6. The enthalpies in Fig. 2B are fitted using thesame pKa

f (6.6). The pKab could not be determined and is greater than 11because there is no return to n=0 in the pH region that is accessible forbinding measurements. The enthalpy of protonation of the ionizablegroup that is affected by radicicol binding in the Hsp90αN-radicicolcomplex was best fit by the value of −8.0 kJ×mol−1. The intrinsic(buffer-independent) enthalpy of radicicol binding can be estimatedfrom the relationship:

ΔbHobs = ΔbHintr + nΔbHcomplex + nΔbHbuffer ð4Þ

As no experimental conditions were found where radicicol wouldbind to the deprotonated protein form, it was not possible to determinethe protonation enthalpy difference between the free and ligandedforms of Hsp90.

Fig. 3 shows the dependencies of the intrinsic binding enthalpies as afunction of temperature. The slopes of the lines give the intrinsic heatcapacities of radicicol binding to Hsp90. The heat capacity of radicicolbinding to Hsp90αN is equal to approximately −620 J×mol−1×K−1.Table 2 lists the intrinsic thermodynamic parameters of radicicolbinding to all tested Hsp90 protein constructs.

The binding of radicicol to Hsp90 is very tight, on the order ofsingle-digit nanomolar or even tighter. Direct measurements bytitrating Hsp90 with radicicol by ITC cannot estimate the bindingconstant because the curve is too steep, as previously discussed [40].The c value (c=C×Kb, C is the molar concentration of the protein, andKb is the binding constant, when the binding stoichiometry is 1:1) foran ITC titration must be between 5 and 500. However, our ITCmeasurements were carried out using Hsp90 concentrations of atleast 4 μM, and the binding constant of radicicol is expected to beapproximately 2×1010 M−1 [40]. Thus, the c value is at least 80,000,significantly above the valueof 500 required for ITC. Therefore, differentmethods are required tomeasure the binding constant accurately in thispH region. This could be done by ITC using displacement titration of aweakly-binding ligand. However, this approach did not work well forradicicol [40]. Themost precise results were obtained using the thermalshift assay.

Table 1Representative ITC data for the binding of radicicol to recombinant human and yeastHsp90 protein constructs as a function of pH, temperature, and buffer ionizationenthalpy.

Construct Temp., °C pH Buffer ΔionizHbufa ΔbHobs

a Kd, nM

Hsp90αN 13 7.5 Pib 7.4 −63.5 1.325 7.5 Pi 5.1 −72.9 0.1537 7.5 Pi 2.9 −81.2 0.4725 5.5 Tris 47.5 −71.1 0.2425 6.0 Tris 47.5 −59.3 0.1425 6.5 Tris 47.5 −55.5 0.3425 7.0 Tris 47.5 −48.7 0.0325 7.5 Tris 47.5 −44.8 0.9125 8.0 Tris 47.5 −40.7 1.425 8.5 Tris 47.5 −37.7 7.425 9.0 Tris 47.5 −35.9 73.225 6.0 Pi 5.1 −68.7 0.2925 8.5 Pi 5.1 −72.8 8.6

Hsp90αF 13 7.5 Pi 7.4 −55.5 1.0625 7.5 Pi 5.1 −68.6 0.3237 7.5 Pi 2.9 −77.4 0.4325 6.0 Tris 47.5 −51.1 1.2325 7.5 Tris 47.5 −40.7 0.2125 8.5 Tris 47.5 −35.6 5.825 6.0 Pi 5.1 −69.0 2.025 8.5 Pi 5.1 −62.3 4.0

Hsp90βN 13 7.5 Pi 7.4 −56.3 2.925 7.5 Pi 5.1 −62.7 0.537 7.5 Pi 2.9 −75.3 3.125 6.0 Tris 47.5 −55.2 0.5525 7.5 Tris 47.5 −35.7 0.3425 8.5 Tris 47.5 −30.9 10.0625 6.0 Pi 5.1 −66.9 0.0725 8.5 Pi 5.1 −60.3 5.91

Hsp90αN 37 7.5 Hepes 21.6 −70.4 0.6Hsp90αF 37 7.5 Hepes 21.6 −69.1 0.59Hsp90βN 37 7.5 Hepes 21.6 −59.3 2.0Hsp90βF 37 7.5 Hepes 21.6 −59.0 3.0yeastHsc82N 37 7.5 Hepes 21.6 −52.5 9.6yeastHsc82F 37 7.5 Hepes 21.6 −53.2 26.0

25 7.5 Tris 47.5 −18.5 2.3813 7.5 Pi 7.4 −42.6 1.4625 7.5 Pi 5.1 −48.8 7.6937 7.5 Pi 2.9 −57.6 13.9

a: units in kJ×mol−1; b: Pi is phosphate buffer. The standard deviations were 3 to5 kJ×mol−1 for the enthalpy and up to 1.6 fold for the Kd, especially when the bindingwas too tight to measure accurately using ITC.

155A. Zubrienė et al. / Biophysical Chemistry 152 (2010) 153–163

Page 5: Author's personal copy · Author's personal copy Thermodynamics of radicicol binding to human Hsp90 alpha and beta isoforms Asta Zubrien a, Malgorzata Gutkowskab,c, Jurgita Matulien

Author's personal copy

2.2. Thermal shift assay of radicicol binding to Hsp90

The thermal shift assay measures the binding constant of a ligandby determining the increase in the melting temperature of the proteinthat is caused by the ligand [25,26]. Protein unfolding is monitoredusing the intrinsic tryptophan fluorescence or the fluorescence of anextrinsic probe (such as 1,8-anilinonaphtalene sulfonate, ANS) uponincreasing the temperature of the protein solution at a constantheating rate. Soluble single-domain globular proteins melt with asingle transition that is affected by ligands. Full-length Hsp90 yields acomplicatedmulti-domainmelting transition profile. However, the N-terminal domain of Hsp90 yields a single transition that can beanalyzed using the thermal shift assay.

Fig. 4 shows ANS fluorescence curves as a function of temperature attwo pH values (pH 7.0 in panel A and pH 10.0 in panel B) and severalradicicol concentrations. The concentration of Hsp90αN was 10 μM.With no radicicol added, andwhen the pH is 7.0, there is a steep increasein fluorescence observed at approximately 50 °C. This increase is due toprotein unfolding that exposes hydrophobic surfaces where ANS canbindandbe excluded fromquenchingby the aqueous environment [41].ANS actually binds using a combination of hydrophobic and ionic

interactions but the ion pairs between the ANS sulfonate groups and theprotein amino groups are not visible using fluorescence [42,43]. Whenthe radicicol concentration is greater than that of the protein, themelting temperature of Hsp90 is increased by more than 10 °C.

