BioMedNet HomeLibraryDatabasesCollaborationsJob ExchangeShopping MallYour Room

How coenzyme B12 radicals are generated: the crystal structure of methylmalonyl-coenzyme A mutase at 2 A resolutionFULL TEXT (TEXT ONLY)
Structure
Vol. 4, No. 3, 15 March 1996

Viewing options: [text only] [+ full figures] [PDF]
You can obtain a free PDF viewer from Adobe Systems Inc.



How coenzyme B12 radicals are generated: the crystal structure of methylmalonyl-coenzyme A mutase at 2 Å resolution
[Research Article]
Filippo Mancia, Nicholas H Keep, Atsushi Nakagawa, Peter F Leadlay, Sean McSweeney, Bjarne Rasmussen, Peter Bösecke, Olivier Diat, Philip R Evans
Structure 1996, 4:339-350.

Outline


Abstract

Backgound

The enzyme methylmalonyl-coenzyme A (CoA) mutase, an alphabeta heterodimer of 150 kDa, is a member of a class of enzymes that uses coenzyme B12 (adenosylcobalamin) as a cofactor. The enzyme induces the formation of an adenosyl radical from the cofactor. This radical then initiates a free-radical rearrangement of its substrate, succinyl-CoA, to methylmalonyl-CoA.


Results

Reported here is the crystal structure at 2 Å resolution of methylmalonyl-CoA mutase from Propionibacterium shermanii in complex with coenzyme B12 and with the partial substrate desulpho-CoA (lacking the succinyl group and the sulphur atom of the substrate). The coenzyme is bound by a domain which shares a similar fold to those of flavodoxin and the B12-binding domain of methylcobalamin-dependent methionine synthase. The cobalt atom is coordinated, via a long bond, to a histidine from the protein. The partial substrate is bound along the axis of a (beta/alpha)8 TIM barrel domain.


Conclusion

The histidine–cobalt distance is very long (2.5 Å compared with 1.95–2.2 Å in free cobalamins), suggesting that the enzyme positions the histidine in order to weaken the metal–carbon bond of the cofactor and favour the formation of the initial radical species. The active site is deeply buried, and the only access to it is through a narrow tunnel along the axis of the TIM barrel domain.



Keywords


Introduction

Adenosylcobalamin, or coenzyme B12 ( Fig. 1a ), has fascinated chemists and biologists ever since its structure was determined by Lenhert and Hodgkin [1] over three decades ago. The principal interest in this molecule stems from the fact that it contains one of the very few naturally occurring metal–carbon bonds, a feature shared with the related molecule, methylcobalamin. The weakness of this bond between the cobalt atom and the 5'-carbon of the 5'-deoxyadenosyl group provides the special properties of this enzyme cofactor. The generally accepted mechanism of adenosylcobalamin-containing enzymes is that the CoIII–C bond splits homolytically to form the 5'-deoxyadenosyl radical, which initiates the enzyme reaction, leaving a pentacoordinated CoII atom in the cobalamin (cob(II)alamin or B12r). (For a review, see [2].)

View Image

Figure 1 Coenzyme B12 and reaction schemes. (a) 5'-Deoxyadenosyl-cobalamin (coenzyme B12), showing part of the atom nomenclature scheme. The dashed lines refer to the bonds between the central cobalt atom and two axial coordinating groups. In the free coenzyme, the 5'-C atom of 5'-deoxyadenosine provides upper axial coordination, while lower axial coordination comes from the N3B atom of a dimethylbenzimidazole linked to the porphyrin-like corrin ring. (b) Minimal mechanistic scheme for coenzyme B12-dependent rearrangements [2]. AdoCH2-B12 is 5'-deoxyadenosyl-cobalamin, Ado-CH2 and Ado-CH3 are respectively the 5'-deoxyadenosyl radical and 5'-deoxyadenosine. X is a generic migrating group. (c) The rearrangement reaction catalyzed by methylmalonyl-CoA mutase. The thioester group (O=C–CoA) is the group that migrates (CoA terminates with a sulphur).

Ten adenosylcobalamin-dependent enzymes have been identified [3]. All but one of these, a ribonucleotide reductase, catalyze unusual 1,2 rearrangements. These are all reactions in which a carbon-bonded hydrogen is exchanged with a group on an adjacent carbon (H and X in Fig. 1b ). The reaction is thought to involve the adenosyl radical abstracting a hydrogen from the substrate to form a substrate radical [2] ( Fig. 1b ). This undergoes rearrangement to form the product radical, which then abstracts a hydrogen from the adenosine to form the product and to re-form the adenosyl radical. The evidence for this mechanism comes mainly from direct spectroscopic observation of free-radical intermediates by electron paramagnetic resonance (EPR) spectroscopy and from experiments with labelled substrates and cofactor [4] [5] [6] [7] [8] [9].

Methylmalonyl-coenzyme A (CoA) mutase is one of the most studied enzymes of this class and is the only one which is present in both mammals and bacteria. In the Gram-positive bacterium Propionibacterium shermanii, methylmalonyl-CoA mutase is involved in the fermentation of pyruvate to propionate, whereas in mammalian liver mitochondria the enzyme is responsible for the conversion of odd-chain fatty acids and branched-chain amino acids via propionyl-CoA to succinyl-CoA for further degradation. The enzyme catalyzes the interconversion between (2R)-methylmalonyl-CoA and succinyl-CoA ( Fig. 1c ) by the migration of the O=C–CoA group [10] [11].

The P. shermanii enzyme is an alphabeta heterodimer of 150 kDa [12] (80 kDa for 728 residues in the alpha chain and 70 kDa for 638 residues in the beta chain [13]), with one binding site per dimer for acyl-CoA [14] and for B12 [15]. In contrast, the human enzyme is an alpha2 homodimer [16], and is highly homologous (60% sequence identity) to the active alpha chain of its bacterial counterpart. The bacterial alpha and beta chains show a lesser degree of homology, with 24% of the residues being identical.

