High order quaternary arrangement confers increased structural stability to Brucella sp. lumazine...

10
High Order Quaternary Arrangement Confers Increased Structural Stability to Brucella sp. Lumazine Synthase* Received for publication, November 3, 2003, and in revised form, November 26, 2003 Published, JBC Papers in Press, December 1, 2003, DOI 10.1074/jbc.M312035200 Vanesa Zylberman‡§, Patricio O. Craig‡, Sebastia ´ n Klinke‡§, Bradford C. Braden, Ana Cauerhff‡, and Fernando A. Goldbaum‡** From the Instituto Leloir, Consejo Nacional de Investigaciones Cientı´ficas y Te ´cnicas and Facultad de Ciencias Exactas y Naturales, Universidad de Buenos Aires, Patricias Argentinas 435 (C1405BWE), Buenos Aires, Argentina and the Department of Natural Sciences, Bowie State University, Bowie, Maryland 20715 The penultimate step in the pathway of riboflavin bio- synthesis is catalyzed by the enzyme lumazine synthase (LS). One of the most distinctive characteristics of this enzyme is the structural quaternary divergence found in different species. The protein exists as pentameric and icosahedral forms, built from practically the same struc- tural monomeric unit. The pentameric structure is formed by five 18-kDa monomers, each extensively con- tacting neighboring monomers. The icosahedrical struc- ture consists of 60 LS monomers arranged as 12 pentam- ers giving rise to a capsid exhibiting icosahedral 532 symmetry. In all lumazine synthases studied, the topolog- ically equivalent active sites are located at the interfaces between adjacent subunits in the pentameric modules. The Brucella sp. lumazine synthase (BLS) sequence clearly diverges from pentameric and icosahedric en- zymes. This unusual divergence prompted us to further investigate its quaternary arrangement. In the present work, we demonstrate by means of solution light scatter- ing and x-ray structural analyses that BLS assembles as a very stable dimer of pentamers, representing a third cat- egory of quaternary assembly for lumazine synthases. We also describe by spectroscopic studies the thermo- dynamic stability of this oligomeric protein and postulate a mechanism for dissociation/unfolding of this macromo- lecular assembly. The higher molecular order of BLS in- creases its stability 20 °C compared with pentameric lum- azine synthases. The decameric arrangement described in this work highlights the importance of quaternary in- teractions in the stabilization of proteins. Riboflavin, an essential cofactor for all organisms, is biosyn- thesized in plants, fungi, and microorganisms. The penulti- mate step in the pathway is catalyzed by the enzyme lumazine synthase (LS). 1 One of the most distinctive characteristics of this enzyme is the structural quaternary divergence found in different species. The protein exists as pentameric and icosa- hedral forms, built from practically the same structural mono- meric unit. The structure of the monomer consists of four repeated -strand/-helix motifs producing a sandwich of four parallel -strands surrounded by four -helices, two on each face of the sheet (1). In all LS studied, the topologically equivalent active sites are located at the interfaces between adjacent subunits in the pentameric modules (2). Bacilliaceae express a bifunctional enzyme complex with lumazine synthase and riboflavin synthase activity. Three -subunits (riboflavin synthase) enclosed by 60 -subunits (lumazine synthase) form a protein particle of 1 MDa (1, 3). The three-dimensional structure of the lumazine synthase/ri- boflavin synthase from Bacillus subtilis complexed with a sub- strate analogue has been determined (4). This structure con- sists of 60 -subunits (lumazine synthase monomers) arranged as 12 pentamers giving rise to a capsid exhibiting icosahedral 532 symmetry. Two other icosahedric LS have been described by x-ray crystallography. Spinach and thermophilic bacterial Aquifex aelicus LS also exhibit icosahedral 532 symmetry. The three proteins form 1-MDa spherical capsids of 60 LS subunits with icosahedral symmetry. The icosahedral LS structures su- perimpose very well, highlighting the conservation of the over- all folding and quaternary arrangement despite the low se- quence homology between them (5). Four structurally characterized LS assemble as pentamers in their native form and do not further associate to form an icosahedral capsid. These include fungal Magnaporthe grisea, yeast Saccharomyces cerevisiae, and Schizosaccharomyces pombe and bacterial Brucella abortus LS (5). Although the four pentameric enzymes fold in a similar arrangement, the postu- lated reasons for the lack of icosahedral order differ. Superpo- sition of the pentameric LS shows that the loop connecting the helices 4 and 5 is critical for preventing the formation of capsids. In all icosahedral LS a pentapeptide kink located in this loop is essential for pentamer-pentamer contacts (5, 6). The yeast and fungal pentameric enzymes have insertions of different length in this loop that change its overall orientation and disrupt potential contacts to neighboring subunits (7). We have previously analyzed the divergence in macromolecular assembly between pentameric and icosahedral LS enzymes. In this regard, a high degree of divergence was observed between the sequence from B. abortus LS and the other pentameric structurally characterized members of the family (8). BLS has a 3-residue insertion between helices 4 and 5 that contribute to form a continuous undistorted helix, unable to form the * This work was supported in part by a Howard Hughes Medical Institute international grant (to F. A. G.), by a grant from the Agencia Nacional de Promocio ´n Cientı ´fica y Tecnolo ´gica, Repu ´ blica Argentina, and by the Fundacio ´n Instituto Leloir. The costs of publication of this article were defrayed in part by the payment of page charges. This article must therefore be hereby marked “advertisement” in accordance with 18 U.S.C. Section 1734 solely to indicate this fact. § Recipients of a fellowship from CONICET. Members of the research career of Consejo Nacional de Investiga- ciones Cientı ´ficas y Te ´cnicas (CONICET). ** To whom correspondence should be addressed. Tel: 54-11-4863- 4012 (ext. 19); Fax: 54-11-4865-2246; E-mail: [email protected]. 1 The abbreviations used are: LS, lumazine synthase; BLS, Brucella sp. lumazine synthase; CD, circular dichroism; TF, tryptophan fluores- cence; ANS, 1-anilino-8-naphthalene sulfonic acid; SLS, static light scattering; SEC, size exclusion chromatography column; LEM, lineal extrapolation method; GdnHCl, guanidine hydrochloride; M w , weight average molecular weight; DTT, dithiothreitol; ASA, accessible surface area. THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 279, No. 9, Issue of February 27, pp. 8093–8101, 2004 © 2004 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A. This paper is available on line at http://www.jbc.org 8093 by guest on April 2, 2016 http://www.jbc.org/ Downloaded from

Transcript of High order quaternary arrangement confers increased structural stability to Brucella sp. lumazine...