Protein melting temperatures were determined by fitting theprotein melting curves (Fig. 4) according to Eq. (5):

y Tð Þ = yF;Tm + mF T−Tmð Þ

+yU;Tm−yF;Tm

� �+ mU−mFð Þ T−Tmð Þ

1 + e ΔUHTm + ΔUCp T−Tmð Þ−T ΔUSTm + ΔUCp ln T =Tmð Þð Þð Þ=RTð5Þ

y(T) is the calculated fluorescence as a function of temperature; yF, Tmis the fluorescence of the probe bound to folded native protein beforethe transition at Tm; yU, Tm is the fluorescence of the probe bound to theunfolded protein after the unfolding transition at Tm;mF is the slope ofthe fluorescence dependence on temperature when the probe isbound to the native protein; mU is the slope of the fluorescencedependence on temperature when the probe is bound to the unfoldedprotein; ΔUHTm is the enthalpy of protein unfolding at Tm; ΔUSTm is theentropy of protein unfolding at Tm; ΔUCp is the heat capacity of protein

A

D

P

PH+ LPH+

H+

LP

L

L

Fpf Fp

b

Fbf

Fbp

-0.1

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

4 5 6 7 8 9 10

pH

n

C

-80

-70

-60

-50

-40

-30

-20

-10

0 10 20 30 40 50

bHob

s , k

J/m

ol n (13 °C) ≈ 0.75

n (25 °C) ≈ 0.72

n (37 °C) ≈ 0.73

ΔbH(13 °C)

ΔaH, kJ/mol

ΔB

-80

-70

-60

-50

-40

-30

4 5 6 7 8 9 10

pH

Tris

PhosphatebHob

s , k

J/m

olΔ

Fig. 2. Demonstration of the linked protonation event observed upon radicicol binding to Hsp90αN. Panel A. Schematic representation of the linked protonation events. Here Fbf

represents ligand binding to unprotonated protein, Fbp represents ligand binding to protonated protein, Fpf represents proton binding to free (unliganded) protein, and Fpb represents

proton binding to ligand-bound protein. Panel B. The observed enthalpies as a function of pH in two buffers: Δ— phosphate, and■— tris. The data points are the enthalpies that wereobserved experimentally using ITC, and the lines are fitted to the single binding-linked protonation model, Eq. (4). Panel C. The observed enthalpies as a function of the bufferdeprotonation enthalpy at various temperatures: ♦— 13 °C,□— 25 °C, and ▲— 37 °C. The data points are the experimentally-observed enthalpies, and the trendlines are linear fits.Their slopes are equal to the binding-linked protonation events. There was little change of n as a function of temperature. The zero intercepts of the lines are equal to the buffer-independent binding enthalpies. The enthalpies are not intrinsic binding enthalpies because they contain the heats of protonation of unknown protein functional groups. Panel D.The number of linked protonation events (n), obtained from the slopes of the lines in Panel C as a function of pH. The line is fitted to the single binding-linked protonation model,Eq. (4). The midpoint of the curve occurring around pH 6.6 indicates the existence of a functional group on the protein that has a pKa ~6.6. Radicicol binds to the protonated form ofthe protein. Therefore the protein must bind a proton to bind radicicol.

156 A. Zubrienė et al. / Biophysical Chemistry 152 (2010) 153–163

Page 6: Author's personal copy · Author's personal copy Thermodynamics of radicicol binding to human Hsp90 alpha and beta isoforms Asta Zubrien a, Malgorzata Gutkowskab,c, Jurgita Matulien

Author's personal copy

unfolding and is assumed to be temperature-independent over thetemperature range studied; R is the universal gas constant; and T isthe absolute temperature (Kelvin).

A number of transition curves at various concentrations of addedligandwere determined and plotted as a function of the concentrationof added ligand, yielding the melting temperature (Tm) dependenciesas a function of the radicicol concentration (Fig. 4C). The curves inFig. 4C were fit globally according to Eq. (6), yielding the pHdependencies of the binding constants.

Lt = KU;Tm−1

� � Pt2KU;Tm

+1

Kb;Tm

!

= e− ΔUHTr + ΔUCp Tm−Trð Þ−Tm ΔUSTr + ΔUCp ln Tm =Trð Þð Þð Þ=RTm−1� �

× ½ Pt2 1

e− ΔUHTr + ΔUCp Tm−Trð Þ−Tm ΔUSTr + ΔUCp ln Tm =Trð Þð Þð Þ=RTm

+1

e− ΔbHT0+ ΔbCp Tm−T0ð Þ−Tm ΔbST0 + ΔbCp ln Tm =T0ð Þ

� �� �=RTm

�ð6Þ

Lt is the total concentration of added ligand, KU, Tm is the proteinunfolding equilibrium constant at Tm; Pt is the total proteinconcentration; Kb, Tm is the ligand binding constant at Tm; ΔUHTr isthe enthalpy of protein unfolding at Tr; Tr is the protein meltingtemperature when no ligand is added; ΔUSTr is the entropy of proteinunfolding at Tr; ΔUCp is the heat capacity of protein unfolding and isassumed to be temperature-independent over the temperature rangestudied; ΔbHT0 is the enthalpy of ligand binding at T0; T0 is thetemperature at which the binding process is studied (usually 37 °C);ΔbST0 is the entropy of ligand binding at T0; and ΔbCp is the heatcapacity of ligand binding and is assumed to be temperature-independent over the temperature range studied.

A

B

-90

-80

-70

-60

-50

10 20 30 40Temperature, oC

Temperature, oC

bHin

tr, k

J/m

ol

ΔCp ≈ -0.62 kJ×mol-1×K-1

ΔCp ≈ -0.86 kJ×mol-1×K-1

ΔCp ≈ -0.76 kJ×mol-1×K-1

ΔCp ≈ -0.62 kJ×mol-1×K-1

Hsp90αN

Hsp90αN

Hsp90αF

-90

-80

-70

-60

-50

-40

10 20 30 40

Hsp90βN

ΔbH

intr

, kJ/

mol

Δ

Fig. 3. Intrinsic binding enthalpies obtained after accounting for the linked protonationevent as a function of temperature for radicicol binding to various proteins. Panel A:■—Hsp90αN, and▲—Hsp90αF. Panel B:■—Hsp90αN, and●—Hsp90βN. The slopesare linear fits to the experimental data and are equal to the intrinsic heat capacities ofradicicol binding. The intrinsic binding parameters are summarized in Table 2.

10

15

20

25

30

35

40

30 40 50 60 70 80Temperature, oC

30 40 50 60 70 80Temperature, oC

Flu

ores

cenc

e, a

rb. u

nits pH 7

0

40

160

10

15

20

25

30

35

40

Flu

ores

cenc

e, a

rb. u

nits pH 10

0

40

160

40

50

60

70

1.E-07 1.E-06 1.E-05 1.E-04 1.E-03Lt, M

A

B

C

Tm

, o C

pH 10.0

pH 7.5

pH 9.0

pH 6.5

Fig. 4. Radicicol binding to Hsp90αN determined from the thermal shift assay. Raw dataat pH 7.0 (Panel A) and pH 10.0 (Panel B). The Hsp90 thermal denaturation transitions(Tm) were increasingly shifted upwards as the concentrations of radicicol increased:■ — 0 μM, × — 40 μM, Δ — 80 μM, and ♦ — 160 μM. Note that both the unliganded andliganded forms of Hsp90αN were destabilized at pH 10.0 and also made the relative Tmshift smaller than at pH 7.0. Panel C. The Tm data from the thermal shift assay as afunction of the concentration of added radicicol at various pH values. The data pointsare obtained from the raw data as in Panels A and B, and the lines are fitted according tothe model, Eq. (6). Note that the protein is most stable at pH 7.5–9.0, but the Tm shiftcaused by the ligand is greatest at pH 6.5. The leftmost data point is obtained when noradicicol is added.