There is one other B12 cofactor, methylcobalamin, in which the carbon of a methyl group serves as the upper axial ligand to the cobalt: this is the cofactor for methionine synthase. In this case the cobalt–carbon bond is cleaved heterolytically, and the cobalt is reduced from a CoIII oxidation state to a highly reactive CoI species, while the methyl group is transferred to the substrate homocysteine to form methionine [17]. The structure of the inactive B12-binding domain of this enzyme has been solved recently [18]. It showed for the first time how the dimethylbenzimidazole base that coordinates the cobalt in the free methylcobalamin is substituted on the enzyme by a histidine residue. Although sequence comparisons reveal relatively few conserved residues [19], extensive structural similarities exist between the B12-binding domains of methionine synthase and methylmalonyl-CoA mutase.

We present the crystal structure to 2 Å resolution of the ternary complex between recombinant methylmalonyl-CoA mutase, coenzyme B12 and the partial substrate desulpho-CoA (CoA with the terminal thiol group replaced by a hydrogen). The implications of the structure for the binding of substrate and cofactor, and for the reaction mechanism are discussed.



Results and discussion


Structure determination and refinement

Crystals of methylmalonyl-CoA mutase with bound coenzyme B12 and in presence of an excess of desulpho-CoA grow under identical conditions in two morphologically indistinguishable and closely related forms, one monoclinic, the other orthorhombic. Each asymmetric unit contains two alphabeta heterodimers of methylmalonyl-CoA mutase. The structure determination proved troublesome, in part because of the weak diffraction of these crystals, but mostly because of poor isomorphism between native crystals. All data were collected from crystals frozen at 95 K. Cryo-cooling was essential both to minimize the radiation damage caused by the high doses of X-ray radiation needed to record the high-resolution reflections, and to obtain complete data sets from single crystals thereby overcoming the problems of non-isomorphism. As one approach to phasing, multiwavelength anomalous diffraction (MAD) data were collected from a single orthorhombic crystal soaked in K2Pt(NO2)4. This experiment alone was not sufficient to produce an interpretable map, even after twofold averaging. This failure was partly due to the large molecular mass in the asymmetric unit (300 kDa), and partly due to a pseudo-centrosymmetric arrangement of the platinum atoms, which leads to ambiguity in the phasing. More useful phases were derived from isomorphous pairs of derivative and native crystals (obtained by cutting the monoclinic crystals into two halves and using one half for collection of native data and the other half for collection of derivative data). Averaging the four copies of the molecule from the maps in the two different space groups gave an interpretable map to 3 Å resolution ( Fig. 2 ).

View Image

Figure 2 Part of the electron density map at 3 Å resolution that was used to build the model, showing the beta sheet of the C-terminal domain of the alpha chain, with the refined model superimposed.

The current model, refined against 2 Å resolution data, consists of amino acid residues 2–728 from the active alpha chain (A2–A728) plus cobalamin and desulpho-CoA, and residues 18–638 from the inactive beta chain (B18–B638).



Architecture

The alpha and beta chains exhibit similar folds. Each consists of two principal domains ( Fig. 3a , Fig. 3b , Fig. 4a and Fig. 4b ): an N-terminal eight-stranded beta/alpha barrel, and a C-terminal alpha/beta domain with five parallel beta-sheet strands (a fold resembling those of flavodoxin and the B12-binding domain of methionine synthase). However, the relative orientation of the domains differs, and the beta chain binds neither cobalamin nor desulpho-CoA.

View Image

Figure 3 Schematic views of the structure (drawn with MOLSCRIPT [52]). (a) The alphabeta dimer, with the active alpha chain at the bottom (dark colours) and the inactive beta chain at the top (pale colours). Domains are coloured differently: the N-terminal arm, which wraps around the other subunit, is red in the alpha chain; the (beta/alpha)8-barrel domain, which binds the green desulpho-CoA, is yellow with the beta sheet strands in orange; the long linker, which wraps around the (beta/alpha)8-barrel is green; and the C-terminal B12-binding domain is blue, with purple beta-sheet strands. The B12 molecule is dark red. The view is along the pseudo-dyad axis which relates the (beta/alpha)8-barrel domains of the alpha and beta chains. The C-terminal domains in the two chains are in very different positions relative to the barrel domain. (b) The active alpha chain, coloured as in (a). (c) The C-terminal domain of the alpha chain. The five-stranded beta sheet has the same topology as the B12-binding domain of methionine synthase [18]. The lower axial ligand to the cobalt is histidine HisA610, which is in turn bound through AspA608 to LysA604. The pseudo-nucleotide tail is bound in a pocket between the beta sheet and the C-terminal helix. This domain consists of residues A560–A728 (B493–B638 in the beta chain). (d) The eight-stranded (beta/alpha)8-barrel domain of the alpha chain, in the same orientation as in (b).

View Image

Figure 4 Stereoviews of Calpha traces. (a) The active alpha chain, with the desulpho-CoA shown in green and B12 in dark red. (b) The inactive beta chain. Every fiftieth residue is numbered and marked by a small sphere. (c) Superposition, based on the TIM-barrel domains, of the Calpha traces of the alpha chain (in red) and the beta chain (in blue). The different relative orientations of the C-terminal domains are shown. The viewpoint is the same as in Fig. 3b and Fig. 3d .