High Order Quaternary Arrangement Confers Increased StructuralStability to Brucella sp. Lumazine Synthase*

Received for publication, November 3, 2003, and in revised form, November 26, 2003Published, JBC Papers in Press, December 1, 2003, DOI 10.1074/jbc.M312035200

Vanesa Zylberman‡§, Patricio O. Craig‡, Sebastian Klinke‡§, Bradford C. Braden¶,Ana Cauerhff‡�, and Fernando A. Goldbaum‡�**

From the ‡Instituto Leloir, Consejo Nacional de Investigaciones Cientıficas y Tecnicas and Facultad de Ciencias Exactas yNaturales, Universidad de Buenos Aires, Patricias Argentinas 435 (C1405BWE), Buenos Aires, Argentina and ¶theDepartment of Natural Sciences, Bowie State University, Bowie, Maryland 20715

The penultimate step in the pathway of riboflavin bio-synthesis is catalyzed by the enzyme lumazine synthase(LS). One of the most distinctive characteristics of thisenzyme is the structural quaternary divergence found indifferent species. The protein exists as pentameric andicosahedral forms, built from practically the same struc-tural monomeric unit. The pentameric structure isformed by five 18-kDa monomers, each extensively con-tacting neighboring monomers. The icosahedrical struc-ture consists of 60 LS monomers arranged as 12 pentam-ers giving rise to a capsid exhibiting icosahedral 532symmetry. In all lumazine synthases studied, the topolog-ically equivalent active sites are located at the interfacesbetween adjacent subunits in the pentameric modules.The Brucella sp. lumazine synthase (BLS) sequenceclearly diverges from pentameric and icosahedric en-zymes. This unusual divergence prompted us to furtherinvestigate its quaternary arrangement. In the presentwork, we demonstrate by means of solution light scatter-ing and x-ray structural analyses that BLS assembles as avery stable dimer of pentamers, representing a third cat-egory of quaternary assembly for lumazine synthases. Wealso describe by spectroscopic studies the thermo-dynamic stability of this oligomeric protein and postulatea mechanism for dissociation/unfolding of this macromo-lecular assembly. The higher molecular order of BLS in-creases its stability 20 °C compared with pentameric lum-azine synthases. The decameric arrangement describedin this work highlights the importance of quaternary in-teractions in the stabilization of proteins.

Riboflavin, an essential cofactor for all organisms, is biosyn-thesized in plants, fungi, and microorganisms. The penulti-mate step in the pathway is catalyzed by the enzyme lumazinesynthase (LS).1 One of the most distinctive characteristics ofthis enzyme is the structural quaternary divergence found in

different species. The protein exists as pentameric and icosa-hedral forms, built from practically the same structural mono-meric unit. The structure of the monomer consists of fourrepeated �-strand/�-helix motifs producing a sandwich of fourparallel �-strands surrounded by four �-helices, two on eachface of the � sheet (1). In all LS studied, the topologicallyequivalent active sites are located at the interfaces betweenadjacent subunits in the pentameric modules (2).

Bacilliaceae express a bifunctional enzyme complex withlumazine synthase and riboflavin synthase activity. Three�-subunits (riboflavin synthase) enclosed by 60 �-subunits(lumazine synthase) form a protein particle of �1 MDa (1, 3).The three-dimensional structure of the lumazine synthase/ri-boflavin synthase from Bacillus subtilis complexed with a sub-strate analogue has been determined (4). This structure con-sists of 60 �-subunits (lumazine synthase monomers) arrangedas 12 pentamers giving rise to a capsid exhibiting icosahedral532 symmetry. Two other icosahedric LS have been describedby x-ray crystallography. Spinach and thermophilic bacterialAquifex aelicus LS also exhibit icosahedral 532 symmetry. Thethree proteins form 1-MDa spherical capsids of 60 LS subunitswith icosahedral symmetry. The icosahedral LS structures su-perimpose very well, highlighting the conservation of the over-all folding and quaternary arrangement despite the low se-quence homology between them (5).

Four structurally characterized LS assemble as pentamersin their native form and do not further associate to form anicosahedral capsid. These include fungal Magnaporthe grisea,yeast Saccharomyces cerevisiae, and Schizosaccharomycespombe and bacterial Brucella abortus LS (5). Although the fourpentameric enzymes fold in a similar arrangement, the postu-lated reasons for the lack of icosahedral order differ. Superpo-sition of the pentameric LS shows that the loop connecting thehelices �4 and �5 is critical for preventing the formation ofcapsids. In all icosahedral LS a pentapeptide kink located inthis loop is essential for pentamer-pentamer contacts (5, 6).The yeast and fungal pentameric enzymes have insertions ofdifferent length in this loop that change its overall orientationand disrupt potential contacts to neighboring subunits (7). Wehave previously analyzed the divergence in macromolecularassembly between pentameric and icosahedral LS enzymes. Inthis regard, a high degree of divergence was observed betweenthe sequence from B. abortus LS and the other pentamericstructurally characterized members of the family (8). BLS hasa 3-residue insertion between helices �4 and �5 that contributeto form a continuous undistorted helix, unable to form the

* This work was supported in part by a Howard Hughes MedicalInstitute international grant (to F. A. G.), by a grant from the AgenciaNacional de Promocion Cientıfica y Tecnologica, Republica Argentina,and by the Fundacion Instituto Leloir. The costs of publication of thisarticle were defrayed in part by the payment of page charges. Thisarticle must therefore be hereby marked “advertisement” in accordancewith 18 U.S.C. Section 1734 solely to indicate this fact.

§ Recipients of a fellowship from CONICET.� Members of the research career of Consejo Nacional de Investiga-

ciones Cientıficas y Tecnicas (CONICET).** To whom correspondence should be addressed. Tel: 54-11-4863-

4012 (ext. 19); Fax: 54-11-4865-2246; E-mail: [email protected] The abbreviations used are: LS, lumazine synthase; BLS, Brucella

sp. lumazine synthase; CD, circular dichroism; TF, tryptophan fluores-cence; ANS, 1-anilino-8-naphthalene sulfonic acid; SLS, static lightscattering; SEC, size exclusion chromatography column; LEM, lineal

extrapolation method; GdnHCl, guanidine hydrochloride; Mw, weightaverage molecular weight; DTT, dithiothreitol; ASA, accessible surfacearea.

THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 279, No. 9, Issue of February 27, pp. 8093–8101, 2004© 2004 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A.

This paper is available on line at http://www.jbc.org 8093

by guest on April 2, 2016

http://ww

w.jbc.org/

Dow

nloaded from

capsid-stabilizing kink (6). Thus, the BLS structure clearlydiverges from pentameric and icosahedric enzymes because ofthe lack of this critical loop. The different orientation of thisstraight helix is compensated by a longer loop bridging thehelix structure with the contiguous sheet. This unusual diver-gence prompted us to further investigate the quaternary ar-rangement of BLS. In the present work, we demonstrate bymeans of solution light scattering and x-ray structural analysisthat BLS assembles as a very stable dimer of pentamers, rep-resenting a third category of quaternary assembly for LS. Wealso describe, by spectroscopic studies, the thermodynamic sta-bility of this oligomeric protein and postulate a mechanism fordissociation/unfolding of its macromolecular assembly.

EXPERIMENTAL PROCEDURES

Expression of the LS Protein

The Brucella sp. LS gene was cloned in pET11a vector (Novagen) asreported previously (9). The plasmid was used to transform BL21(DE3)strain Escherichia coli-competent cells (Stratagene, La Jolla, CA). Am-picillin-resistant colonies were grown until A600 � 1.0 in LB mediumcontaining 100 �g/ml ampicillin, at 37 °C with agitation (300 rpm). Fivemilliliters of this culture was diluted to 500 ml and grown to reach anA600 of 1.0. At this point the culture was induced adding 1 mM isopropyl-1-thio-�-D-galactopyranoside and incubated for 4 h at 37 °C with agita-tion (300 rpm). The bacteria were centrifuged at 15,000 � g during 20min at 4 °C.

