12.07.2015 Views

download (pdf, 719 kb) - Institute of Pharmacology and Toxicology

download (pdf, 719 kb) - Institute of Pharmacology and Toxicology

download (pdf, 719 kb) - Institute of Pharmacology and Toxicology

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

1050 The 2.5 Å Structure <strong>of</strong> M. tuberculosis Rv2740Table 1. Kinetic characteristics <strong>of</strong> Rv2740 with different substratesSubstrate/enzyme V max (nmol mg K1 min K1 ) K m (mM) k cat /K m (M K1 s K1 )Styrene 7,8-oxide (LEH) n.d. (1700G250) O5000 (1400G280) 3G0.6 (340G30)Cholesterol 5,6-oxide 1.1G0.4 60G30 5.5G0.59,10-Epoxystearic acid 30G10 45G15 210G65V max for styrene 7,8-oxide is recorded as not determined (n.d.); saturation was not obtained at substrate concentrations as high as 5 mM,<strong>and</strong> so only k cat /K m could be estimated with confidence. Values measured for LEH with the same substrate 5 are shown in parenthesesfor comparison. No turnover could be detected for LEH with cholesterol-5,6-oxide <strong>and</strong> 9,10-epoxystearic acid as substrates.Table 2. Data collection <strong>and</strong> refinement statisticsNative dataSe-Met dataA. Data collectionBeamline (wavelength, Å) MaxLab 711 (1.134) ESRF ID14-EH2 (0.934)Resolution range (Å) 30.0–3.0 (3.05–3.0) 70.9–2.50 (2.64–2.50)Unit cell dimensions (Å) 80.6, 80.6, 118.7 81.9, 81.9, 117.0Number <strong>of</strong> observed/unique reflections 70,441/9338 348,194/16,247Completeness (%) 99.8 (99.8) 100.0 (100.0)Multiplicity/anomalous multiplicity 7.5 (7.5)/– 21.2 (21.5)/10.8 (11.4)R merge 0.042 (0.136) 0.095 (0.33)hIi/hs(I)i 26.4 (7.7) 29.0 (8.8)B. RefinementR cryst (%) – 22.2aR free – 25.4Number <strong>of</strong> atoms (Average B (Å 2 )) – 3285 (42.1)Number <strong>of</strong> waters (Average B (Å 2 )) – 90 (35.9)Bond RMSD from ideal values b (Å) – 0.023Angle RMSD from ideal values b (deg.) – 1.96Ramach<strong>and</strong>ran plot outliers c (%) – 0.83Values in parentheses are for the highest resolution shell.a R free , calculated from a r<strong>and</strong>omly chosen 5% <strong>of</strong> unique reflections.b Ideal values from Engh & Huber. 36c Defined according to Kleywegt & Jones. 37Overall structureStatistics for the X-ray data <strong>and</strong> the final refinedmodel <strong>of</strong> the selenomethionine (Se-Met) substitutedform <strong>of</strong> Rv2740 are summarized in Table 2. Thestructure is characterized by a curved six-str<strong>and</strong>edb-sheet that packs against three antiparallela-helices, creating a wide pocket lined primarilywith hydrophobic side-chains. A fourth long helixcompletes the fold (Figure 2). The asymmetric unitcontains three chemically equivalent moleculeswith very similar conformations; residues 16–140<strong>of</strong> each chain can be superimposed with a rootmean-square(RMS) distance <strong>of</strong> less than 0.1 Å.Two<strong>of</strong> the molecules form a dimer with subunits relatedby a nearly perfect non-crystallographic (NCS)2-fold axis. The third forms a dimer with aneighboring molecule, its 2-fold rotational axiscoinciding with a crystallographic dyad.The subunit interface <strong>of</strong> each dimer has a totalburied surface area <strong>of</strong> 2000 Å 2 , within the rangeexpected for homodimers (1700 (G1100) Å 2 ). 8 Threemajor segments are involved in quaternary interactionsbetween the two subunits. W105, Y123,W105 0 <strong>and</strong> Y123 0 form a hydrophobic surface; theprime indicates that a residue belongs to theadjacent polypeptide chain. Two sets <strong>of</strong> doublebidentate salt-bridges, R79–E111 0 –E90–R121 0 , alsostabilize the interaction. Although C107 <strong>and</strong> C107 0are close enough to form a disulfide bridge acrossthe dimer interface, they are clearly reduced in thecrystal structure, contributing instead to the hydrophobicinteractions between the subunits.Figure 2. The Rv2740 dimer. The two subunits are colorcodedgoing from red to blue, beginning with the Nterminus <strong>of</strong> the A molecule <strong>and</strong> ending with the Cterminus <strong>of</strong> the B molecule.


