6
Flexibility of the metal-binding region in apo-cupredoxins María-Eugenia Zaballa a , Luciano A. Abriata a , Antonio Donaire b,1 , and Alejandro J. Vila a,1 a Instituto de Biología Molecular y Celular de Rosario, Facultad de Ciencias Bioquímicas y Farmacéuticas, Universidad Nacional de Rosario, Suipacha 531, S2002LRK Rosario, Argentina; and b Departamento de Química Inorgánica, Facultad de Química, Universidad de Murcia, Campus Universitario de Espinardo, Apdo. 4021, 30100 Murcia, Spain Edited by Harry B. Gray, California Institute of Technology, Pasadena, CA, and approved April 18, 2012 (received for review November 30, 2011) Protein-mediated electron transfer is an essential event in many biochemical processes. Efficient electron transfer requires the reor- ganization energy of the redox event to be minimized, which is ensured by the presence of rigid donor and acceptor sites. Electron transfer copper sites are present in the ubiquitous cupredoxin fold, able to bind one or two copper ions. The low reorganization en- ergy in these metal centers has been accounted for by assuming that the protein scaffold creates an entatic/rack-induced state, which gives rise to a rigid environment by means of a preformed metal chelating site. However, this notion is incompatible with the need for an exposed metal-binding site and proteinprotein inter- actions enabling metallochaperone-mediated assembly of the cop- per site. Here we report an NMR study that reveals a high degree of structural heterogeneity in the metal-binding region of the nonme- tallated Cu A -binding cupredoxin domain, arising from microsecond to second dynamics that are quenched upon metal binding. We also report similar dynamic features in apo-azurin, a paradigmatic blue copper protein, suggesting a general behavior. These findings reveal that the entatic/rack-induced state, governing the features of the metal center in the copper-loaded protein, does not require a preformed metal-binding site. Instead, metal binding is a major contributor to the rigidity of electron transfer copper centers. These results reconcile the seemingly contradictory requirements of a rigid, occluded center for electron transfer, and an accessible, dynamic site required for in vivo copper uptake. metalloproteins nuclear magnetic resonance protein dynamics L ong-range electron transfer (ET) in proteins is a key chemical event in many essential biological processes such as cellular respiration and photosynthesis (13). According to Marcussemi- classical theory, the outstanding efficiency of these processes is based on the maximization of the superexchange coupling be- tween donor and acceptor and the minimization of the reorgani- zation energy, given the low driving forces found in most biological systems (46). Thus, with driving forces as low as 0.1 eV and distances larger than 10 Å, efficient long-range electron transfer is only possible if the nuclear reorganization energy of the reactants is below 1 eV (4, 7). Transition metal ions such as copper and iron are ubiquitous in ETchains because their redox potentials and electronic structures can be tuned by the protein environment to match the require- ments of different biological redox events. Cycling between the coordination geometries preferred by these metals in each redox state would in principle lead to high reorganization energies (7). Electron transfer iron centers avoid this by using rigid cofactors such as iron-sulfur clusters and heme groups (810). Instead, copper ions in ET centers are only bound to protein residues, as observed in type 1 (blue) copper and Cu A centers; and thus the protein fold around the metal ion is expected to impart rigidity to these otherwise flexible metal centers (1012). Particularly, out- er-sphere coordination involving hydrogen bonding networks has been proposed as responsible for the low reorganization energies observed in copper centers in proteins (13, 14). The strain intro- duced by the protein fold is then predicted to be responsible for the unusual functional and spectroscopic features observed in type 1 copper and Cu A centers (15). This scenario, known as the entatic/rack-induced state hypothesis, was originally formulated for protein-bound cofactors by Lumry and Eyring in 1954 (16) and then extended to electron transfer metalloproteins by Malmström and Williams in the late 1960s (17, 18). This concept has been matter of debate along many years (7, 15, 1926). In particular, Ryde et al. have reported in vacuum quantum calcula- tions suggesting the absence of any strain in the geometry adopted by the copper in the protein framework (21, 22). The entatic/rack-induced state hypothesis applied to ET copper proteins was strongly supported by several X-ray struc- tures of type 1 copper proteins revealing that the conformation of the copper ligands in the metal-depleted (apo) form was iden- tical to that found in the holoproteins (2731). These results prompted the idea that the protein fold determines the coordina- tion geometry of the metal center by creating a preformed chelating site with very little flexibility, thus minimizing confor- mational changes that would normally be exhibited during Cu (II)/Cu(I) redox cycling (19). This idea is reinforced by the fact that all ETcopper centers are bound to a rigid Greek-key β-barrel domain, known as the cupredoxin fold (3134). Instead, iron- sulfur and heme centers are found in different folding motifs, with diverse intrinsic mobilities (3537). The notion of a preformed rigid binding site is conflicting with the finding that copper centers are loaded by specific metallocha- perones (38, 39). The concept of metallochaperones is relatively new and it has replaced the idea that metal uptake depends entirely on the chelating properties of the apoprotein. Several copper chaperones have been identified and characterized so far, requiring specific proteinprotein interactions that can be metal- mediated in some cases (40). In the latter situation, the metal ion is simultaneously bound to ligands from both the metallochaper- one and the target protein during the transfer step (40). In any case, the assumption of a highly rigid metal site, which should also be hidden from the bulk solvent to prevent unwanted side reac- tions, does not favor metal delivery by another protein. Thus, the demands for efficient electron transfer and in vivo metallation are hard to reconcile given the current knowledge in the field. We have recently elucidated the chaperone-mediated mechan- ism of copper uptake by the Cu A -binding cupredoxin domain in Thermus thermophilus ba 3 cytochrome c oxidase (41). In that work, we observed that the NMR spectra of the apo and metal- Author contributions: M.-E.Z., L.A.A., A.D., and A.J.V. designed research; M.-E.Z., L.A.A., and A.D. performed research; M.-E.Z., L.A.A., A.D., and A.J.V. analyzed data; and M.-E.Z., L.A.A., A.D., and A.J.V. wrote the paper. The authors declare no conflict of interest. This article is a PNAS Direct Submission. Data deposition: The atomic coordinates have been deposited in the Protein Data Bank, www.pdb.org (PDB ID code 2LLN). The NMR chemical shifts have been deposited in the BioMagResBank, www.bmrb.wisc.edu (accession nos. 18081 and 18254). 1 To whom correspondence may be addressed. E-mail: [email protected] or [email protected]. This article contains supporting information online at www.pnas.org/lookup/suppl/ doi:10.1073/pnas.1119460109/-/DCSupplemental. 92549259 PNAS June 12, 2012 vol. 109 no. 24 www.pnas.org/cgi/doi/10.1073/pnas.1119460109