Protein Purification and Refolding

BLS protein was successfully expressed as inclusion bodies by trans-formation of strain BL21(DE3) E. coli-competent cells (9). The inclusionbodies were solubilized in 50 mM Tris, 5 mM EDTA, 8 M urea pH 8.0 atroom temperature overnight with agitation. The solubilized materialwas refolded by dialysis during 72 h against phosphate-buffered salinecontaining 1 mM dithiothreitol. This preparation was purified in aMono-Q column in a fast-protein liquid chromatography apparatus(Amersham Biosciences, Uppsala, Sweden) using a linear gradient ofbuffer B (50 mM Tris, 1 M NaCl, pH 8.5). The peak enriched with BLSwas further purified on a Superdex-200 column with phosphate-buff-ered saline buffer, 1 mM DTT. The purity of the BLS preparation wasdetermined on SDS-15% (w/v) polyacrylamide gels. Purified BLS wasconcentrated (10 mg/ml), frozen in liquid N2, and stored at �20 °C.

Circular Dichroism

BLS samples were diluted in 50 mM sodium phosphate, pH 7.0, 1 mM

DTT, with increasing concentration of denaturants (urea or GdnHCl).All experiments were performed at 25 °C, and samples were incubatedat least 2 h before taking CD measurements. Spectra were measured ona spectropolarimeter (JASCO J-810) using either 0.1- or 0.5-cm pathlength quartz cells. Unfolding was monitored by far-UV CD (260–200nm) and expressed as the percentage of molar ellipticity at 222 nm as afunction of denaturant concentration. The molar ellipticity of the pro-tein incubated without GdnHCl was taken as 100%. For renaturationtests, BLS was first denatured in high concentration of GdnHCl (6 M),and renaturation was induced by overnight dialysis against 50 mM

sodium phosphate, pH 7.0, 1 mM DTT. After this treatment, the CD andSLS signals typical of the native protein were recovered.

Intrinsic Fluorescence Measurements

Emission spectra were carried out by excitation of the samples at 295nm, and data collection from 300 to 450 nm, using 3-nm band passes forboth excitation and emission. Experiments were carried out in 50 mM

sodium phosphate, pH 7.0, 1 mM DTT in the presence of increasingconcentrations of urea or GdnHCl. Samples were incubated for at least 2 hprior to taking fluorescence measurements. All fluorescence emissionspectra were measured at 25 °C on a Jasco FP-770 spectrofluorometer.

Thermal Denaturation Monitored by CD

BLS samples were incubated in 50 mM sodium phosphate, 1 mM DTT,pH 7.0 (in the presence or absence of 2 M GdnHCl) or in 50 mM buffercitrate, 1 mM DTT, pH 4.5. Thermal denaturation was conducted byslowly increasing the temperature with a Peltier system (Jasco). Therange of temperature scanning was 25–95 °C at a speed of 4 °C/min.Molar ellipticity at 220 nm was measured every 0.5 °C. Fast or slowcooling back to 25 °C (from 95 to 25 °C at a speed of 1 °C/min) did notshow a recovery of ellipticity demonstrating the irreversibility of the

thermal unfolding. Thus the temperature midpoint of the thermal tran-sition was considered as an apparent Tm.

pH Dependence of LS Unfolding

pH-induced BLS dissociation and unfolding was evaluated by dilut-ing protein samples (5 �M monomers) in 50 mM citrate, 1 mM DTT, pH6.0–2.5 or in 50 mM sodium phosphate, 1 mM DTT, pH 7.5- 7.0. After1–2 h of incubation at room temperature, CD and SLS signals weredetermined. ANS fluorescence emission was measured in the samebuffer conditions as CD and SLS experiments, aggregating to the sam-ples 50 �M ANS. Values were normalized to ANS fluorescence emissionin the same buffer. Excitation and emission were set at 365 and 470nm, respectively.

Determination of Molecular Weight of BLS by StaticLight Scattering

The weight average molecular weight (Mw) of BLS under differentconditions was determined on a Precision Detectors PD2010 light-scattering instrument tandemly connected to an high-performance liq-uid chromatography system and an LKB 2142 differential refractome-ter. In general, 20–100 �l of LS (0.3–1 mg/ml) was loaded on a Superdex200 HR-10/30 (24 ml) or a Sephadex G-25 (1 ml) column and eluted with50 mM phosphate buffer, 1 mM DTT under different pH, urea, GdnHCl,and NaCl conditions. The 90° light scattering and refractive indexsignals of the eluting material were recorded on a PC computer andanalyzed with the Discovery32 software supplied by Precision Detec-tors. The 90° light scattering detector was calibrated using bovineserum albumin (M

w: 66.5 kDa) as a standard. Prior to the injection in a

size exclusion chromatography column (SEC), each sample was prein-cubated for 1–2 h at room temperature in the elution buffer.

Thermodynamic Parameters

Thermodynamics evaluation of GdnHCl-induced unfolding of BLSKu1 Ku2

was fitted to a two-step model: N10 ¢O¡ (2)N5 ¢O¡ 5U. The first step(N10 7 2N5) represents the dissociation of decameric BLS (N10) in twofolded pentamers (N5), whereas the second step (N5 7 5U) representsthe concomitant dissociation and unfolding of the pentameric structure(N5) in monomeric subunits (U). Because both steps are well resolvedfrom each other, their thermodynamic parameters were independentlyanalyzed using the lineal extrapolation method (Equation 1) assuminga two-state transition model for each step (10),

�GU � �GH2O � m�D� (Eq. 1)

where �GU is the free energy of unfolding of a protein at a givendenaturant concentration, �GH2O is the free energy of unfolding in theabsence of denaturant, and m the dependence of the free energy ondenaturant concentration ([D]).

Step 1—The first step was monitored by SLS in the range of 1.5–2.2M GdnHCl. The concentration at equilibrium of the N10 and N5 speciesof BLS were calculated using Equations 4 and 5 derived from rear-rangement of Equations 2 and 3,

Mw � �Mw10�ccN10 Mw5�ccN5)/ccT (Eq. 2)

ccT � ccN10 ccN5 (Eq. 3)

[N10] � ccT(Mw � Mw5)/(Mw10 � Mw5)Mw10 (Eq. 4)

[N5] � (ccT � ccN10)/Mw5 (Eq. 5)

where Mw represents the weight average molecular weight of BLS.ccN10, ccN5, and ccT represent the decamer, pentamer, and total proteinconcentration in millgrams/ml, respectively. [N10] and [N5] representthe molar concentration and Mw10 and Mw5 the molecular weight of thedecameric (174.4 kDa) and pentameric (87.2 kDa) species ofBLS, respectively.

The equilibrium constant KU(1) and the free energy change �GU(1) forStep 1 are defined in Equations 6 and 7.