The 2.5 Å Structure <strong>of</strong> M. tuberculosis Rv2740 1051Comparison with known structuresA DALI search 9 <strong>of</strong> the Protein Data Bank (PDB 10 ),using the Rv2740 A subunit as a probe, identifiedseveral structural homologues (Table 3). Rathersurprisingly, the most similar structure is that <strong>of</strong>Bal32a, 11 a thermostable protein <strong>of</strong> unknownfunction from an unidentified bacterium. Strongsimilarity to Rv2740 is observed in both the subunit<strong>and</strong> dimer. However, Bal32a has a long loop thatcovers the entrance <strong>of</strong> the active-site pocket, turningit into a closed cavity. Robinson et al. 11 speculatethat the interior <strong>of</strong> the protein may be accessible tosubstrates via movements <strong>of</strong> a flexible cap. As forRv2740, a pair <strong>of</strong> cysteine residues is observed in thereduced state at the subunit–subunit interface in thecrystal structure. Oxidation results in an increase inthe melting temperature <strong>of</strong> Bal32a from 65 8C to80 8C, strongly suggesting a role for the disulfide instabilization <strong>of</strong> its dimer.As expected, Rv2740 was found to be similar instructure to LEH. 5 Similarities were also noted to D5-3-ketosteroid isomerase, 12 scytalone dehydratase 13,14<strong>and</strong> a carotenoid-binding protein <strong>of</strong> unknownfunction. 15 The D5-3-ketosteroid isomerase forms adimer with a similar construction to that <strong>of</strong> Rv2740,LEH <strong>and</strong> Bal32a. The scytalone dehydratase forms atrimer along similar principles, while the correspondingregion <strong>of</strong> the carotenoid-binding protein representsa single domain <strong>of</strong> a larger protein.Active siteThe three residues mainly responsible forcatalysis in LEH have been identified through sitedirectedmutagenesis. 5 In the mycobacterial epoxidehydrolase, these side-chains correspond to R91,D93 <strong>and</strong> D122, which are situated at the bottom <strong>of</strong>the active-site pocket (Figure 3(a)). Two otherresidues, Y46 <strong>and</strong> N48 (in Rv2740 numbering),have been suggested to be associated with the activewater molecule <strong>of</strong> the hydrolase reaction. The activesites <strong>of</strong> LEH <strong>and</strong> the mycobacterial enzyme havealmost identical hydrogen-bonding networks, theonly exception being N55 in LEH, which isstabilized by a single hydrogen bond. The correspondingside-chain <strong>of</strong> Rv2740, N48, is coordinatedboth by Y46 <strong>and</strong> S52. In Bal32a, the equivalents <strong>of</strong>R91, D93 <strong>and</strong> D122 are changed to H105, N107 <strong>and</strong>E136, respectively, suggesting that its function isdifferent from that <strong>of</strong> Rv2740, Furthermore, thecounterparts <strong>of</strong> Y46 <strong>and</strong> N48 are completely absent.Although the catalytic residues <strong>of</strong> LEH <strong>and</strong>Rv2740 are very similar, the shape <strong>of</strong> the pocketthat contains them is quite different (Figure 3(b)).The deeper substrate-binding site <strong>of</strong> the mycobacterialstructure is in part explained by the differingorientations <strong>of</strong> L35 <strong>and</strong> L28, as well as I80 <strong>and</strong> F73,in the LEH <strong>and</strong> Rv2740 structures, respectively.Substitutions such as M78 to V71 <strong>and</strong> L74 to M67,combined with the different packing <strong>of</strong> the third<strong>and</strong> fourth helices, make room in the mycobacterialenzyme for substrates bulkier than limonene-1,2-epoxide. The side-chains <strong>of</strong> L95, M129, I97, L140,V138, L137, L134 <strong>and</strong> F130 form a large solventexposedhydrophobic patch at the upper part <strong>of</strong> thepocket; this feature seems to be well conservedthroughout Mycobacteriaceae.As in LEH, an endogenous lig<strong>and</strong> <strong>of</strong> unknownorigin was found in the active-site cavity <strong>of</strong> each <strong>of</strong> thethree molecules in the asymmetric unit, bound close tothe catalytic residues (Figure 3(c) <strong>and</strong> (d)). The electrondensity is essentially identical in averaged <strong>and</strong>unaveraged maps at 2.5 Å resolution, suggesting asingle compound with a single well-defined bindingmode. The character <strong>of</strong> this lig<strong>and</strong> is markedlydifferent from the heptanamide modeled in the LEHcase, being more consistent with a cyclohexane ringwith a single attached aliphatic chain. The shape <strong>of</strong> thedensity does not resemble that <strong>of</strong> any