Table 2The intrinsic thermodynamic parameters of radicicol binding to Hsp90 at 25 °C.

Protein Kd,nM

ΔbHintr,kJ×mol−1

ΔbGintr,kJ×mol−1

TΔbSintr,kJ×mol−1

ΔbSintr,J×mol−1×K−1

ΔbCp,J×mol−1×K−1

Hsp90αN 0.04 −70.7 −59.4 −11.4 −38 −620Hsp90αF 0.04 −66.8 −59.4 −7.5 −25 −860Hsp90βN 0.15 −60.6 −56.1 −4.6 −15 −760Hsp90βF 0.15 −61.1 −56.1 −5.0 −17 N.D.Hsc82F 0.25 −46.7 −54.8 8.1 27.1 −620Uncertainties ±1.6-

fold±4 ±2.6 ±4.7 ±16 ±140

157A. Zubrienė et al. / Biophysical Chemistry 152 (2010) 153–163

Page 7: Author's personal copy · Author's personal copy Thermodynamics of radicicol binding to human Hsp90 alpha and beta isoforms Asta Zubrien a, Malgorzata Gutkowskab,c, Jurgita Matulien

Author's personal copy

The binding constant at the physiological temperature T0 isdetermined using Eq. (7):

Kb;T0= e− ΔbHT0

−T0ΔbST0

� �=RT0 ð7Þ

The radicicol dosing curves in Fig. 4C show the greatest shift in Tmat pH 6.5, but the unliganded protein is most stable at pH 7.5. At pH9.0, the shift is significantly smaller than near neutral pH, indicating asignificant decrease in the binding constant.

To compare radicicol binding to the two human Hsp90 isoforms,we carried out a thermal shift assay at various pH values. Fig. 5compares selected fluorescence profiles at pH 7.0 forα and βHsp90N.The melting temperature shift is greater for the alpha isozyme. Asimilar result is seen in Fig. 5B, where the radicicol dosing curves areshown at pH 8.0 for the two isozymes. Therefore, both the ITC andthermal shift assays show stronger radicicol binding to the alphaisozyme rather than the beta isozyme.

Fig. 6 shows the pH dependencies of the binding constants asdetermined using ITC and TSA. Both methods clearly show that theobserved binding constant decreases at pH values greater than approx-imately 7.0. The line in Fig. 6 is fitted using the pKa

f of 6.6 and the Kintr of2.5×1010 M−1. At pH values below 8.5, the observed binding constant isdetermined correctly only using the TSA, and it is underestimated usingITC because the c value is above 500 (as described earlier).

2.3. Thermodynamics of radicicol binding to the Hsp90 andHsc82 isozymes

After dissecting the proton linkage, the intrinsic thermodynamicparameters of radicicol binding to four recombinant human Hsp90

constructs are shown in Fig. 7. All intrinsic thermodynamic para-meters of radicicol binding, including binding to yeast Hsc82, arelisted in Table 2. Most importantly, it appears that radicicol bindsabout 4–5 times more strongly to Hsp90α than to Hsp90β. Truncatingthe protein did not affect the binding constant significantly. However,it should be kept inmind that it wasmore difficult to determine the Kd

for the full protein as the TSA method was not used due to thecomplicated unfolding T-curve. Therefore, the data are based solely onITC comparisons at pH values where reliable ITC data could beobtained (i.e., pH 8.0–9.0).

0

5

10

15

20

25

30

35

35 45 55 65 75Temperature, oC

Flu

ores

cenc

e, a

rb.u

nits pH 7.0

Hsp90βN

Hsp90αN

47

51

55

59

63

1,E-07 1,E-06 1,E-05 1,E-04 1,E-03

Hsp90αN

Hsp90βN

pH = 8.0

A

B

Tm

,o C

Lt, M

Fig. 5. Radicicol binding to the Hsp90 α and β isoforms at pH 8.0 from the thermal shiftassay. Panel A. The raw data. Filled symbols — Hsp90αN, open symbols — Hsp90βN. Thetriangles show the melting curves in the absence of added radicicol, whereas the squaresshow theproteinmelting curves after 20 μMradicicol is added. Panel B. The dependence ofthe melting temperature of Hsp90αN (■) and Hsp90βN (▲) on the concentration ofradicicol. Themelting temperature shift is greater for theα than for theβ isozyme.Analysisof the data shows that the binding constant is greater for the α than for the β isozyme.

1,E+06

1,E+07

1,E+08

1,E+09

1,E+10

1,E+11

5 6 7 8 9 10pH

ITC Tris

ITC Phosphate

TSAITC upper limit

Kb,

M-1

Fig. 6. The observed binding constants of the interaction of radicicol with Hsp90αNobtained using three experimental approaches: Δ — isothermal titration calorimetry inphosphate buffer, ■ — isothermal titration calorimetry in Tris buffer, and ×— thethermal shift assay, all at 25 °C. The line is fitted according to Eq. (1) for a linkedprotonation event using pKa

f =6.6. The pKab is somewhere greater than 11, but the exact

value is unknown. There is a clear decrease in the binding strength at higher pH. Thedashed line shows the upper limit of accurate Kb determination by ITC (where the cvalue (c=C×Kb) is 500). TSA provides more accurate values at low pH than does ITC.

-90-80-70-60-50-40-30-20-10

010

10 20 30 40Temperature, oC

Temperature, oC

bEne

rgy i

ntr,

kJ/m

ol

Hsp90αN

TΔb S intr

Hsp90αF

Hsp90αN

Hsp90αN

Hsp90βN

Hsp90αF

-90-80-70-60-50-40

-30-20-10

010

10 20 30 40

Δb G intr

Δb G intr

Δb H intr

Δb H intr

TΔb S intr

A

Hsp90αN

Hsp90βN

ΔbE

nerg

y int

r, kJ

/mol

Δ

B

Fig. 7. The intrinsic thermodynamic parameters (squares — enthalpies, triangles —

entropies, and circles— Gibbs free energies) of radicicol binding to various proteins as afunction of temperature. Panel A: Hsp90αN (filled symbols) and Hsp90αF (opensymbols). Panel B: Hsp90αN (filled symbols) and Hsp90βN (open symbols). Theintrinsic binding constants and the Gibbs free energies were essentially indistinguish-able for the two Hsp90α constructs (Panel A). Therefore, ΔbG is shown as a single linefor both constructs.