The N-terminal barrel is preceded by an extended segment which wraps around the other subunit. It is followed by a long linker region (about 160 residues) that encloses the barrel domain and connects it to the C-terminal domain. The linker region in the alpha chain consists of two helices packed against one side of the beta/alpha barrel, followed by a poorly ordered loop stretched across the surface of the barrel, then another group of four helices packed against the other side of the barrel domain before it leads into the C-terminal domain. The linker region in the beta chain is much shorter (95 residues). In the alpha chain, the C-terminal domain packs on one end of the beta/alpha barrel, sandwiching the corrin ring of the cobalamin and forming the active-site cavity. In the beta chain, the C-terminal domain is swung away from the beta/alpha barrel domain, and no active site is formed ( Fig. 4c ). The beta/alpha barrel domains of the alpha and beta chains are related by a local twofold axis, with two helices from each subunit packed together around the dyad. The approximate dyad symmetry breaks down about halfway through the linker region, at residues A482 and B477.



Binding of coenzyme B12

The flavodoxin-like C-terminal domain of the alpha chain is responsible for cofactor binding ( Fig. 3c ). This domain exhibits only slight similarity in amino acid sequence to the B12-binding domain of methylcobalamin-dependent methionine synthase, but has a similar tertiary structure. In fact, the C-terminal domain of the alpha chain superimposes on the B12-binding domain of methionine synthase better (with a root mean square [rms] deviation of 1.7 Å for 110 Calpha atoms) than it does on the corresponding domain of the beta chain (rms deviation of 2.1 Å for 83 Calpha atoms). The cobalamin also binds in a very similar way to the two enzymes. In free coenzyme B12, the lower axial ligand of the cobalt atom comes from the N3B nitrogen of the dimethylbenzimidazole group, which is linked to the corrin ring via the pseudo-nucleotide tail (see Fig. 1a ). In contrast, in methylmalonyl-CoA mutase, as in methionine synthase, the tail is completely displaced away from the cobalt and is bound in a deep pocket between the beta sheet and the C-terminal helix. The cobalt ligand is formed by the epsi-nitrogen (Nepsi2) of histidine A610; the Ndelta1 atom of this histidine starts a hydrogen-bonded chain to AspA608 and LysA604. This lysine is completely buried.

EPR spectroscopy has provided independent confirmation of a histidine ligand to the cobalt atom [18] [20]. The histidine and aspartate residues are conserved between methylmalonyl-CoA mutases from different species, and between methylmalonyl-CoA mutase and 2-methyleneglutarate mutase, glutamate mutase and methionine synthase [19]. The lysine, on the other hand, is only present in the methylmalonyl-CoA mutases. These three conserved residues form part of a B12-binding motif, which is not apparent in the sequences of ethanolamine ammonia-lyase and diol dehydrase, two of the other adenosylcobalamin-dependent enzymes.

On the upper side of the cobalamin we would expect to see the adenosyl group above the cobalt, but no such group is present in the current structure. Electron-density maps ( Fig. 5a ) show no evidence of any ligand in this position. This pentacoordinated state of the cobalt is consistent with a reduced CoII species (cob(II)alamin or B12r), rather than CoIII, which would be expected to be hexacoordinated. Absorption spectra from crystals grown under similar conditions ( Fig. 6 ) show that they contain substantial and variable amounts of CoII: the crystal used to collect the 2 Å resolution data was probably a mixture of CoII and CoIII species, with CoII predominating. The CoII species could be formed by diffusion of the adenosyl radical away from the enzyme.

View Image

Figure 5 The active site and substrate binding. (a) The active-site pocket, with the SIGMAA-weighted 2Fo–Fc map [44] at 2 Å resolution after refinement. The upper axial ligand position on the cobalt atom is unoccupied. The position that would be occupied by the succinyl or methylmalonyl group of the substrate or product, is filled by a tightly bound water molecule (orange sphere, B=13.5 Å2), also hydrogen bonded to Nepsi2 of HisA244 (behind the water, not shown), and by density that has been interpreted as a glycerol molecule. (b) A cross-section through the (beta/alpha)8 barrel, showing some of the short hydrophilic side chains, typically serines, which create the hole in the centre of the barrel. A number of water molecules also participate in the binding of the substrate. (c) Surface representations of the (beta/alpha)8-barrel domain (drawn with GRASP [53]). From left to right: the alpha chain, viewed from the active-site pocket, with the green desulpho-CoA visible through the hole; the same view with the desulpho-CoA removed, showing the hole; the equivalent view of the beta chain, in which the hole has been filled principally by larger hydrophilic side chains.

View Image

Figure 6 Absorption spectra taken (a) from crystals and (b) in solution. In (a), the two solid lines represent data recorded from two crystals of the complex between methylmalonyl-CoA, coenzyme B12 and desulpho-CoA. The dashed line represents data recorded from a crystal grown in the absence of any CoA analogue. In (b), the spectrum typical of enzyme-bound CoIII-adenosylcobalamin (dashed line) was recorded from a 2.5 mg ml-1 solution of methylmalonyl-CoA mutase with bound cofactor. The spectrum typical of cob(II)alamin (solid line) was obtained by addition of the non-physiological substrate succinyl-(carbadethia)-CoA to the solution used for the CoIII spectrum [9].

Some details of the stereochemistry of the cobalt atom and the corrin are given in Table 1, both for methylmalonyl-CoA mutase and for various free cobalamin structures. The most notable feature in the mutase is the very long bond length, about 2.5 Å, between the cobalt atom and nitrogen atom of the lower axial ligand (Nepsi2 of HisA610). Comparison of the axial bond lengths in different cobalamins shows that a strongly bonded upper ligand such as CN or OH leads to a short Co–N bond to the lower axial ligand (1.93–2.01 Å), whereas the methyl and adenosyl groups, which have a Co–alkyl bond, have a longer lower axial bond (2.19–2.21 Å), similar in length to this bond in the pentacoordinated CoII of reduced B12r (2.16 Å).