chemical usedin either purification or crystallization.Related sequencesIn addition to the group <strong>of</strong> sequences previouslyidentified as potential LEHs, 5 at least four moresequences from actinobacteria <strong>and</strong> proteobacteriahave a highly conserved catalytic center (Figure 4).A few differences from the originally defined LEHfamily are evident. For example, Y46 in themycobacterial enzyme is replaced by a tryptophanin both Ralstonia solanacearum (GenBank accessionnumber 17547309; partial sequence) <strong>and</strong> Erwiniacaratova (50120522). The latter protein also lacks ahydrogen bond between N48 <strong>and</strong> S52 (Rv2740numbering), leaving the residue corresponding toN48 solely responsible for positioning the putativecatalytic water molecule.Four additional sequences are more distantlyrelated, but likely to have similar structure. TheTable 3. Comparisons with DALIPDB entryZRMS differenceResiduesalignedLength <strong>of</strong>sequence Identical (%) Protein1TUH 15.8 2.7 122 131 19 Bal32a1OPY 14.9 2.1 115 123 16 D5-3-Ketosteroid isomerase1NU3 13.9 2.2 117 141 28 LEH3STD 13.3 2.4 121 162 5 Scytalone dehydratase1M98 12.0 2.2 106 316 11 Unknown function, orangecarotenoid proteinOnly hits with Z-scores greater than 12 are listed.


1052 The 2.5 Å Structure <strong>of</strong> M. tuberculosis Rv2740Figure 3. The active site. (a) Residues lining the cavity <strong>of</strong> Rv2740. (b) Comparison <strong>of</strong> the size <strong>of</strong> the substrate-bindingpockets <strong>of</strong> Rv2740 (blue) <strong>and</strong> LEH (red). Solvent-accessible surfaces were generated in O, using a probe radius <strong>of</strong> 1.3 Å.Residues <strong>of</strong> Rv2740 are labeled. (c) <strong>and</strong> (d) The Rv2740 active site with electron density for endogenous lig<strong>and</strong> (12F obs KF calc j contoured at 0.4 eÅ 3 ), using views that are similar to those in (a) <strong>and</strong> (b), respectively.protein from Burkholderia cepacia (46319940), as well astwo from the Mycobacterium avium subspecies paratuberculosis(41407700 <strong>and</strong> 41409656), is missing one <strong>of</strong>the catalytic aspartate residues (D93). In the M. tuberculosisRv0141c (15607283), as well as both <strong>of</strong> theM. avium sequences, R91 is substituted by histidine orasparagine. All four have replaced N48 with variousother hydrophilic residues, suggesting that theirfunction may be different from that <strong>of</strong> Rv2740.DiscussionStudies <strong>of</strong> LEH suggest that it uses a single-step,push–pull mechanism that is distinct from thetwo-step mechanism used by the a/b-hydrolaseclass <strong>of</strong> EHs. 5 The single-step mechanism featuresan Asp-Arg-Asp triad, i.e. residues R91, D93 <strong>and</strong>D122 in Rv2740 (Figure 1(c)). We suggest that D122activates a water molecule for its attack on theepoxide ring, while the ring is simultaneouslyencouraged to open by a proton donated by D93.The arginine side-chain interacts with <strong>and</strong> orientsthe carboxylate groups <strong>of</strong> the catalytic aspartateresidues, while Y46 <strong>and</strong> N48 help position thecatalytic water.We show here that Rv2740 can hydrolyze threestructurally diverse epoxides, albeit with moderatecatalytic efficiency (Table 1). If Rv2740, like LEH,metabolizes carbon compounds obtained from thesurroundings, it would be expected to have a higheractivity on one or more preferred substrates. This isin contrast to the broad specificity <strong>and</strong> slowerreaction rates associated with a function in detoxification(a major function <strong>of</strong> mammalian a/b EHs).Obviously, the physiological substrates for Rv2740have yet to be identified, <strong>and</strong> the observedefficiency on these non-optimal substrates is underst<strong>and</strong>ablylower. The hydrophobic character <strong>of</strong> thesubstrate binding cavity <strong>and</strong> the low K m value withepoxystearic acid <strong>and</strong> cholesterol 5,6-epoxidesuggest that the most relevant substrates may befatty acid or steroid derivatives. However, themetabolism <strong>of</strong> such compounds in M. tuberculosisremains largely unexplored. In this context, theobserved electron density for an endogenous lig<strong>and</strong>


Figure 4. Structure-based sequence alignment <strong>of</strong> Rv2740 <strong>and</strong> LEH, together with alignment <strong>of</strong> other related sequences. The last five residues <strong>of</strong> the first M. avium sequence areomitted for simplicity. Asterisks denote residues lining the active-site cavity.


1054 The 2.5 Å Structure <strong>of</strong> M. tuberculosis Rv2740in the active site <strong>of</strong> Rv2740 is extremely interesting.Although the compound probably originated fromthe E. coli expression host, it provides leads for thetask <strong>of</strong> identifying the true substrates for Rv2740.This lig<strong>and</strong> could only be partially displaced byvalpromide (K I w100 mM), even at the highestsoluble concentration, 10 mM; its apparent persistencethroughout purification also indicates tightbinding. The shape <strong>of</strong> the electron density issuggestive <strong>of</strong> a cyclohexane ring with an attachedaliphatic chain (Figure 3(c) <strong>and</strong> (d)). Limoneneitself, one <strong>of</strong> the physiological substrates <strong>of</strong> LEH,has some similarities in shape (Figure 1), but thewider site <strong>of</strong> Rv2740 (Figure 3(b)) would allow alarger substrate. The relevant portion <strong>of</strong> thecarotenoid in the related structure describedabove (Table 3) is placed in a very similar location<strong>and</strong> orientation within that binding-site pocket.However, residues lining the corresponding site inRv2740 are not conserved; indeed, sequence identitybetween the two proteins is very low, <strong>and</strong> thesimilarities could only be uncovered based oncomparisons <strong>of</strong> the three-dimensional structures.The ability <strong>of</strong> Rv2740 to hydrolyze cholesterol 5,6-epoxide is intriguing for reasons <strong>of</strong> broader interest.On the basis <strong>of</strong> an analysis <strong>of</strong> LEH, we havepreviously speculated that it could be a homologue<strong>of</strong> the human cholesterol epoxide hydrolase, a keyenzyme in the metabolism <strong>of</strong> cholesterol oxidationproducts which has, so far, evaded attempts atpurification <strong>and</strong> cloning. 5 Although LEH itself didnot show any turnover with cholesterol epoxide, thepresent study shows that the wider active site <strong>of</strong>Rv2740 does allow such a reaction to occur.Another interesting finding is the high enantioselectivity<strong>of</strong> Rv2740 with a fatty acid epoxidesubstrate. Although the enantioselectivity <strong>of</strong> thisclass <strong>of</strong> enzymes has been shown for the rigidsubstrate limonene 1,2-epoxide, 16 the phenomenonis rather unexpected with 9,10-epoxystearic acid,since the groups that could direct such selectivityare quite distant from the catalytic center (Figure 1).In recent years, there has been a strong interest inepoxide hydrolases for the production <strong>of</strong> enantiopureepoxides <strong>and</strong> diols, mainly as building blocksin the fine chemical <strong>and</strong> drug industries. 17 Indeed,mining for such novel enzymes was a goal in thestudies with Bal32a referred to here 11 as well assimilar ones with a/b-type epoxide hydrolases. 18Given its unusual properties, Rv2740 will provide auseful addition to the toolbox <strong>of</strong> enzymes availablefor biocatalytic production <strong>of</strong> fine chemicals.Experimental ProceduresRadiolabeled compounds[26- 14 C]Cholesterol 5,6-oxide (53 mCi/mmol), racemic[1- 14 C]9,10-epoxystearic acid (56 mCi/mmol) <strong>and</strong> racemic[ 14 C]styrene 7,8-oxide (15 mCi/mmol) were synthesized asdescribed. 19Cloning, over-expression <strong>and</strong> purificationThe sequence corresponding to the open reading frame<strong>of</strong> the Rv2740 gene was amplified by PCR from total DNA<strong>of</strong> M. tuberculosis strain H37Rv, using the high-fidelitypolymerase Pfu Turbo (Stratagene) with the primersATGGCTCATCATCATCATCATTGTGCCGAGCTGACCGAAACATC (forward) <strong>and</strong> CTACAGCGTTGCCTTCAGCGATG (reverse). The forward primer includedthe sequence for a histidine-tag immediately upstream <strong>of</strong>the gene. After addition <strong>of</strong> a 3 0 A overhang by incubationwith Taq polymerase (Roche), the DNA fragment wasligated into the pCR T7 TOPO vector (Invitrogen), <strong>and</strong>used to transform E. coli TOP10 cells (Invitrogen); theconstruct was verified by DNA sequence analysis.Expression was performed in E. coli Rosetta (DE3) cellswith a pLacI plasmid (Novagen) or in E. coli BL21-AI cells(Invitrogen). The cells were cultured in LB medium at37 8C. At A 600 Z0.5–1.0, the temperature was lowered to24 8C <strong>and</strong> expression <strong>of</strong> the target gene induced withIPTG (100 mg/l) for three hours. After harvesting bycentrifugation, the cell pellet was resuspended in lysisbuffer (50 mM NaH 2 PO 4 , 300 mM NaCl, 10 mM imidazole,10%(v/v) glycerol, 0.5%(v/v) Triton X-100 (pH 8.0))with 1 mg/ml <strong>of</strong> lysozyme, 1 mM PMSF, 0.01 mg/ml <strong>of</strong>RNase A <strong>and</strong> 0.02 mg/ml <strong>of</strong> DNase I, <strong>and</strong> lysed using aOne Shot cell disruptor (Constant Systems Ltd.). Thesoluble fraction was incubated with 1 ml <strong>of</strong> pre-equilibratedNi-NTA agarose (Qiagen) slurry at 4 8C for onehour, then poured into a column. After washing with 20column volumes <strong>of</strong> buffer (20 mM imidazole, 50 mMNaH 2 PO 4 , 300 mM NaCl, 10% glycerol (pH 8.0)), theprotein was eluted with four column volumes <strong>of</strong> 250 mMimidazole in the same buffer. Fractions containing theprotein were pooled <strong>and</strong> further purified on a HiLoad 16/60 Superdex 75 prep grade column (Amersham Biosciences)equilibrated in 20 mM Tris–HCl (pH 7.5),150 mM NaCl, 10% glycerol. The protein eluted in twopeaks, apparently a monomer <strong>and</strong> a dimer. The twosamples were concentrated separately to 4–5 mg/mlusing Vivaspin 6 concentrators (10,000 MWCO).For expression <strong>of</strong> the Se-Met labeled protein, it wasnecessary to efficiently suppress the background level <strong>of</strong>expression. Therefore, the gene was transferred to thepET101D vector (Invitrogen) <strong>and</strong> the expression performedin BL21-AI cells. Se-Met was introduced into theprotein by metabolic inhibition. 20 Fermentation wasperformed in minimal medium supplemented withSe-Met, lysine, threonine, phenylalanine, leucine, isoleucine<strong>and</strong> valine according to published procedures.Expression was performed at room temperature overnight.The purification was carried out as described forthe native protein, except that 10 mM b-mercaptoethanolwas included in all buffers to prevent Se-Met oxidation.Enzyme assaysThe enzymatic activity <strong>of</strong> native Rv2740 with racemic[ 14 C]styrene 7,8-oxide as the substrate was determinedessentially as described. 5 For the assay with [1- 14 C]-cis-9,10-epoxystearic acid, 10–15 mg <strong>of</strong> pure enzyme wasincubated with the substrate in 100 mM sodium phosphate,(pH 7.4), 50 mM NaCl in a total volume <strong>of</strong> 50 ml<strong>and</strong> incubated for 15 minutes at 37 8C. The substrate wasadded from a stock solution in acetonitrile, resulting in afinal solvent concentration in the assay mixture <strong>of</strong> 2%; theorganic solvent has no effect on the enzyme activity. Thereaction was terminated <strong>and</strong> 200 ml <strong>of</strong> ethylacetate was


1056 The 2.5 Å Structure <strong>of</strong> M. tuberculosis Rv27405. Ar<strong>and</strong>, M., Hallberg, B. M., Zou, J., Bergfors, T.,Oesch, F., van der Werf, M. J. et al. (2003). Structure <strong>of</strong>Rhodococcus erythropolis limonene-1,2-epoxide hydrolasereveals a novel active site. EMBO J. 22, 2583–2592.6. van der Werf, M. J., Swarts, H. J. & de Bont, J. A.(1999). Rhodococcus erythropolis DCL14 contains anovel degradation pathway for limonene. Appl.Environ. Microbiol. 65, 2092–2102.7. Pacifici, G. M., Franchi, M., Bencini, C. & Rane, A.(1986). Valpromide inhibits human epoxide hydrolase.Br. J. Clin. Pharmacol. 22, 269–274.8. Jones, S. & Thornton, J. M. (1996). Principles <strong>of</strong> protein–protein interactions. Proc. Natl Acad. Sci. USA, 93, 13–20.9. Holm, L. & S<strong>and</strong>er, C. (1993). Protein structurecomparison by alignment <strong>of</strong> distance matrices. J.Mol. Biol. 233, 123–138.10. Berman, H. M., Westbrook, J., Feng, Z., Gillil<strong>and</strong>, G.,Bhat, T. N., Weissig, H. et al. (2000). The Protein DataBank. Nucl. Acids Res. 28, 235–242.11. Robinson, A., Wu, P. S., Harrop, S. J., Schaeffer, P. M.,Dosztanyi, Z., Gillings, M. R. et al. (2005). Integronassociatedmobile gene cassettes code for folded proteins:the structure <strong>of</strong> Bal32a, a new member <strong>of</strong> the adaptablealphaCbeta barrel family. J. Mol. Biol. 346, 1229–1241.12. Wu, Z. R., Ebrahimian, S., Zawrotny, M. E.,Thornburg, L. D., Perez-Alvarado, G. C., Brothers, P.et al. (1997). Solution structure <strong>of</strong> 3-oxo-delta5-steroidisomerase. Science, 276, 415–418.13. Lundqvist, T., Rice, J., Hodge, C. N., Basarab, G. S.,Pierce, J. & Lindqvist, Y. (1994). Crystal structure <strong>of</strong>scytalone dehydratase–a disease determinant <strong>of</strong> the ricepathogen, Magnaporthe grisea. Structure, 2, 937–944.14. Chen, J. M., Xu, S. L., Wawrzak, Z., Basarab, G. S. &Jordan, D. B. (1998). Structure-based design <strong>of</strong> potentinhibitors <strong>of</strong> scytalone dehydratase: displacement <strong>of</strong> awater molecule from the active site. Biochemistry, 37,17735–17744.15. Kerfeld, C. A., Sawaya, M. R., Brahm<strong>and</strong>am, V., Cascio,D., Ho, K. K., Trevithick-Sutton, C. C. et al. (2003). Thecrystal structure <strong>of</strong> a cyanobacterial water-solublecarotenoid binding protein. Structure, 11, 55–65.16. van der Werf, M. J., Orru, R. V. A., Overkamp, K. M.,Swarts, H. J., Osprian, I., Steinreiber, A. et al. (1999).Substrate specificity <strong>and</strong> stereospecificity <strong>of</strong> limonene-1,2-epoxidehydrolase from Rhodococcus erythropolisDCL14; an enzyme showing sequential <strong>and</strong>enantioconvergent substrate conversion. Appl. Microbiol.Biotechnol. 52, 380–385.17. Archelas, A. & Furstoss, R. (2001). Synthetic applications<strong>of</strong> epoxide hydrolases. Curr. Opin. Chem. Biol. 5, 112–119.18. Zhao, L., Han, B., Huang, Z., Miller, M., Huang, H.,Malashock, D. S. et al. (2004). Epoxide hydrolasecatalyzedenantioselective synthesis <strong>of</strong> chiral 1,2-diolsvia desymmetrization <strong>of</strong> meso-epoxides. J. Am. Chem.Soc. 126, 11156–11157.19. Müller, F., Ar<strong>and</strong>, M., Frank, H., Seidel, A., Hinz, W.,Winkler, L. et al. (1997). Visualization <strong>of</strong> a covalentintermediate between microsomal epoxide hydrolase,but not cholesterol epoxide hydrolase, <strong>and</strong> theirsubstrates. Eur. J. Biochem. 