158 A. Zubrienė et al. / Biophysical Chemistry 152 (2010) 153–163

Page 8: Author's personal copy · Author's personal copy Thermodynamics of radicicol binding to human Hsp90 alpha and beta isoforms Asta Zubrien a, Malgorzata Gutkowskab,c, Jurgita Matulien

Author's personal copy

Dissection of the radicicol binding thermodynamics into theirenthalpic and entropic contributions shows that radicicol-Hsp90binding is driven by enthalpy and opposed by entropy. However, theentropic opposition accounts for approximately only 10% of the energyopposing binding. About 90% of the binding energy comes from thefavorable enthalpic contribution. The thermodynamic parameters ofbinding to the full-length protein and the N-terminal domain are verysimilar. The difference is within the error of the measurements(Table 2).

The heat capacity change during radicicol-Hsp90 binding isapproximately equal to −700±140 J×mol×K−1. The slight differ-ence between the N-terminal and full-length forms of Hsp90, andbetween the alpha and beta isoforms is most likely due to the error ofthe measurements. The negative heat capacity is often considered asignature of hydrophobic interactions, and the value is approximatelyof the expected magnitude for a molecule that has the MW ofradicicol.

There is a small but statistically significant difference between theenthalpy and the Gibbs free energy of radicicol binding to Hsp90α andHsp90β. The binding is slightly weaker to Hsp90β than to Hsp90α.This reduction in affinity is caused primarily by less favorableenthalpic rather than entropic contributions. The reduction in thefavorable Gibbs free energy was approximately equal to 4 kJ/mol,whereas the reduction in the favorable exothermic enthalpy wasapproximately equal to 7 kJ/mol.

However, the binding of radicicol to yeast Hsc82N and yeastHsc82F differed significantly from the binding to human Hsp90. Theenthalpy of binding to the yeast forms of both protein constructs wasless exothermic than was the binding to the human protein (Table 1).Full protonation linkage analysis determined that the enthalpy ofprotonation of an unknown protein group linked to radicicol bindingwas essentially the same as in both human isozymes indicating thatthe same residue is responsible for the linked protonation event.

However, the most striking difference between yeast and humanisozymes is in the intrinsic enthalpy of binding. The enthalpy ofradicicol binding to yeast Hsc82F was 20.1 kJ/mol less exothermicthan Hsp90αF isozyme and 14.4 kJ/mol less exothermic than Hsp90βFisozyme. This large difference in enthalpies combined with relativelysmall difference in the intrinsic Gibbs free energies yields opposingcontributions from the entropy of binding. It appears that radicicolbinding to human Hsp90 isozymes is entropically opposed whilebinding to yeast Hsc82 is entropically favored process. However,overall enthalpic contribution to the binding reaction is significantlygreater than the entropic contribution.

Despite the difficulties involved and the large number of ITCtitrations required to carry out the full thermodynamic proton linkageanalysis, it is important to dissect the linkage to obtain the intrinsicbinding parameters of any ligand. The binding of radicicol to Hsp90 isa good example of a single binding-linked protonation event forwhich the thermodynamic characterization of binding would yieldmisleading results without the linkage analysis.

3. Discussion

Radicicol is an important natural inhibitor of Hsp90 that has beenused asa lead compound todesign inhibitors for therapeuticpurposes. Itis now at preclinical development at the Memorial Sloan-KetteringCancer Center [12]. Single amino acid mutations in Hsp90may increaseresistance to radicicol through an increased affinity ofmutatedHsp90 tocochaperoneAha1 [44] andarenot related to change in radicicol affinity.Other mutants, however, directly affect radicicol binding affinity at theactive site [16].

Radicicol binding to Hsc82 was characterized structurally bydetermining the structures of various protein constructs with radicicolusing X-ray crystallography. A detailed mechanism of resistance toradicicol was recently determined for the fungus Humicola fuscoatra,

whichproduces radicicol [5]. TheHsp90 fromthe fungus contains aminoacid substitutions that reduce radicicol affinity but do not affect ATPaseactivity. However, despite detailed structural studies, we lacked adetailed energetic characterization of the interaction of radicicol withHsp90. Various studies only use several ITC experiments and determinea limited number of association constants, without reporting theintrinsic binding enthalpies, entropies, and heat capacities.

To correlate the structural features of binding (such as theintermolecular van der Waals contacts, hydrogen bonds and presenceof water molecules) with the energetics of binding one needs to carryout a detailed thermodynamic study of the binding reaction and todetermine various dependencies (such as the pH-, buffer-, andtemperature-dependencies). If there are any binding-linked proton-ation reactions, it is necessary to dissect their input to everythermodynamic parameter of the binding. Even after this analysis, itmay not be possible to assign energetic parameters to all of thestructural parameters. The binding reaction may contain a number ofother linked events, such as any protein conformational changes thatmay occur upon ligand binding [45].

The dissection of linked protonation reactions and the applicationof two biophysical techniques, i.e., ITC and TSA, determined the Kd

values of radicicol with Hsp90α to be approximately 0.04 nM, withHsp90β to be approximately 0.15 nM, and with Hsc82 to beapproximately 0.25 nM. These values are significantly tighter thandetermined previously using ITC [15] and without the dissection ofthe linked protonation reactions. This approach was more accuratethan the displacement ITC experiments, as previously discussed. [40]The combination of ITC and TSA helps to determine ultra-tight bindingconstants by analyzing the pH-dependence (Fig. 6). Interestingly,geldanamycin was shown to inhibit Hsp90α about 3-fold tighter thanHsp90β [46].

The main uncertainty of the TSA method is the precision of theenthalpy of Hsp90 protein unfolding. A 10% uncertainty in the value ofunfolding enthalpy yields an uncertainty of the binding constant of~2-fold [26]. The enthalpy of unfolding can be determined usingseveral biophysical techniques such as DSC and acid-titration ITC[47,48], and by fitting raw TSA curves or ligand-dosing curves [40].The enthalpy of unfolding was accurately measured by all thesemethods only for the single-domain Hsp90N, but could not be done asprecisely for the full-length multi-domain Hsp90 protein. Therefore,radicicol binding to Hsp90F was not determined by TSA. However, thebinding constants for the full-length protein were shown to be thesame as for the N-terminal Hsp90 domain by the ITC titrations at pH8.0–9.0 and thus assumed to be the same at pH 7.0, where binding istoo tight to be determined by ITC.

There were no such conditions found at high pH where therewould be no linked protonation event. Therefore, it was not possibleto determine the pKa shift of the amino acid in Hsp90 responsible forthe proton uptake at high pH. As the pKa value was approximately 6.6,it is natural to assume that it is a histidine residue that binds a protonfrom aqueous solution upon radicicol binding (i.e., that a histidineresidue is required to be protonated for the Hsp90 to bind radicicol).

However, there are no histidines in human Hsp90 in the proximityof the radicicol binding site. Therefore, the linkage may be related toaspartic acid 93 (corresponding to D79 in yeast Hsc82), which islocated in the immediate vicinity of the radicicol binding site.We haveconstructed D93A and D93E mutants in order to test this possibility.However, both mutants did not express well or went to inclusionbodies indicating that the protein with such mutation has verydifferent structure and stability. As a control, several other adjacent Dmutants expressed and purifiedwell. This indicates that human D93 iscrucial for the protein. Radicicol binding measurements could not becarried out with these mutants.