Table 1. Cobalamin stereochemistry.
Structure Upper axial ligand Co–upper axial ligand Co–N lower axial ligand Co–N plane N lower axial–N plane Fold angle (°) Tilt C7–C8 (°)§ Ref.
Vitamin B12 CN 1.86 2.01 +0.02 2.02 14.1 32.7 [48]
Imidazolyl-B12 CN 1.86 1.97   0.00 1.96   7.3 31.1 [48]
Hydroxycobalamin OH 1.95 1.93 +0.05 1.97 17.5 39.6 [49]
B12r (CoII) 2.16 +0.15 2.30 11.5 31.2 [50]
Methyl-cobalamin CH3 1.99 2.19 +0.02 2.20 14.8 30.1 [51]
Coenzyme B12 AdoCH2* 1.97 2.21 –0.01 2.20 10.0 30.0 (C Kratky, pers. comm.)
Methylmalonyl-CoA mutase:
   With no restraints on Co–N bonds 2.53 +0.11 2.64 8.3 6.8 This work
   With Co–N axial bond restrained to 2.20 Å 2.30 +0.13 2.43 8.4 7.6 This work
Distances between the cobalt atom and axial ligands are given (in Å). *AdoCH2 is the 5'-deoxyadenosyl group. Distance from the Co to the mean plane of the pyrrole nitrogens: a positive distance means a displacement towards the lower axial ligand. The fold is along the Co–C10 line, defined as the angle between two linked conjugated planes in the corrin ring: C4 C5 C6 C7 C8 C9 C10 N21 N22; and C10 C11 C12 C13 C14 C15 C16 N23 N24. §The tilt is the angle between the C7–C8 bond and the plane of the pyrrole nitrogens (i.e. the out of plane twist of the C6 C7 C8 C9 N22 pyrrole ring). #Imidazolyl-B12 has an imidazole group replacing the dimethylbenzimidazole of vitamin B12.
Return to table reference [1] [2] [3]

To examine the proposition that the Co–Nepsi2(HisA610) bond has a more conventional length, this bond was added as a restraint in the refinement, with a length of 2.20 Å and it was given the same weight as the other covalent bonds in the molecule (ensuring that there were no repulsion restraints between the histidine and the corrin). This reduced the bond length to 2.30 Å, but distorted the histidine ring. Very similar values for this bond length were obtained when restraints were applied within the programs REFMAC (which uses the same restraints as in PROLSQ, see the Materials and methods section) and X-PLOR [21] (which uses pseudo-energy restraints), showing that the bond length is insensitive to the type of stereochemical restraints used in the refinement. We interpret these results as validating the unusual bond length, and that the protein holds the corrin and the histidine further apart than their ideal positions.

The corrin ring in the mutase is flatter than in most cobalamins (Table 1): for instance, the fold angle of the conjugated systems about the Co–C10 line is similar to that in imidazolyl-B12 rather than to that in the normal cobalamins in which dimethylbenzimidazole is linked to the cobalt. The bulkier base of the free coenzyme bends the corrin upwards more than the smaller imidazole or histidine. In the mutase, pyrrole ring B is substantially less twisted from the corrin plane than in any of the free cobalamins. This can be seen from the angle between the C7–C8 bond and the plane of the pyrrole nitrogens ('tilt' in Table 1). It may be significant that the acetamide group attached to C7 is hydrogen bonded to TyrA89 ( Fig. 7a ), which is one of the putative active-site residues. Ring B is also less well ordered than the rest of the corrin, as indicated by higher temperature factors (data not shown).

View Image

Figure 7 Schematic representation of ligand binding. (a) The interactions between protein and coenzyme B12. The hydrophobic pocket for the dimethylbenzimidazole is lined by residues IleA617, TyrA705, GlyA685 and SerA655: this last forms a hydrogen bond to the N3B nitrogen atom of the base, which is the atom that coordinates the cobalt atom in the free coenzyme. LeuA657 stacks against HisA610, the residue that coordinates the cobalt, and forms hydrophobic interactions with the beginning of the pseudo-nucleotide tail (C56) and with the C20 methyl group of the corrin. (b) The interactions between protein and desulpho-CoA. TyrA75 stacks on the adenine ring. Note that ArgB45 is the only interaction between the beta chain and the substrate. CoA would have an additional thiol group attached to the left-hand end.

Details of the B12-binding site are shown schematically in Fig. 7a . The pseudo-nucleotide tail runs through a cavity filled with well-ordered water molecules, then buries the dimethylbenzimidazole group in a tight hydrophobic pocket. The oxygens of the phosphate group interact only with two water molecules which are in turn bonded to other waters, to the nitrogen of GlyA613 and to the amide oxygen O63 of the corrin. The side-chain amides attached to the corrin interact with both the (beta/alpha)8-barrel domain and the C-terminal domain.



Binding of coenzyme A

The partial substrate, desulpho-CoA, is bound along the axis of the (beta/alpha)8-barrel domain, with the ADP-3'-phosphate moiety at the N terminus of the beta strands, partly exposed to solvent, and the pantetheine chain completely buried between the beta strands ( Fig. 3d ). In the true substrate or product, the succinyl or methylmalonyl group would protrude into the active-site cavity (see below). The pantetheine chain of CoA is approximately the same length as the beta strands of the barrel.