Flexibility of the metal-binding region in apo-cupredoxins

Embed Size (px)

Citation preview

Page 1: Flexibility of the metal-binding region in apo-cupredoxins

Flexibility of the metal-binding regionin apo-cupredoxinsMaría-Eugenia Zaballaa, Luciano A. Abriataa, Antonio Donaireb,1, and Alejandro J. Vilaa,1

aInstituto de Biología Molecular y Celular de Rosario, Facultad de Ciencias Bioquímicas y Farmacéuticas, Universidad Nacional de Rosario, Suipacha 531,S2002LRK Rosario, Argentina; and bDepartamento de Química Inorgánica, Facultad de Química, Universidad de Murcia, Campus Universitario deEspinardo, Apdo. 4021, 30100 Murcia, Spain

Edited by Harry B. Gray, California Institute of Technology, Pasadena, CA, and approved April 18, 2012 (received for review November 30, 2011)

Protein-mediated electron transfer is an essential event in manybiochemical processes. Efficient electron transfer requires the reor-ganization energy of the redox event to be minimized, which isensured by the presence of rigid donor and acceptor sites. Electrontransfer copper sites are present in the ubiquitous cupredoxin fold,able to bind one or two copper ions. The low reorganization en-ergy in these metal centers has been accounted for by assumingthat the protein scaffold creates an entatic/rack-induced state,which gives rise to a rigid environment by means of a preformedmetal chelating site. However, this notion is incompatible with theneed for an exposed metal-binding site and protein–protein inter-actions enabling metallochaperone-mediated assembly of the cop-per site. Herewe report an NMR study that reveals a high degree ofstructural heterogeneity in the metal-binding region of the nonme-tallated CuA-binding cupredoxin domain, arising from microsecondto second dynamics that are quenched upon metal binding. Wealso report similar dynamic features in apo-azurin, a paradigmaticblue copper protein, suggesting a general behavior. These findingsreveal that the entatic/rack-induced state, governing the featuresof themetal center in the copper-loaded protein, does not require apreformed metal-binding site. Instead, metal binding is a majorcontributor to the rigidity of electron transfer copper centers.These results reconcile the seemingly contradictory requirementsof a rigid, occluded center for electron transfer, and an accessible,dynamic site required for in vivo copper uptake.

metalloproteins ∣ nuclear magnetic resonance ∣ protein dynamics

Long-range electron transfer (ET) in proteins is a key chemicalevent in many essential biological processes such as cellular

respiration and photosynthesis (1–3). According to Marcus’ semi-classical theory, the outstanding efficiency of these processes isbased on the maximization of the superexchange coupling be-tween donor and acceptor and the minimization of the reorgani-zation energy, given the low driving forces found in mostbiological systems (4–6). Thus, with driving forces as low as 0.1 eVand distances larger than 10 Å, efficient long-range electrontransfer is only possible if the nuclear reorganization energy ofthe reactants is below 1 eV (4, 7).

Transition metal ions such as copper and iron are ubiquitous inETchains because their redox potentials and electronic structurescan be tuned by the protein environment to match the require-ments of different biological redox events. Cycling between thecoordination geometries preferred by these metals in each redoxstate would in principle lead to high reorganization energies (7).Electron transfer iron centers avoid this by using rigid cofactorssuch as iron-sulfur clusters and heme groups (8–10). Instead,copper ions in ETcenters are only bound to protein residues, asobserved in type 1 (blue) copper and CuA centers; and thus theprotein fold around the metal ion is expected to impart rigidity tothese otherwise flexible metal centers (10–12). Particularly, out-er-sphere coordination involving hydrogen bonding networks hasbeen proposed as responsible for the low reorganization energiesobserved in copper centers in proteins (13, 14). The strain intro-duced by the protein fold is then predicted to be responsible for

the unusual functional and spectroscopic features observed intype 1 copper and CuA centers (15). This scenario, known as theentatic/rack-induced state hypothesis, was originally formulatedfor protein-bound cofactors by Lumry and Eyring in 1954 (16)and then extended to electron transfer metalloproteins byMalmström and Williams in the late 1960s (17, 18). This concepthas been matter of debate along many years (7, 15, 19–26). Inparticular, Ryde et al. have reported in vacuum quantum calcula-tions suggesting the absence of any strain in the geometryadopted by the copper in the protein framework (21, 22).

The entatic/rack-induced state hypothesis applied to ETcopper proteins was strongly supported by several X-ray struc-tures of type 1 copper proteins revealing that the conformationof the copper ligands in the metal-depleted (apo) form was iden-tical to that found in the holoproteins (27–31). These resultsprompted the idea that the protein fold determines the coordina-tion geometry of the metal center by creating a preformedchelating site with very little flexibility, thus minimizing confor-mational changes that would normally be exhibited during Cu(II)/Cu(I) redox cycling (19). This idea is reinforced by the factthat all ETcopper centers are bound to a rigid Greek-key β-barreldomain, known as the cupredoxin fold (31–34). Instead, iron-sulfur and heme centers are found in different folding motifs,with diverse intrinsic mobilities (35–37).