The enthalpy of protonation of an unknown groupwas determinedto be equal to−8.0 kJ×mol−1. This is too small for a histidine residue(about−29 kJ/mol) or phenolic groupof radicicol (about−21 kJ/mol).

159A. Zubrienė et al. / Biophysical Chemistry 152 (2010) 153–163

Page 9: Author's personal copy · Author's personal copy Thermodynamics of radicicol binding to human Hsp90 alpha and beta isoforms Asta Zubrien a, Malgorzata Gutkowskab,c, Jurgita Matulien

Author's personal copy

However, the value is close to the enthalpy of Asp protonation which isequal to −5.6±1.6 kJ/mol (based on values determined at 25 °C asreviewed in [36]). Therefore, we consider it to be highly likely thatAsp93 is responsible for the linked protonation effect. However, thecertain identity of the amino acid remains to be determined.

Fig. 8A shows the structural arrangement of important radicicolinteractions bound to Hsc82 (PDB ID 1bgq). Radicicol forms hydrogenbonds only with Asp79 and several water molecules. Other directcontacts between radicicol and protein are identical for the threeprotein isoforms. Therefore, other amino acids are not shown. There isno crystal structure available in the PDB of radicicol bound to anyhuman Hsp90 isozyme. Therefore, a computational model of humanHsp90α in the radicicol-bound form was constructed on the basis of

the high-resolution (1.55 Å) structure of human Hsp90α complexedwith dihydroxyphenyl amide inhibitor (PDB code: 3EKO). First, thehuman Hsp90α structure (3EKO) was superimposed with the nativeand mutant yeast Hsc82 structures (PDB: 1BGQ and 2WER, respec-tively) that were solved with radicicol bound. Superposition of thestructures revealed that the aromatic rings of the inhibitors super-imposed very closely in all three structures. Therefore, we obtainedthe human Hsp90α complex with radicicol by simply substituting thedihydroxyphenyl amide inhibitor with radicicol from any of the super-imposed yeast structures. The radicicol-bound Hsp90β model wasobtained using the model of Hsp90α bound to radicicol, constructed asdescribed above. The few side chains that differ between Hsp90α andHsp90β were replaced in the Hsp90α structure and positioned

Fig. 8. A structural model of radicicol binding to Hsp90. Panel A. Experimental crystal sctructure of radicicol bound to the yeast Hsc82 (pdb ID 1bgq). Panels B and C. Computationalmodels of radicicol bound to the human Hsp90α (Panel B) and Hsp90β (Panel C), prepared as explained in the text. The human Hsp90α and β isozymes differ only by one amino acid(position 52 in α) in the vicinity of radicicol binding site, i.e., Ser in α and Ala in β. The serine OH group can make three additional hydrogen bonds that cannot be formed in the betaisozyme, where the serine is replaced with alanine. This is essentially the only difference between the two modeled structures, and is therefore likely to be responsible for thedifferences in the binding. Gibbs free energy and enthalpy. Yeast Hsc82 contains Ala in the structurally equivalent position 38 (as in human Hsp90β).

160 A. Zubrienė et al. / Biophysical Chemistry 152 (2010) 153–163

Page 10: Author's personal copy · Author's personal copy Thermodynamics of radicicol binding to human Hsp90 alpha and beta isoforms Asta Zubrien a, Malgorzata Gutkowskab,c, Jurgita Matulien

Author's personal copy

automatically using SCWRL3 [49]. Water molecules that occupiedessentially the same position in the human X-ray structure (3EKO) andboth yeast structures (1BGQ and 2WER) were retained in the models.

Fig. 8B and C compares the hydrogen bond arrangements in thevicinity of radicicol, Asp93 and Ser52 (Hsp90α) or Ala (Hsp90β) thatare most likely to account for the difference in the radicicol bindingenergetics to the alpha and beta isoforms. There is only one aminoacid (Ser52 in Hsp90α or Ala47 in Hsp90β) that differs in the vicinityof the radicicol binding site. There is no direct contact between Ser orAla and radicicol. However, as described previously [16], watermolecules play a crucial role in mediating the radicicol–Hsp90interaction. When radicicol binds to the alpha isozyme, Ser52 makesthree additional contacts that cannot be formedwith the beta isozyme(in which Ala lacks the OH group). These three additional hydrogenbonds are likely to be responsible for ~7 kJ/mol of the enthalpy and~4 kJ/mol of the Gibbs free energy difference between the radicicolbinding to the alpha and beta isozymes. The binding is enthalpicallymore favorable when the additional hydrogen bonds are formed.

However, the structure of radicicol–Hsc82 complex is very similarto Hsp90β, but the energetics of binding is significantly different. Theenthalpy of radicicol binding to Hsc82 is 14.4 kJ/mol less exothermicthan for Hsp90β. This difference could not be assigned to any obviousstructural features of the protein-radicicol complex. Therefore, addi-tional possibilities such as linked folding or conformational changereaction coupled with the binding reaction must be considered as thepossible source of the significant difference in binding enthalpy.

Heat capacity changes were calculated using five published empiricalrelationships between changes in solvent-accessible surface area and ΔCpto compare with experimentally measured values (Table 3). Calculatedheat capacitieswere significantly smaller in their absolute values than theexperimental ΔCp values. This indicates a conformational changeoccurring during radicicol binding to the chaperone similarly to theconformational change occurring upon ATP binding [45].

Radicicol binding to Hsp90 is interesting in many respectsregarding drug design. First, the binding reaction is very tight (i.e., ithas a very favorable Gibbs free energy). Second, the binding reactionhas a very favorable enthalpy of binding. Third, there are few directcontacts between Hsp90 and radicicol that could account for such alarge binding energy. Fourth, water molecules play an essential role inthe recognition and binding.

4. Materials and methods

4.1. Materials

Radicicol was purchased fromA.G. Scientific, Inc., dissolved in DMSOat 50 mM, and stored at −20 °C. The concentration of radicicol wasdetermined spectrophotometrically using an extinction coefficient of14,700 M−1 cm−1 in methanol at 265 nm [50]. Partial degradation ofradicicol in DMSO solution occurred if stored over amonth, especially inthe presence of reductants. It was important to prepare fresh solutionsfor all measurements.

4.2. Hsp90 constructs

Hsp90αΝ — The N-terminal domain of alpha Hsp90. The geneencoding full-length human Hsp90α protein was purchased fromRZPD, Deutsches Ressourcenzentrum für Genomforschung GmbH(Germany). For protein expression, the N-terminal fragment of theHsp90α gene, corresponding to amino acids 3–241, was inserted intothe pET-15b vector (Novagen, Madison, WI, USA) using the XhoI/BamHI restriction sites, thereby fusing a His6 tag to the N-terminus ofthe protein.

Hsp90αF — Full length alpha Hsp90 was cloned and purified aspreviously described [51].