KU(1) � [N5]2/[N10] (Eq. 6)

�GU(1) � �RT ln KU(1) (Eq. 7)

The �GH2O(1) and m(1) values (Equation 1) for this transition wereobtained from the extrapolation to zero denaturant concentration andfrom the slope of the linear regression fit of the �GU(1) values calculated

Structural Stability of Lumazine Synthase8094

by guest on April 2, 2016

http://ww

w.jbc.org/

Dow

nloaded from

as a function of GdnHCl concentration in the range of 1.5 to 2.2 M. Thedissociation constant of the decameric arrangement (KD) was estimatedfrom the �GH2O value of this transition using the equation, KD �e(��GH2O/(RT)).

Step 3—The second step was monitored by CD and FT in the rangeof 2.4 and 3.5 M GdnHCl. The equilibrium constant KU(2) and the freeenergy change �GU(2) for this transition are defined in Equations 8and 9.

KU(2) � [U]5/[N5] (Eq. 8)

�GU(2) � �RT ln KU(2) (Eq. 9)

The total protein concentration in monomer units (PT) and the frac-tional population in the native (FN) and unfolded (FU) states werecalculated using Equations 10–12,

PT � 5[N5] [U] (Eq. 10)

FN � 5[N5]/PT (Eq. 11)

FU � 1 � FN � �U�/PT � �mf �D� � F � Y/��mf �D� � F � �mu�D� � U�

(Eq. 12)

where Y is the experimental spectroscopic value, [D] is the GdnHClconcentration, F and U are the intercepts, and mf and mu are the slopesof the pre- and post-unfolding baselines, respectively. Combining Equa-tions 1 and 8–12, we obtain the general equation (Equation 13)as follows.

5� �mf �D� F) � Y(mf[D] F) � (mu[D] U)�

5

PT4

� (mf[D] F) � Y(mf[D] F) � (mu[D] U)

� 1�e(�GH20 � m(2)[D])/RT � 0 (Eq. 13)

�GH2O(2) and m(2) values of Step 2 were obtained fitting the experimen-tal data measured by CD and FT to Equation 13 by a non-linear squarefit method. The midpoint of this transition [D]50% was calculated as inEquation 14.

�D�50% � �RT ln�5/16 PT4 � �GH2O]/m (Eq. 14)

Crystallization, Data Collection, and Structure Determination

In addition to the previously obtained BLS crystals, which led to theresolution of its three-dimensional structure (6), two further crystal

forms of BLS were obtained in this work by means of the hanging drop,vapor diffusion method. The first form was diamond-like crystals ob-tained using 30% (w/v) polyethylene glycol 400, 0.1 M sodium acetate in0.1 M MES buffer, pH 6.5, which diffracted to 2.9 Å and belongs to thetrigonal space group P3121. Furthermore, we obtained plate-like crys-tals using 12% (w/v) polyethylene glycol 4000, 0.1 M sodium acetate in0.1 M HEPES buffer, pH 7.5, which diffracted to 3.0 Å and belongs to themonoclinic space group P21. X-ray diffraction data were collected bothat our in-house x-ray source, a Bruker M18XH6 MAC Science rotatinganode interfaced to a Siemens X-1000 multiwire area detector and atthe D03B protein crystallography beamline at the Laboratorio Nacionalde Luz Sıncrotron, Campinas, Brazil (11). Data reduction and process-ing were carried out with the programs MOSFLM, Scala, and Truncatefrom the CCP4 suite (12). Crystal packings were determined using themolecular replacement procedure as implemented in the AMoRe pack-age (12), with the previously solved BLS structure as a search model.Crystallographic symmetry construction was carried out using the pro-gram O (13).

Structural Analysis

Total accessible surface area and buried surface areas of interaction(�ASA) between monomers and pentamers in pentamer and decamerstructures, respectively, were calculated with Surface Racer 1.2 pro-gram (14) using an implementation of the Lee and Richards (15) algo-rithm and a probe radius of 1.7 Å. Intermolecular polar and non-polarinteractions were calculated with the Molmol 2k.2 (16) and Contacts ofStructural Units (17) programs. Any pair of atoms is considered to be incontact if the distance between them is less than 4 Å.

RESULTS

BLS Is a Stable Dimer of Pentamers in Solution—Previousstudies (6, 9) have shown that BLS does not assemble as a1-MDa capsid oligomer, having a retention time in SEC com-patible with an apparent molecular mass of 90 kDa. This be-havior suggested that the oligomeric structure of this protein insolution was a pentamer. However, the estimation of the mo-lecular weight of a protein by its retention time on a SECcolumn is a method prone to artifacts. Additionally, the se-quence divergence of BLS, compared with other pentameric LS,prompted us to re-analyze the quaternary structure of theprotein by spectroscopic techniques such as static light scatter-ing (SLS).

SLS experiments show that the protein has a molecular mass

FIG. 1. Stability of BLS in presence of urea or GdnHCl monitored by FT and CD. A, percentage of fluorescence emission at 343 nm asa function of denaturant concentration. BLS samples (2.5 �M of monomers) were incubated in 50 mM sodium phosphate, pH 7.0, 1 mM DTT, withincreasing concentration of urea (●) or GdnHCl (E). Emission intensity of the sample incubated without GdnHCl was taken as 100%. The datarepresent the average of triplicate experiments. The estimated errors are below 10% of the average. B, far-UV CD spectra of native LS in 50 mM

sodium phosphate, pH 7.0 (thin line), after incubation with urea 8 M (thick line), or after incubation with 6 M GdnHCl in the same buffer(dotted line).

Structural Stability of Lumazine Synthase 8095

by guest on April 2, 2016

http://ww

w.jbc.org/

Dow

nloaded from

of 180 kDa in solution, corresponding to an assembly of twopentamers (decameric arrangement of the 18-kDa polypeptidechain). To characterize this new quaternary arrangement, westudied the thermodynamic stability of BLS, evaluating theunfolding of the protein induced by the common chemical de-naturants (urea and GdnHCl) and pH.

Preincubation of BLS with increasing concentrations of ureashows that the enzyme remains as a stable dimer of pentamers,with no detectable change in its quaternary structure (180kDa, determined by SLS, data not shown) as well as in itstertiary and secondary structures as followed by tryptophanfluorescence and CD (Fig. 1). The absence of structuralchanges, even in 8 M urea, indicates that the quaternary ar-rangement of BLS is very stable.

Conversely, GdnHCl produces a cooperative and reversiblechange in the tertiary structure reflected by a decrease intryptophan fluorescence emission (Fig. 1A). This measuresenses the environment of Trp-22, the unique tryptophan inBLS monomers that is located on the active site at the mono-mer-monomer interface (6). In addition, GdnHCl incubation (6M) produces a complete loss of secondary structure of BLS asmonitored by CD spectra (Fig. 1B). The differential effect ofGdnHCl and urea cannot be explained by their differences inionic strength. The behavior of BLS in 8.0 M urea and in thepresence of 1 M NaCl is superimposable with its describedstability in absence of salt (data not shown), implying thatunfolding with GdnHCl seems to be due to more specific inter-actions of the guanidinium cation with the protein (18, 19).Thus, we used GdnHCl-induced denaturation to study themechanism of BLS unfolding.

