15
Comparative Analysis of the Effects of a-Crystallin and GroEL on the Kinetics of Thermal Aggregation of Rabbit Muscle Glyceraldehyde-3-Phosphate Dehydrogenase Kira A. Markossian Nikolay V. Golub Natalia A. Chebotareva Regina A. Asryants Irina N. Naletova Vladimir I. Muronetz Konstantin O. Muranov Boris I. Kurganov Published online: 20 November 2009 Ó Springer Science+Business Media, LLC 2009 Abstract Effects of a-crystallin and GroEL on the kinetics of thermal aggregation of rabbit muscle glycer- aldehyde-3-phosphate dehydrogenase (GAPDH) have been studied using dynamic light scattering and analytical ultracentrifugation. The analysis of the initial parts of the dependences of the hydrodynamic radius of protein aggregates on time shows that in the presence of a-crys- tallin or GroEL the kinetic regime of GAPDH aggregation is changed from the regime of diffusion-limited cluster– cluster aggregation to the regime of reaction-limited clus- ter–cluster aggregation, wherein the sticking probability for the colliding particles becomes lower the unity. In contrast to a-crystallin, GroEL does not interfere with formation of the start aggregates which include denatured GAPDH molecules. On the basis of the analytical ultracentrifuga- tion data the conclusion has been made that the products of dissociation of GAPDH and a-crystallin or GroEL play an important role in the interactions of GAPDH and chaper- ones at elevated temperatures. Keywords Glyceraldehyde-3-phosphate dehydrogenase a-Crystallin GroEL Aggregation Dynamic light scattering Sedimentation velocity Abbreviations DLCA Diffusion-limited cluster–cluster aggregation DLS Dynamic light scattering DSC Differential scanning calorimetry GAPDH Glyceraldehyde-3-phosphate dehydrogenase RLCA Reaction-limited cluster–cluster aggregation 1 Introduction The mechanism of thermal denaturation of proteins involves partial unfolding and conformational modification of the protein molecule. The interaction of exposed hydrophobic surfaces during protein unfolding is accom- panied by formation of aggregates. The molecular chap- erones, which belong to the family of heat-shock proteins, can interact with non-native or partially folded polypeptide chains and prevent irreversible protein aggregation during heat stress [1, 2]. Chaperonin GroEL from Escherichia coli [3, 4] and small heat-shock proteins (sHsps) [5, 6] are among the well-studied chaperones. GroEL, a 14 subunit, double-ring oligomer of *800 kDa [3, 4], has some hydrophobic sites exposed to the solvent [7] and interacts transiently with intermediates of protein fold- ing, prevents their hydrophobic surfaces from ‘‘incorrect’’ intermolecular and intramolecular interactions, and favors correct folding of polypeptide chains [2, 811]. A range of in vitro studies directly demonstrated that GroEL can rapidly and efficiently bind to non-native K. A. Markossian (&) N. V. Golub N. A. Chebotareva B. I. Kurganov Bach Institute of Biochemistry, Russian Academy of Sciences, 119071 Moscow, Russia e-mail: [email protected] R. A. Asryants I. N. Naletova V. I. Muronetz Belozersky Institute of Physico-Chemical Biology, Moscow State University, 119992 Moscow, Russia V. I. Muronetz Faculty of Bioengineering and Bioinformatics, Moscow State University, 119992 Moscow, Russia K. O. Muranov Emanuel Institute of Biochemical Physics, Russian Academy of Sciences, 119991 Moscow, Russia 123 Protein J (2010) 29:11–25 DOI 10.1007/s10930-009-9217-9

Comparative Analysis of the Effects of α-Crystallin and GroEL on the Kinetics of Thermal Aggregation of Rabbit Muscle Glyceraldehyde-3-Phosphate Dehydrogenase

  • Upload
    ras

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Comparative Analysis of the Effects of a-Crystallin and GroELon the Kinetics of Thermal Aggregation of Rabbit MuscleGlyceraldehyde-3-Phosphate Dehydrogenase

Kira A. Markossian • Nikolay V. Golub • Natalia A. Chebotareva •

Regina A. Asryants • Irina N. Naletova • Vladimir I. Muronetz •

Konstantin O. Muranov • Boris I. Kurganov

Published online: 20 November 2009

� Springer Science+Business Media, LLC 2009

Abstract Effects of a-crystallin and GroEL on the

kinetics of thermal aggregation of rabbit muscle glycer-

aldehyde-3-phosphate dehydrogenase (GAPDH) have been

studied using dynamic light scattering and analytical

ultracentrifugation. The analysis of the initial parts of the

dependences of the hydrodynamic radius of protein

aggregates on time shows that in the presence of a-crys-

tallin or GroEL the kinetic regime of GAPDH aggregation

is changed from the regime of diffusion-limited cluster–

cluster aggregation to the regime of reaction-limited clus-

ter–cluster aggregation, wherein the sticking probability for

the colliding particles becomes lower the unity. In contrast

to a-crystallin, GroEL does not interfere with formation of

the start aggregates which include denatured GAPDH

molecules. On the basis of the analytical ultracentrifuga-

tion data the conclusion has been made that the products of

dissociation of GAPDH and a-crystallin or GroEL play an

important role in the interactions of GAPDH and chaper-

ones at elevated temperatures.

Keywords Glyceraldehyde-3-phosphate dehydrogenase �a-Crystallin � GroEL � Aggregation � Dynamic light

scattering � Sedimentation velocity

Abbreviations

DLCA Diffusion-limited cluster–cluster aggregation

DLS Dynamic light scattering

DSC Differential scanning calorimetry

GAPDH Glyceraldehyde-3-phosphate dehydrogenase

RLCA Reaction-limited cluster–cluster aggregation

1 Introduction

The mechanism of thermal denaturation of proteins

involves partial unfolding and conformational modification

of the protein molecule. The interaction of exposed

hydrophobic surfaces during protein unfolding is accom-

panied by formation of aggregates. The molecular chap-

erones, which belong to the family of heat-shock proteins,

can interact with non-native or partially folded polypeptide

chains and prevent irreversible protein aggregation during

heat stress [1, 2]. Chaperonin GroEL from Escherichia coli

[3, 4] and small heat-shock proteins (sHsps) [5, 6] are

among the well-studied chaperones.

GroEL, a 14 subunit, double-ring oligomer of *800 kDa

[3, 4], has some hydrophobic sites exposed to the solvent [7]

and interacts transiently with intermediates of protein fold-

ing, prevents their hydrophobic surfaces from ‘‘incorrect’’

intermolecular and intramolecular interactions, and favors

correct folding of polypeptide chains [2, 8–11].

A range of in vitro studies directly demonstrated that

GroEL can rapidly and efficiently bind to non-native

K. A. Markossian (&) � N. V. Golub � N. A. Chebotareva �B. I. Kurganov

Bach Institute of Biochemistry, Russian Academy of Sciences,

119071 Moscow, Russia

e-mail: [email protected]

R. A. Asryants � I. N. Naletova � V. I. Muronetz

Belozersky Institute of Physico-Chemical Biology,

Moscow State University, 119992 Moscow, Russia

V. I. Muronetz

Faculty of Bioengineering and Bioinformatics,

Moscow State University, 119992 Moscow, Russia

K. O. Muranov

Emanuel Institute of Biochemical Physics,

Russian Academy of Sciences, 119991 Moscow, Russia

123

Protein J (2010) 29:11–25

DOI 10.1007/s10930-009-9217-9

proteins and suppress their aggregation, as has been shown

for ribulose-1,5-bisphosphate oxygenase-carboxylase [12],

a-glucosidase [13], glutamine synthase [14], dihydrofolate

reductase [15], rhodanese, citrate synthase [16] and malate

dehydrogenase [17].

It has been accepted that GroEL prevents protein

aggregation by encapsulating individual chains within the

so-called ‘Anfinsen cage’ provided by the GroEL–GroES

complex [2, 18].

However, GroEL is able to prevent protein aggregation

not only by encapsulating individual protein chains inside

its oligomeric structure [2, 19, 20], but also outside the

cage [21, 22]. In some cases the presence of GroEL either

alone or additionally to ADP or ATP is enough for effec-

tive GroEL-assisted protein folding [23, 24].

It has been suggested that hydrophobic interactions

represent the main factor stabilizing the GroEL complex

with nonnative proteins [25, 26]. However, electrostatic

interactions also play an important role in the complexing

of GroEL with non-native proteins [27, 28].

Recently Semisotnov et al. [29–32] proposed the model

of GroEL functioning as a molecular chaperone. According

to this model, GroEL binds strongly ‘‘hydrophobic’’ states

of proteins or polypeptides decreasing their concentrations

in solution thereby decreasing their non-specific interac-

tions. Bound polypeptides can sometimes dissociate from

the GroEL surface and adopt native conformation in

solution and lose possibility to interact with GroEL.

Chaperonin function consists in ‘‘keeping’’ of hydrophobic

protein intermediates and thereby in decreasing probability

of their non-specific association [31, 32]. According to the

model of the GroEL action proposed by these authors

GroEL-assisted protein folding proceeds out of the chap-

eronin cavity.

Ybarra and Horowitz [33] were the first to demonstrate

the formation of GroEL monomers under conditions

commonly used for preparation of chaperonin. The essen-

tial requirements are the simultaneous presence of nucle-

otides such as Mg-ATP or Mg-ADP and a solid-phase

anion-exchange medium. Surin et al. [34] used the same

procedure to obtain the GroEL monomer. According to the

results obtained by these authors, the change in pH of the

buffer solution from 7.5 to 9.0 resulted in the increase in

the yield of the monomeric form. The monomeric GroEL is

a globular protein with a pronounced secondary and a rigid

tertiary structure. The affinity of the monomeric form to the

hydrophobic probe, 8-anilino-1-naphthalenesulfonic acid,

is an order of magnitude higher than that for the subunit in

the oligomeric particles, suggesting that the monomeric

form is characterized by the high degree of exposure of the

hydrophobic clusters of the protein molecule [34].

Belonging to the family of small heat shock proteins,

a-crystallin is a large oligomeric complex ranging in size

from 200 to 800 kDa and possessing chaperone-like

activity. a-Crystallin binds unfolded proteins and forms

stable, soluble complexes with denatured proteins pre-

venting their aggregation and precipitation [35, 36]. The

ability of a-crystallin to suppress heat-induced aggrega-

tion of proteins is a result of hydrophobic interactions

with these denatured proteins, and this ability increases

when a-crystallin is heated [37–40]. The mechanism of

suppression of thermal aggregation of a number of pro-

teins by a-crystallin has also been studied by Kurganov

et al. [41–47]. It has been shown that suppression of

aggregation is due to the diminution of the size of start

aggregates in the presence of a-crystallin, increase in the

duration of the latent period over which the start aggre-

gates are formed and transition of the aggregation process

from the kinetic regime of ‘‘diffusion-limited cluster–

cluster aggregation’’ (DLCA) into the regime of ‘‘reac-

tion-limited cluster–cluster aggregation’’ (RLCA),

wherein the sticking probability for the colliding particles

becomes less than unity.

Process of thermal denaturation and aggregation of

glyceraldehyde-3-phosphate dehydrogenase (GAPDH; EC

1.2.1.12), a tetrameric enzyme of 144 kDa, is the subject

of a number of studies [48–50]. The secondary structure

of the enzyme is relatively heat stable, showing change

only at pH 8.85. Under these conditions, the enzyme first

dissociates within several minutes, probably into dimers,

and with prolonged heating it becomes irreversibly

aggregated [48]. The dissociative mechanism of thermal

denaturation of GAPDH was supported using differential

scanning calorimetry and analytical ultracentrifugation

[43, 51].

Glyceraldehyde-3-phosphate dehydrogenase has been

used as a target protein to examine the effects of GroEL

[15, 52–54]. Naletova et al. [54] have demonstrated that

GroEL does not prevent thermal denaturation of GAPDH,

but effectively binds the thermally denatured enzyme, thus

preventing the aggregation of the denatured molecules. The

complex GroEL–GAPDH is characterized by increased

thermal stability compared to GroEL.

Earlier we studied the effect of a-crystallin on thermal

denaturation of GAPDH and aggregation at the regime of

temperature elevation with a constant rate [43]. When

aggregation of GAPDH was studied in the presence of

a-crystallin, the start aggregates of lesser size were

observed. In the present work we have compared the

effects of GroEL and a-crystallin on the aggregation pro-

cess of GAPDH monitored at a fixed temperature of 45�C.

We have shown using dynamic light scattering (DLS) that

in the presence of each chaperone GAPDH aggregation

proceeds in the RLCA regime. However, these chaperones

have different effect on the formation of start aggregates

involving unfolded GAPDH molecules.

12 K. A. Markossian1 et al.

123

2 Materials and Methods

2.1 Isolation of GAPDH

Glyceraldehyde-3-phosphate dehydrogenase was isolated

from rabbit skeletal muscles as described by Scopes and

Stoter [55] with an additional purification step using gel-

filtration on Sephadex G-100. GAPDH concentration was

determined spectrophotometrically at 280 nm, using the

absorption coefficient A1%1cm of 10.6 [56].

2.2 Isolation of a-Crystallin

Freshly excised lenses from 2-year-old steers were

obtained from a local slaughterhouse and stored frozen at

-20 �C. Purification of a-crystallin was performed mainly

according to the procedure described earlier [57]. The de-

capsulated lens cortex was homogenized at 0 �C in 40 mM

sodium phosphate, pH 6.8, containing 100 mM NaCl,

1 mM EDTA, and 3 mM NaN3. The homogenate was

centrifuged at 27,000g for 1 h at 4 �C, and the supernatant

containing the soluble crystallin was fractionated by gel

filtration using TSK-gel HW-55 (Sigma) column

(2.5 9 90 cm). The top of the fraction containing a-crys-

tallin was collected by measuring absorption at 280 nm.

a-Crystallin fraction was rechromatographied using TSK-

gel HW-55 columns. a-Crystallin peak corresponding to

molecular mass of 750 kDa was symmetrical. The fraction

containing HMW a-crystallin was eluted in the void vol-

ume. TSK-gel HW-55 column was calibrated with high

molecular mass kit: thyroglobulin (660 kDa), ferritin

(440 kDa), catalase (220 kDa), aldolase (158 kDa) and

BSA (68 kDa) (Sigma). Finally, top of peak corresponding

to a-crystallin was collected and concentrated by ultrafil-

tration (Millipore PTTK disk membrane (Sigma), NMWL

30000). Thus, the procedure used for isolation and purifi-

cation of a-crystallin provided complete separation of

HMW fraction of a-crystallin. Concentration of a-crystallin

was determined spectrophotometrically at 280 nm, using

the absorption coefficient A1%1cm of 8.5 [40].

2.3 Isolation of GroEL

GroEL was expressed in Esherichia coli strain W3110

transfected with the plasmid pOF39. The expression

product was purified as described by Corrales and Fersht

[26] with some modifications [54]. The resulting prepara-

tion was concentrated to 4 mg/ml using YM-10 Centriprep

centrifugal filter units and dialyzed against 10 mM

K-phosphate buffer, pH 7.5. The solution of GroEL was

stored at –70 �C. GroEL concentration was determined

spectrophotometrically at 280 nm, using the absorption

coefficient A1%1cm of 1.8 [58].

2.4 GAPDH Assay

Glyceraldehyde-3-phosphate dehydrogenase activity was

measured by monitoring the increase in the absorbance at

340 nm, caused by the formation of NADH, using a UV

1601 Schimadzu spectrophotometer (Japan). The reaction

was carried out at 20 �C and initiated by the addition of

2–6 lg of the enzyme to the reaction mixture (1 ml) con-

taining 100 mM glycine, 100 mM KH2PO4, pH 8.9, 5 mM

EDTA, 1 mM NAD and 1 mM glyceraldehyde-3-phos-

phate. The dehydrogenase activity constituted 70 U/mg of

the protein.

2.5 Thermal Inactivation of GAPDH

Thermal inactivation of GAPDH was studied at 45 �C in

10 mM Na-phosphate buffer, pH 7.5, containing 1 mM

EDTA and 1 mM dithiothreitol. The inactivation process

was initiated by the addition of a portion of the GAPDH

solution to the buffer preheated for 10 min. At fixed time

intervals, aliquots were withdrawn to determine the enzy-

matic activity in the standard mixture.

2.6 Dynamic Light Scattering Studies

For light scattering measurements a commercial instrument

Photocor Complex (Photocor Instruments Inc., USA

www.photocor.com) was used with a He–Ne laser

(Coherent, USA, Model 31-2082, 632.8 nm, 10 mW) as a

light source. The temperature of sample cell was controlled

by the proportional integral derivative (PID) temperature

controller to within ±0.1 �C. A quasi-cross correlation

photon counting system with two photomultiplier tubes

(PMT) was used to increase the accuracy of particle sizing

in the range from 1.0 nm to 5.0 lm. DLS data have been

accumulated and analyzed with multifunctional real-time

correlator Photocor-FC. DynaLS software (Alango, Israel)

was used for polydisperse analysis of DLS data.

The diffusion coefficient D of the particles is directly

related to the decay rate sc of the time-dependent correla-

tion function for the light-scattering intensity fluctuations:

D = 1/2sck2. In this equation k is the wave number of the

scattered light, k = (4pn/k)sin(h/2), where n is the refrac-

tive index of the solvent, k is the wavelength of the incident

light in vacuum and h is the scattering angle. The mean

hydrodynamic radius of the particles, Rh, can then be

calculated according to the Stokes–Einstein equation:

D = kBT/6pgRh, where kB is Boltzmann’s constant, T is the

absolute temperature and g is the shear viscosity of the

solvent.

The kinetics of thermal aggregation of GAPDH was

studied by DLS in 10 mM Na-phosphate buffer, pH 7.5.

All solutions for the experiments were prepared using

Comparative Analysis of the Effects of a-Crystallin 13

123

deionized water obtained with Easy-Pure II RF system

(Barnstead, USA). The buffer was placed in a cylindrical

cell with a diameter of 6.3 mm and pre-incubated for

10 min at appropriate temperature. Cells with stoppers

were used for incubation at high temperature to avoid

evaporation. The aggregation process was initiated by the

addition of an aliquot of GAPDH. To study the effect of a-

crystallin or GroEL on aggregation of GAPDH, aliquots of

both proteins (protein substrate and chaperone) were

simultaneously added into the cell up to the final volume of

0.5 mL. The scattering light was collected at 90� scattering

angle and the accumulation time of the autocorrelation

function was 30 s.

Analysis of the dependence of the hydrodynamic radius

of protein aggregates on time allows discriminating

between the DLCA and RLCA regimes of aggregation. In

our previous works [41, 42, 51, 59] the conclusion about

the validity of the DLCA regime for thermal aggregation of

proteins was made on the fact that from a certain value of

time (t [ t*) the dependence of the hydrodynamic radius

(Rh) of protein aggregates on time follows the power law:

Rh ¼ R�h 1þ K1 t � t�ð Þ½ �1=df; ð1Þ

where R�h is the Rh value at t = t*, K1 is a constant and df is

the fractal dimension of the aggregates formed as a result

of random aggregation. For DLCA regime, the parameter

df has a universal value: df = 1.8 [60, 61].

The following equation may be used for analysis of the

initial parts of the dependence of Rh on time (see [51]):

Rh ¼ Rh;0 1þ 1

t2Rt � t3ð Þ

� �; ð2Þ

where Rh,0 is the hydrodynamic radius of start aggregates,

parameter t3 is the moment of time at which the start

aggregates appear (we have conserved the notation t3which was used in our previous work [62]) and t2R is the

time interval over which the Rh value increases from Rh,0 to

2Rh,0.

Previously we have shown that when protein aggrega-

tion proceeds in the presence of a-crystallin, the depen-

dence of the hydrodynamic radius on time follows the

exponential law [41–44]:

Rh ¼ Rh;0 expln 2

t2Rt � t3ð Þ

� �� �; ð3Þ

where Rh,0 is the hydrodynamic radius of the start aggre-

gates, t3 is the moment of time at which the start aggregates

appear and t2R is the time interval over which the hydro-

dynamic radius of aggregates is doubled. The reciprocal

value of parameter t2R characterizes the rate of aggregation.

The higher 1/t2R, the higher is the aggregation rate. The

exponential dependence of Rh on time is typical of the

RLCA regime [63].

2.7 Analytical Ultracentrifugation

Sedimentation velocity studies of GAPDH, a-crystallin,

GroEL and the mixture of GAPDH with a-crystallin or

GroEL were carried out at 21 and 45–46 �C in a Model E

analytical ultracentrifuge (Beckman), equipped with

absorbance optics, a photoelectric scanner, a monochro-

mator, a computer on line and special facilities (additional

chamber with heater) for high temperatures. A four-hole

rotor An-F Ti and 12 mm double sector cells were used.

The rotor was preheated at elevated temperature during a

night in the thermostat before the run. Sedimentation pro-

files of GAPDH and a-crystallin were recorded by mea-

suring absorbance at 280 nm. All cells were scanned

simultaneously. The time interval between scans was

3 min. The sedimentation coefficients were estimated from

the differential sedimentation coefficient distribution [c(s)

vs. s] or [c(s, f/f0) vs. s] which were analyzed using

SEDFIT program [64, 65]. The sedimentation coefficient

distribution c(s) of Lamm equation solutions is based on

the approximation of a single, weight-average frictional

coefficient of all particles, determined from the experi-

mental data, which scales the diffusion coefficient to the

sedimentation coefficient consistent with the traditional

s*M2/3 power law. It provides a high hydrodynamic res-

olution, where diffusion broadening of the sedimentation

boundaries is deconvoluted from the sedimentation coef-

ficient distribution. A generalization of c(s) to a two-

dimensional distribution of sedimentation coefficient and

frictional ratio, c(s, f/f0), which is representative of a more

general set of size-and-shape distributions, including mass-

Stokes radius distributions, c(M, RS), and sedimentation

coefficient-molar mass distributions c(s, M) has recently

been proposed [64]. Even though the additional dimension

describing f/f0 (or M, respectively) does not offer high

resolution, the hydrodynamic resolution in sedimentation

coefficient, c(s, *) is usually close to the c(s) resolution

[64]. The c(s, *) distribution can be considered as a more

general version of c(s) distribution, which does not make

any assumptions on the diffusional properties of macro-

molecular ensemble, yet still deconvolutes diffusional

broadening on the basis of the experimentally measured

sedimentation boundary shapes [65].

The c(s) analysis was performed with regularization at a

confidence level of 0.68 and a floating frictional ratio. The

sedimentation coefficients were corrected to the standard

conditions (a solvent with the density and viscosity of

water at 20 �C).

2.8 Calculations

Origin 7.0 software (OriginLab Corporation, USA) was

used for the calculations.

14 K. A. Markossian1 et al.

123

3 Results

3.1 Kinetics of Thermal Aggregation of GAPDH

Effects of a-crystallin and GroEL on thermal aggregation of

GAPDH were studied at 45 �C. First of all, we analyzed the

kinetics of aggregation of GAPDH (0.4 mg/mL, 10 mM Na-

phosphate, pH 7.5) at this temperature using DLS. Figure 1a

and b shows the dependences of the light scattering intensity

(I) and hydrodynamic radius (Rh) of the protein aggregates

on time, respectively. To analyze the dependence of Rh on

time, we constructed the light scattering intensity versus

hydrodynamic radius plot (Fig. 1c). The length cut off on the

abscissa axis by the initial linear dependence of I on Rh

corresponds to the hydrodynamic radius of the protein

aggregates (Rh,0) which are formed in the system at the

moment when the light scattering intensity begins to

increase. We have called such primary clusters the start

aggregates [41]. The Rh,0 value characterizing the size of the

start aggregates is equal to 65.5(±1.3) nm. With a knowledge

of the Rh,0 value we can analyze the initial part of the

dependence of Rh on time using Eq. (2). Parameter t3 (the

moment of time at which the start aggregates appear) was

found to be 4.5(± 0.2) min (Table 1). The reciprocal value

of parameter t2R (t2R is the time interval over which the Rh

value increases from Rh,0 to 2Rh,0) is also given in Table 1.

At rather high values of time (t [ t* = 13 min) the

dependence of Rh on time follows the power law (1) with

df = 1.86(± 0.14). The obtained value of df means that

GAPDH aggregation proceeds in the DLCA regime [63].

This result is consistent with the data we obtained previ-

ously [51].

3.2 Effect of a-Crystallin on Thermal Inactivation

of GAPDH

Since the process of GAPDH thermal aggregation involves

the initial stage of unfolding of the protein molecule, it was

of interest to study the effect of a-crystallin on stability of

GAPDH. Figure 2 shows the kinetics of thermal inactiva-

tion of GAPDH (0.4 mg/mL) at 45 �C. As can be seen

from this Figure, a-crystallin accelerates thermal inactiva-

tion of GAPDH. The accelerating effect of a-crystallin

increases with increasing a-crystallin concentration and

reaches maximum at a-crystallin concentration of 0.2 mg/

mL. At higher concentrations of a-crystallin the shape of

the dependences of the relative enzymatic activity on time

remained practically unchanged. The destabilizing effect of

a-crystallin is consistent with the data obtained by differ-

ential scanning calorimetry (DSC). In our previous work

[43] we showed that the position of maximum on the DSC

profile was shifted towards the lower temperatures in the

presence of a-crystallin.

3.3 Effect of a-Crystallin on Kinetics of Thermal

Aggregation of GAPDH

Figure 3 demonstrates the effect of a-crystallin on the

kinetics of GAPDH aggregation at 45 �C. Addition of

a-crystallin reduced the increment in light scattering

intensity and prevented precipitation of GAPDH.

Fig. 1 Kinetics of thermal aggregation of GAPDH (0.4 mg/mL) at

45 �C (10 mM Na-phosphate buffer, pH 7.5). The dependences of the

light scattering intensity and hydrodynamic radius of the protein

aggregates (Rh) on time (panels A and B, respectively). The solidcurve in panel B was calculated from Eq. (2). Inset in panel B shows

the initial part of the dependence of Rh on time. The horizontal dottedline corresponds to the Rh,0 value (Rh,0 = 65 nm). c The light

scattering intensity versus the hydrodynamic radius plot

Comparative Analysis of the Effects of a-Crystallin 15

123

Calculations of the hydrodynamic radius show that

aggregates of lesser size are formed in the presence of

a-crystallin (Fig. 4a–d). To carry out the quantitative

analysis of the dependences of Rh on time obtained in

the presence of a-crystallin, we estimated the hydro-

dynamic radius of the initial particles (Rh,0) occurring in

the system at the instant when the initial increase in the

light scattering intensity was registered. For this purpose

we constructed the light scattering intensity versus

hydrodynamic radius plots (Fig. 5). The values of Rh,0

determined from the length cut off on the abscissa axis

by the straight line in these coordinates for a-crystallin

concentrations of 0.025, 0.05 and 0.1 mg/mL were

found to be 9.2–9.3 nm (Table 1). It should be noted

that this value is close to the hydrodynamic radius of

a-crystallin measured at 45 �C (10.1(± 0.2) nm) [43].

Therefore, we assumed that the initial particles regis-

tered in the system at the point in time, at which the

light scattering intensity begins to increase (i.e., the

particles with Rh = Rh,0), are the complexes of a-crys-

tallin with denatured GAPDH.

The initial parts of the dependences of Rh on time

obtained in the presence of a-crystallin at the concentra-

tions of 0.025, 0.05 and 0.1 mg/mL follow the exponential

law. Solid curves in insets in Fig. 4a, b and in Fig. 4c are

calculated from Eq. (3). Parameter t3 and reciprocal value

of parameter t2R are given in Table 1. The diminution of

the 1/t2R value characterizes the decrease in the aggrega-

tion rate.

As in the case of thermal aggregation of aL-crystallin

[41], aggregation of GAPDH in the presence of a-crys-

tallin (0.025–0.1 mg/mL) is accompanied by splitting the

aggregate population into two components at rather high

values of time (t [ tcrit). The values of tcrit and the values

of Rh at t = tcrit (Rcrit) are given in Table 1.

It is noteworthy that at higher concentrations of

a-crystallin (0.4 mg/mL) the increase in the Rh value in

time is insignificant and distribution of aggregates by size

remained unimodal over the entire period of process

registration (1,100 min; Fig. 4d).

Fig. 2 Effect of a-crystallin on thermal inactivation of GAPDH

(0.4 mg/mL) at 45 �C (10 mM Na-phosphate buffer, pH 7.5,

containing 1 mM EDTA and 1 mM dithiothreitol). The dependences

of the relative enzymatic activity A/A0 on time (A0 and A are the

initial and current values of the enzymatic activity). Concentrations of

a-crystallin were as follows: (1) 0, (2) 0.1, (3) 0.2 and (4) 0.4 mg/mL

Fig. 3 Effect of a-crystallin on GAPDH (0.4 mg/mL) aggregation at

45 �C. The dependences of the light scattering intensity on time

obtained in the absence of a-crystallin (1) and in the presence of

various concentrations of a-crystallin: (2) 0.025, (3) 0.05, (4) 0.1 and

(5) 0.4 mg/mL

Table 1 Parameters Rh,0, t3, 1/t2R, tcrit and Rh,crit for the dependences of Rh on time in the case of aggregation of GAPDH (0.4 mg/mL) registered

at 45 �C in the absence and in the presence of a-crystallin

Concentration of a crystallin

(mg/mL)

Rh,0 (nm) t3 (min) 1/t2R (min-1) tcrit (min) Rh,crit (nm)

0 65.5(± 1.3) 4.5(± 0.2) 0.53(± 0.06) – –

0.025 9.3(± 0.7) 16.4(± 0.4) 0.0111(± 0.0007) 390 125

0.05 9.2(± 0.8) 16.2(± 1.0) 0.0057(± 0.0009) 650 110

0.1 9.3(± 0.6) 15.7(± 1.4) 0.0017(± 0.0001) 790 25

Equations (2) and (3) were used for calculations of parameters t3 and 1/t2R for GAPDH aggregation studied in the absence and in the presence of

a-crystallin, respectively

16 K. A. Markossian1 et al.

123

3.4 Effect of a-Crystallin on Heat-Induced

Dissociation of GAPDH

To control the oligomeric state of GAPDH and a-crystallin,

we carried out the sedimentation velocity studies of

GAPDH and a-crystallin at 21 �C. The major peak in the

c(s) distribution for GAPDH (Fig. 6a) with sedimentation

coefficient of 8.6(± 0.6) S corresponds to the tetrameric

form of GAPDH. Figure 6b shows the differential sedi-

mentation coefficient distribution c(s) for the mixture of

GAPDH (0.4 mg/mL) and a-crystallin (0.4 mg/mL; dashed

line). The sedimentation coefficient (8.6(± 0.6) S) and the

Fig. 5 The light scattering intensity versus the hydrodynamic radius

plots for thermal aggregation of GAPDH (0.4 mg/mL) in the presence

of a-crystallin at 45 �C. The concentrations of a-crystallin were as

follows: a 0.025, b 0.05 and c 0.1 mg/mL

Fig. 4 Effect of a-crystallin on the size of aggregates formed in the

course of heating of GAPDH (0.4 mg/mL) at 45 �C. The dependences

of Rh on time obtained in the presence of various concentrations of

a-crystallin: a 0.025, b 0.05, c 0.1 and d 0.4 mg/mL. Insets show the

initial parts of the dependences of Rh on time. Solid curves are

calculated from Eq. (3)

Comparative Analysis of the Effects of a-Crystallin 17

123

shape of the main peak in c(s) distribution for mixture were

the same as in the case of GAPDH alone. The similar result

was obtained for the mixture of GAPDH (0.4 mg/mL) and

a-crystallin (1.0 mg/mL; solid line). For clarity sake, inset

in Fig. 5 shows the c(s) distribution for a-crystallin alone

(1 mg/mL).

Comparison of the sedimentation profiles of GAPDH

(0.4 mg/mL; Fig. 7a) and the mixture of GAPDH with

a-crystallin (0.4 mg/mL; Fig. 7b) indicates that levels of

optical densities corresponding to absorbance of GAPDH

are identical. These results suggest that there is no inter-

action of GAPDH with a-crystallin at 21 �C.

Earlier we showed that incubation of GAPDH at 45 �C

induced dissociation of the protein [51]. Differential sedi-

mentation coefficient distributions for GAPDH and mixture

of GAPDH with a-crystallin are presented in Fig. 8. Panel A

shows general c(s, *) distribution obtained for GAPDH at

45 �C and corrected to the standard conditions. Total incu-

bation time of was 1.5 h. The dissociated enzyme forms

with the sedimentation coefficients s20,w = 3.3(± 0.1) and

5.4(± 0.1) S were observed. The main peak (3.3 S) corre-

sponds to the monomeric form and minor peak (5.4 S) corre-

sponds to the dimeric form of the enzyme. Figure 8b–d shows

the c(s) distributions for mixtures of GAPDH (0.4 mg/mL)

and a-crystallin at the concentrations of 0.1, 0.4 or 1.0 mg/

mL, respectively. All distributions were obtained at 45 �C

and corrected to the standard conditions. Total incubation

time was 1.5 h. The disappearance of the peak corresponding

to the monomeric form of GAPDH with s20,w = 3.3(± 0.1)

S (Fig. 8b–d) was observed. The peak with sedimentation

coefficient of 5.8(± 0.1) S registered for the mixture of

GAPDH with a-crystallin (0.1 mg/mL, Fig. 8b), probably,

corresponds to the complex of GAPDH dimer and products

of dissociation of a-crystallin. When the concentration of

a-crystallin was increased, this peak was shifted towards

the higher values of sedimentation coefficient, namely, to

7.2(± 0.4) S at a-crystallin concentration of 0.4 mg/mL

(Fig. 8c) and 13.0(± 0.2) S at a-crystallin concentration of

1.0 mg/mL (Fig. 8d). Simultaneously the peaks with the

higher sedimentation coefficients appear on c(s) distribu-

tions obtained for mixtures containing a-crystallin 0.4 and

1.0 mg/mL (Fig. 8c, s20,w = 12.7(± 0.3) and 14.0(± 0.3)

S; Fig. 8d, s20,w = 14.7(± 0.2) and 16.5(± 0.3) S). These

peaks correspond to free a-crystallin. This conclusion can be

made from the comparison of c(s) distributions presented in

Fig. 8c and d with the corresponding distributions obtained

for a-crystallin heated for 1.5 h at 45 �C (Fig. 9a, s20,w =

12.9(± 0.3) and 14.0(± 0.3) S; Fig. 9d, s20,w = 14.9(± 0.2)

S). It should be noted that Fig. 9a and b demonstrate disso-

ciation of a-crystallin at elevated temperatures. The char-

acter of the c(s) distribution for a-crystallin at the

temperature of 21 �C was indicative of its polydispersity, the

sedimentation coefficients for main peaks being in the range

20–25 S (inset in Fig. 6b). The c(s) distribution for a-crys-

tallin (1 mg/ml) heated for 1.5 h at 45 �C contains two peaks

with s20,w = 10.2(± 0.2) and 14.9(± 0.2) S (Fig. 9b).

Decrease in a-crystallin concentration to 0.4 mg/mL resul-

ted in further dissociation of the protein. This is evidenced by

the increase in the fraction of the peak with s20,w = 10.2 S

and transformation of the peak with s20,w = 14.9 S to two

peaks with lesser values of s20,w (14.0 and 12.9 S). Such a

stimulation of dissociation of a-crystallin at the decrease in

its concentration we have observed earlier at 48 �C [44].

3.5 Effect of GroEL on the Kinetics of Thermal

Aggregation of GAPDH

To compare the effects of a-crystallin and GroEL on

thermal aggregation of GAPDH, we have studied the

kinetics of GAPDH aggregation at 45 �C in the presence of

Fig. 6 Sedimentation behavior of GAPDH (10 mM Na-phosphate

buffer, pH 7.5, containing 10 mM NaCl) in the absence and in the

presence of a-crystallin at 21 �C. a Differential sedimentation

coefficient distribution c(s) of GAPDH, b c(s) distributions of the

mixture of GAPDH (0.4 mg/mL) and a-crystallin at the concentration

of 0.4 mg/mL (dashed curve) or 1 mg/mL (solid curve). Figure near

the curve is the value of sedimentation coefficient. Inset shows c(s)

distribution for a-crystallin (1.0 mg/mL)

18 K. A. Markossian1 et al.

123

various concentrations of GroEL. The addition of GroEL

resulted in the reduction of the initial increment in the light

scattering intensity in the course of GAPDH aggregation

(Fig. 10). Figure 11 shows the time dependences of the

hydrodynamic radius of particles obtained in the course of

aggregation of GAPDH (0.4 mg/mL or 0.5 lM) at con-

centrations of GroEL as follows: 0.15, 0.3, 0.6 and 1.2 mg/mL

(0.031, 0.062, 0.125 and 0.25 lM, respectively). The

distinctive property of these dependences is that the

hydrodynamic radius of protein aggregates remained con-

stant over certain time interval. This circumstance makes it

possible to size the initial protein aggregates. The calcu-

lations show that the average values of Rh of the initial

protein aggregates (52.5, 58.7, 57.1 and 61.9 nm for the

above concentrations of GroEL) are very close to the size

of the start aggregates registered in the absence of GroEL

(Rh,0 = 65.5 nm). Above a certain point in time (t [ t3) the

increase in the Rh value with time follows the exponential

law. Solid curves in Fig. 11a–d were calculated from Eq.

(3). The use of Eq. (3) allows estimating parameters t3 and

t2R. The values of these parameters are given in Table 2.

The t3 value (the moment of time at which the Rh value of

protein aggregates begins to increase) rises with increasing

the concentration of GroEL. The decrease in the 1/t2R value

with increasing GroEL concentration is indicative of the

diminution of the aggregation rate.

3.6 Analysis of Possibility of Interaction between

GAPDH and GroEL

To check a possibility of interaction of GAPDH and GroEL

under non-denaturing conditions, we studied the sedimen-

tation behavior of GAPDH, GroEL and their mixture at

room temperature (0.1 M NaCl, 10 mM Na-phosphate

buffer, pH 7.5; Fig. 12). The main peak in the differential

sedimentation coefficient distribution c(s) for GAPDH

[s = 7.7(± 0.5) S] in Fig. 12a) corresponds to tetramer.

The major peak in c(s) distribution for GroEL

[s = 22.4(± 0.6) S] in Fig. 12b corresponds to the homo

14-mer with molecular mass of *800 kDa. The additional

peaks with s = 5.4(± 0.5), 10.5(± 0.5) and 15.1(± 0.5) S

correspond to the dissociated forms of GroEL. On the basis

of the obtained values of sedimentation coefficients and the

values of the frictional ratio we estimated the molecular

masses for the peaks with s = 5.4 and s = 15.1 S (72 and

402 kDa). One can infer that these peaks correspond to

the monomeric and heptameric forms, respectively. The

appearance of the dissociated forms in the preparation of

GroEL is a not a surprising result, because the procedures

used for isolation and purification of GroEL can promote

the formation of GroEL monomers [12, 13]. Figure 12c

shows the sedimentation coefficients distribution c(s) for

the mixtures of GAPDH and GroEL. Analysis of c(s) dis-

tribution indicates that there is no interaction between

GAPDH and GroEL at room temperature. The major peak

with s = 7.8(± 0.3) S corresponds to the tetrameric form

of GAPDH. The peaks with s = 6.0(± 0.3), 15.0(± 0.2)

and 21.7(± 0.3) S represent the oligomeric forms of

GroEL, which are identical to the corresponding oligo-

meric forms registered for individual GroEl (Fig. 12b).

Figure 13 shows the c(s) distributions for two concen-

trations of GroEL heated for 2 h at 46 �C. When the GroEL

concentration was 1 mg/mL, c(s) distribution included three

peaks (Fig. 13a). The main peak with s20,w = 22.8(± 0.4) S

corresponds to the 14-mer of GroEL. One can assume that

the peak with s20,w = 4.0(± 0.3) S corresponds to the

unfolded form of the GroEL monomer and peak with

Fig. 7 Sedimentation profiles

of GAPDH (0.4 mg/mL) in the

absence (a) and in the presence

(b) of a-crystallin (0.4 mg/mL)

at 21 �C. The rotor speed was

44,000 rpm

Comparative Analysis of the Effects of a-Crystallin 19

123

s20,w = 6.4(± 0.3) S corresponds to aggregates formed by

this species. It is significant that the monomeric form of

GroEL is characterized by markedly lower thermostability

than the tetradecameric form [13]. The increase in the GroEL

concentration to 1.7 mg/mL (Fig. 13b) resulted in the disap-

pearance of the peaks with s20,w = 4.0 and 6.4 S. However,

the higher molecular mass species with s20,w = 10.8(± 0.6),

16.5(± 0.6 S) and 19.8(± 0.4) S appeared. It should be noted

that apart from 23.5 S-peak (14-mer) the c(s) distribution

contained large-sized particles with s20,w = 27.5((0.3) S.

This species probably corresponds to the complex of tet-

radecameric GroEL molecule with the unfolded GroEL

subunits. One can speculate that the appearance of the

peak with s20,w = 19.8 S is connected with the formation

of the complex between single-ring chaperonin and

denatured GroEL. Thus, heating of GroEL may result

in dissociation of the tetradecameric form of GroEL to

monomers, denaturation of the monomeric form and

complexation of the products of denaturation with single-

and double-rings.

Fig. 9 Sedimentation behavior of a-crystallin at 45 �C. Distributions

c(s) were obtained for a-crystallin (0.4 mg/mL (a) and 1.0 mg/mL

(b)) heated for 1.5 h at 45 �C and corrected to the standard

conditions. The rotor speed was 30,000 rpm

Fig. 8 Differential sedimentation coefficient distributions c(s, f/f0) for

GAPDH (0.4 mg/ml; a) and c(s) for the mixture of GAPDH (0.4 mg/mL)

and a-crystallin at the concentrations of 0.1, 0.4 and 1.0 mg/mL (b–d,

respectively) heated for 1.5 h at45 �C. Distributions c(s, f/f0) andc(s) were

obtained at 45 �C and corrected to the standard conditions. The rotor speed

was 30,000 rpm

20 K. A. Markossian1 et al.

123

In the case of GAPDH, heating resulted in more pro-

nounced dissociation of the initial oligomeric form into the

oligomeric forms of lesser size. Figure 14a shows the

sedimentation behavior of GAPDH (0.2 mg/mL) heated for

2 h at 46 �C before the run. The c(s) distribution contains a

single peak with s20,w = 4.6(± 0.3) S. The frictional ratio

was equal to 1.9. This peak corresponds to the dissociated

form of the tetrameric GAPDH (the dimeric form with the

changed conformation).

Figure 14b shows the sedimentation behavior of the

mixture of GAPDH (0.2 mg/mL) and GroEL (1.0 mg/mL).

There are major peaks with s20,w = 6.6 ± (0.6), s20,w =

12.0 ± (0.6) and 19.8(± 0.3) S and also several minor

peaks in the c(s) distribution. When the GroEL concen-

tration in the mixture with GAPDH was 1.7 mg/mL

(Fig. 14c), the c(s) distribution contained the peaks with

s20,w = 6.7(± 0.6), 10.3(± 0.6) and 27.3(± 0.3) S. It

should be noted that the peak corresponding to the 14-mer

GroEL is lacking in both mixtures of GAPDH with GroEL

(Fig. 14b, c; confer with Fig. 13a, b). The peak with

s20,w = 27.5 S registered in the mixture of GAPDH

(0.2 mg/mL) and GroEL (1.7 mg/mL; Fig. 14c) is proba-

bly identical to the corresponding peak in the c(s) distri-

bution obtained for GroEL (1.7 mg/m) heated for 2 h at

46 �C Fig. 13b. As assumed, this species is the complex of

the 14-mer GroEL and products of GroEL denaturation.

One can assume that the species with s20,w = 27.5 in the

mixture of GAPDH with GroEL (Fig. 14c) is not a result of

interaction of the 14-mer GroEL with denatured GAPDH.

The c(s) distribution shown in Fig. 14c does not contain the

peak corresponding to dissociated form of GAPDH

(s20,w = 4.5; Fig. 14a). Thus, denatured GAPDH becomes

in complex with the dissociated forms of GroEL and is

represented by 6.6 S-peak or 10.3-peak (Fig. 14c). It

Fig. 10 Effect of GroEL on thermal aggregation of GAPDH (0.4 mg/

mL) at 45 �C. The dependences of the light scattering intensity on

time obtained without GroEL (1) and in the presence of various

concentrations of GroEL: (2) 0.15, (3) 0.3, (4) 0.6 and (5) 1.2 mg/mL

Fig. 11 The dependences of the hydrodynamic radius of the protein

aggregates on time for aggregation of GAPHD (0.4 mg/mL) heated at

45 �C in the presence of GroEL at the following concentrations:

a 0.15, b 0.3, c 0.6 and d 1.2 mg/mL. Solid curves are calculated from

Eq. (3)

Comparative Analysis of the Effects of a-Crystallin 21

123

should be noted that the c(s) distribution for heated GroEL

contains 10.8 S-peak (Fig. 13b), whereas 6.7 S-peak is

lacking. Taking into account this result we can assume that

6.7 S-peak (Fig. 14c) corresponds to the main complex of

dissociated forms of GAPDH and GroEL. This conclusion

is supported by the observation that 6.6 S-peak is a major

Table 2 Parameters Rh,0, t3 and 1/t2R for the dependences of Rh on time for aggregation of GAPDH (0.4 mg/mL) registered at 45 �C in the

absence and in the presence of GroEL

Concentration

of GroEL (mg/mL)

Rh,0 (nm) t3 (min) 1/t2R (min-1)

0 65.5(± 1.3) 4.5(± 0.2) 0.53(± 0.06)

0.15 52.5(± 1.9) 7.7(± 0.4) 0.22(± 0.02)

0.3 58.7(± 2.3) 12.3(± 1.3) 0.132(± 0.017)

0.6 57.1(± 1.3) 49.2(± 1.5) 0.069(± 0.007)

1.2 61.9(± 1.5) 99(± 2) 0.015(± 0.001)

Equations (2) and (3) were used for calculations of parameters t3 and 1/t2R for GAPDH aggregation studied in the absence and in the presence of

GroEL, respectively

Fig. 12 Differential sedimentation coefficient distributions c(s) for

GAPDH (0.4 mg/mL; a), GroEL (1.4 mg/mL; b) and their mixture at

the same concentrations (c) obtained at 21 �C. The rotor speed was

48,000 rpm

Fig. 13 Sedimentation behavior of GroEL at 46 �C. Distributions

c(s) for GroEL at the concentrations: a 1.0 mg/mL and b 1.7 mg/ml.

The solutions of GroEL were heated for 2 h at 46 �C before the run.

The rotor speed was 44,000 rpm

22 K. A. Markossian1 et al.

123

component in the c(s) distribution of the mixture of

GAPDH (0.2 mg/mL) and GroEL (1.0 mg/mL; Fig. 13b).

The fact that the monomeric form of GroEL is character-

ized by the higher degree of exposure of the hydrophobic

clusters of the protein molecule in comparison with the

tetradecameric form [12, 13] can serve as an explanation of

the preferential binding of the denatured GAPDH with the

GroEL monomer.

4 Discussion

One of the functions of molecular chaperones is stabilization

of unfolded proteins [66]. Complexation of chaperones with

unfolded protein substrates prevents aggregation of the lat-

ter. The challenge now is to carry out a comparative analysis

of the anti-aggregating ability of chaperones of different

classes. In the present work we compared the anti-aggre-

gating ability of chaperones of two classes. One chaperone

(a-crystallin) is a representative of the family of small heat

shock proteins, and the other chaperone (GroEL) belongs to

the family of Hsp60 chaperones, also termed chaperonins.

Our experiments on thermal aggregation of GAPDH have

shown that the kinetic mechanisms of suppressive action of

a-crystallin and GroEL are similar. The exponential char-

acter of the dependence of the hydrodynamic radius of pro-

tein aggregates on time obtained in the presence of a-

crystallin and GroEL indicates that these chaperones induce

the transition of the kinetic regime of GAPDH aggregation

from the DLCA regime, wherein the rate of aggregation is

limited by diffusion of the interacting particles (the start

aggregates and aggregates of higher order) and sticking

probability for the colliding particles is equal to unity, to the

RLCA regime, wherein the sticking probability for the col-

liding particles becomes lower the unity. As for a-crystallin,

this conclusion is in agreement with the results of our pre-

vious papers [41, 42, 67]. However, there is a distinction

between a-crystallin and GroEL in their influence on the first

stage of thermal aggregation of GAPDH, namely the stage of

formation of the start aggregates. Complexation of denatured

form of GAPDH with a-crystallin blocks the formation of the

start aggregates. Sticking of the complexes denatured

GAPDH-a-crystallin proceeds in the RLCA regime. The

denatured GAPDH-a-crystallin complex is the initial parti-

cle participating in the aggregation process. When the sys-

tem contains GroEL, the start aggregates are identical in size

to those registered in the absence of GroEL. However,

incorporation of GroEL in the start aggregates results in such

alterations of their surface which diminish their sticking

probability.

Sedimentation velocity and DSC analysis show that at

the elevated temperatures a-crystallin or GroEL can

interact with the intermediates of unfolding of GAPDH.

Recent investigations show that, when interpreting the anti-

aggregating function of the chaperones, one must take into

account dissociation-association processes both for protein

substrates and chaperones. The data obtained in the present

work support the idea that the products of dissociation of

both proteins play an important role in the interactions of

these proteins at elevated temperatures. The similar results

were obtained for interaction of a-crystallin with glycogen

phosphorylase b at 48 �C [44].

Fig. 14 Sedimentation behavior of GAPDH (0.2 mg/mL) and the

mixture of GAPDH with GroEL at 46 �C. a Differential sedimenta-

tion coefficient distribution c(s) for GAPDH. b and c c(s) distributions

for the mixture of GAPDH (0.2 mg/mL) with GroEL at the

concentrations of 1.0 and 1.7 mg/mL, respectively. All samples were

heated for 2 h at 46 �C before the run. The rotor speed was

44,000 rpm

Comparative Analysis of the Effects of a-Crystallin 23

123

Scheme 1 demonstrates the interaction of a-crystallin

with the denatured form of GAPDH. In this scheme, D is

the denatured form of GAPDH, A is the start aggregate

containing n denatured GAPDH molecules, C is a-crys-

tallin and A is the start aggregate containing GroEL.

Strictly speaking, this scheme should be a complex one.

The fact that a-crystallin accelerates thermal inactivation

of GAPDH (the data of the present work) and reduces

thermostability of the enzyme (according to the DSC data

presented in our previous work [42]) is indicative of its

ability to interact with the intermediates of GAPDH

unfolding. As for GroEL, this chaperone does not affect

thermal inactivation and denaturation of GAPDH [68]. It is

interesting to note that according to the DSC data GAPDH

enhances thermostability of GroEL: in the presence of

GAPDH the position of maximum on the DSC profile for

GroEL is shifted from 60.6 to 63.6 �C (the ratio for molar

concentrations [GAPDH]:[GroEL] was varied from 1:1 to

4:1) [68]. Our results show that when GAPDH denaturation

proceeds in the presence of GroEL the latter becomes

incorporated in the start aggregates formed by the dena-

tured molecules of GAPDH. One may imply that it is not

simply complexation of GroEL with denatured GAPDH,

but incorporation of GroEL in the start aggregates, which is

responsible for stabilization of GroEL in the presence of

GAPDH. The multiple contacts of GroEL molecule with

the denatured molecules of GAPDH may be realized in the

start aggregate. It should be noted that diminution of the

size of the start aggregates, i.e., their destruction in the

presence of a-crystallin, was also observed when studying

thermal aggregation of aL-crystallin from bovine lens [41],

yeast alcohol dehydrogenase I [42], glycogen phosphory-

lase b from rabbit skeletal muscles [44, 67] and aspartate

aminotransferase from pig heart mitochondria [45]. The

fact that a-crystallin and GroEL have different effects on

thermostability of GAPDH is indicative of different char-

acter of interaction of these chaperones with the protein

substrate under study. The interaction of a-crystallin with

unfolded GAPDH blocks the formation of the start aggre-

gates. As for GroEL, the interaction of this chaperone with

unfolded GAPDH does not affect the formation of the start

aggregates. However, GroEL incorporated in the start

aggregates reduces the sticking probability for the colliding

start aggregates. Thus, one can assume that a-crystallin and

GroEL interact with different regions of unfolded GAPDH

molecule.

Acknowledgments This study was funded by the Russian Foun-

dation for Basic Research (grants 08-04-00666-a, 08-04-00231_a and

08-08-00540_a), the Program ‘‘Molecular and Cell Biology’’ of the

Presidium of the Russian Academy of Sciences and CNTP (Russia,

02.512.11.2249).

References

1. Ellis RJ (1999) Curr Biol 9:137–139

2. Hartl FU, Hayer-Hartl M (2002) Science 295:1852–1858

3. Hartl FU (1996) Nature 381:571–579

4. Ranson NA, White HE, Saibil HR (1998) Biochem J 333:233–

242

5. Muchowski PJ, Bassuk JA, Lubsen NH, Clark JI (1997) J Biol

Chem 272:2578–2582

6. Haslbeck M, Franzmann T, Weinfurtner D, Buchner J (2005) Nat

Struct Mol Biol 12:842–846

7. Surin AK, Kotova NV, Kashparov IA, Marchenkov VV,

Marchenkova SY, Semisotnov GV (1997) FEBS Lett 405:260–

262

8. Bukau B, Horwich AL (1998) Cell 92:351–366

9. Fink AL (1999) Physiol Rev 79:425–449

10. Frydman J (2001) Annu Rev Biochem 70:603–647

1) GAPDH aggregation in the absence of chaperones

The stage of formation of start aggregates

nD →

The stage of aggregate growth

+ → 2

. . . . . . . . . . . . .

i + jDLCAregime⎯⎯⎯→ i+j

. . . . . . . . . . . . .

2) GAPDH aggregation in the presence of ααα-crystallin

The stage of complexation of denatured protein with α-crystallin

D + C → D⋅C

The stage of aggregate growth

(D⋅C) + (D⋅C) → (D⋅C)2

. . . . . . . . . . . . . . . . . . .

(D⋅C)i + (D⋅C)jRLCAregime⎯⎯⎯→ (D⋅C)i+j

. . . . . . . . . . . . . . . . . . .

3) GAPDH aggregation in the presence of GroEL

The stage of formation of start aggregates

nD + GroEL →

The stage of aggregate growth

+ → 2

. . . . . . . . . . . . .

i + jRLCAregime⎯⎯⎯→ i+j

. . . . . . . . . . . . .

A

A A A

A A A

A A A

A

A A A

Scheme 1 Illustrates mechanisms of suppression of GAPDH aggre-

gation by a-crystallin and GroEL

24 K. A. Markossian1 et al.

123

11. Clare DK, Bakkes PJ, van Heerikhuizen H, van der Vies SM,

Saibil HR (2009) Nature 457:107–110

12. Goloubinoff P, Gatenby AA, Lorimer GH (1989) Nature 337:

44–47

13. Holl-Neugebauer B, Rudolph R, Schmidt M, Buchner J (1991)

Biochemistry 30:11609–11614

14. Fisher MT (1992) Biochemistry 31:3955–3963

15. Martin J, Horwich AL, Hartl FU (1992) Science 258:995–998

16. Mendoza JA, Manson M, Joves F, Ackermann E (1996) Bio-

technol Tech 10:535–540

17. Hartman DJ, Surin BP, Dixon NE, Hoogenraad NJ, Hoj PB

(1993) Proc Natl Acad Sci U S A 90:2276–2280

18. Ellis RJ, Hartl FU (1999) Curr Opin Struct Biol 9:102–110

19. Ellis RJ (1996) Fold Des 1:9–15

20. Ellis RJ (2003) Protein folding. Curr Biol 13:R881–R883

21. Chaudhuri TK, Farr GW, Fenton WA, Rospert S, Horwich AL

(2001) Cell 107:235–246

22. Farr GW, Fenton WA, Chaudhuri TK, Clare DK, Saibil HR,

Horwich AL (2003) EMBO J 22:3220–3230

23. Jaenicke R (1995) Philos Trans R Soc Lond B Biol Sci 348:

97–105

24. Corrales FJ, Fersht AR (1996) Fold Des 1:265–273

25. Fenton WA, Kashi Y, Furtak K, Horwich AL (1994) Nature

371:614–619

26. Lin Z, Schwartz FP, Eisenstein E (1995) J Biol Chem 270:1011–

1014

27. Aoki K, Taguchi H, Shindo Y, Yoshida M, Ogasahara K, Yutani

K, Tanaka N (1997) J Biol Chem 272:32158–32162

28. Magonet E, Delaive E, Martin B, Remacle J (1992) Biochem Int

28:603–612

29. Marchenkov VV, Sokolovskii IV, Kotova NV, Galzitskaya OV,

Bochkareva ES, Girshovich AS, Semisotnov GV (2004) Biofizika

(Mosc) 49:987–994

30. Marchenko NY, Marchenkov VV, Kaysheva AL, Kashparov IA,

Kotova NV, Kaliman PA, Semisotnov GV (2006) Biochemistry

(Mosc) 71:1357–1364

31. Marchenkov VV, Semisotnov GV (2009) Int J Mol Sci 10:2066–

2083

32. Marchenkov VV, Marchenko NY, Marchenkova SY, Semisotnov

GV (2006) Usp Biol Khim (Mosc) 46:279–302

33. Ybarra J, Horowitz PM (1995) J Biol Chem 270:22962–22967

34. Surin AK, Kotova NV, SYu Marchenkova, Marchenkov VV,

Semisotnov GV (1999) Bioorg Khim 25:358–364

35. Horwitz J (1992) Proc Natl Acad Sci U S A 89:10449–10453

36. Carver JA, Guerreiro N, Nicholls KA, Truscott RJ (1995) Bio-

chim Biophys Acta 1252:251–260

37. Wang K, Spector A (1994) J Biol Chem 269:13601–13608

38. Burgio MR, Kim CJ, Dow CC, Koretz JF (2000) Biochem Bio-

phys Res Commun 268:426–432

39. Abgar S, Vanhoudt J, Aerts T, Clauwaert J (2001) Biophys J

80:1986–1995

40. Putilina T, Skouri-Panet F, Prat K, Lubsen NH, Tardieu A (2003)

J Biol Chem 278:13747–13756

41. Khanova HA, Markossian KA, Kurganov BI, Samoilov AM,

Kleimenov SY, Levitsky DI, Yudin IK, Timofeeva AC, Muranov

KO, Ostrovsky MA (2005) Biochemistry 44:15480–15487

42. Markossian KA, Kurganov BI, Levitsky DI, Khanova HA,

Chebotareva NA, Samoilov AM, Eronina TB, Fedurkina NV,

Mitskevich LG, Merem’yanin AV, Kleymenov SY, Makeeva VF,

Muronets VI, Naletova IN, Shalova IN, Asryants RA, Schmal-

hausen EV, Saso L, Panyukov YV, Dobrov EN, Yudin IK,

Timofeeva AC, Muranov KO, Ostrovsky MA (2006) In: Oba-

linsky TR (ed) Protein folding: new research , Nova Science

Publishers Inc., New York, pp 89–171

43. Khanova HA, Markossian KA, Kleimenov SY, Levitsky DI,

Chebotareva NA, Golub NV, Asryants RA, Muronetz VI, Saso L,

Yudin IK, Muranov KO, Ostrovsky MA, Kurganov BI (2007)

Biophys Chem 125:521–531

44. Meremyanin AV, Eronina TB, Chebotareva NA, Kurganov BI

(2008) Biopolymers 89:124–134

45. Golub NV, Markossian KA, Sholukh MV, Muranov KO,

Kurganov BI (2009) Eur Biophys J 38:547–556

46. Markossian KA, Golub NV, Kleymenov SY, Muranov KO,

Sholukh MV, Kurganov BI (2009) Int J Biol Macromol 44:441–

446

47. Markossian KA, Yudin IK, Kurganov BI (2009) Intern J Mol Sci

10:1314–1345

48. Lin YZ, Liang SJ, Zhou JM, Tsou CL, Wu PQ, Zhou ZK (1990)

Biochim Biophys Acta 1038:247–252

49. Levashov P, Orlov V, Boschi-Muller S, Talfournier F, Asryants

R, Bulatnikov I, Muronetz V, Branlant G, Nagradova N (1999)

Biochim Biophys Acta 1433:294–306

50. Shalova IN, Asryants RA, Sholukh MV, Saso L, Kurganov BI,

Muronetz VI, Izumrudov VA (2005) Macromol Biosci 5:1184–

1192

51. Markossian KA, Khanova HA, Kleimenov SY, Levitsky DI,

Chebotareva NA, Asryants RA, Muronetz VI, Saso L, Yudin IK,

Kurganov BI (2006) Biochemistry 45:13375–13384

52. Bhattacharyya AM, Horowitz PM (2002) Biochemistry 41:2248–

2421

53. Melkani GC, Zardeneta G, Mendoza JA (2005) Int J Biochem

Cell Biol 37:1375–1385

54. Naletova IN, Muronetz VI, Schmalhausen EV (2006) Biochim

Biophys Acta 1764:831–838

55. Scopes RK, Stoter A (1982) Methods Enzymol 90:479–490

56. Kirschenbaum DM (1972) Int J Protein Res 4:63–73

57. Chiou SH, Azari P, Himmel ME, Squire PG (1979) Int J Pept

Protein Res 13:409–417

58. Walters C, Errington N, Rowe AJ, Harding SE (2002) Biochem J

364:849–855

59. Panyukov Y, Yudin I, Drachev V, Dobrov E, Kurganov B (2007)

Biophys Chem 127:9–18

60. Weitz DA, Huang JS, Lin MY, Sung J (1984) Phys Rev Lett

53:1657–1660

61. Lin MY, Lindsay HM, Weitz DA, Ball RC, Klein R, Meakin P

(1989) Proc R Soc Lond 423:71–87

62. Golub N, Meremyanin A, Markossian K, Eronina T, Chebotareva

N, Asryants R, Muronets V, Kurganov B (2007) FEBS Lett

581:4223–4227

63. Weitz DA, Huang JS, Lin MY, Sung J (1985) Phys Rev Lett

54:1416–1419

64. Brown PH, Schuck P (2006) Biophys J 90:4651–4661

65. Brown PH, Balbo A, Schuck P (2007) Biomacromolecules

8:2011–2024

66. Saibil H (2000) Curr Opin Struct Biol 10:251–258

67. Meremyanin AV, Eronina TB, Chebotareva NA, Kleimenov SY,

Yudin IK, Muranov KO, Ostrovsky MA, Kurganov BI (2007)

Biochemistry (Mosc) 72:518–528

68. Naletova IN, Shmal’gauzen EV, Shalova IN, Pleten AP, Tsi-

riul’nikov K, Haertle T, Muronets VI (2006) Biomed Khim

(Mosc) 52:518–525

Comparative Analysis of the Effects of a-Crystallin 25

123