The notion of a preformed rigid binding site is conflicting withthe finding that copper centers are loaded by specific metallocha-perones (38, 39). The concept of metallochaperones is relativelynew and it has replaced the idea that metal uptake dependsentirely on the chelating properties of the apoprotein. Severalcopper chaperones have been identified and characterized so far,requiring specific protein–protein interactions that can be metal-mediated in some cases (40). In the latter situation, the metal ionis simultaneously bound to ligands from both the metallochaper-one and the target protein during the transfer step (40). In anycase, the assumption of a highly rigid metal site, which should alsobe hidden from the bulk solvent to prevent unwanted side reac-tions, does not favor metal delivery by another protein. Thus, thedemands for efficient electron transfer and in vivo metallation arehard to reconcile given the current knowledge in the field.

We have recently elucidated the chaperone-mediated mechan-ism of copper uptake by the CuA-binding cupredoxin domain inThermus thermophilus ba3 cytochrome c oxidase (41). In thatwork, we observed that the NMR spectra of the apo and metal-

Author contributions: M.-E.Z., L.A.A., A.D., and A.J.V. designed research; M.-E.Z., L.A.A.,and A.D. performed research; M.-E.Z., L.A.A., A.D., and A.J.V. analyzed data; and M.-E.Z.,L.A.A., A.D., and A.J.V. wrote the paper.

The authors declare no conflict of interest.

This article is a PNAS Direct Submission.

Data deposition: The atomic coordinates have been deposited in the Protein Data Bank,www.pdb.org (PDB ID code 2LLN). The NMR chemical shifts have been deposited in theBioMagResBank, www.bmrb.wisc.edu (accession nos. 18081 and 18254).1To whom correspondence may be addressed. E-mail: [email protected] or [email protected].

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1119460109/-/DCSupplemental.

9254–9259 ∣ PNAS ∣ June 12, 2012 ∣ vol. 109 ∣ no. 24 www.pnas.org/cgi/doi/10.1073/pnas.1119460109

Page 2: Flexibility of the metal-binding region in apo-cupredoxins

lated forms of this protein show significant differences, suggestingthat there might be nonnegligible structural perturbations uponmetal binding, consistent with the finding that copper uptake ismediated by a metallochaperone (Fig. S1). Here we report a de-tailed NMR study showing the solution structure and dynamicfeatures of this cupredoxin domain in the nonmetallated form.We have found that not only is the metal-binding site disorderedin the apoprotein but it also exhibits significant dynamics in themicrosecond to second timescale. Moreover, to prove that ourresults can be extended to other cupredoxins, we report an NMR-based study of Pseudomonas aeruginosa apo-azurin, which hasbeen considered as the paradigm of blue copper proteins andlong-range ET (42–44). Also in this case, we show there is signif-icant dynamics in the metal-binding region only in the absence ofthe copper ion. Thus, our findings indicate that the rigid coppercenter needed for efficient long-range ET is not preformed in thenonmetallated cupredoxin. Instead, metal binding does signifi-cantly contribute to the rigidity of these centers and the metal-binding region in the apoprotein is flexible enough to allow invivo metallation.

ResultsThe CuA-Binding Cupredoxin Domain. Subunit II of T. thermophilusba3 oxidase consists of a soluble CuA-binding domain anchored tothe membrane through an N-terminal transmembrane helix (32,45). This domain adopts a cupredoxin fold similar to the onefound in type 1 copper proteins, with a longer loop that allowsbinding of an additional metal ion (12). The two copper ions ofthe CuA site are bridged by two cysteine ligands (Cys149 andCys153), giving rise to a rigid Cu2S2 core. Each copper ion is alsocoordinated to a terminal histidine residue (His114 and His157)and a weak axial ligand: a methionine sulfur (Met160) and a back-bone carbonyl (Gln151), respectively. The soluble fragment here-in studied comprises only the soluble cupredoxin domain ofsubunit II, where the transmembrane helix is not part of the pur-ified protein. This fragment retains the structure observed in thewhole oxidase (32, 45) and it is competent for ETwith its biolo-gical redox partner (46, 47). Given that residues 1–42 are absentin our protein construct, residue numbering starts at 43 in theresults presented here, following that employed in the X-raystructure of oxidized holoCuA [Protein Data Bank (PDB) ID2CUA] (32).*

Resonance Assignment and Solution Structure of apoCuA.1H, 13C,

and 15N resonances were assigned for most residues in the non-metallated form of the soluble domain of T. thermophilus ba3oxidase subunit II (apoCuA, hereafter) (Table S1). Resonancescorresponding to eight non-proline residues were missing inthe 1H, 15N-heteronuclear single quantum coherence (HSQC)spectrum of apoCuA, whereas 19 residues presented duplicatedNH cross-peaks (see below). Five of the residues with missingresonances (Cys149-Gln151, His157, and Gln158) are locatedat the loop bearing five out of the six copper ligands (ligand loop,hereafter) whereas the remaining correspond to His114 (theN-terminal histidine coordinated to one of the copper ions), itsadjacent residue Gly115, and the N-terminal Met43. These eightresidues are thus involved in an exchange process whosefrequency lies right on an intermediate timescale that broadenssignals beyond detection. Among residues with duplicated reso-nances in the 1H, 15N-HSQC spectrum, five are located in theN-terminal region (Val44, Ile45, Ala47, Lys49, and Leu50) andduplication here arises from the cis-trans isomerization of Pro46in a slow regime, as reported for the reduced holoprotein [Bio-MagResBank (BMRB) entry 5819; ref. 48]. Most of the remain-

ing residues with duplicated correlations are located around themetal-binding site (Ala85, Ala87, Tyr90, Val112, Ile113, His117,Gly120, Asn124, Gly154, Gly156, Met160-Thr163). Comparisonof the assignments obtained for apoCuA with those availablefor the reduced holoprotein revealed that most chemical shiftperturbations also map close to the metal-binding site (Fig. 1and Fig. S1).

The calculated three-dimensional structure of apoCuA is welldefined by 2,147 (including 879 long-range) NOE-based distancerestraints and 215 experimental torsion angle restraints (Fig. S2and Table S2). The atomic rmsd values for the 20 models inthe refined structure are 1.0� 0.2 and 1.4� 0.1 Å for the back-bone and all heavy atoms, respectively. Considering only residuesfrom the rigid β-barrel domain (residues 53–148 and 162–168),the rmsd values for backbone and all heavy atoms are 0.49� 0.07and 1.01� 0.08 Å, respectively.