Hsp90βF — Full-length beta Hsp90 was cloned from the gene thatwas amplified from the pBAD24Hsp90beta vector (the kind gift of Dr.Dawid Walerych) using the primers 5′-GGGGGTCTAGACATATGCCT-GAGGAAGTGC-3′ and 5′-GCTAAGTGTCGACCTAATCGACTTCTTCCA-3′.This was then cloned using the NdeI and SalI restriction sites into thepET28b expression vector.

Hsp90βN — The N-terminal domain of beta Hsp90. The sequencecoding for the N-domain (amino acid residues 1–239) of humanHsp90beta was amplified from the same template using the same forwardprimer and the reverse primer 5′-CATAAGCTTTATCTTACTCTTCTTTCT-CACC-3′. This was then cloned using the NdeI and HindIII restrictionsites into the pET28b expression vector.

Yeast Hsc82F — The yeast (S. cerevisiae) Hsc82 gene was amplifiedfrom cDNA (the kind gift of Dr. Pawel Bieganowski) using the primers:5′-CGGCACATATGGCTGGTGAAACTTTTGAAT-3′ and 5′-GCCGGATCCT-TAATCAACTTCTTCCATCTCGG-3′. This was then cloned using the NdeIand BamHI restriction sites into the pET28b expression vector.

Yeast Hsc82N — Sequence coding for the yeast Hsc82 N-terminaldomain (amino acid residues 1–222) was amplified from the Hsc82plasmid (pET28b) using the same forward primer and the reverseprimer 5′-GCCGGATCCTTATTCTGGAATTGGAACTTCCTTTTC-3′. Thiswas then cloned using the NdeI and BamHI restriction sites into thepET28b expression vector.

4.3. Protein expression and purification

His6-tagged Hsp90αN protein was expressed in the Escherichia colistrain BL21 (DE3). The bacterial cultures were grown until an A550 of0.5–0.8 was reached, and the expression was induced by the additionof 1 mM isopropyl-1-thio-β-D-galactopyranoside (final concentra-tion). Protein was purified according to Richter et al. [52]. The cellswere lysed by sonication, and the soluble protein was purified using aNi-IDA affinity column, followed by an anion exchange chromatog-raphy column (Amersham Biosciences). The eluted protein wasdialyzed into a storage buffer containing 20 mM Hepes (pH 7.5) and50 mM NaCl. The purity of the Hsp90αN preparations was analyzedusing SDS-PAGE and determined to be higher than 98%. The proteinconcentrations were determined by UV–VIS spectrophotometry andconfirmed using the standard Bradford method. The protein stocksolution was stored at −70 °C.

Other proteins (except Hsp90αN and Hsp90αF) were expressed intheE. coliBL21RIL strain (Stratagene).All proteins carried anN-terminalHis6 tag and were purified on His-selective HF resin (Sigma) accordingto a standard protocol, before being further purified on a Resource Qcolumn (Biorad) in a 0.05–0.50 M KCl gradient. The proteins weredialyzed against buffer containing 25 mM Hepes, pH 7.5, 150 mM KCland 10% glycerol, and frozen in liquid nitrogen.

4.4. Isothermal titration calorimetry

The protein solutions (2–5 μM) were loaded into the VP-ITCisothermal titration calorimeter (Microcal, Inc.) cell (approximately2 ml, with an active cell volume of 1.4 ml). The titration syringe (250 μl)was filled with 20–50 μM (usually 40) ligand solution. Titrations were

Table 3Experimental and calculated ΔCp values for the binding of radicicol to Hsp90α, Hsp90β,and Hsc82. The ΔCp values were calculated using area coefficients reviewed in [54].

Solvent accessible surface area and heat capacitydifference upon binding

Hsp90α Hsp90β Hsc82

ΔAnon-polar-calculated, Å2 −191 −193 −182,4ΔApolar-calculated, Å2 −58,6 −57,4 −71,9ΔCp-experimental, J×mol−1×K−1 −620 −760 −620ΔCp-calc-Spolar, J×mol−1×K−1 −221 −225 −202ΔCp-calc-Murphy–Freire, J×mol−1×K−1 −295 −300 −265ΔCp-calc-Meyers, J×mol−1×K−1 −201 −204 −186ΔCp-calc-Makhatadze–Privalov, J×mol−1×K−1 −357 −363 −327ΔCp-calc-Robertson–Murphy, J×mol−1×K−1 −157 −157 −158

161A. Zubrienė et al. / Biophysical Chemistry 152 (2010) 153–163

Page 11: Author's personal copy · Author's personal copy Thermodynamics of radicicol binding to human Hsp90 alpha and beta isoforms Asta Zubrien a, Malgorzata Gutkowskab,c, Jurgita Matulien

Author's personal copy

carried out using 25 injections of 10 μl each, injected at 3 to 4 minuteintervals. The stirring speedwas 260 rpm. Titrationswere carried out ata constant temperature in the 13–37 °C temperature range. The ligandsolutionswere prepared in the same buffer as the protein solutions, andcontained the same concentration of DMSO (usually 0.5–-1%). The totalnumber of independent ITC titration experiments was 140. Variation ofthe protein concentration (within 1–10 μM range) or the time intervalsbetween injections (within 3–5 min range) did not affect themeasuredenthalpies to an extent greater than the standard error.

4.5. Protein denaturation experiments

The thermal shift assay was performed using the iCycler iQ RealTime Detection System (Bio-Rad, Hercules, CA), a Corbett Rotor-Gene6000 (QIAGEN Rotor-Gene Q), and the ISS PC1 spectrofluorimeterequipped with a temperature-controlled water bath. The temperaturereproducibility was measured to be ±0.2 °C. The protein concentra-tions were measured spectrophotometrically (ε280 (Hsp90αN)=15,930 M−1 cm−1). The protein concentration was usually 10 μMand the ligand concentrations varied from 0 to 200 μM. Universalbuffer containing 50 mM sodium phosphate, 50 mM sodium acetate,25 mM sodium borate, and 50 mM sodium chloride was usually usedfor the pH-dependence experiments of the protein denaturation.DMSO was added to make up 0.5–1% (v/v) of the solution in eachmeasurement. Unfolding of the protein was monitored by measuringthe fluorescence of the solvatochromic fluorescent dye Dapoxyl™sulfonic acid sodium salt (iCycler) or ANS, both usually at 50 μM. Thetotal volume of the reaction was 10 μl when using the iCycler orCorbett instruments and 3 ml for the ISS PC1 cuvette that was coveredto prevent evaporation. Samples in the iCycler or Corbett instrumentswere overlaid with 2.5 μl of silicone oil DC 200. For this assay, 96-welliCycler iQ PCR plates were used. Samples in the ISS PC1 were excitedwith 380±5 nm UV light, and the ANS fluorescence emission wasregistered at 510±5 nm. The samples were heated at a rate of 1 °C/min. The binding constants obtained by the thermal shift assay havestandard error better than 1.6 fold. The greatest source of error is inthe determination of the enthalpy of unfolding. The Tm is determinedwith ±0.2 °C standard error.

4.6. Structure analysis and calculation of ΔCp

Solvent accessible surface areas were calculated using Naccesscomputer program [53]. The differences in surface areas upon radicicolbinding were estimated by subtracting the areas in the presence andabsence of radicicol in the structures of 1bgq (Hsc82N) and modeledHsp90αNandHsp90βN.Heat capacitieswere calculatedusing empiricalequation relating it to changes in non-polar and polar solvent-accessiblesurface areas (ΔAnon-polar and ΔApolar respectively):

ΔCP = k1ΔAnon�polar + k2ΔApolar

where k1 and k2 are empirical coefficients determined for severalprotein sets reviewed in [54].

Acknowledgments

The project was supported in part by EEA and Norway Grants2004-LT0019-IP-1EEE and the Lithuanian Government.

References

[1] T. Taldone, W. Sun, G. Chiosis, Discovery and development of heat shock protein90 inhibitors, Bioorg. Med. Chem. 17 (2009) 2225–2235.

[2] S.K. Wandinger, K. Richter, J. Buchner, The Hsp90 chaperone machinery, J. Biol.Chem. 283 (2008) 18473–18477.

[3] L. Neckers, M. Mollapour, S. Tsutsumi, The complex dance of the molecularchaperone Hsp90, Trends Biochem. Sci. 34 (2009) 223–226.

[4] L. Neckers, S. Tsutsumi, M. Mollapour, Visualizing the twists and turns of amolecular chaperone, Nat. Struct. Mol. Biol. 16 (2009) 235–236.

[5] M.P. Mayer, C. Prodromou, J. Frydman, The Hsp90 mosaic: a picture emerges, Nat.Struct. Mol. Biol. 16 (2009) 2–6.

[6] D. Walerych, M.B. Olszewski, M. Gutkowska, A. Helwak, M. Zylicz, A. Zylicz, Hsp70molecular chaperones are required to support p53 tumor suppressor activityunder stress conditions, Oncogene 28 (2009) 4284–4294.

[7] V.M. Banz, M. Medova, A. Keogh, C. Furer, Y. Zimmer, D. Candinas, D. Stroka, Hsp90transcriptionally and post-translationally regulates the expression of NDRG1 andmaintains the stability of its modifying kinase GSK3beta, Biochim. Biophys. Acta1793 (2009) 1597–1603.

[8] R.L. van Montfort, P. Workman, Structure-based design of molecular cancertherapeutics, Trends Biotechnol. 27 (2009) 315–328.

[9] S.Y. Sharp, K. Boxall, M. Rowlands, C. Prodromou, S.M. Roe, A. Maloney, M. Powers,P.A. Clarke, G. Box, S. Sanderson, L. Patterson, T.P. Matthews, K.M. Cheung, K. Ball,A. Hayes, F. Raynaud, R. Marais, L. Pearl, S. Eccles, W. Aherne, E. McDonald, P.Workman, In vitro biological characterization of a novel, synthetic diaryl pyrazoleresorcinol class of heat shock protein 90 inhibitors, Cancer Res. 67 (2007)2206–2216.

[10] M. Sgobba, G. Rastelli, Structure-based and in silico design of Hsp90 inhibitors,ChemMedChem 4 (2009) 1399–1409.

[11] A. Zuehlke, J.L. Johnson, Hsp90 and co-chaperones twist the functions of diverseclient proteins, Biopolymers 93 (2010) 211–217.

[12] Y. Fukuyo, C.R. Hunt, N. Horikoshi, Geldanamycin and its anti–cancer activities,Cancer Lett. (2009) 24–35.

[13] P. Delmotte, J. Delmotte-Plaque, A new antifungal substance of fungal origin,Nature 171 (1953) 344.

[14] T.W. Schulte, S. Akinaga, S. Soga, W. Sullivan, B. Stensgard, D. Toft, L.M. Neckers,Antibiotic radicicol binds to the N-terminal domain of Hsp90 and sharesimportant biologic activities with geldanamycin, Cell Stress Chaperones 3(1998) 100–108.

[15] S.M. Roe, C. Prodromou, R. O'Brien, J.E. Ladbury, P.W. Piper, L.H. Pearl, Structuralbasis for inhibition of the Hsp90molecular chaperone by the antitumor antibioticsradicicol and geldanamycin, J. Med. Chem. 42 (1999) 260–266.

[16] C. Prodromou, J.M. Nuttall, S.H. Millson, S.M. Roe, T.S. Sim, D. Tan, P. Workman, L.H.Pearl, P.W. Piper, Structural basis of the radicicol resistance displayed by a fungalHsp90, ACS Chem. Biol. 4 (2009) 289–297.

[17] S. Nilapwar, E. Williams, C. Fu, C. Prodromou, L.H. Pearl, M.A. Williams, J.E.Ladbury, Structural-thermodynamic relationships of interactions in the N-terminal ATP-binding domain of Hsp90, J. Mol. Biol. 392 (2009) 923–936.

[18] A. Velazquez Campoy, E. Freire, ITC in the post-genomic era...? Priceless, Biophys.Chem. 115 (2005) 115–124.

[19] A. Velazquez-Campoy, H. Ohtaka, A. Nezami, S. Muzammil, E. Freire, Isothermaltitration calorimetry, Curr. Protoc. Cell Biol. (2004) Chapter 17, Unit 17 8.

[20] E. Freire, Isothermal titration calorimetry: controlling binding forces in leadoptimization, Drug Discov. Today: Technologies 1 (2004) 295–299.

[21] M.C. Lo, A. Aulabaugh, G. Jin, R. Cowling, J. Bard, M. Malamas, G. Ellestad,Evaluation of fluorescence-based thermal shift assays for hit identification in drugdiscovery, Anal. Biochem. 332 (2004) 153–159.

[22] F.H. Niesen, H. Berglund, M. Vedadi, The use of differential scanning fluorimetry todetect ligand interactions that promote protein stability, Nat. Protoc. 2 (2007)2212–2221.

[23] M.W. Pantoliano, E.C. Petrella, J.D. Kwasnoski, V.S. Lobanov, J. Myslik, E. Graf, T.Carver, E. Asel, B.A. Springer, P. Lane, F.R. Salemme, High-density miniaturizedthermal shift assays as a general strategy for drug discovery, J. Biomol. Screen. 6(2001) 429–440.

[24] M.J. Todd, F.R. Salemme, Direct binding assays for pharma screening, Genetic Eng.News 23 (2003) 28–29.

[25] P. Cimmperman, L. Baranauskiene, S. Jachimoviciute, J. Jachno, J. Torresan, V.Michailoviene, J. Matuliene, J. Sereikaite, V. Bumelis, D. Matulis, A quantitativemodel of thermal stabilization and destabilization of proteins by ligands, Biophys.J. 95 (2008) 3222–3231.

[26] D. Matulis, J.K. Kranz, F.R. Salemme, M.J. Todd, Thermodynamic stability ofcarbonic anhydrase: measurements of binding affinity and stoichiometry usingThermoFluor, Biochemistry 44 (2005) 5258–5266.