The (beta/alpha)8-thinspace or TIM-barrel structure is one of the most frequently observed folds in proteins, with at least 45 examples in the Brookhaven Protein Database. In these structures, the centre of the barrel is typically filled by large, often branched, hydrophobic, side chains [22]. In the alpha chain of methylmalonyl-CoA mutase, the barrel is lined by hydrophilic, mainly small, side chains (ThrA85, SerA114, SerA162, SerA164, SerA166, ThrA195, SerA239, SerA285 and SerA362), creating a hole through which the CoA can thread, and providing hydrogen bonds to the amide groups of CoA, some via a network of water molecules ( Fig. 5b thinspaceand Fig. 7b ). In the beta chain, the hole is filled, partly by the replacement of short side chains by longer ones (e.g. LeuB137 replaces SerA114, GluB165 replaces SerA164 and AsnB281 replaces SerA285) and partly by movement of longer side chains, such as ArgB110, to obstruct the hole ( Fig. 5c ). The side chains are still largely hydrophilic. The ADP-3'-phosphate group is bound by five arginines and one lysine. One of the arginines, ArgB45, is contributed by the beta chain.

The structures of at least eight CoA-binding proteins are known, in addition to methylmalonyl-CoA mutase, but they are completely unrelated and have no common motif for binding CoA. Binding of a ligand along the axis of a (beta/alpha)8-barrel fold has not, to our knowledge, been observed previously. It seems particularly appropriate in this case, not only because it provides a narrow access channel to a deeply buried active site, allowing the enzyme to protect its reactive intermediates from side reactions, but also because the dimensions of the barrel fit CoA derivatives well. The hole is created by the absence of large side chains: the shape of the barrel is not significantly different from that in other proteins with this fold.



The active site

The present structure lacks both the adenosyl group of the cofactor and the rearranged part of the substrate. A detailed description of the active site, and hypotheses regarding the mechanism of rearrangement will therefore have to await further structures of substrate and inhibitor complexes. However, some observations can be made. The active-site cavity is formed by the loops at the C termini of the beta strands of the (beta/alpha)8-barrel domain, and it is closed off by the corrin ring of the B12 coenzyme and its associated amide side chains. The active site is completely inaccessible to solvent, except through the CoA channel along the barrel axis. Such a deeply buried active site would protect the reactive radical intermediates from undesired side reactions, and a similar inaccessibility has been seen in a number of other free-radical enzymes, such as galactose oxidase [23], copper amine oxidase [24], and prostaglandin H2 synthase [25].

Modelling succinyl-CoA or methylmalonyl-CoA into the active site places the cobalt atom too far from the substrate for its direct participation in the rearrangement reaction. The carbonyl oxygen of the thioester comes close to the position of a water molecule tightly bound to HisA244. This histidine is in a position to bind to the carbonyl oxygen in the rearrangement reaction, and is likely to be the major catalytic residue. Other groups which may be involved in substrate binding include ArgA207, TyrA89 and the C7-acetamide of the corrin.



The beta chain

The common ancestor of the heterodimeric bacterial methylmalonyl-CoA mutases and the homodimeric mammalian enzymes was presumably a homodimer, as the subunits exhibit substantial sequence homology. The inactive beta chain has no obvious major function, although it contributes one residue, ArgB45, to the substrate-binding site. The beta chain is significantly less well ordered than the alpha chain, as shown by the higher average temperature factor (38 Å2 and 37 Å2 for the two alpha chains, compared with 43 Å2 and 53 Å2 for the two beta chains in the asymmetric unit), and its larger number of poorly ordered loops. Many of the residues which are important for cofactor and substrate binding have been lost in the beta chain. The beta chain seems to be an evolutionary relic, although we cannot exclude the possibility of it having other functions.



Catalysis of radical formation

An important role of the enzyme is to favour the breaking of the Co–C bond when the substrate is bound, in order to initiate the reaction. In the free coenzyme, the dissociation energy for homolysis of this bond is about 100–125 kJ mol-1 [26] [27], with an estimated dissociation rate of about 10-7 s-1 at 30°C [2] [3]. This rate is much too slow to explain the reaction rates of the order of 102 s-1that have been reported for several of these reactions, including the one catalyzed by methylmalonyl-CoA mutase [28]. Thus, one of the roles of the enzyme is to favour the formation of the initial cofactor-derived radical, both thermodynamically and kinetically, and to do so in response to substrate binding, which weakens the Co–C bond beyond that in the holoenzyme [4] [7] [8] [9]. The strength of the Co–C bond with respect to homolysis could be modulated by the lower axial ligand (trans effect), either by steric or electronic effects. Alternatively, the crowding of the adenosyl group by protein groups or by the upward folding of the corrin ring (cis effect) could influence the Co–C bond. We propose that the major effect in this enzyme is trans and steric, in that the protein holds the HisA610 side chain too far from the corrin, putting the cobalt into a strained or 'entatic' state [29]. The 2.5 Å long Co–N bond would stabilize the CoII species relative to CoIII, thus favouring the formation of the adenosyl radical. At the same time, the formation of hydroxycobalamin (which would lead to irreversible inactivation) is disfavoured. In order to hold the histidine in this position, the enzyme provides a strong hydrogen bond to the negatively charged AspA608. However, the aspartate will induce a negative charge on the histidine, which will favour CoIII with respect to CoII, and strengthen the Co–C bond. This undesired side effect of the structural anchor on the histidine is compensated for by the positively charged LysA604. On binding of the substrate, the position of the histidine relative to the corrin may be altered by conformational changes, such as relative movements of the domains. The charge state of the histidine is less likely to alter on substrate binding, because the buried LysA604 will lock the His–Asp–Lys system into a state with net zero charge. The proposed modulation of the Co–C bond strength by movement of the Co-linked histidine is reminiscent of the regulation of oxygen binding in haemoglobin by positioning the proximal histidine, which is linked to the R–T state allosteric transition [30].