Dissociation of BLS Dimer of Pentamers Is Tight and Pre-cedes the Unfolding of the Pentamer—SLS analysis of theGdnHCl-induced unfolding of BLS shows a biphasic behavior

FIG. 2. Dissociation and unfolding of BLS monitored by circu-lar dichroism and SLS. A, BLS samples (2.5 �M of monomers) wereincubated in 50 mM sodium phosphate, pH 7.0, containing differentamounts of GdnHCl. After 2 h of incubation at room temperature, thefollowing parameters were measured: percentage of molar ellipticity at222 nm (f) and molecular weight determined by SLS using a Sephadex

FIG. 3. Dissociation and unfolding of BLS as a function of thepH. BLS samples (5 �M of monomers) were incubated in 50 mM acetate/citrate (pH 6.0–2.5) or sodium phosphate (pH 7.0–6.0). After 1-h incu-bation at room temperature, the following parameters were measured:f, molecular weight determined by SLS using a Sephadex G-25 column;�, percentage of molar ellipticity at 222 nm; or E, percentage of ANSfluorescence intensity. ANS binding was monitored by 365-nm excita-tion and 470-nm emission.

G-25 column (�). The values represent the average of at least threeindependent experiments. The estimated errors are below 10%. The CDdata were adjusted to a two-state equation describing the unfoldingfrom pentamer to monomers as described under “Experimental Proce-dures.” B, elution profiles obtained for BLS in 50 mM sodium phosphate,pH 7.0, containing increasing [GdnHCl] as indicated on the graph.Samples were incubated for 2 h prior to injection into a pre-equilibratedSuperdex-200 column. The arrows indicate the weight average molec-ular weight (Mw) of the peaks as determined by SLS in a flow-cellfashion.

Structural Stability of Lumazine Synthase8096

by guest on April 2, 2016

http://ww

w.jbc.org/

Dow

nloaded from

(Fig. 2A). In the first step, observed between 1.5 and 2.2 M

GdnHCl, the SLS signal intensity of the protein is reduced tohalf the value measured in the absence of denaturant. How-ever, no changes in the far-UV CD spectra and tryptophanfluorescence of the protein are observed in this range (Figs. 1Aand 2A). Thus, intrinsic tryptophan fluorescence is completelyinsensitive to the change in the quaternary structure of theprotein, indicating that the dissociation does not modify signif-icantly the environment of Trp-22. These results point to theexistence of a dissociation phenomenon of the decameric struc-ture of BLS into two pentameric subunits, with no furtherchanges in tertiary and secondary structure. On the otherhand, the second step observed between 2.4 and 3.5 M GdnHCl,shows a 5-fold decrease in the SLS signal of BLS concomitantlyto the disappearance of its far-UV CD signal at 222 nm, and asignificant decrease in the fluorescence intensity of its singletryptophan residue (Figs. 1A and 2A). These results are inter-preted as an unfolding and loss of the tertiary and secondarystructures of the protein coupled to the dissociation of its pen-tameric arrangement in five monomeric subunits (18 kDa). Theoverlapping of the changes observed in the second step by SLS,CD (Fig. 2A) and fluorescence (Fig. 1A) clearly supports ourmodel. These results were further verified by SLS coupled togel filtration chromatography (Fig. 2B), where it was possibleto isolate a 90-kDa intermediate at 2 M GdnHCl. The fact thatan intermediate of 90 kDa is stable enough to be detected bythis methodology suggests that the decameric assembly is com-posed of two previously associated pentamers.

The dissociation of the decamer to folded pentamers is alsoobserved when BLS is incubated at acidic pH (Fig. 3). Thisphenomenon is evidenced by a 2-fold decrease in the SLS signalintensity in the range of pH 4.0–5.0 as compared with the valuemeasured at pH 7.0. ANS binds to exposed hydrophobic sur-faces in partially folded intermediates with higher affinity thanto native or completely unfolded proteins (20–22). This bindingresult in a marked increase in fluorescence emission comparedwith the free ANS. We found that the dissociation of BLS at pHin the range 6.0–4.0 is a reversible process that occurs withoutsignificant exposure of hydrophobic patches and changes insecondary structure as monitored by ANS fluorescence and CDspectra (Fig. 3).

At pHs below 4.0 the protein shows an irreversible unfolding,evidenced by a loss of ellipticity at 222 nm and a dramaticincrease in binding to ANS. Clearly, both changes are coupled,implying that at acidic pHs the protein exposes hydrophobicsurfaces when secondary structure is lost. The light scatteringresponse cannot be accurately monitored below pH 4.0, pre-sumably because of aggregation.

Thermodynamic Stability of BLS Measured by Chemical De-naturation—To describe the thermodynamic stability of BLS,we studied the dissociation and unfolding steps taking advan-tage of the differential effect of GdnHCl at distinct concentra-tion ranges. Both transitions (Steps 1 and 2) were shown to behighly cooperative and reversible. The dissociation step wascharacterized by SLS using GdnHCl up to 2.2 M, in a conditionthat does not disturb the tertiary and secondary structures ofthe pentamer. This assumption is supported by the facts thatthe circular dichroism spectrum is not modified (see Fig. 2A)and the intrinsic tryptophan fluorescence does not vary uponaddition of GdnHCl (see Fig. 1A). The SLS determination of theequilibrium between the dimer of pentamers and the pentamergives a �G of 90 � 20 kJ/mol decamer as estimated by the LEMmethod assuming a two-state transition (see “ExperimentalProcedures”). This value indicates that the protein remains asa decamer under physiological conditions (estimated KD �2.48 � 10�16 M).

On the other hand, the transition from folded pentamer tounfolded monomers is a highly cooperative process that can bemeasured by GdnHCl denaturation at higher concentrationsfollowed by tryptophan fluorescence (Fig. 1A), SLS, and circu-lar dichroism (Fig. 2A). These signals show a sharp and over-lapping change around 2.4–3.5 M GdnHCl. Thus, all threespectroscopic analyses rule out the existence of a populatedintermediate during this transition. The dependence of [D]50%

with protein concentration (Table I) clearly supports the modelof a two-state transition from a folded pentamer to unfoldedmonomers. Thermodynamic analysis of this equilibrium showsthat the �G is 330 � 30 kJ/mol pentamer (Table I; see “Exper-imental Procedures” for details). In agreement with this anal-ysis, GdnHCl denaturation of BLS previously incubated at pH5.0 (as a dissociated pentamer, see Fig. 3), gives a �G of 300 �30 kJ/mol pentamer, clearly indicating the accuracy of thedetermined thermodynamic parameter for the stability of thepentamer.

Thermal Denaturation of BLS—The stability of BLS to ther-mal denaturation was followed by measuring the molar ellip-ticity at 222 nm as a function of increasing temperature, asshown in Fig. 4. The enzyme shows a sharp decrease of ellip-ticity between 85 and 95 °C with an apparent Tm of 88 � 2 °C.The loss of secondary structure is not recovered after slowcooling of the samples, indicating that an irreversible unfolding

FIG. 4. Thermal denaturation of BLS monitored by CD. BLSsamples (2.5 �M of monomers) were incubated for 2 h at room temper-ature in 50 mM sodium phosphate (dashed line), 50 mM sodium phos-phate with 2 M GndHCl (solid line), and 50 mM acetate/citrate pH 4.5(dotted line). Thermal denaturation was followed by CD and expressedas a percentage of molar ellipticity at 222 nm as a function of increasingtemperature.

TABLE IEffect of the protein concentration on the thermodynamic parameters

for the unfolding of the intermediate BLS pentamer

Protein �D�50%a �Gb

�M M kJ/mol

0.25 2.60 322 � 300.50 2.70 344 � 272.50 2.70 330 � 165.0 2.76 347 � 1825.0 2.88 317 � 30

a �D�50% is the denaturation midpoint for the unfolding of the inter-mediate BLS pentamer (concentration of GdnHCl required to achieve a50% decrease in molar ellipticity). �D�50% values were calculated usingEquation 13.

b �GH2O is the free energy of unfolding of the intermediate BLSpentamer in the absence of denaturant. Values were derived from theaverage of three independent GdnHCl denaturation experiments mon-itored by CD spectroscopy. Free energies were determined by non-linearregression analysis (see “Experimental Procedures”).

Structural Stability of Lumazine Synthase 8097

by guest on April 2, 2016

http://ww

w.jbc.org/

Dow

nloaded from

phenomenon of BLS takes place under these conditions. Ther-mal denaturation of the protein, previously incubated in twodifferent conditions that produce the dissociation of thedecamer into pentamers (2.0 M GdnHCl and pH 4.5), producesa very similar decrease of about 20 °C in the thermal stabilityof BLS (see Fig. 4), further demonstrating that dissociation anddenaturation can be uncoupled under these conditions. Thus,this shift of 20 °C can be attributed to the contribution of thepentamer-pentamer interface to the overall stability of BLS.

Quaternary Arrangement of BLS Is Confirmed by X-ray Crys-tallography—X-ray structure analysis confirmed the quater-nary arrangement of BLS. Brucella sp. LS has been crystallizedin three different forms (Refs. 6 and 23 and this work). One ofthese forms (space group P3121) has a high content of water(around 70%) and diffracts x-rays to medium resolution (2.9 Å)with a single pentamer in the asymmetric unit. Modifying thereagent conditions we obtained a second crystalline form (spacegroup P21) that has two pentamers in the asymmetric unit. Thepreviously obtained crystal form (space group R32) is moredensely packed and diffracted to 2.7-Å resolution with a singlepentamer in the asymmetric unit, allowing for the resolution ofthe three-dimensional structure of the enzyme (RCSB ProteinData Bank code 1DI0). Despite the different types of crystalarrangement (Fig. 5A), all crystal forms show the same qua-ternary arrangement. The previously described pentamerforms a tightly packed dimer of pentamers (Fig. 5B), with the Nterminus of the straight � helix (composed of helices �4 and �5,

as named in other LS) and the loop connecting the contiguoussheet making a protuberant surface that tightly fits in theneighboring pentamer. This region of the polypeptide chain hasa very high content of histidines and phenylalanines (7 His and4 Phe on a stretch of 18 residues) (Fig. 5C).

Tables II and III show the structural analysis about thenature of the monomer-monomer and pentamer-pentamer in-terfaces. Each of the monomers that gives rise to the decamericstructure buries �45% of its ASA, established mainly by hy-drophobic contacts with two neighboring monomers (pentamerassembly) and between pentamers (decamer assembly). Eachmonomer buries 35.5% of its ASA to form the pentamer assem-bly. The resulting pentamer is an intertwined structure withthe interface predominantly making non-polar contacts (64.8%of non-polar �ASA, 245 van der Waals contacts per monomer).Each monomer also makes 6 hydrogen bonds and 1 salt bridgethat stabilize the pentamer assembly. Two to three bridgingwaters per monomer make an additional 6 hydrogen bonds.Each monomer buries 9.2% of its ASA to form the decamerassembly. This interface is also mainly hydrophobic (61.5% ofnon-polar �ASA, 92 van der Waals contacts per pentamer).Each pentamer also makes 2 hydrogen bonds that stabilize thedecamer assembly. Noteworthy, 10 phosphates make 20 bridg-ing hydrogen bonds stabilizing this interface (the details to bepublished elsewhere).

Integrated Mechanism of Decameric BLS Unfolding—Thecrystallographic structural analysis, together with the solution

FIG. 5. Quaternary assembly of Bru-cella sp. LS. A, three different crystalforms of BLS show the same quaternaryarrangement. BLS folds as a dimer of pen-tamers (each decamer is depicted in a dif-ferent color) with spool-like shape. B,three-dimensional structure of thedecameric BLS. Three monomers areshown in different colors to highlight thenature of this oligomeric intertwinedstructure. The ribbons and surface dia-grams clearly show how the continuoushelix 4 protrudes into the adjacent penta-mer. C, the surface of BLS forming thepentamer-pentamer interface is shown.Histidine residues are colored in red, phe-nylalanines in yellow. The 5 C-terminaltips of helix 4 are highlighted in the pe-riphery, whereas the internal circle of his-tidines is shown in the center, close to thepore that goes through the center of thepentamer.

Structural Stability of Lumazine Synthase8098

by guest on April 2, 2016

http://ww

w.jbc.org/

Dow

nloaded from

studies described above, allowed us to postulate a comprehen-sive model for BLS unfolding (Fig. 6). As shown, the decamericassembly dissociates in two different conditions to stable foldedpentamers, with an estimated �G of 90 � 20 kJ/mol decamer.Thus, BLS would be capable to shift from a decameric to apentameric quaternary state at acidic pH, conserving the sta-bility of the protein. The effect of the pH on dissociation wouldbe explained by the high density of histidines at the pentamer-pentamer interface (see Fig. 5C). Histidine residues becomeprotonated at pH levels below 6.0, thus producing charge re-pulsion and loss of contacts resulting in the dissociation of thedimer of pentamers. In contrast, the acidic pH does not disturbthe stability of the monomer-monomer interface, because�35% of the accessible surface area of the monomeric polypep-tide is buried in monomer-monomer contacts of hydrophobicnature (Tables II and III). In agreement, GdnHCl-induced pen-tamer unfolding at neutral and acidic pHs gives approximatelythe same value of �G (300–330 kJ/mol). This high value of freeenergy would be the product of the tight and intertwined na-ture of the monomer-monomer interface, stabilized for a very

high number of van der Waals contacts (Table III) resistant tothe effect of the low pH.

DISCUSSION

Common cellular proteins exhibit only marginal stabilities,with free energies of stabilization of the order of 50 kJ/mol (24).In contrast, quaternary interactions have a dominating role inthe stabilization of small oligomeric proteins (25). Unfoldingpathways for oligomeric proteins have been shown to varysignificantly. Subunit dissociation could occur before or afterpolypeptide unfolding, or the two reactions could occur simul-taneously without significant population of equilibrium inter-mediates (26). Examples of dissociation to folded monomersfollowed by monomer unfolding have been described (27). Sev-eral examples of dissociation coupled to unfolding have alsobeen described (28–30). The decameric BLS contains elementsof both types of mechanisms. The dissociation of the dimer ofpentamers and the isolation of a pentameric intermediate pop-ulation with conserved folded structure, detected around 2.0 M

GdnHCl chloride and at acidic pHs, would be the consequence

FIG. 6. Proposed mechanism of dissociation and unfolding of BLS. The dissociation of BLS renders intermediate folded pentamers,without any significant change in the secondary and tertiary structures. Dissociation is induced by 2 M [GdnHCl] at neutral pH; further increaseof [GdnHCl] produces the concerted dissociation of the pentamer and unfolding of the 18-kDa monomer. Incubation at pH 5.0 without GdnHClproduces the same pentameric intermediate, with approximately the same thermodynamic stability as judged by GdnHCl denaturation. Ther-modynamic estimates of dissociation and unfolding are shown.

TABLE IIAnalysis of the surface areas at the monomer-monomer and the pentamer-pentamer interfaces

Polar and non-polar buried surface areas (�ASA) of a monomer in monomer-monomer and pentamer-pentamer interactions are shown. All thecalculations were made using a probe radius of 1.7Å. The percentage of the buried surface area of each monomer in a hypothetical situation being“free” in solution with respect to the total accessible surface area is represented in parentheses.

Total ASA, monomer�ASA

Pentamer Pentamer-pentamer Decamer

Å2 Å2 Å2 Å2

Total area 8026.8 2847.7 (35.5%) 736.0 (9.2%) 3583.7 (44.6%)Polar area 3262.8 (40.6%) 1003.2 (35.2%) 283.2 (38.5%) 1286.4 (35.9%)Non-polar area 4764.0 (59.4%) 1844.5 (64.8%) 452.8 (61.5%) 2297.3 (64.1%)

TABLE IIIPolar and non-polar contacts, bridging phosphate ions, and water molecules in monomer-monomer and pentamer-pentamer interfaces

An acceptor and a donor atom form a hydrogen bond if the distance between them is up to 3.4 Å with an angle of more than 90°. Oppositelycharged atoms in close proximity are defined as forming a salt bridge if they are less than or equal to 3.9 Å apart. All the non-polar contacts havebeen calculated using 4.0-Å cut-off.

Total contacts Hydrogen bonds Salt bridges Van der Waals contactsBridging

phosphatesBridgingwaters

Number HB Number HB

Monomer-monomer interface 252 6 1 245 0 0 2/3 6Pentamer-pentamer interface 94 2 0 92 10 20 0 0

Structural Stability of Lumazine Synthase 8099

by guest on April 2, 2016

http://ww

w.jbc.org/

Dow

nloaded from

of structural adaptation during quaternary divergence frompentameric to decameric lumazine synthases. In this sense, thecontinuous undistorted helix 4 of BLS (see introduction) pro-duces a complementary surface between pentamers composedof hydrophobic interactions, in which there is a marked pres-ence of histidine and phenylalanine residues (Fig. 5C). Addi-tionally, phosphate ions help to cement this interface by meansof bridging hydrogen bonds. The fact that we have been able toisolate the intermediate folded pentamer in different condi-tions would allow for detailed mutagenesis and kinetic foldingstudies of this unusual protein interface. It would allow also forprotein engineering procedures in the use of BLS as proteincarrier for the development of immunogens (see below).

In contrast, the pentamer presents an intertwined structurewith each monomer interacting extensively with its adjacentmonomers via multiple hydrophobic, hydrogen bond, and saltbridge interactions. As a consequence, dissociation of the pen-tamer and unfolding of the resulting monomers occurs as aconcerted mechanism. Even though the macroscopic methodsemployed in this study support a two-state model, we cannotrule out the existence of intermediates. Because residues ofneighboring monomers form the active site of the enzyme,there are no structural reasons for the existence of foldedmonomers.

The apparent transition temperature (Tm) of BLS is veryhigh, typical of a protein from a thermophilic organism. Unex-pectedly high Tm from other LS were previously described (5).Thermal unfolding of BLS shows an intermediate behaviorbetween pentameric and icosahedric LS, as expected for itsdecameric order. The pentameric S. cerevisiae LS melting curvehas a maximum at 74.1 °C, whereas icosahedric B. subtilisshowed a sharp melting transition at 93.3 °C as determined bydifferential scanning calorimetry (DSC) (5). Circular dichroismexperiments described in this work show that the melting pointof BLS is 88 � 2 °C, indicating a thermal stability closer to thatof higher order LS. Preliminary differential scanning calorim-etry experiments show a similar melting transition tempera-ture (results not shown). Zhang et al. (5) postulated that thepredominantly hydrophobic nature of BLS monomer-monomerinterface as compared with other LS, would produce a higherthermal stability. In addition, previous dissociation of BLS topentamers (see Figs. 2 and 3) yields a Tm of 70 � 2 °C, close tothe value of the pentameric S. cerevisiae LS. Thus, additionalthermal stability of BLS can be attributed to the formation ofthe pentamer-pentamer interface. Our results and those ofZhang et al. show a clear relationship between quaternaryorder of pentameric assemblies and thermal stability. More-over, the results shown here make BLS an excellent model tostudy the mechanisms of folding and thermodynamic stabilityof an oligomeric stable protein.

The new quaternary arrangement presented here raises thequestion about the biological implications of decameric LS.Moreover, a close inspection on the sequences of LS (8) suggeststhat the homologues from Rhodococcus erythropolis, Sinorhi-zobium meliloti, Mesorhizobium loti, Pseudomonas fluorescens,and possibly others are also decameric LS. A striking feature ofthe BLS structure is that the association of pentamers createsan internal space, delimitated by the external pores (8.5 Å ofinternal diameter) and two internal pores of around 23 Å ofdiameter (see Fig. 5C). These pores delimitate an empty spacewith the shape of a truncated cone and with a height of 60 Å.Although substrate channeling has to be ruled out because ofthe lack of an internal riboflavin synthase as in the case oficosahedric enzymes (31), the channels along the 5-fold axes(pores of 8.5 Å of diameter) could allow the passage of sub-strates but appear too narrow for the exit of lumazine (32).

Thus, this structure could be acting as a reservoir of recentlysynthesized 6,7-dimethyl-8-ribityllumazine. The fact that BLSdissociates at acidic pHs, like those encountered at phagolyso-somes, suggest a putative role of BLS as a transporter andreservoir of lumazine for riboflavin synthesis in the demandingconditions for surviving found inside eukaryotic cells (33). Ex-perimental work is needed to demonstrate the validity of thishypothesis, but the low activity of this enzyme (9) and theexistence of fluorescent products associated to recombinantlyexpressed purified BLS (resistant to urea dissociation) supportthis idea.

BLS folds as a highly stable dimer of pentamers and is ahighly immunogenic protein (23, 34, 35). These characteristicsresemble that of the B subunits of E. coli heat-labile entero-toxin (EtxB) and cholera toxin (CtxB). Both toxins assemble invivo into exceptionally stable homopentameric complexes,which maintain their quaternary structure in a range of con-ditions that would normally be expected to cause protein dena-turation (36). These remarkable stability properties, as well asthe inherent immunogenicity of EtxB and CtxB pentamers,have prompted considerable interest in their use as vaccinedelivery vehicles (36). Noteworthy, BLS is more stable to ther-mal denaturation than EtxB and CtxB toxins (Tm of 84 and75 °C, respectively) (37, 38). Thus, BLS is a promising potentialcarrier for the polymeric delivery of antigens or epitopes. Thepresence of ten sites of linkage and the natural disorderedconformation of its N termini (6) linked to its high stabilityindicate that BLS is a potential candidate for the developmentof subunit vaccines.

Acknowledgments—We acknowledge the support of Fundacion Insti-tuto Leloir, Dr. Gonzalo Prat Gray for the critical reading of the manu-script, and D. Laplagne for help in the early steps of the work.

REFERENCES

1. Ladenstein, R., Schneider, M., Huber, R., Bartunik, H. D., Wilson, K., Schott,K., and Bacher, A. (1988) J. Mol. Biol. 203, 1045–1070

2. Gerhardt, S., Haase, I., Steinbacher, S., Kaiser, J. T., Cushman, M., Bacher,A., Huber, R., and Fischer, M. (2002) J. Mol. Biol. 318, 1317–1329

3. Bacher, A. (1986) Methods Enzymol. 122, 192–1994. Ritsert, K., Huber, R., Turk, D., Ladenstein, R., Schmidt-Base, K., and Bacher,

A. (1995) J. Mol. Biol. 253, 151–1675. Zhang, X., Meining, W., Fischer, M., Bacher, A., and Ladenstein, R. (2001) J.

Mol. Biol. 306, 1099–11146. Braden, B. C., Velikovsky, C. A., Cauerhff, A. A., Polikarpov, I., and Goldbaum,

F. A. (2000) J. Mol. Biol. 297, 1031–10367. Meining, W., Mortl, S., Fischer, M., Cushman, M., Bacher, A., and Ladenstein,

R. (2000) J. Mol. Biol. 299, 181–1978. Fornasari, M. S., Laplagne, D. A., Frankel, N., Cauerhff, A. A., Goldbaum,

F. A., and Echave, J. (2003) Mol. Biol. Evol. 10.1093/molbev/msg12449. Goldbaum, F. A., Velikovsky, C. A., Baldi, P. C., Mortl, S., Bacher, A., and

Fossati, C. A. (1999) J. Med. Microbiol. 48, 833–83910. Greene, R. F., Jr., and Pace, C. N. (1974) J. Biol. Chem. 249, 5388–539311. Polikarpov, I., Oliva, G., Castellano, E. E., Garratt, R. C., Arruda, P., Leite, A.,

and Craievich, A. (1998) Nucl. Instruments Methods Physics Res. A 405,159–164

12. Navaza, J. A. C. (1994) Acta Crystallogr. Sect. A 50, 157–16313. Jones, T. A., Zou, J. Y., Cowan, S. W., and Kjeldgaard, M. (1991) Acta Crys-

tallogr. Sect. A 47, 110–11914. Tsodikov, O. V., Record, M. T., Jr., and Sergeev, Y. V. (2002) J. Comput. Chem.

23, 600–60915. Lee, B., and Richards, F. M. (1971) J. Mol. Biol. 55, 379–40016. Koradi, R., Billeter, M., and Wuthrich, K. (1996) J. Mol. Graph. 14, 51–55,

29–3217. Sobolev, V., Sorokine, A., Prilusky, J., Abola, E. E., and Edelman, M. (1999)

Bioinformatics 15, 327–33218. Botelho, M. G., Gralle, M., Oliveira, C. L., Torriani, I., and Ferreira, S. T.

(2003) J. Biol. Chem. 278, 34259–3426719. Kohn, W. D., Monera, O. D., Kay, C. M., and Hodges, R. S. (1995) J. Biol.

Chem. 270, 25495–2550620. Semisotnov, G. V., Rodionova, N. A., Razgulyaev, O. I., Uversky, V. N., Gripas,

A. F., and Gilmanshin, R. I. (1991) Biopolymers 31, 119–12821. Daniel, E., and Weber, G. (1966) Biochemistry 5, 1893–190022. D’Alfonso, L., Collini, M., and Baldini, G. (1999) Biochim. Biophys. Acta 1432,

194–20223. Goldbaum, F. A., Polikarpov, I., Cauerhff, A. A., Velikovsky, C. A., Braden,

B. C., and Poljak, R. J. (1998) J. Struct. Biol. 123, 175–17824. Jaenicke, R. (2000) J. Biotechnol. 79, 193–20325. Boudker, O., Todd, M. J., and Freire, E. (1997) J. Mol. Biol. 272, 770–77926. Guidry, J. J., Moczygemba, C. K., Steede, N. K., Landry, S. J., and Wittung-

Stafshede, P. (2000) Protein Sci. 9, 2109–2117

Structural Stability of Lumazine Synthase8100

by guest on April 2, 2016

http://ww

w.jbc.org/

Dow

nloaded from

27. Reddy, G. B., Bharadwaj, S., and Surolia, A. (1999) Biochemistry 38,4464–4470

28. Dams, T., and Jaenicke, R. (1999) Biochemistry 38, 9169–917829. Panse, V. G., Swaminathan, C. P., Aloor, J. J., Surolia, A., and Varadarajan, R.

(2000) Biochemistry 39, 2362–236930. Mok, Y. K., de Prat Gay, G., Butler, P. J., and Bycroft, M. (1996) Protein Sci.

5, 310–31931. Kis, K., Volk, R., and Bacher, A. (1995) Biochemistry 34, 2883–289232. Ladenstein, R., Ritsert, K., Huber, R., Richter, G., and Bacher, A. (1994) Eur.

J. Biochem. 223, 1007–101733. Kohler, S., Michaux-Charachon, S., Porte, F., Ramuz, M., and Liautard, J. P.

(2003) Trends Microbiol. 11, 215–21934. Goldbaum, F. A., Leoni, J., Wallach, J. C., and Fossati, C. A. (1993) J. Clin.

Microbiol. 31, 2141–214535. Velikovsky, C. A., Cassataro, J., Giambartolomei, G. H., Goldbaum, F. A.,

Estein, S., Bowden, R. A., Bruno, L., Fossati, C. A., and Spitz, M. (2002)Infect. Immun. 70, 2507–2511

36. Ruddock, L. W., Coen, J. J., Cheesman, C., Freedman, R. B., and Hirst, T. R.(1996) J. Biol. Chem. 271, 19118–19123

37. Ruddock, L. W., Webb, H. M., Ruston, S. P., Cheesman, C., Freedman, R. B.,and Hirst, T. R. (1996) Biochemistry 35, 16069–16076

38. Bhakuni, V., Xie, D., and Freire, E. (1991) Biochemistry 30, 5055–5060

Structural Stability of Lumazine Synthase 8101

by guest on April 2, 2016

http://ww

w.jbc.org/

Dow

nloaded from

Cauerhff and Fernando A. GoldbaumVanesa Zylberman, Patricio O. Craig, Sebastián Klinke, Bradford C. Braden, Ana

sp. Lumazine SynthaseBrucellaHigh Order Quaternary Arrangement Confers Increased Structural Stability to

doi: 10.1074/jbc.M312035200 originally published online December 1, 20032004, 279:8093-8101.J. Biol. Chem. 

  10.1074/jbc.M312035200Access the most updated version of this article at doi:

 Alerts:

  When a correction for this article is posted• 

When this article is cited• 

to choose from all of JBC's e-mail alertsClick here

  http://www.jbc.org/content/279/9/8093.full.html#ref-list-1

This article cites 37 references, 7 of which can be accessed free at

by guest on April 2, 2016

http://ww

w.jbc.org/

Dow

nloaded from