The apoprotein in solution adopts a β-barrel fold where mostregions are well-defined except the N terminus and loops 86–90and 149–161 (ligand loop) (Fig. 1 and Fig. S3). The disorderobserved at the ligand loop might be due to the lack of experi-mental constraints (Fig. S2) and/or true mobility of this region inthe absence of the metal ions. Indeed, the low number of con-straints in this region parallels the missing correlations in the1H, 15N-HSQC, suggestive of some degree of flexibility (seebelow). It is then clear that the polypeptide presents a similarGreek-key β-barrel fold both in the apo- and copper-loadedforms, and that the structural perturbations between both formsof the protein are confined to the metal-binding region, with theligand loop showing the highest degree of disorder.

Dynamics of apoCuA and Reduced holoCuA. Subnanosecond timescale.The dynamics of apo- and reduced holoCuA in the pico-to-nano-second timescale was studied by measuring 15N R1 and R2 and1H-15N heteronuclear NOE. Average correlation times of 8.3�0.1 ns (apoCuA) and 8.2� 0.1 ns (holoCuA) were estimated andused for relaxation data analysis with the model-free approach(49–51). Order parameter values (S2) for apo- and holoCuAare remarkably high along most of the sequence, averaging0.9� 0.1 in both proteins and confirming the high level of rigidityof the β-barrel (Fig. S4). As expected, the order parameter valuesfor both protein forms drop below 0.6 at the N terminus, due tothe high mobility usually observed in this region. S2 values lowerthan 0.6 are also observed for residues 154–156 in apoCuA. How-ever, a slight decrease in S2 values is also found in this region for

Fig. 1. NMR-based solution structure of T. thermophilus apoCuA. (A) Car-toon representation of the X-ray structure of oxidized holoCuA (from PDBID 2CUA; ref. 32). (B) Ribbon representation of the best 20 out of 80 structurescalculated by CYANA for apoCuA in solution (this work, PDB ID 2LLN). Regionswith significant chemical shift perturbations (Eq. S1) between apo- andholoCuA and structural disorder in the calculated structure of the apoproteinare highlighted in blue (ligand loop, residues 149–161), and orange (residues86–90).

*The PDB coordinates (PDB ID 2LLN) and resonance assignments (BioMagResBank entry18081) for apoCuA obtained during this work are deposited using the linear 1–126 num-bering.

Zaballa et al. PNAS ∣ June 12, 2012 ∣ vol. 109 ∣ no. 24 ∣ 9255

CHEM

ISTR

YBIOCH

EMISTR

Y

Page 3: Flexibility of the metal-binding region in apo-cupredoxins

holoCuA. Overall, these data do not reveal significantly differentdynamic features in the pico-to-nanosecond timescale betweenthe apo and holo forms of the CuA-binding domain.

Microsecond to millisecond timescale. Protein motions in the micro-to-millisecond timescale were probed by constant-time Carr–Purcell–Meiboom–Gill (CPMG) relaxation dispersion experi-ments (52, 53). Fig. 2A shows the ΔR2 profiles obtained forapoCuA and reduced holoCuA, where ΔR2 is the differencebetween R2 values measured at CPMG frequencies of 33 and966 Hz, validated by fitting of the dispersion curves to Eq. S2(Fig. 2, Inset and Fig. S5). The ΔR2 profile for apoCuA showstwo regions experiencing significant exchange in comparison tothe holoprotein. Both regions map close to the metal-binding site,as shown in Fig. 3A. Thus, residues in the metal-binding regionexhibit significant micro-to-millisecond dynamics in the apopro-tein, and this flexibility is lost upon binding of the metal ions.

Protein compactness.Protein compactness was studied by a 1H, 15Nheterogeneity-band-selective optimized-flip-angle short-transient-heteronuclear multiple quantum coherence (HET-SOFAST-HMQC) experiment with irradiation of aliphatic protons. TheλNOE parameter calculated from this experiment reports on theproton density around the probed amide, with low λNOE values(<0.3) indicating well-structured compact segments and highervalues indicating a less compact environment (which might bedue in turn to enhanced dynamics) (54). Fig. 2B shows the λNOEprofiles obtained for apo- and reduced holoCuA. Both proteinforms present a similar global compactness along most of the se-quence, except for a significant difference observed at the ligandloop, where the higher λNOE values obtained for the apoproteinsuggest a less compact structure in the absence of the copper ions.

Millisecond to second timescale.The absence (for eight non-prolineresidues) and duplication (for 19 residues) of NH cross-peaks inthe 1H, 15N-HSQC spectrum of apoCuA reveals chemical ex-change in the intermediate and slow regimes on the chemical shifttimescale, respectively. As detailed above, most of these residues

correspond to copper ligands and/or map to the surroundingsof the metal-binding site. We note that only one correlation isobserved for all these residues in the 1H, 15N-HSQC spectrumof reduced holoCuA (BMRB entry 5819). Thus, slow dynamicfeatures on the copper ligands and surrounding residues are onlyobserved in the absence of the copper ions (Fig. 3B).

Dynamics of apo-azurin. In order to extend these results to othercupredoxins, we have studied the solution dynamics of P. aerugi-nosa apo-azurin. The blue copper protein azurin has been exten-sively studied as a paradigmatic protein for electron transfer(42–44), and Cu(I)-azurin has been the subject of several NMRstudies (55–58). Instead, apo-azurin has not been characterizedby NMR.

Backbone 1H and 15N resonances were assigned for mostresidues of P. aeruginosa apo-azurin. Resonances correspondingto 17 non-proline residues were missing in the 1H, 15N-HSQCspectrum, indicating they are involved in an intermediate ex-change process on the chemical shift timescale. These residuesare Ala1-Cys3, Gly9-Asp11, Asn18, His35-Leu39, Gly45, His46,Gly88-Glu91, most of them mapping to the surroundings of themetal-binding site (Fig. 4B). This behavior contrasts with that ofCu(I)-azurin, whose 1H, 15N-HSQC spectrum shows one cross-peak for every non-proline residue in the protein (except theN-terminal Ala1) (56), indicating that this slow dynamic processtakes place only in the nonmetallated form, in agreement withthe findings on apoCuA. Moreover, most chemical shift perturba-tions resulting from metal removal in azurin cluster close to thecopper center (Fig. 4A and Fig. S6), again resembling the situa-tion met in the CuA-binding domain.

Protein dynamics in the micro-to-millisecond timescale wereevaluated by relaxation dispersion experiments in apo-azurin.As observed in Fig. 4C, the ΔR2 profile, validated by fitting ofthe dispersion curves to Eq. S2 (Fig. S7), shows three definedregions in the micro-to-millisecond timescale. Residues in theseregions are next to residues with missing correlations in the 1H,15N-HSQC of apo-azurin, thus evidencing a common exchangeprocess. Comparison of these data with those reported by

Fig. 2. Protein dynamics and compactness of apoCuA (black squares) and reduced holoCuA (gray triangles). (A) ΔR2 profiles showing dynamics in themicro-to-millisecond timescale. Fits of experimental data to Eq. S2 for some apoCuA residues are shown as examples in the inset of the figure (Gln91, circles;Met160, triangles; and Gly162, squares, showing significant relaxation dispersion; and Ala101, diamonds, located in a region showing no exchange). (B) λNOE

profiles showing proton density around the backbone amides.

9256 ∣ www.pnas.org/cgi/doi/10.1073/pnas.1119460109 Zaballa et al.

Page 4: Flexibility of the metal-binding region in apo-cupredoxins

Orekhov and coworkers for Cu(I)-azurin (55, 57) reveals an in-creased mobility at the metal-binding region in the absence of thecopper ion.

DiscussionIn this work, we present an NMR study on apoCuA, revealing thatthe metal-binding region (mainly the ligand loop) shows a disor-dered and less compact structure in the apoprotein compared tothe holo form (Figs. 1 and 2B) and that this disorder, in turn, isdue to a dynamic behavior of this region in the microsecond tosecond timescale (Figs. 2A and 3). The blue copper protein azurindisplays similar dynamic features in its apo form which also mapto the metal-binding region (Fig. 4). The absence of such dynamicfeatures both in holoCuA and Cu(I)-azurin allows us to concludethat this region is significantly flexible in the apoproteins and thatonly metal binding quenches the dynamics of these loops in thecupredoxin fold. Given the different loop length and structure inCuA and azurin, these results strongly suggest that this behavior isa general feature of cupredoxins.

Dynamics of Apo-Cupredoxins and the Entatic/Rack-Induced State.The entatic/rack-induced state in ET copper proteins refers tothe organized protein structure around the metal-binding sitethat ensures a low reorganization energy for efficient electrontransfer (7). There have been numerous reports and reviews sup-porting the entatic/rack-induced state hypothesis (7, 13, 19, 20,26, 59). On the other hand, theoretical and experimental studieshave questioned the role of the protein matrix in defining thegeometry of the copper center (21, 22, 24, 25). X-ray studies onthe apo form of several cupredoxins, showing no changes in theligands’ conformations upon removal of the copper ion, wereconsidered as experimental evidence of the entatic/rack-induced

Fig. 3. Mapping of the dynamic features in apoCuA. Residues experiencingdynamics in the absence of the copper ions (i.e., in apoCuA) are highlighted inthe structure of holoCuA (PDB ID 2CUA, ref. 32, used to evidence the proxi-mity of these residues to the metal ions). (A) Dynamics in the micro-to-milli-second timescale. Residues with ΔR2 < 1.3 s−1, ΔR2 > 1.3 s−1, and missingresidues are shown in blue, orange, and gray, respectively. (B) Slow dynamics.Residues with zero (missing), one or two (duplicated) correlations in the 1H,15N-HSQC spectrum of apoCuA are shown in red, gray, and yellow, respec-tively.

Fig. 4. NMR results on apo-azurin. Residues showing significant chemical shift perturbations or enhanced dynamic features are mapped on the tridimensionalstructure of Cu(II)-azurin (PDB ID 4AZU; ref. 33). (A) Chemical shift perturbations between apo-azurin and Cu(I)-azurin backbone amide resonances. Residueswith chemical shift perturbations ðCSPÞ < 0.2, 0.2 < CSP < 0.5, CSP > 0.5 are shown in blue, yellow, and red, respectively. Residues with missing NH correlationsare shown in gray. (B) Microsecond to second dynamics of apo-azurin. Residues with ΔR2 < 2 s−1, ΔR2 > 2 s−1 are shown in blue and orange, respectively.Residues with missing correlations in the 1H, 15N-HSQC spectrum are shown in red. In both representations, prolines are shown in gray. (C)ΔR2 profiles showingdynamics in themicro-to-millisecond timescale. Fits of experimental data to Eq. S2 for some apo-azurin residues are shown as examples in the inset of the figure(Ile87, circles; Met44, squares, showing significant relaxation dispersion; and Thr61, triangles, located in a region showing no exchange).

Zaballa et al. PNAS ∣ June 12, 2012 ∣ vol. 109 ∣ no. 24 ∣ 9257

CHEM

ISTR

YBIOCH

EMISTR

Y

Page 5: Flexibility of the metal-binding region in apo-cupredoxins

hypothesis for ETcopper proteins (27–31). In addition, those re-sults put forward the idea that the protein matrix creates a pre-formed rigid chelating site in the absence of the metal ions, thusentirely determining the geometry of the copper center (19).

The present NMR-based results provide compelling evidenceof structural disorder, flexibility, and conformational exchange inthe metal-binding region of the apo form of the CuA-bindingcupredoxin domain. Flexibility and slow chemical exchange werealso observed here for apo-azurin, supporting the extension ofthese findings to other cupredoxins and contrasting the men-tioned X-ray studies. This discrepancy may be attributed to thefact that most of the reported structures of apo-cupredoxins wereobtained by crystallization of the holoprotein and subsequentmetal removal from the crystal by soaking with a chelating agent(28–30). Instead, direct crystallization of apo-azurin gave rise tosome (minor) structural heterogeneity selectively on the copperligands (27). However, analysis of the structure of apo-azurindoes not reveal a significant increase in the B factors in the metal-binding region nor the lack of electron density, suggesting that theabsence of dynamics in these cases is due to crystal packing.

We note that our results do not necessarily support nor refutethe entatic/rack-induced state hypothesis because they apply tothe nonmetallated protein and do not assess the strain the proteinenvironment exerts on the Cu(I) and Cu(II) centers. Instead, ourfindings show that the entatic/rack-induced state, when operativeon the metallated protein, does not require a preformed rigidchelating site. This conception is in line with the results reportedby Wittung-Stafshede and coworkers, showing that a rigid, ET-functional blue copper center can be obtained from Cu(II) bind-ing to unfolded azurin, suggesting that the constraints imposed bythe protein to the geometry of the metal center do not require apreorganized binding site defined in the polypeptide (60–62).

Dynamics of Apo-Cupredoxins and in Vivo Copper Uptake.The coppercargo in the CuA-binding cupredoxin domain of T. thermophilusoxidase is due to the action of a specific Cu(I)-metallochaperone(PCuAC) (41). Similar copper-loading mechanisms have beenproposed for the type 1 protein plastocyanin (63). Copper trans-fer mediated by chaperones generally takes milliseconds to sec-onds, overlapping with the fluxional timescales measured in thiswork for the metal-binding region of apoCuA and apo-azurin(although no copper chaperones have been reported for azurinto date). This observation suggests that the observed flexibilityof this region in the apo form of the protein allows the cupredoxinfold to sample different conformations, enabling copper transfer.

Our unique picture of a folded apo-cupredoxin consisting in aβ-barrel with disordered flexible loops is fully consistent with theneed of specific protein–protein interactions and defines an ex-

posed, dynamic metal-binding site allowing for efficient metallo-chaperone-mediated copper binding in vivo. In turn, the fullyconserved Greek-key β-barrel scaffold would be able to rendera rigid and efficient ET site once the copper ions are bound.Our results thus allow us to reconcile the seemingly contradictoryrequirements of a rigid, occluded center for electron transfer, andan accessible, dynamic site for in vivo copper uptake. Nature hassolved this problem by exploiting the cupredoxin fold, which isendowed at the same time with rigid and dynamic features thatare finely tuned by the binding of the metal ions.

Materials and MethodsProtein Preparation. Samples of azurin and apo- and holoCuA proteins wereuniformly labeled with 15N or 13C and 15N (Cambridge Isotope Laboratories,Inc.) and produced as described elsewhere (40, 47, 64). The apoCuA proteinwas expressed and purified in the absence of copper, whereas azurin waspurified as Cu(II)-azurin and the apoprotein was obtained by dialysis against50 mM KCN, 200 mM Tris·HCl pH 8.5, followed by several dialysis steps with-out the complexing agent. Protein samples for NMR experiments were pre-pared in 100 mM phosphate buffer pH 7.0 for azurin and pH 6.0 for the CuA

domain, adding 100 mM KCl in the case of the holoprotein or 2 mM DTT forthe apoproteins in order to avoid oxidation of the cysteine ligands. Proteinconcentration was about 1 mM.

Nuclear Magnetic Resonance Spectroscopy.NMR experiments were carried outon a 600 MHz Bruker Avance II Spectrometer equipped with a triple reso-nance inverse (TXI) probe head and on a 900 MHz Bruker Avance II Spectro-meter equipped with a triple resonance inverse (TCI) cryoprobe. Allexperiments were carried out at 298 K using standard techniques, as de-scribed in SI Text. The obtained family of structures for apoCuA is depositedat the Protein Data Bank under PDB ID 2LLN. Chemical shifts for all 1H, 13C,and 15N nuclei assigned in this work are deposited at the Biological MagneticResonance Data Bank under accession numbers 18081 and 18254 for apoCuA

and apo-azurin, respectively. Resonance assignments for the reduced metal-lated proteins, holoCuA and Cu(I)-azurin, were taken from the literature (48,56) and transferred by us to the sample conditions used in this work.

ACKNOWLEDGMENTS. The authors thank J. H. Richards (Caltech) for providingthe plasmid for azurin expression. The Centro di Risonanze Magnetiche ofFlorence, Italy, and the European Program Bio-NMR (BIO-NMR-00012) areacknowledged for the access to high magnetic field spectrometers. A.D.also thanks the Spanish Ministerio de Educación for a grant supporting afour months’ stay in Rosario, Argentina (PR2009-0479). M.E.Z. and L.A.A. aredoctoral and postdoctoral fellows from Consejo Nacional de InvestigacionesCientíficas y Técnicas (CONICET), respectively. A.J.V. is an Howard HughesMedical Institute International Research Scholar and a staff member fromCONICET. The Bruker Avance II 600 MHz was purchased with funds fromAgencia Nacional de Promoción Científica y Tecnológica (ANPCyT) and CON-ICET. This research was funded by ANPCyT, Argentina (Proyecto de Investiga-ción Científica y Tecnológica 2007-314), the Spanish Ministerio de Economía yCompetitividad, and Fundación Séneca de la Región deMurcia, Spain (projectnumbers SAF2011-26611 and 15354/PI/10, respectively).

1. Ramirez BE, Malmström BG, Winkler JR, Gray HB (1995) The currents of life: The term-inal electron-transfer complex of respiration. Proc Natl Acad Sci USA 92:11949–11951.

2. Beratan DN, et al. (2009) Steering electrons in moving pathways. Acc Chem Res42:1669–1678.

3. Beratan DN, Betts JN, Onuchic JN (1991) Protein electron transfer rates set by the brid-ging secondary and tertiary structure. Science 252:1285–1288.

4. Marcus RA, Sutin N (1985) Electron transfers in chemistry and biology. Biochim BiophysActa 811:265–322.

5. Gray HB, Winkler JR (1996) Electron transfer in proteins. Annu Rev Biochem65:537–561.

6. Gray HB, Winkler JR (2005) Long-range electron transfer. Proc Natl Acad Sci USA102:3534–3539.

7. Gray HB, Malmstrom BG, Williams RJ (2000) Copper coordination in blue proteins.J Biol Inorg Chem 5:551–559.

8. Peterson-Kennedy SE, McGourty JL, Kalweit JA, Hoffman BM (1986) Temperaturedependence of and ligation effects on long-range electron transfer in complementary(zinc, ferric) hemoglobin hybrids. J Am Chem Soc 108:1739–1746.

9. Babini E, et al. (2000) Bond-mediated electron tunneling in ruthenium-modified high-potential iron-sulfur protein. J Am Chem Soc 122:4532–4533.

10. Winkler JR, Wittung-Stafshede P, Leckner J, Malmström BG, Gray HB (1997) Effects offolding on metalloprotein active sites. Proc Natl Acad Sci USA 94:4246–4249.

11. DiBilio AJ, et al. (1997) Reorganization energy of blue copper: Effects of temperatureand driving force on the rates of electron transfer in ruthenium- and osmium-modifiedazurins. J Am Chem Soc 119:9921–9922.

12. Farver O, Lu Y, Ang MC, Pecht I (1999) Enhanced rate of intramolecular electron trans-fer in an engineered purple CuA azurin. Proc Natl Acad Sci USA 96:899–902.

13. Lancaster KM, et al. (2011) Electron transfer reactivity of type zero Pseudomonasaeruginosa azurin. J Am Chem Soc 133:4865–4873.

14. Yanagisawa S, Banfield MJ, Dennison C (2006) The role of hydrogen bonding atthe active site of a cupredoxin: The Phe114Pro azurin variant. Biochemistry45:8812–8822.

15. Randall DW, Gamelin DR, LaCroix LB, Solomon EI (2000) Electronic structure contribu-tions to electron transfer in blue Cu and Cu(A). J Biol Inorg Chem 5:16–29.

16. Lumry R, Eyring H (1954) Conformation changes in proteins. J Phys Chem 58:110–120.17. Vallee BL, Williams RJP (1968) Metalloenzymes: The entatic nature of their active sites.

Proc Natl Acad Sci USA 59:498–505.18. Malmström BG (1965) Two forms of copper in copper-containing oxidases. Oxidases

and Related Redox Systems, eds TE King, HS Mason, and M Morrison (Wiley, NewYork), Vol 1, pp 207–216.

19. Malmström BG (1994) Rack-induced bonding in blue copper proteins. Eur J Biochem223:711–718.

20. Williams RJP (1995) Energised (entatic) states of groups and of secondary structures inproteins and metalloproteins. Eur J Biochem 234:363–381.

21. Ryde U, Olsson MHM, Pierloot K, Roos BO (1996) The cupric geometry of blue copperproteins is not strained. J Mol Biol 261:586–596.

22. Ryde U, OlssonMHM, Roos BO, De Kerpel JOA, Pierloot K (2000) On the role of strain inblue copper proteins. J Biol Inorg Chem 5:565–574.

9258 ∣ www.pnas.org/cgi/doi/10.1073/pnas.1119460109 Zaballa et al.

Page 6: Flexibility of the metal-binding region in apo-cupredoxins

23. Larsson S (2000) Energy saving electron pathways in proteins. J Biol Inorg Chem5:560–564.

24. Messerschmidt A, et al. (1998) Rack-induced metal binding vs. flexibility: Met121Hisazurin crystal structures at different pH. Proc Natl Acad Sci USA 95:3443–3448.

25. Buning C, et al. (2000) Loop-Directed mutagenesis of the blue copper protein amicya-nin from Paracoccus versutus and its effect on the structure and the activity of thetype-1 copper site. J Am Chem Soc 122:204–211.

26. LaCroix LB, et al. (1996) Electronic structure of the perturbed blue copper site in nitritereductase: Spectroscopic properties, bonding, and implications for the entatic/rackstate. J Am Chem Soc 118:7755–7768.

27. Nar H, Messerschmidt A, Huber R, van de KampM, Canters GW (1992) Crystal structureof Pseudomonas aeruginosa apo-azurin at 1.85A resolution. FEBS Lett 306:119–124.

28. Baker EN, et al. (1991) The relative rigidity of the type 1 copper site in azurin, as seen inhigh resolution X-ray analyses of various forms of the protein. J Inorg Biochem 43:162.

29. Garrett TP, Clingeleffer DJ, Guss JM, Rogers SJ, Freeman HC (1984) The crystal structureof poplar apoplastocyanin at 1.8-A resolution. The geometry of the copper-bindingsite is created by the polypeptide. J Biol Chem 259:2822–2825.

30. Petratos K, Papadovasilaki M, Dauter Z (1995) The crystal structure of apo-pseudoa-zurin from Alcaligenes faecalis S-6. FEBS Lett 368:432–434.

31. Durley R, Chen L, Lim LW, Mathews FS, Davidson VL (1993) Crystal structure analysis ofamicyanin and apoamicyanin from Paracoccus denitrificans at 2.0 A and 1.8 A resolu-tion. Protein Sci 2:739–752.

32. Williams PA, et al. (1999) The CuA domain of Thermus thermophilus ba3-type cyto-chrome c oxidase at 1.6 A resolution. Nat Struct Biol 6:509–516.

33. Nar H, Messerschmidt A, Huber R, van de KampM, Canters GW (1991) Crystal structureanalysis of oxidized Pseudomonas aeruginosa azurin at pH 5.5 and pH 9.0. J Mol Biol221:765–772.

34. Guss JM, Bartunik HD, Freeman HC (1992) Accuracy and precision in protein structureanalysis: Restrained least- squares refinement of the structure of poplar plastocyaninat 1.33 A resolution. Acta Crystallogr B 48:790–811.

35. Kissinger CR, Sieker LC, Adman ET, Jensen LH (1991) Refined crystal structure offerredoxin II from Desulfovibrio gigas at 1.7 A. J Mol Biol 219:693–715.

36. Fetrow JS, Baxter SM (1999) Assignment of 15N chemical shifts and 15N relaxationmeasurements for oxidized and reduced iso-1-cytochrome c. Biochemistry38:4480–4492.

37. Liu W, Rumbley JN, Englander SW, Wand AJ (2009) Fast structural dynamics in reducedand oxidized cytochrome c. Protein Sci 18:670–674.

38. Robinson NJ, Winge DR (2010) Copper metallochaperones. Annu Rev Biochem79:537–562.

39. Huffman DL, O’Halloran TV (2001) Function, structure, and mechanism of intracellularcopper trafficking proteins. Annu Rev Biochem 70:677–701.

40. Banci L, et al. (2006) The Atx1-Ccc2 complex is a metal-mediated protein–proteininteraction 10. Nat Chem Biol 2:367–368.

41. Abriata LA, et al. (2008) Mechanism of Cu(A) assembly. Nat Chem Biol 4:599–601.42. Van Amsterdam IM, et al. (2002) Dramatic modulation of electron transfer in protein

complexes by crosslinking. Nat Struct Biol 9:48–52.43. Shih C, et al. (2008) Tryptophan-accelerated electron flow through proteins. Science

320:1760–1762.44. Langen R, et al. (1995) Electron tunneling in proteins: Coupling through a beta strand.

Science 268:1733–1735.45. Soulimane T, et al. (2000) Structure and mechanism of the aberrant ba 3 -cytochrome

oxidase from Thermus thermophilus. EMBO J 19:1766–1776.

46. Maneg O, Ludwig B, Malatesta F (2003) Different interaction modes of two cyto-chrome-c oxidase soluble CuA fragments with their substrates. J Biol Chem278:46734–46740.

47. Slutter CE, et al. (1996) Water-Soluble, recombinant Cu A-domain of the cytochromeba 3 subunit II from Thermus thermophilus. Biochemistry 35:3387–3395.

48. Mukrash MD, et al. (2004) Complete 1H, 15N and 13C assignment of the solubledomain of the ba3 oxidase subunit II of Thermus thermophilus in the reduced state.J Biomol NMR 28:297–298.

49. Cordier F, et al. (1998) Solution structure, rotational diffusion anisotropy and localbackbone dynamics of Rhodobacter capsulatus cytochrome c2. J Mol Biol 281:341–361.

50. Lipari G, Szabo A (1982) Model-free approach to the interpretation of nuclear mag-netic resonance relaxation in macromolecules. 1. Theory and range of validity. J AmChem Soc 104:4546–4559.

51. Lipari G, Szabo A (1982) Model-free approach to the interpretation of nuclear mag-netic resonance relaxation in macromolecules. 2. Analysis of experimental results.J Am Chem Soc 104:4559–4570.

52. Loria JP, Rance M, Palmer AGI (1999) A relaxation-compensated Carr-Purcell-Meiboom-Gill sequence for characterizing chemical exchange by NMR spectroscopy.J Am Chem Soc 121:2331–2332.

53. Millet O, Loria JP, Kroenke CD, Pons M, Palmer AGI (2000) The static magnetic fielddependence of chemical exchange linebroadening defines the NMR chemical shifttimescale. J Am Chem Soc 122:2867–2877.

54. Schanda P, Forge V, Brutscher B (2006) HET-SOFAST NMR for fast detection of structuralcompactness and heterogeneity along polypeptide chains. Magn Reson Chem 44:S177–S184.

55. Korzhnev DM, Karlsson BG, Orekhov VY, Billeter M (2003) NMR detection of multipletransitions to low-populated states in azurin. Protein Sci 12:56–65.

56. van de Kamp M, et al. (1992) Complete sequential 1H and 15N nuclear magneticresonance assignments and solution secondary structure of the blue copper proteinazurin from Pseudomonas aeruginosa. Biochemistry 31:10194–10207.

57. Zhuravleva AV, et al. (2004) Gated electron transfers and electron pathways in azurin:A NMR dynamic study at multiple fields and temperatures. J Mol Biol 342:1599–1611.

58. Kalverda AP, et al. (1999) Backbone dynamics of azurin in solution: Slow confor-mational change associated with deprotonation of histidine 35. Biochemistry38:12690–12697.

59. Lancaster KM, Sproules S, Palmer JH, Richards JH, Gray HB (2010) Outer-sphereeffects on reduction potentials of copper sites in proteins: The curious case of highpotential type 2 C112D/M121E Pseudomonas aeruginosa azurin. J Am Chem Soc132:14590–14595.

60. Pozdnyakova I, Wittung-Stafshede P (2001) Biological relevance of metal bindingbefore protein folding. J Am Chem Soc 123:10135–10136.

61. Wittung-Stafshede P (2004) Role of cofactors in folding of the blue-copper proteinazurin. Inorg Chem 43:7926–7933.

62. Zong C, et al. (2007) Establishing the entatic state in folding metallated Pseudomonasaeruginosa azurin. Proc Natl Acad Sci USA 104:3159–3164.

63. Castruita M, et al. (2011) Systems biology approach in Chlamydomonas revealsconnections between copper nutrition and multiple metabolic steps. Plant Cell23:1273–1292.

64. Karlsson BG, Pascher T, Nordling M, Arvidsson RH, Lundberg LG (1989) Expression ofthe blue copper protein azurin from Pseudomonas aeruginosa in Escherichia coli. FEBSLett 246:211–217.

Zaballa et al. PNAS ∣ June 12, 2012 ∣ vol. 109 ∣ no. 24 ∣ 9259

CHEM

ISTR

YBIOCH

EMISTR

Y