Biological implications

Coenzyme B12 (5'-deoxyadenosylcobalamin) consists of a central cobalt atom in a porphyrin-like corrin ring, bonded to the 5'-C of 5'-deoxyadenosine on one side and to a base on the other side. This molecule is the cofactor for a series of enzymes, the mutases, which catalyze unusual rearrangement reactions. The reactions take place via radical intermediates, formed indirectly following breakage of the cobalt–carbon bond. It has been a long-standing puzzle how the enzymes catalyze the breaking of this bond.

Methylmalonyl-CoA mutase is the only member of this class found in mammals as well as in bacteria. The enzyme catalyzes the interconversion between succinyl-CoA and methylmalonyl-CoA. In bacteria such as Propionibacterium shermanii, it is involved in the fermentation of pyruvate to propionate, a pathway which helps balance the redox state in anaerobic growth. In mammals, the enzyme is required for the degradation of odd-chain fatty acids, to convert propionyl-CoA to succinyl-CoA.

The crystal structure of methylmalonyl-CoA mutase in complex with coenzyme B12 and the partial substrate desulpho-CoA (lacking the succinyl group and the terminal sulphur atom of the substrate succinyl-CoA) shows that the base which coordinates the cobalt in the free coenzyme B12 is replaced by a histidine side chain from the protein. Both the mode of binding and the fold of the B12-binding domain are very similar to those in the B12-binding fragment of methionine synthase [18], which binds the other B12 cofactor, methylcobalamin. Methylcobalamin provides a methyl group for methylation reactions as a positively charged group rather than as a radical, and it remains unclear why the cobalt–carbon bonds in the two cofactors are split in different ways (i.e. homolytically versus heterolytically). In the structure of the complex presented here, the coenzyme is largely in the reduced CoIIform, having lost the adenosyl radical. The bond between the cobalt and the histidine is significantly longer than in model compounds, and this stretched bond explains how the enzyme stabilizes the reduced state, favouring the formation of the adenosyl radical.

In methylmalonyl-CoA mutase, the substrate is bound through a narrow tunnel along the axis of a (beta/alpha)8-thinspace or TIM-barrel domain. This tunnel provides the only direct access to an active-site cavity, which, in common with other enzymes that produce reactive radical intermediates, is protected from unwanted side reactions. The mode of substrate binding is completely different from that in the other 45 known structures with this common fold.



Materials and methods


Crystallization and characterization of the crystals

Recombinant P. shermanii methylmalonyl-CoA mutase was prepared from an overexpressing clone of Escherichia coli as described elsewhere [31]. Crystals of the complex between methylmalonyl-CoA mutase, coenzyme B12, and desulpho-CoA were grown by vapour diffusion at 23°C, by equilibrating against a reservoir solution containing 14% (w/v) polyethylene glycol (PEG) 4000 and 20% (v/v) glycerol in 100 mM Tris-HCl buffer at pH 7.5 a 10 mul drop consisting of equal volumes of the reservoir solution and of a solution of 20 mg ml-1 protein, 2 mM coenzyme B12 and 12 mM desulpho-CoA in Tris-HCl buffer at pH 7.5. Crystals grow within three weeks to a size of up to 1.0×0.4×0.3 mm3. The crystals were stabilized in a solution containing 16% PEG 4000 and 20% glycerol in 100 mM Tris-HCl buffer at pH 7.5. The crystals grow in two closely related and morphologically indistinguishable forms, one monoclinic (crystal form 1 [32]) the other orthorhombic (crystal form 2). Each form contains two molecules in the asymmetric unit, with a solvent content of 48%.



Data collection

Data collection statistics are given in Table 2. The crystals were suspended in a thin liquid film of stabilizing solution on a rayon loop [33] and frozen at 95 K directly in the cold nitrogen gas stream with a 600 Series Cryostream Cooler (Oxford Cryosystems, Oxford, UK) [34]. In order to overcome severe problems of non-isomorphism of derivatives in crystal form 1, each crystal was cut in two halves with a thin glass blade. The two halves were then transferred into two separate stabilizing solutions, one containing a heavy-atom compound. Native A and derivative A1, and Native B and derivative B, are isomorphous pairs made in such a way. Derivative A2 was found to merge as well with native A as with its own matching half, and so it was paired to native A. Native A and derivative A1 data sets were collected at a wavelength of 1.000 Å on station 9.5 at the Synchrotron Radiation Source (SRS), Daresbury, UK. Derivative A2, native B and derivative B data sets were collected using CuKalpha radiation from an Elliot GX13 rotating anode generator with a 100 mum focal spot, collimated by a Supper double-mirror system (Natick, MA). Native D data set was collected at a wavelength of 0.900 Å on Beamline 4 at the European Synchrotron Radiation Facility (ESRF), Grenoble, France. MAD data were collected from a single orthorhombic crystal (crystal form 2), pre-reacted by soaking in 2 mM K2Pt(NO2)4 for 48 h, on station 9.5 at the SRS, Daresbury, UK. The crystal was aligned with the b* axis approximately along the rotation axis to allow the near simultaneous recording of Bijvoet pairs of reflections and six data sets were collected on and around the platinum LIII absorption edge. Native C data were collected at a wavelength of 0.898 Å on station 9.6 at the SRS, Daresbury, UK. All the data were recorded on either 180 mm (derivative A2, native B, derivative B, MAD data) or 300 mm diameter MAR Research image plates (Hamburg, Germany) and integrated with MOSFLM [35]. Scaling and processing were performed using the CCP4 suite of programs [36].

Table 2. X-ray structure determination.
Data set Resolution (Å) Reflections measured/unique Completeness (%) Rmerge* (%) Riso (%) Phasing power RCullis§
MIR analysis (crystal form 1 #)
Native A 40.0–3.0 244thinspace564/65thinspace599 99.9 (99.8) 4.0 (16.5)
Derivative A1 20.0–3.0 243thinspace462/65thinspace391 99.6 (99.5) 6.3 (30.9) 12.8 (30.8) 0.60 0.85
Derivative A2 23.4–3.2 189thinspace492/51thinspace778 95.8 (94.2) 4.9 (16.9) 17.8 (24.0) 0.93 0.79
Native B 21.7–3.2 201thinspace038/54thinspace041 99.5 (99.1) 6.9 (25.6)
Derivative B 25.8–3.2 247thinspace860/53thinspace012 97.6 (96.6) 6.2 (23.0) 12.6 (24.1) 0.60 0.84
MIR analysis (crystal form 2 #)
lambda1 (1.0400 Å) 30.0–2.8 307thinspace915/81thinspace302 98.0 (96.8) 6.1 (25.2) 6.4 (16.6)
lambda2 (1.0718 Å) 30.0–2.8 302thinspace043/81thinspace091 97.7 (95.3) 5.6 (22.2) 5.9 (16.1)
lambda3 (1.0723 Å) 30.0–2.8 301thinspace922/81thinspace066 97.7 (95.1) 5.7 (22.6) 5.9 (16.5)
lambda4 (1.1200 Å) 30.0–2.95   263thinspace817/69thinspace453 97.7 (95.0) 4.2 (11.4) 4.8 (12.0)
lambda5 (1.0726 Å) 30.0–2.8 283thinspace685/80thinspace350 96.9 (91.4) 5.0 (19.7)
lambda6 (1.0723 Å) 30.0–2.8 280thinspace614/80thinspace318 96.8 (90.9) 4.5 (17.5) 5.7 (15.7)
Data used for density modification (crystal form 2#)
Native C 20.0–3.2 148,048/54,499 97.5 (97.7) 7.0 (19.6)
Data used for refinement (crystal form 1#)
Native D 28.5–2.0 1,008,268/217,377 99.8 (99.8) 5.1 (13.8)
*Rmerge=SigmaSigmai|I(h)–I(h)i|/SigmaSigmaiI(h)i, where I(h) is the mean intensity after rejections. Riso=Sigma||FPH|–|FP||/Sigma|FP|, where FPH and FP are the structure factors for the derivative and native, respectively. For the MAD analysis, Riso is calculated between data sets at two different wavelengths: wavelength lambda5 is treated as the reference in this calculation. Phasing power is rms (|FH|/E), where |FH| is the heavy-atom structure factor amplitude and E is the residual lack of closure (|FPH–FP|–|FH|). §RCullis=Sigma|E|/||FPH|–|FP||. The summation is over centric reflections only. #Crystal form 1 is monoclinic, spacegroup P21 (a=119.8 Å, b=161.3 Å, c=88.4 Å, beta=105.07°). Crystal form 2 is orthorhombic, spacegroup P212121 (a=123.4, Å b=162.0 Å, c=166.2 Å). Derivative A1 (two heavy-atom sites) was soaked for 20 h in 0.2 mM ethylmercury thiosalicylate. Derivative A2 (four heavy-atom sites) was soaked for 24 h in 0.4 mM methylmercury. Derivative B (eight heavy-atom sites) was soaked for 24 h in 15 mM KAu(CN)2. The crystal used for MAD data collection (eight heavy-atom sites) was soaked in 2 mM K2Pt(NO2)4 for 48 h..
Return to table reference [1]


Phasing

The isomorphous difference and anomalous difference Patterson maps were solved with the SHELXS-90 program [37]. Minor sites were subsequently found in difference Fourier maps. Heavy-atom parameter refinement and phase calculations were carried out for both crystal forms using the program MLPHARE [38]. In crystal form 1 two phase sets were derived: one for native A to 3 Å resolution, with the contributions from two derivatives, derivative A1 and A2; the other for native B to 3.2 Å resolution, with one derivative, derivative B. Phase probability coefficients were then combined and applied to the structure-factor amplitudes of native A. The derivative phase set (FPH) derived from the MAD analysis was transferred to a native set of amplitudes (native C) by vectorial subtraction of the calculated heavy-atom contribution. An initial molecular envelope was determined in crystal form 1 by performing a local correlation search with the DEMON package [39]. The density inside the molecular mask was then used as a molecular replacement model for crystal form 2 with a combination of MAVE [40], AMoRe [41] and TFFC [42]. The mask was subsequently improved by iterative averaging. The initial electron-density maps were improved within each crystal form by real-space solvent flattening and twofold averaging using SOLOMON [43], and between the forms by reciprocal-space phase combination using SIGMAA [44]. The procedure was carried out at 3.0 Å resolution for crystal form 1 and at 3.2 Å resolution for crystal form 2 and resulted in an interpretable 3.0 Å resolution map.



Model building and refinement

The atomic model was built using the graphics program O [45]. The model was fitted to the high-resolution native D data by rigid-body refinement with the program TNT [46], and the resolution extended in four steps from 3 Å to 2 Å. Refinement has been carried out with TNT and REFMAC (G Murshudov, unpublished program), which refines a maximum-likelihood target, using stereochemical restraints as in PROLSQ [47]. Initially, a tight non-crystallographic symmetry restraint was applied to keep the two molecules in the same conformation. At a later stage of refinement, this restraint was released to allow some small but significant differences between the molecules.

At the current stage of refinement the model has an R-factor of 21.9% for 95% of the data between 20 Å and 2 Å, including 1532 water molecules and 9 glycerol molecules, a total of 22683 atoms. The free R-factor for the remaining 5% of the data within this resolution range is 27.4%. The rms deviation from standard values of bond lengths is 0.017 Å.

The coordinates and structure factors have been deposited in the Protein Data Bank, ID code 1REQ.