245, 490–496.20. Van Duyne, G. D., St<strong>and</strong>aert, R. F., Karplus, P. A.,Schreiber, S. L. & Clardy, J. (1993). Atomic structures<strong>of</strong> the human immunophilin FKBP-12 complexes withFK506 <strong>and</strong> rapamycin. J. Mol. Biol. 229, 105–124.21. Page, R., Grzechnik, S. K., Canaves, J. M., Spraggon, G.,Kreusch, A., Kuhn, P. et al. (2003). Shotgun crystallizationstrategy for structural genomics: an optimized twotieredcrystallization screen against the Thermotogamaritima proteome. Acta Crystallog. sect. D, 59, 1028–1037.22. Otwinowski, Z. & Minor, W. (1997). Processing <strong>of</strong>X-ray diffraction data collected in oscillation mode.Methods Enzymol. 276, 307–326.23. Matthews, B. W. (1968). Solvent content <strong>of</strong> proteincrystals. J. Mol. Biol. 33, 491–497.24. Knight, S. D. (2000). RSPS version 4.0: a semiinteractivevector-search program for solving heavyatomderivatives. Acta Crystallog. sect. D, 56, 42–47.25. de La Fortelle, E. & Bricogne, G. (1997). Maximumlikelihoodheavy-atom refinement in the MIR <strong>and</strong>MAD methods. Methods Enzymol. 276, 472–494.26. Jones, T. A., Zou, J.-Y., Cowan, S. W. & Kjeldgaard, M.(1991). Improved methods for building protein modelsin electron density maps <strong>and</strong> the location <strong>of</strong> errors inthese models. Acta Crystallog. sect. A, 47, 110–119.27. Kleywegt, G. J. & Jones, T. A. (1994). Halloween.masks <strong>and</strong> bones. In From First Map to Final Model(Bailey, S., Hubbard, R. & Waller, D., eds), pp. 59–66,SERC Daresbury Laboratory, Warrington, UK.28. Cowtan, K. & Main, P. (1998). Miscellaneous algorithmsfor density modification. Acta Crystallog. sect. D, 54, 487–493.29. Jones, T. A. (2004). Interactive electron-density mapinterpretation: from INTER to O. Acta Crystallog. sect.D, 60, 2115–2125.30. Brünger, A. T., Adams, P. D., Clore, G. M., DeLano, W.L., Gros, P., Grosse-Kunstleve, R. W. et al. (1998).Crystallography <strong>and</strong> NMR system (CNS): a news<strong>of</strong>tware suite for macromolecular structure determination.Acta Crystallog. sect. D, 54, 905–921.31. Murshudov, G. N., Vagin, A. A. & Dodson, E. J. (1997).Refinement <strong>of</strong> macromolecular structures by themaximum-likelihood method. Acta Crystallog. sect.D, 53, 240–255.32. Read, R. J. (1986). Improved Fourier coefficients formaps using phases from partial structures witherrors. Acta Crystallog. sect. A, 42, 140–149.33. Perrakis, A., Sixma, T. K., Wilson, K. S. & Lamzin, V. S.(1997). wARP: improvement <strong>and</strong> extension <strong>of</strong>crystallogaphic phases by weighted averaging <strong>of</strong>multiple-refined dummy atomic models. Acta Crystallog.sect. D, 53, 448–455.34. Lopez, R., Silventoinen, V., Robinson, S., Kibria, A. &Gish, W. (2003). WU-Blast2 server at the EuropeanBioinformatics <strong>Institute</strong>. Nucl. Acids Res. 31, 3795–3798.35. Harris, M. & Jones, T. A. (2001). Molray - a webinterface between O <strong>and</strong> the POV-Ray ray tracer. ActaCrystallog. sect. D, 57, 1201–1203.36. Engh, R. A. & Huber, R. (1991). Accurate bond <strong>and</strong>angle parameters for X-ray protein structure refinement.Acta Crystallog. sect. A, 47, 392–400.37. Kleywegt, G. J. & Jones, T. A. (1996). Phi/Psi-cology:Ramach<strong>and</strong>ran revisited. Structure, 4, 1395–1400.Edited by M. Guss(Received 29 April 2005; received in revised form 17 June 2005; accepted 21 June 2005)Available online 7 July 2005

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!