[27] T.M. Mezzasalma, J.K. Kranz, W. Chan, G.T. Struble, C. Schalk-Hihi, I.C. Deckman,B.A. Springer, M.J. Todd, Enhancing recombinant protein quality and yield byprotein stability profiling, J. Biomol. Screen. 12 (2007) 418–428.

[28] U.B. Ericsson, B.M. Hallberg, G.T. Detitta, N. Dekker, P. Nordlund, Thermofluor-based high-throughput stability optimization of proteins for structural studies,Anal. Biochem. 357 (2006) 289–298.

[29] K.A. Krukenberg, U.M. Bottcher, D.R. Southworth, D.A. Agard, Grp94, theendoplasmic reticulum Hsp90, has a similar solution conformation to cytosolicHsp90 in the absence of nucleotide, Protein Sci. 18 (2009) 1815–1827.

[30] C.K. Vaughan, P.W. Piper, L.H. Pearl, C. Prodromou, A common conformationallycoupled ATPase mechanism for yeast and human cytoplasmic HSP90s, FEBS J. 276(2009) 199–209.

[31] L.H. Pearl, C. Prodromou, Structure and mechanism of the hsp90 molecularchaperone machinery, Annu. Rev. Biochem. 75 (2006) 271–294.

[32] M. Mickler, M. Hessling, C. Ratzke, J. Buchner, T. Hugel, The large conformationalchanges of Hsp90 are only weakly coupled to ATP hydrolysis, Nat. Struct. Mol. Biol.16 (2009) 281–286.

[33] M. Hessling, K. Richter, J. Buchner, Dissection of the ATP-induced conformationalcycle of themolecular chaperone Hsp90, Nat. Struct. Mol. Biol. 16 (2009) 287–293.

[34] E. Freire, Do enthalpy and entropy distinguish first in class from best in class?Drug Discov. Today 13 (2008) 869–874.

162 A. Zubrienė et al. / Biophysical Chemistry 152 (2010) 153–163

Page 12: Author's personal copy · Author's personal copy Thermodynamics of radicicol binding to human Hsp90 alpha and beta isoforms Asta Zubrien a, Malgorzata Gutkowskab,c, Jurgita Matulien

Author's personal copy

[35] M.L. Doyle, G. Louie, P.R. Dal Monte, T.D. Sokoloski, Tight binding affinitiesdetermined from thermodynamic linkage to protons by titration calorimetry,Methods Enzymol. 259 (1995) 183–194.

[36] J.J. Christensen, L.D. Hansen, R.M. Izatt, Handbook of Proton Ionizations Heats,Wiley-Interscience, 1976.

[37] B.M. Baker, K.P. Murphy, Evaluation of linked protonation effects in proteinbinding reactions using isothermal titration calorimetry, Biophys. J. 71 (4) (1996)2049–2055.

[38] B.M. Baker, K.P. Murphy, Dissecting the energetics of a protein–proteininteraction: the binding of ovomucoid third domain to elastase, J. Mol. Biol. 268(2) (1997) 557–569.

[39] J.M. Bradshaw, G. Waksman, Calorimetric investigation of proton linkage bymonitoring both the enthalpy and association constant of binding: application tothe interaction of the Src SH2 domain with a high-affinity tyrosyl phosphopep-tide, Biochemistry 37 (1998) 15400–15407.

[40] A. Zubriene, J. Matuliene, L. Baranauskiene, J. Jachno, J. Torresan, V. Michailoviene,P. Cimmperman, D. Matulis, Measurement of nanomolar dissociation constants bytitration calorimetry and thermal shift assay — radicicol binding to Hsp90 andethoxzolamide binding to CAII, Int. J. Mol. Sci. 10 (2009) 2662–2680.

[41] S.R. Anderson, G. Weber, Fluorescence polarization of the complexes of 1-anilino-8-naphthalenesulfonate with bovine serum albumin. Evidence for preferentialorientation of the ligand, Biochemistry 8 (1969) 371–377.

[42] D. Matulis, C.G. Baumann, V.A. Bloomfield, R.E. Lovrien, 1-anilino-8-naphthalenesulfonate as a protein conformational tightening agent, Biopolymers 49 (1999)451–458.

[43] D. Matulis, R. Lovrien, 1-Anilino-8-naphthalene sulfonate anion-protein bindingdepends primarily on ion pair formation, Biophys. J. 74 (1998) 422–429.

[44] A. Zurawska, J. Urbanski, J. Matuliene, J. Baraniak, M.P. Klejman, S. Filipek, D.Matulis,P. Bieganowski, Mutations that increase both Hsp90 ATPase activity in vitro andHsp90 drug resistance in vivo, Biochim. Biophys. Acta 1803 (2010) 575–583.

[45] A. Dehner, J. Furrer, K. Richter, I. Schuster, J. Buchner, H. Kessler, NMR chemicalshift perturbation study of the N-terminal domain of Hsp90 upon binding of ADP,AMP-PNP, geldanamycin, and radicicol, Chembiochem 4 (2003) 870–877.

[46] L.T. Gooljarsingh, C. Fernandes, K. Yan, H. Zhang, M. Grooms, K. Johanson, R.H.Sinnamon, R.B. Kirkpatrick, J. Kerrigan, T. Lewis, M. Arnone, A.J. King, Z. Lai, R.A.Copeland, P.J. Tummino, A biochemical rationale for the anticancer effects ofHsp90 inhibitors: slow, tight binding inhibition by geldanamycin and itsanalogues, Proc. Natl Acad. Sci. USA 103 (2006) 7625–7630.

[47] S. Nakamura, S. Kidokoro, Isothermal acid-titration calorimetry for evaluating thepH dependence of protein stability, Biophys. Chem. 109 (2004) 229–249.

[48] L. Baranauskiene, J. Matuliene, D. Matulis, Determination of the thermodynamicsof carbonic anhydrase acid-unfolding by titration calorimetry, J. Biochem.Biophys. Methods 70 (2008) 1043–1047.

[49] A.A. Canutescu, A.A. Shelenkov, R.L. Dunbrack Jr., A graph-theory algorithm forrapid protein side-chain prediction, Protein Sci. 12 (2003) 2001–2014.

[50] H.P. Sigg, W. Loeffler, Production of radicicol, USA, 1969.[51] I. Cikotiene, E. Kazlauskas, J. Matuliene, V. Michailoviene, J. Torresan, J. Jachno, D.

Matulis, 5-Aryl-4-(5-substituted-2, 4-dihydroxyphenyl)-1, 2, 3-thiadiazoles asinhibitors of Hsp90 chaperone, Bioorg. Med. Chem. Lett. 19 (2009) 1089–1092.

[52] K. Richter, P. Muschler, O. Hainzl, J. Buchner, Coordinated ATP hydrolysis by theHsp90 dimer, J. Biol. Chem. 276 (2001) 33689–33696.

[53] S.J. Hubbard, J.M. Thornton, ‘NACCESS’ Computer Program, 1993.[54] N.V. Prabhu, K.A. Sharp, Heat capacity in proteins, Annu. Rev. Phys. Chem. 56 (2005)

521–548.

163A. Zubrienė et al. / Biophysical Chemistry 152 (2010) 153–163