Acknowledgements

We dedicate this paper to the memory of Dorothy Hodgkin. We are extremely grateful to P Williams and J Hajdu for recording the spectra from the crystals. We thank S Knight for his contributions earlier in the project, C Cardin and A Murzin for discussions, T Woollard, R Read and I Fearnley for their assistance, and C Kratky for cobalamin coordinates. We thank the EU for funds to travel to ESRF Grenoble, the staff of EMBL Grenoble, especially A Åberg, for assistance in experiments performed at the ESRF on BL4, and the staff of the Synchrotron Radiation Facility at Daresbury for their help. FM was supported by the Italian Consiglio Nazionale delle Ricerche (CNR), and by Sigma-Tau industrie farmaceutiche riunite s.p.a., Pomezia, Italy. NHK was supported by a Wellcome Trust Studentship and a Denman-Baynes Studentship from Clare College, Cambridge. AN was supported by the Japan Society for the Promotion of Science.



References
  1. Lenhert P.G. , Hodgkin D.C. : 1961,
    Structure of the 5,6-dimethylbenzimidazolylcobamide coenzyme.
    Nature 192: 937–938.
    Return to citation reference [1]

  2. Halpern J. : 1985,
    Mechanisms of coenzyme B12-dependent rearrangements.
    Science 227: 869–875. [Medline]
    Return to citation reference [1] [2] [3] [4]

  3. Pratt J.M. : 1993,
    Nature's design and use of catalysis based on Co and the macrocyclic corrin ligand: 4×109 years of coordination chemistry.
    Pure & Appl. Chem. 65: 1513–1520.
    Return to citation reference [1] [2]

  4. Rétey J. , Arigoni D. : 1966,
    Coenzym B12 als gemeinsamer wasserstoffüberträger der dioldehydrase und der methylmalonyl-CoA mutase reaktion.
    Experientia 22: 783–784.
    Return to citation reference [1] [2]

  5. Cardinale G.J. , Abeles R.H. : 1967,
    Mechanistic similarities in the reactions catalysed by diol dehydrase and methylmalonyl-CoA mutase.
    Biochim. Biophys. Acta 132: 517–518.
    Return to citation reference [1]

  6. Cockle S.A. , Hill H.A.O. , Williams R.J.P. , Davies S.P. , Foster M.A. : 1972,
    The detection of intermediates during the conversion of propane-1,2-diol to propionaldehyde by glycerol dehydrase, a coenzyme B12 dependent reaction.
    J. Am. Chem. Soc. 94: 275–276.
    Return to citation reference [1]

  7. Zhao Y. , Such P. , Rétey J. : 1992,
    Radical intermediates in the coenzyme B12dependent methylmalonyl-CoA mutase reaction shown by ESR spectroscopy.
    Angew. Chem. Int. Ed. Engl. 31: 215–216.
    Return to citation reference [1] [2]

  8. Zhao Y. , Abdend A. , Kuntz M. , Such P. , Rétey J. : 1994,
    Electron paramagnetic resonance studies of the methylmalonyl-CoA mutase reaction.
    Eur. J. Biochem. 225: 891–896. [Medline]
    Return to citation reference [1] [2]

  9. Keep N.H. , Smith G.A. , Evans M.C.W. , Diakun G.P. , Leadlay P.F. : 1993,
    The synthetic substrate succinyl(carbadethia)-CoA generates cob(II)alamin on adenosylcobalamin-dependent methylmalonyl-CoA mutase.
    Biochem. J. 295: 387–392. [Medline]
    Return to citation reference [1] [2] [3]

  10. Rétey J. , Lynen F. : 1964,
    The absolute configuration of methylmalonyl-CoA mutase.
    Biochem. Biophys. Res. Commun. 16: 358–361.
    Return to citation reference [1]

  11. Eggerer H. , Overath P. , Lynen F. , Stadtman E.R. : 1960,
    On the mechanism of the cobamide coenzyme dependent isomerization of methylmalonyl-CoA to succinyl-CoA.
    J. Am. Chem. Soc. 82: 2643–1644.
    Return to citation reference [1]

  12. Francalanci F. , Davis N.K. , Fuller J.Q. , Murfitt D. , Leadlay P.F. : 1986,
    The subunit structure of methylmalonyl-CoA mutase from Propionibacterium shermanii.
    Biochem. J. 236: 489–494. [Medline]
    Return to citation reference [1]

  13. Marsh E.N. , McKie N. , Davis N.K. , Leadlay P.F. : 1989,
    Cloning and structural characterization of the genes coding for adenosylcobalamin-dependent methylmalonyl-CoA mutase from Propionibacterium shermanii.
    Biochem. J. 260: 345–352. [Medline]
    Return to citation reference [1]

  14. Wölfle K. , Michenfelder M. , König A. , Hull W.E. , Rétey J. : 1986,
    On the mechanism of action of methylmalonyl-CoA mutase. Change of the steric course on isotopic substitution.
    Eur. J. Biochem. 156: 545–554. [Medline]
    Return to citation reference [1]

  15. Keep N.H. : 1992,
    Studies on the structure and mechanism of methylmalonyl-CoA mutase. [PhD thesis].
    University of Cambridge, Cambridge, UK:
    Return to citation reference [1]

  16. Jansen R. , Kalousek F. , Fenton W.A. , Rosenberg L.E. , Ledley F.D. : 1989,
    Cloning of full-length methylmalonyl-CoA mutase from a cDNA library using the polymerase chain reaction.
    Genomics 4: 198–205. [Medline]
    Return to citation reference [1]

  17. Drennan C.L. , Matthews R.G. , Ludwig M.L. : 1994,
    Cobalamin-dependent methionine synthase: the structure of a methylcobalamin-binding fragment and implications for other B12-dependent enzymes.
    Curr. Opin. Struct. Biol. 4: 919–929. [Medline]
    Return to citation reference [1]

  18. