Stereoselectivity and conformational stability of haloalkane dehalogenase DbjA from Bradyrhizobium...

11
Stereoselectivity and conformational stability of haloalkane dehalogenase DbjA from Bradyrhizobium japonicum USDA110: the effect of pH and temperature Radka Chaloupkova 1,2 , Zbynek Prokop 1,2 , Yukari Sato 3 , Yuji Nagata 3 and Jiri Damborsky 1,2 1 Loschmidt Laboratories, Department of Experimental Biology, Faculty of Science, Masaryk University, Brno, Czech Republic 2 International Clinical Research Center, St Anne’s University Hospital Brno, Czech Republic 3 Graduate School of Life Sciences, Tohoku University, Sendai, Japan Keywords activity; enantioselectivity; haloalkane dehalogenase; oligomerization; pH; structure; thermostability Correspondence J. Damborsky, Loschmidt Laboratories, Department of Experimental Biology, Faculty of Science, Masaryk University, Kamenice 5 A13, 625 00 Brno, Czech Republic Fax: +420549496302 Tel: +420549493467 E-mail: [email protected] (Received 6 February 2011, revised 15 May 2011, accepted 31 May 2011) doi:10.1111/j.1742-4658.2011.08203.x The effect of pH and temperature on structure, stability, activity and enantioselectivity of haloalkane dehalogenase DbjA from Bradyrhizobium japonicum USDA110 was investigated in this study. Conformational changes have been assessed by circular dichroism spectroscopy, functional changes by kinetic analysis, while quaternary structure was studied by gel filtration chromatography. Our study shows that the DbjA enzyme is highly tolerant to pH changes. Its secondary and tertiary structure was not affected by pH in the ranges 5.3–10.3 and 6.2–10.1, respectively. Oligomeri- zation of DbjA was strongly pH-dependent: monomer, dimer, tetramer and a high molecular weight cluster of the enzyme were distinguished in solu- tion at different pH conditions. Moreover, different oligomeric states of DbjA possessed different thermal stabilities. The highest melting tempera- ture (T m = 49.1 ± 0.2 °C) was observed at pH 6.5, at which the enzyme occurs in dimeric form. Maximal activity was detected at 50 °C and in the pH interval 7.7–10.4. While pH did not have any effect on enantiodiscri- minination of DbjA, temperature significantly altered DbjA enantioselectiv- ity. A decrease in temperature results in significantly enhanced enantioselectivity. The temperature dependence of DbjA enantioselectivity was analysed with 2-bromobutane, 2-bromopentane, methyl 2-bromopropi- onate and ethyl 2-bromobutyrate, and differential activation parameters D RS DH z and D RS DS z were determined. The thermodynamic analysis revealed that the resolution of b-bromoalkanes was driven by both enthal- pic and entropic terms, while the resolution of a-bromoesters was driven mainly by an enthalpic term. Unique catalytic activity and structural stabil- ity of DbjA in a broad pH range, combined with high enantioselectivity with particular substrates, make this enzyme a very versatile biocatalyst. Enzyme EC 3.8.1.5 haloalkane dehalogenase. Abbreviations CD, circular dichroism; MRE, mean residue ellipticity. 2728 FEBS Journal 278 (2011) 2728–2738 ª 2011 The Authors Journal compilation ª 2011 FEBS

Transcript of Stereoselectivity and conformational stability of haloalkane dehalogenase DbjA from Bradyrhizobium...

Stereoselectivity and conformational stability ofhaloalkane dehalogenase DbjA fromBradyrhizobium japonicum USDA110: the effect of pHand temperatureRadka Chaloupkova1,2, Zbynek Prokop1,2, Yukari Sato3, Yuji Nagata3 and Jiri Damborsky1,2

1 Loschmidt Laboratories, Department of Experimental Biology, Faculty of Science, Masaryk University, Brno, Czech Republic

2 International Clinical Research Center, St Anne’s University Hospital Brno, Czech Republic

3 Graduate School of Life Sciences, Tohoku University, Sendai, Japan

Keywords

activity; enantioselectivity; haloalkane

dehalogenase; oligomerization; pH;

structure; thermostability

Correspondence

J. Damborsky, Loschmidt Laboratories,

Department of Experimental Biology,

Faculty of Science, Masaryk University,

Kamenice 5 ⁄ A13, 625 00 Brno, Czech

Republic

Fax: +420549496302

Tel: +420549493467

E-mail: [email protected]

(Received 6 February 2011, revised 15 May

2011, accepted 31 May 2011)

doi:10.1111/j.1742-4658.2011.08203.x

The effect of pH and temperature on structure, stability, activity and

enantioselectivity of haloalkane dehalogenase DbjA from Bradyrhizobium

japonicum USDA110 was investigated in this study. Conformational

changes have been assessed by circular dichroism spectroscopy, functional

changes by kinetic analysis, while quaternary structure was studied by gel

filtration chromatography. Our study shows that the DbjA enzyme is

highly tolerant to pH changes. Its secondary and tertiary structure was not

affected by pH in the ranges 5.3–10.3 and 6.2–10.1, respectively. Oligomeri-

zation of DbjA was strongly pH-dependent: monomer, dimer, tetramer and

a high molecular weight cluster of the enzyme were distinguished in solu-

tion at different pH conditions. Moreover, different oligomeric states of

DbjA possessed different thermal stabilities. The highest melting tempera-

ture (Tm = 49.1 ± 0.2 �C) was observed at pH 6.5, at which the enzyme

occurs in dimeric form. Maximal activity was detected at 50 �C and in the

pH interval 7.7–10.4. While pH did not have any effect on enantiodiscri-

minination of DbjA, temperature significantly altered DbjA enantioselectiv-

ity. A decrease in temperature results in significantly enhanced

enantioselectivity. The temperature dependence of DbjA enantioselectivity

was analysed with 2-bromobutane, 2-bromopentane, methyl 2-bromopropi-

onate and ethyl 2-bromobutyrate, and differential activation parameters

DR�SDHz and DR�SDSz were determined. The thermodynamic analysis

revealed that the resolution of b-bromoalkanes was driven by both enthal-

pic and entropic terms, while the resolution of a-bromoesters was driven

mainly by an enthalpic term. Unique catalytic activity and structural stabil-

ity of DbjA in a broad pH range, combined with high enantioselectivity

with particular substrates, make this enzyme a very versatile biocatalyst.

Enzyme

EC 3.8.1.5 haloalkane dehalogenase.

Abbreviations

CD, circular dichroism; MRE, mean residue ellipticity.

2728 FEBS Journal 278 (2011) 2728–2738 ª 2011 The Authors Journal compilation ª 2011 FEBS

Introduction

Haloalkane dehalogenases (EC 3.8.1.5) make up an

important class of enzymes which are able to cleave car-

bon–halogen bonds in a broad range of halogenated ali-

phatic compounds. The hydrolytic dehalogenation

catalysed by these enzymes proceeds by nucleophilic

substitution of a halogen atom with a hydroxyl group

forming corresponding alcohols [1]. Haloalkanes, halo-

alcohols and alcohols are valuable building blocks in

organic and pharmaceutical synthesis [2–4], making

haloalkane dehalogenases potentially applicable in bio-

catalysis. We have recently shown that newly isolated

haloalkane dehalogenase DbjA from Bradyrhizobium ja-

ponicum USDA110 [5] possesses new substrate specific-

ity with high catalytic activity towards b-methylated

haloalkanes and sufficient enantioselectivity for indus-

trial scale synthesis of optically pure compounds [6].

Interestingly, the haloalkane dehalogenase DbjA (a) can

kinetically discriminate between enantiomers of two dis-

tinct groups of substrates, a-bromoesters and b-bro-moalkanes; (b) has enantioselectivity based on distinct

molecular interactions, which can be modified sepa-

rately by engineering of a surface loop; and (c) can

adopt an inverse temperature dependence of enantiose-

lectivity for b-bromoalkanes, but not a-bromoesters, by

mutating this surface loop and a flanking residue [7].

Use of enzymes in biocatalytic preparation of opti-

cally pure substances has been rapidly expanding in

recent years [8]. The efficient utilization of enzymes in

industrial processes requires that a number of criteria

are fulfilled, e.g. high activity, stability under process

conditions, appropriate substrate specificity and enanti-

oselectivity [9–11]. The manipulation of the physical

environment is an attractive way to provide additional

control of enzyme stereochemistry and catalytic func-

tionality alongside other methods, such as protein

engineering and directed evolution [12–14]. Under-

standing the effect of physical parameters on the struc-

ture and activity of an enzyme is important for

optimization of the operational conditions of a biocat-

alytic process, while knowledge of the structure–func-

tion relationships provides an essential theoretical

framework for modification of a biocatalyst by

rational protein design [15].

In this work we have systematically examined the

effects of pH and temperature on the stability, oligo-

merization state and functionality of the DbjA enzyme

using CD spectroscopy, size exclusion chromatogra-

phy, activity and enantioselectivity assays. Thermody-

namic analysis has been used to address the origin of

enantiomeric discrimination by determining differential

activation enthalpy and entropy for the enzymatic

reaction with racemic substrates 2-bromobutane,

2-bromopentane, ethyl 2-bromopropionate and methyl

2-bromobutyrate.

Results and Discussion

Conformational changes

CD spectroscopy was used for investigation of the sec-

ondary and tertiary structure of the DbjA enzyme at

pH conditions ranging from 1.7 to 11.5 in the far UV

and near UV spectral regions, respectively. The far UV

CD spectrum of native enzyme, measured in 50 mM

potassium phosphate buffer (pH 7.5 at 4 �C), exhibitedtwo negative features at 208 and 222 nm characteristic

of a-helical content (Fig. 1A, red bold curve). Similar

spectral features were found throughout the pH range

5.3–10.3, suggesting that enzyme secondary structure

remained preserved under these conditions. Calculated

Fig. 1. Far UV (A) and near UV (B) CD spectra of DbjA as a func-

tion of pH. The spectra shown represent the average of 10 consec-

utive scans.

R. Chaloupkova et al. Stereochemistry and conformational stability of DbjA

FEBS Journal 278 (2011) 2728–2738 ª 2011 The Authors Journal compilation ª 2011 FEBS 2729

a-helical content as a function of pH using the method

of Chen et al. [16], which is based on far UV CD data

at 222 nm, is presented in Fig. 2. Predicted a-helicalcontent at pH 5.3–10.3 was about 30.5%. The second-

ary structure of DbjA remains intact within five pH

units. At lower pH levels (pH < 5.0), the enzyme visu-

ally aggregates with simultaneous loss of UV signal.

On the other hand, at pH 11.0–11.4, the enzyme stays

in solution showing approximately a 42% loss in

a-helical content in comparison with its native state.

A strong negative band at 204 nm and a weak band

at 220 nm suggest that DbjA enzyme conformation

starts to be disordered at these extremely alkaline

conditions.

The near UV CD spectrum of the native state of the

enzyme reveals three negative ellipticity peaks at 259,

265 and 285 nm and a positive peak at 292 nm

(Fig. 1B, red bold curve). The ellipticity values at these

wavelengths remain preserved within the pH range

6.2–10.1. In acidic conditions, pH < 6.2, the CD

intensity at 285 and 292 nm slightly increases as a

result of the decreasing pH. The positive ellipticity at

292 nm can be attributed to a tryptophan environment,

since this region corresponds to the absorption band

for this residue [17]. The intensity changes observed at

292 nm might be related to a change in the tryptophan

environment as a result of the loss of some tertiary

interactions. This indicates that the enzyme starts to

lose its tertiary interactions without any secondary

structure loss before complete aggregation. In alkaline

conditions, pH > 10.7, the protein loses most of its

tertiary structure. A considerable increase in the ellip-

ticity at pH ‡ 10.7 is observed at 259 nm. This could

be caused by sudden exposure of phenylalanine resi-

dues in the extreme alkaline pH region. Comparison of

both near UV and far UV CD spectra determined at

various pH conditions revealed similar pH regions at

which the enzyme is structurally stable.

Changes in the structure could be attributed to a

change of ionization state of the enzyme at pH condi-

tions close to its isoelectric point (pI). The predicted pI

of DbjA is 5.89. Although many proteins demonstrate

a state of minimal solubility at their pI conditions,

DbjA remains soluble with a preserved secondary

structure. When pH is decreased below 5.3, the enzyme

suddenly passes from a nearly native state which is sol-

uble to a completely aggregated state. On the other

hand, alkalic denaturation of DbjA is accompanied by

significant modification of both secondary and tertiary

structure. At pH conditions 10.3–11.5, the enzyme

occurs in disordered conformation and remains

soluble.

Temperature dependence of conformational stability

was evaluated by performing a thermal unfolding

experiment at different pH conditions. Dependence of

the melting temperature on pH was monitored by CD

spectroscopy at 222 nm (Fig. 2). All thermal transi-

tions obtained were irreversible, possibly because of

the aggregation phenomena in the denatured state

where visible aggregates were observed after heating of

the enzyme sample up to 80 �C. The pH dependence

exhibits a bell-shaped curve with the highest Tm

(49.1 ± 0.2 �C) at pH 6.5. A decrease in DbjA ther-

mostability at pH below 6.5 possibly corresponds to

the loss of tertiary interactions, as indicated by CD

spectra determined in the near UV spectral region. On

the other hand, the decrease in the enzyme thermosta-

bility at a pH above 6.5 could be attributed to the

changes in the protonation state of the enzyme, since

no changes in enzyme structure were observed in this

pH region. Generally, two major factors are known to

determine optimal pH for protein stability: amino acid

composition and tertiary structure [18]. In addition, we

suggest that quaternary structure can also influence

thermal stability of proteins.

Oligomerization

Analytical gel filtration was used to quantitatively

assess the effect of pH on the oligomerization state of

DbjA. Monomer, dimer, tetramer and high molecular

weight clusters were distinguished by enzyme elution

Fig. 2. pH-dependent dissociation, deactivation and denaturation of

DbjA: , melting temperature evaluated from measured changes in

ellipticity at 222 nm with increasing temperature; m, relative activity

(in %) representing the portion of the maximal detected specific

activity (lmolÆs)1Æmg)1) at a particular pH; , near UV CD at

259 nm; h, a-helical content calculated by the method of Chen

et al. [16] based on far UV CD at 222 nm.

Stereochemistry and conformational stability of DbjA R. Chaloupkova et al.

2730 FEBS Journal 278 (2011) 2728–2738 ª 2011 The Authors Journal compilation ª 2011 FEBS

volume at different pH conditions (Fig. 3). While the

pure monomer of DbjA is found under the lowest tested

pH conditions (pH 5.9), the dimeric form is a dominant

species at pH conditions equal to or higher than 6.1. As

pH increases, both dimeric and tetrameric forms are

present in solution. Abundance of the tetrameric form

gradually increases until it prevails at pH 9.6 (Fig. S1).

A high molecular weight cluster appears in solution as

another oligomeric form of DbjA at pH conditions

higher than 9.6. The presence of this cluster most prob-

ably corresponds to change in the conformation of the

enzyme detected by CD spectroscopy. In these alkaline

conditions, the DbjA enzyme occurs in a predominantly

unordered conformation which leads to association of

the enzyme to a high molecular weight cluster. Associa-

tion of oligomeric proteins at extreme conditions proba-

bly represents protection against aggregation.

These results demonstrate that oligomerization of

DbjA in solution strongly depends on the pH of the

surrounding environment. One of the major driving

forces for oligomerization comes from shape comple-

mentarity between the associating molecules, brought

about by a combination of hydrophobic and polar

interactions [19]. As determined by gel filtration, the

enzyme is monomeric at conditions close to its pI

(5.89). This suggests that the monomer is predomi-

nantly favoured at a pH where the net charge of the

enzyme is equal to zero. Under these conditions, all

oligomer-forming residues contribute to the overall

enzyme electronegativity via intramolecular interac-

tions and for that reason they do not contribute to the

formation of oligomers. As pH increases above pI, the

enzyme starts to be more and more negatively charged

and its oligomeric form is favoured. The enzyme

occurs in different ratios of dimeric and tetrameric

forms in the pH range 6.5–9.6. Under these conditions,

charged interface residues may establish intermolecular

interactions leading to the formation of a DbjA oligo-

mer. Crystallographic analysis [7] of the DbjA struc-

ture revealed that subunits of the dimer interact

predominantly in two regions: the C-terminal part of

the last helix (R292–P306) and the b-strand 8 region

(R269–L275).

As was evident from measured Tm at different pH

conditions, oligomeric states of DbjA obviously influ-

ence its thermal stability at different pH conditions.

The highest thermostability of the enzyme was detected

at pH 6.5, when the dimeric form predominates in

solution. With increasing occurrence of the tetrameric

form in solution, thermal stability of the enzyme

decreases. DbjA thermostability also slightly decreases

at pH 5.7–5.9, when the enzyme is monomeric. This

suggests that different forms of DbjA have different

thermostabilities: Tm (dimer) > Tm (monomer) > Tm

(tetramer). The stability of a high molecular weight

cluster present in solution above pH 9.6 is not dis-

cussed because its occurrence is accompanied by con-

formational changes which naturally lead to

destabilization of the protein structure.

Fig. 3. Gel filtration chromatograms of solutions with DbjA at dif-

ferent pH conditions. The peaks marked I, II, III and IV represent

monomer, dimer, tetramer and a high molecular weight cluster,

respectively. Molecular weight (MW) standards (Fig. S2) included

ribonuclease A (13.7 kDa, line 1), ovalbumin monomer (43.0 kDa,

line 2), albumin monomer (67.0 kDa, line 3), ovalbumin dimer

(86.0 kDa, line 4) and albumin dimer (134.0 kDa, line 5). Blue Dex-

tran (line 6) was used for determination of the dead volume of the

gel filtration column.

R. Chaloupkova et al. Stereochemistry and conformational stability of DbjA

FEBS Journal 278 (2011) 2728–2738 ª 2011 The Authors Journal compilation ª 2011 FEBS 2731

pH profile

Measurement of DbjA activity was performed to

explore whether catalytic function directly relates to

conformational stability at various pH values. Experi-

ments were done under different pH conditions and

saturated concentrations of substrate 1-iodohexane for

which DbjA exhibited the highest catalytic efficiency

[5]. The activity profile of this enzyme shows a maxi-

mum at pH 9.7 (Fig. 2). However, the enzyme retains

at least 90% of its maximum activity at pH conditions

ranging from 7.7 to 10.4. DbjA thus possesses the

broadest pH optimum compared with other biochemi-

cally characterized haloalkane dehalogenases (Fig. 4).

This phenomenon is most likely related to the fact that

DbjA occurs as oligomer. The melting temperatures of

DbjA detected at optimal pH represent only 79.1% of

maximal Tm. For this reason, the pH interval at which

the enzyme possesses the highest activity and the high-

est thermostability simultaneously is narrowed to

between pH 7.4 and 8.7 (Fig. 2). DbjA activity

decreases below pH 7.0 and above pH 10.4 with no

activity detected below pH 5.0 and above pH 11.0.

These results correlate well with the conformational

stability as a function of pH observed by CD spectros-

copy. The loss of enzymatic activity at highly alkalic

conditions is caused by change from native to predom-

inantly disordered conformation. The drop in activity

below pH 7.0 is not induced by the structural changes

but by change in the protonation state of catalytic

amino acids.

Catalytic residues of DbjA comprise five key resi-

dues forming the so-called catalytic pentad [1]. The

catalytic pentad of DbjA consists of three residues

involved in the catalytic reaction, Asp103, Glu127 and

His280, and two H-bond donating residues, Asn38 and

Trp104, involved in stabilization of a halogen group

of the substrate. With respect to particular dissociation

constants of catalytic residues, pKAspa = 3.90

(b-COOH), pKGlua = 4.07 (c-COOH), pKHis

a = 6.04

(imidazol) [20], it is evident that the residue affecting

the enzyme activity below pH 7.0 is His280. At pH

6.1, the enzyme retains 50% of its maximal activity

which nicely corresponds to pKHisa . Under these condi-

tions, 50% of histidine is protonated and thus non-

reactive and 50% is still reactive. The imidazol ring of

His becomes protonated and the enzyme loses its activ-

ity when the pH decreases further. Knowledge of the

pH interval at which the enzyme retains its structure

but loses most of its activity due to protonation of cat-

alytic histidine is interesting for further detailed deter-

mination of its catalytic mechanism. An alkyl–enzyme

intermediate can be captured by protein crystallogra-

phy at these pH conditions as has been previously

described for the haloalkane dehalogenase DhlA [21].

The effect of pH on enantioselectivity

The dependence of DbjA enantioselectivity on pH was

tested in a reaction with 2-bromopentane. Although

the effect of pH on enzyme enantioselectivity has

already been described for both charged [22] and

uncharged [23] substrates, in the case of DbjA no sig-

nificant change in enantioselectivity was observed at

pH values ranging from 6.7 to 10.1 (data not shown).

Results indicated that ionization of the alkyl–enzyme

intermediate is the same for both enantiomers at all

tested pH values and corresponds with the theoretical

Enzymes

pH

6 7 8 9 10

DhlAa

DhaAb

LinBc

DhmAd

DmbAe

DmbBe

DmbCf

DrbAf

DbjAg

Fig. 4. Comparison of the pH profiles of biochemically characterized haloalkane dehalogenases. Enzyme activity was quantified as the spe-

cific enzyme activity in units of lmolÆs)1Æmg)1 under conditions corresponding to initial velocity measurements. Black boxes represent maxi-

mal dehalogenating activity. Grey boxes represent retained dehalogenating activity at the level of at least 90% of the maximal enzymatic

activity. a DhlA from Xanthobacter autotrophicus GJ10 [24]; b DhaA from Rhodococcus sp. [25]; c LinB from Sphingobium japonicum UT26

[39]; d DhmA from Mycobacterium avium N85 [26]; e DmbA and DmbB from Mycobacterium bovis 5033 ⁄ 66 [27]; f DmbC from Mycobacte-

rium bovis 5033 ⁄ 66 and DrbA from Rhodopirellula baltica SH1 [28]; g DbjA from Bradyrhizobium japonicum USDA110, this study.

Stereochemistry and conformational stability of DbjA R. Chaloupkova et al.

2732 FEBS Journal 278 (2011) 2728–2738 ª 2011 The Authors Journal compilation ª 2011 FEBS

rule that pH dependence of stereoselectivity can only

be observed around the pK values of groups in the

active site whose ionization controls the enzyme activ-

ity [23]. Ionization of the catalytic His of DbjA could

be reflected at tested pH conditions, although this

effect on enantioselectivity was not observed. The pKa

values of other catalytic amino acids of DbjA, i.e.

nucleophile Asp and catalytic acid Glu, are lower than

the pH conditions at which the enzyme aggregates.

Temperature profile

Measurement of enzymatic activity at different temper-

atures was carried out to study the effect of tempera-

ture on the rate of the dehalogenation reaction. The

enzyme exhibited the highest activity at 50 �C,although above this temperature it became rapidly

inactivated. This observation is in good agreement

with similar experiments previously described for other

haloalkane dehalogenases possessing the highest activ-

ity at temperatures ranging from 35 to 50 �C [24–28].

Thermodynamic analysis of enantioselectivity

The temperature dependence of DbjA enantioselectivi-

ty was studied to determine differential activation

parameters, enthalpy (DR�SDHz) and entropy

(DR�SDSz), contributing to the kinetic resolution of

selected b-bromoalkanes (2-bromobutane and 2-brom-

opentane) and a-bromoesters (methyl 2-bromopropio-

nate and ethyl 2-bromobutyrate). The temperature

dependence of DbjA enantioselectivity was measured

in the temperature range from 20 to 50 �C. The E val-

ues and the thermodynamic components of enantiose-

lectivity determined based on the linear relation of

ln E and T )1 are summarized in Table 1. Although

the studied temperature interval was relatively small,

highly significant changes in DbjA enantioselectivity

were observed. Variation of the reaction temperature

from 20 to 50 �C caused a decrease in E value of DbjA

from 174 to 13 in the reaction with 2-bromopentane,

from 474 to 197 with ethyl-2-bromopropionate and

from 225 to 83 with methyl 2-bromobutyrate. Since

enzyme enantioselectivity is defined as the ratio of the

specificity constants for (R)) and (S)) enantiomers,

the E value does not depend on the degree of conver-

sion or variation of the reaction mechanism of individ-

ual enantiomers with temperature. It should be noted

that the enthalpic and the entropic components of dif-

ferential activation free energy (DR�SDGz) both con-

tribute to the overall success of the kinetic resolution

of enantiomers [29,30]. All substrates have a racemic

temperature significantly above the experimental tem-

perature indicating that the entropic component coun-

teracts the enthalpic component of enantiomeric

discrimination. The linearity between ln E and T )1

observed from 20 to 50 �C suggested that a single tran-

sition state structure is held in this temperature range

for all tested substrates.

Enantiomeric discrimination of 2-bromobutane was

not observed at any tested temperature (Fig. 5). This

Table 1. Thermodynamic components for the dehalogenation of selected halogenated compounds catalyzed by DbjA. Errors were calculated

from the standard errors of the linear regression ln E versus T)1. Tr is the racemic temperature at which no stereochemical discrimination of

the enzyme between the (R)) and (S)) enantiomers occurs, E = 1 and DR�SDGz = 0. It is defined by the ratio of the differential activation

enthalpy and entropy, Tr ¼ DR�SDHz=DR�SDSz, and is constant for a particular racemic substrate converted by a particular enzyme [29,31].

No enantioselectivity was observed for 2-bromobutane.

Substrate E, 298 K

DR�SDHz(kJÆmol)1)

DR�SDSz(JÆmol)1ÆK)1)

TDR�SDSz,298 K (kJÆmol)1)

DR�SDGz,298 K (kJÆmol)1) Tr (�C)

2-Bromobutane 1 – – – – –

2-Bromopentane 132 )69.5 ± 2.6 )193.8 ± 8.4 )57.8 ± 2.5 )11.7 86

Ethyl 2-bromopropionate 392 )24.1 ± 1.8 )31.2 ± 5.9 )9.3 ± 1.8 )14.8 497

Methyl 2-bromobutyrate 209 )25.8 ± 2.2 )42.3 ± 7.2 )12.6 ± 2.1 )13.2 337

Fig. 5. The temperature dependence of enantiomeric ratios

determined for dehalogenation of selected b-bromoalkanes (2-bromo-

butane, 2-bromopentane) and a-brominated esters (ethyl 2-bromopro-

pionate, methyl 2-bromobutyrate) catalyzed by DbjA.

R. Chaloupkova et al. Stereochemistry and conformational stability of DbjA

FEBS Journal 278 (2011) 2728–2738 ª 2011 The Authors Journal compilation ª 2011 FEBS 2733

result excludes the possibility that the absence of DbjA

enantioselectivity towards 2-bromobutane is due to the

fact that the initial E value was determined at a tem-

perature (20 �C) close to the racemic temperature for

this particular enzymatic resolution. If this were the

case, the enantioselectivity of DbjA could be increased

with increasing reaction temperature, changing also the

enantio-preference of the enzyme [31]. However, our

measurements confirm that the absence of 2-bromobu-

tane discrimination is the effect of zero DR�SDGz at

all tested temperatures. (R)) and (S)) enantiomers of

this simple chiral molecule are probably too similar

to each other to be kinetically recognized by the

enzyme. Surprisingly, adding a single carbon atom to

a substrate molecule provided enough structural dif-

ference for high enantiomeric discrimination as was

seen in the case of 2-bromopentane (Fig. 5). This

finding indicates the importance of the length of the

b-substituted bromo-n-alkanes for their kinetic resolu-

tion. The temperature dependence of DbjA enantiose-

lectivity for 2-bromopentane revealed that both

thermodynamic parameters, DR�SDHz and DR�SDSz,where the entropic term represents 83% of the enthal-

pic term, are important for enantiodiscrimination

(Table 1). The high contribution of entropy indicates

the importance of solvation, conformational degrees

of freedom of the protein, or restriction of substrate

motion in the transition state of the reaction. b-bro-moalkanes display high flexibility within the enzyme

active site which is related to the significant influence

of DR�SDSz for their kinetic resolution by the DbjA

enzyme. This implies that enantiomeric recognition of

b-bromoalkanes by DbjA is mediated by the differen-

tial conformational freedom of enantiomers upon

binding and ⁄or a displacement of a different number

of active site water molecules by the (R)) and (S))enantiomer [32,33].

The temperature dependence of DbjA enantioselec-

tivity with ethyl 2-bromopropionate and methyl 2-bro-

mobutyrate revealed that differential activation

enthalpy represents a major contribution to their

discrimination (Table 1). The high contribution of

enthalpy is related to differences in the complementar-

ity of each enantiomer in the transition state compris-

ing steric and electrostatic interactions between the

enzyme active site, its substrate and the solvent.

a-bromoesters obviously possess limited flexibility

inside the active site cavity due to their ability to form

an additional hydrogen bond of a carboxylic oxygen

with halide stabilizing residues. This implies that DbjA

enantioselectivity towards a-bromoesters is due to

different interactions of individual enantiomers with

the residues of the enzyme active site in the Michaelis

complex and ⁄or the transition state of the dehalogen-

ation reaction [34].

The thermodynamic analysis showed that DbjA

enantioselectivity towards b-bromoalkanes and

a-bromoesters is differently influenced by individual

thermodynamic contributions, differential activation

enthalpy and entropy. The resolution of b-bromoalk-

anes was found to be driven by both enthalpic and

entropic terms, while the resolution of a-bromoesters

was driven mainly by an enthalpic term. These results

correspond well with the proposal that enantioselectivi-

ty of DbjA with b-bromoalkanes and a-bromoesters is

based on two distinct molecular interactions [7].

Conclusions

Here we show that DbjA possesses unusually high

structural and functional stability towards a broad

range of pH conditions. Oligomerization of DbjA is

strongly pH dependent. Monomer, dimer, tetramer and

a high molecular weight cluster of the enzyme were

distinguished in solution at different pH conditions and

each oligomeric state demonstrated different stability.

The highest thermostability occurred at pH conditions

when the enzyme occurs in its dimeric form. Tempera-

ture significantly alters enantioselectivity, but an effect

of pH on DbjA enantiodicrimination was not observed.

Lowering the temperature results in considerable

enhancement of enantioselectivity. The results from

thermodynamic analysis are in good agreement with the

proposal that enantiomeric discrimination of b-bromi-

nated alkanes and a-brominated esters by DbjA is

controlled by distinct molecular interactions [7]. These

results indicate unique properties of DbjA compared

with other known and characterized members of haloal-

kane dehalogenases. Catalytic activity and structural

stability in a broad range of pH conditions combined

with high enantioselectivity with selected substrates

make DbjA a very versatile biocatalyst.

Experimental procedures

Enzyme preparation

The His-tagged DbjA was overexpressed in Escherichia coli

BL21 using a previously described method [5] and purified

using the HighTrap Chelating HP 5-mL column charged

with Ni2+ ions (GE Healthcare, Uppsala, Sweden). The

enzyme was bound to the resin in equilibrating buffer

(20 mM potassium phosphate buffer, pH 7.5, containing

0.5 M sodium chloride and 10 mM imidazole). Unbound

and weakly bound proteins were washed out with the buf-

fer containing 10 mM imidazole. The target enzyme was

Stereochemistry and conformational stability of DbjA R. Chaloupkova et al.

2734 FEBS Journal 278 (2011) 2728–2738 ª 2011 The Authors Journal compilation ª 2011 FEBS

eluted by a buffer containing 500 mM imidazole. The active

fractions were pooled and dialysed against a 50 mM potas-

sium phosphate buffer (pH 7.5). The enzyme was kept at

4 �C during the purification procedure and stored in 50 mM

phosphate buffer at 4 �C until use.

CD spectroscopy

CD spectra were recorded at room temperature (22 �C)using a Jasco J-810 spectrometer (Jasco, Tokyo, Japan). All

the spectra were obtained at an interval of 0.1 nm with a

scanning speed of 100 nmÆmin)1, 1 s response time and

2 nm bandwidth. Cuvettes of 0.1 and 1 cm path length

were used in the far and near UV regions, respectively. The

protein concentrations for the far UV and the near UV

spectra acquisition were 0.23 mgÆmL)1 and 1.15 mgÆmL)1,

respectively. Each spectrum shown is the average of 10 indi-

vidual scans and has been corrected for baseline noise. CD

spectra were expressed in millidegrees. The a-helical contentof the enzyme was calculated from the mean residue ellip-

ticity (MRE) value at 222 nm using the following equation

as described by Chen et al. [16]:

a -helix% ¼MRE222�234030300� 100

ð1Þ

Thermal denaturation

Thermal unfolding of DbjA was followed at different pH

conditions by monitoring the ellipticity at 222 nm over the

temperature range 20–80 �C, with a resolution 0.2 �C, at aheating rate 0.5 �CÆmin)1. Recorded thermal denaturation

curves were roughly normalized to represent signal changes

between � 1 and 0, and fitted to sigmoidal curves using

software ORIGIN 6.1 (OriginLab, Northampton, MA, USA).

The melting temperatures (Tm) were evaluated as the mid-

point of the normalized thermal transition.

Prediction of the isoelectric point

The theoretical isoelectric point (pI) of DbjA was predicted

based on the amino acid sequence by using EXPASY SERVER

[35–37].

Effect of pH

DbjA activity and enantioselectivity were measured at dif-

ferent pH conditions. Britton–Robinson buffer solutions

were used to cover the pH range 1.7–11.5. The solutions

were prepared by mixing 0.04 M phosphoric, boric and ace-

tic acid with the appropriate volume of sodium hydroxide

(0.2 M) and sodium perchlorate monohydrate to get a con-

stant ionic strength of 0.15 M. The assays were performed

with 1-iodohexane as the substrate for activity measurement

at 37 �C or 2-bromopentane as the substrate for enantiose-

lectivity measurement at 25 �C.

Effect of temperature

The effect of temperature on DbjA activity and enantiose-

lectivity was determined by performing activity and enanti-

oselectivity assays at different temperatures. The activity

measurements were evaluated at temperatures ranging from

20 to 60 �C and the enantioselectivity of the DbjA enzyme

was monitored in the temperature range 20–50 �C, both in

50 mM glycin buffer at pH 8.6. Activity measurements were

performed with 1-iodohexane, and enantioselectivity mea-

surements with 2-bromobutane, 2-bromopentane, methyl

2-bromopropionate and ethyl 2-bromobutyrate.

Gel filtration chromatography

The molecular mass of DbjA enzyme at different pH condi-

tions was analysed using the FPLC system AKTA (GE

Healthcare) equipped with UV280 detection (GE Healthcare,

Uppsala, Sweden) and Superdex� 200 10 ⁄ 300 GL column

(GE Healthcare, Uppsala, Sweden). A total volume of

100 lL of each protein sample was applied to the column

and separated at a constant flow rate of 0.5 mLÆmin)1.

Britton–Robinson buffer with an appropriate pH value was

used as the mobile phase. The molecular weight standards

from the Gel Filtration Calibration Kit (GE Healthcare,

Uppsala, Sweden) included ribonuclease A (13.7 kDa), oval-

bumin monomer (43.0 kDa), albumin monomer (67.0 kDa),

ovalbumin dimer (86.0 kDa) and albumin dimer (134.0 kDa).

The dead volume of the Superdex� 200 10 ⁄ 300 GL column

was determined using the Blue Dextran of the calibration kit.

All protein standards as well as enzyme samples were trans-

ferred into the Britton–Robinson buffer by using a 5-mL

HighTrap Desalting Sephadex G-25 Superfine column (GE

Healthcare, Uppsala, Sweden).

Activity assay

DbjA activity was assayed by the colorimetric method

developed by Iwasaki et al. [38]. The halide ions released

were analysed after a reaction with mercuric thiocyanate

and ferric ammonium sulfate spectrophotometrically at

460 nm using the Sunrise microplate reader (Tecan,

Grodig ⁄ Salzburg, Austria). The dehalogenation reaction

was performed in 25-mL Reacti flasks closed by Miniert

valves at various temperatures. The reaction mixture was

composed of 15 mL of buffer and 2 lL of substrate 1-iod-

ohexane. The reaction was initiated by the addition of

enzyme in a final concentration of 0.15 lM. The reaction

was monitored by withdrawing 1 mL samples at 10, 20,

30, 40, 50 and 60 min from the reaction mixture. The reac-

tion mixture samples were immediately mixed with 0.1 mL

35% nitric acid to terminate the reaction. Dehalogenation

activity was quantified as a rate of product formation in

time. Each activity was measured in three to five indepen-

dent replicates and represented as mean values of relative

R. Chaloupkova et al. Stereochemistry and conformational stability of DbjA

FEBS Journal 278 (2011) 2728–2738 ª 2011 The Authors Journal compilation ª 2011 FEBS 2735

activity with plotted standard errors. Relative activities

represented a percentage of maximal specific activity

detected.

Enantioselectivity assay

Enantioselectivity was analysed in 25-mL Reacti flasks

closed by Miniert valves containing 20 mL of glycin buffer

(100 mM, pH 8.6). Chiral substrates were added to a final

concentration of 0.5–3.0 mM with regard to enzyme affinity.

The enzymatic reaction was initiated by the addition of

appropriate amounts of the DbjA enzyme depending on

enzyme activity (final concentration 0.2–2.0 lM). The reac-

tion was monitored by periodical withdrawing of 0.5 mL

sample aliquots from the reaction mixture. The reaction

was stopped by mixing the sample with 1 mL of diethyl

ether containing 1,2-dichloroethane as an internal standard.

After extraction, diethyl ether was anhydrated on a glass

column with sodium sulphate. The samples were automati-

cally analysed by using Hewlett-Packard 6890 gas chro-

matograph (Agilent, Santa Clara, USA) equipped with a

flame ionization detector and chiral capillary column Chi-

raldex B-TA and Chiraldex G-TA (Alltech, Deerfield,

USA). Michaelis–Menten parameters were derived by fitting

the progress curves obtained from kinetic resolution experi-

ments into a competitive kinetic pattern by numerical

integration using the software MICROMATH SCIENTIST

(ChemSW, Fairfield, USA). Enantioselectivity was deter-

mined as the enantiomeric ratio (E) defined by

E ¼ kRcat=KRm

kScat=KSm

ð2Þ

where kcat and Km represent the Michaelis–Menten parame-

ters of the two enantiomers.

Thermodynamic analysis

The difference in activation enthalpy and entropy between

enantiomers was determined by studying the variation of

the enzyme enantiomeric ratio with temperature:

lnE ¼ �DR�SDHzR

� 1

Tþ DR�SDSz

Rð3Þ

The enantiomeric ratio (or rather lnE) varied with recipro-

cal temperature to an extent determined by the enthalpic

term (the slope of Eqn 3, DR�SDHz ⁄R), at a level deter-

mined by the entropic term (the intercept of Eqn 3,

DR�SDSz ⁄R). A racemic temperature (Tr) was determined

as the ratio of the differential activation enthalpy and

entropy:

Tr ¼DR�SDHz

DR�SDSzð4Þ

Acknowledgements

This work was financially supported by the Grant

Agency of the Czech Academy of Sciences

(IAA401630901 to J.D.), the Czech Ministry of Educa-

tion (MSM0021622412 and LC06010 to J.D.), the

Grant Agency of the Czech Republic (203 ⁄ 08 ⁄ 0114 to

R.Ch.) and the European Regional Development Fund

(project FNUSA-ICRC no. CZ.1.05 ⁄ 1.1.00 ⁄ 02.0123 to

Z.P.). The authors thank Eva Chovancova for the pre-

diction of DbjA quaternary structure and Monika

Strakova for assistance with protein expression and

purification.

References

1 Janssen DB (2004) Evolving haloalkane dehalogenases.

Curr Opin Chem Biol 8, 150–159.

2 Fetzner S & Lingens F (1994) Bacterial dehalogenases:

biochemistry, genetics, and biotechnological applica-

tions. Microbiol Rev 58, 641–685.

3 Patel RN (2004) Biocatalytic synthesis of chiral phar-

maceutical intermediates. Food Technol Biotechnol 42,

305–325.

4 Patel RN (2006) Biocatalysis: synthesis of chiral inter-

mediates for drugs. Curr Opin Drug Discov Devel 9,

741–764.

5 Sato Y, Monincova M, Chaloupkova R, Prokop Z,

Ohtsubo Y, Minamisawa K, Tsuda M, Damborsky J &

Nagata Y (2005) Two rhizobial strains, Mesorhizobi-

um loti MAFF303099 and Bradyrhizobium japonicum

USDA110, encode haloalkane dehalogenases with novel

structures and substrate specificities. Appl Environ

Microbiol 71, 4372–4379.

6 Prokop Z, Damborsky J, Nagata Y & Janssen DB

(2009) Method of production of optically active

halohydrocarbons and alcohols using hydrolytic

dehalogenation catalysed by haloalkane dehalogenases.

US7, 632, 666.

7 Prokop Z, Sato Y, Brezovsky J, Mozga T, Chaloupkova

R, Koudelakova T, Jerabek P, Stepankova V, Natsume

R, Leeuwen JGE et al. (2010) Enantioselectivity of

haloalkane dehalogenases and its modulation by

surface loop engineering. Angew Chem Int Ed 49,

6111–6115.

8 Schoemaker HE, Mink D & Wubbolts MG (2003) Dis-

pelling the myths – biocatalysis in industrial synthesis.

Science 299, 1694–1697.

9 Bornscheuer UT & Pohl M (2001) Improved biocata-

lysts by directed evolution and rational protein design.

Curr Opin Chem Biol 5, 137–143.

10 Pollard DJ & Woodley JM (2007) Biocatalysis for

pharmaceutical intermediates: the future is now. Trends

Biotechnol 25, 66–73.

Stereochemistry and conformational stability of DbjA R. Chaloupkova et al.

2736 FEBS Journal 278 (2011) 2728–2738 ª 2011 The Authors Journal compilation ª 2011 FEBS

11 Woodley JM (2008) New opportunities for biocatalysis:

making pharmaceutical processes greener. Trends

Biotechnol 26, 321–327.

12 Jaeger KE & Eggert T (2004) Enantioselective biocatal-

ysis optimized by directed evolution. Curr Opin Biotech-

nol 15, 305–313.

13 Kazlauskas RJ (2005) Enhancing catalytic promiscuity

for biocatalysis. Curr Opin Chem Biol 9, 195–201.

14 Bornscheuer UT (2005) Trends and challenges in enzyme

technology. Adv Biochem Eng Biotechnol 100, 181–203.

15 Hult K & Berglund P (2003) Engineered enzymes for

improved organic synthesis. Curr Opin Biotech 14,

395–400.

16 Chen Y-H, Yang JT & Martinez HM (1972) Determi-

nation of the secondary structures of proteins by

circular dichroism and optical rotatory dispersion.

Biochemistry 11, 4120–4131.

17 Fasman GD (1996) Circular Dichroism and Conforma-

tional Analysis of Biomolecules. Plenum Press,

New York, NY.

18 Alexov E (2004) Numerical calculation of the pH of

maximal protein stability. Eur J Biochem 271, 173–185.

19 Shallom D, Golan G, Shoham G & Shoham Y (2004)

Effect of dimer dissociation on activity and thermostabil-

ity of the a-glucuronidase from Geobacillus stearothermo-

philus: dissecting the differnt oligomeric forms of family

67 glycoside hydrolases. J Bacteriol 186, 6928–6937.

20 Fersht A (1999) Structure and Mechanism in Protein

Science: A Guide to Enzyme Catalysis and Protein Fold-

ing. W.H. Freeman, New York, NY.

21 Verschueren KHG, Seljee F, Rozeboom HJ, Kalk KH

& Dijkstra BW (1993) Crystallographic analysis of the

catalytic mechanism of haloalkane dehalogenase. Nature

363, 693–698.

22 Lummer K, Rieks A, Galunsky B & Kasche V (1999)

pH dependence of penicillin amidase enantioselectivity

for charged substrates. Biochim Biophys Acta 1433,

327–334.

23 Secundo F & Phillips RS (1996) Effects of pH on enan-

tiospecificity of alcohol dehydrogenases from Thermo-

anaerobacter ethanolicus and horse liver. Enzyme

Microb Technol 19, 487–492.

24 Keuning S, Janssen DB & Witholt B (1985) Purification

and characterization of hydrolytic haloalkane dehalo-

genase from Xanthobacter autotrophicus GJ10. J Bacte-

riol 163, 635–639.

25 Yokota T, Omori T & Kodama T (1987) Purification

and properties of haloalkane dehalogenase from Coryne-

bacterium sp. strain m15-3. J Bacteriol 169, 4049–4054.

26 Jesenska A, Bartos M, Czernekova V, Rychlik I, Pavlik

I & Damborsky J (2002) Cloning and expression of

the haloalkane dehalogenase gene dhmA from

Mycobacterium avium N85 and preliminary character-

ization of DhmA. Appl Environ Microbiol 68,

3724–3730.

27 Jesenska A, Pavlova M, Strouhal M, Chaloupkova R,

Tesinska I, Monincova M, Prokop Z, Bartos M, Pavlik

I, Rychlik I et al. (2005) Mycobacterial haloalkane

dehalogenases: cloning, biochemical properties and

distribution. Appl Environ Microbiol 71, 6736–6745.

28 Jesenska A, Monincova M, Chrobakova T, Hasan K,

Chaloupkova R, Prokop Z, Geerlof A & Damborsky J

(2009) Isolation and biochemical characterization of

haloalkane dehalogenases DrbA and DmbC: representa-

tives of novel subfamily. Appl Environ Microbiol 75,

5157–5160.

29 Phillips RS (1996) Temperature modulation of the

stereochemistry of enzymatic catalysis: prospects for

exploitation. Trends Biotechnol 14, 13–16.

30 Ottosson J, Fransson L & Hult K (2002) Substrate

entropy in enzyme enantioselectivity: an experimental

and molecular modeling study. Protein Sci 11, 1462–

1471.

31 Pham VT, Phillips RS & Ljungdahl LG (1989) Temper-

ature-dependent enantiospecificity of secondary alcohol

dehydrogenase from Thermoanaerobacter ethanolicus.

J Am Chem Soc 111, 1935–1936.

32 Overbeeke PL, Orrenius C, Jongejan JA & Duine JA

(1998) Enthalpic and entropic contributions to lipase

enantioselectivity. Chem Phys Lipids 93, 81–93.

33 Phillips RS (2002) How does active site water affect

enzymatic stereorecognition? J Mol Catal B Enzym

19–20, 103–107.

34 Galunsky B, Ignatova S & Kasche V (1997) Tempera-

ture effects on S1- and S’1-enantioselectivity of

alpha-chymotrypsin. Biochim Biophys Acta 1343,

130–138.

35 Bjellqvist B, Hughes GJ, Pasquali C, Paquet N, Ravier

F, Sanchez J-C, Frutiger S & Hochstrasser DF (1993)

The focusing positions of polypeptides in immobilized

pH gradients can be predicted from their amino acid

sequences. Electrophoresis 14, 1023–1031.

36 Bjellqvist B, Basse B, Olsen E & Celis JE (1994)

Reference points for comparisons of two-dimensional

maps of proteins from different human cell types

defined in a pH scale where isoelectric points correlate

with polypeptide compositions. Electrophoresis 15, 529–

539.

37 Gasteiger E, Hoogland C, Gattiker A, Duvaud S,

Wilkins MR, Appel RD & Bairoch A (2005) The Prote-

omics Protocols Handbook. Humana Press, Totowa, NJ.

38 Iwasaki I, Utsumi S & Ozawa T (1952) New colorimet-

ric determination of chloride using mercuric thiocyanate

and ferric ion. Bull Chem Soc Jpn 25, 226.

39 Nagata Y, Miyauchi K, Damborsky J, Manova K,

Ansorgova A & Takagi M (1997) Purification and

characterization of haloalkane dehalogenase of a new

substrate class from a c-hexachlorocyclohexane-degrad-ing bacterium, Sphingomonas paucimobilis UT26. Appl

Environ Microbiol 63, 3707–3710.

R. Chaloupkova et al. Stereochemistry and conformational stability of DbjA

FEBS Journal 278 (2011) 2728–2738 ª 2011 The Authors Journal compilation ª 2011 FEBS 2737

Supporting information

The following supplementary material is available:

Fig. S1. Distribution of various forms of DbjA in

solution at different pH conditions.

Fig. S2. Gel filtration chromatogram of ribonuclease A

(13.7 kDa, line 1), ovalbumin monomer (43.0 kDa, line

2), albumin monomer (67.0 kDa, line 3), ovalbumin

dimer (86.0 kDa, line 4) and albumin dimer (134.0 kDa,

line 5) used as molecular weight standards. Blue Dex-

tran (line 6) was used for determination of dead volume

of the gel filtration column.

This supplementary material can be found in the

online version of this article.

Please note: As a service to our authors and readers,

this journal provides supporting information supplied

by the authors. Such materials are peer-reviewed and

may be reorganized for online delivery, but are not

copy-edited or typeset. Technical support issues arising

from supporting information (other than missing files)

should be addressed to the authors.

Stereochemistry and conformational stability of DbjA R. Chaloupkova et al.

2738 FEBS Journal 278 (2011) 2728–2738 ª 2011 The Authors Journal compilation ª 2011 FEBS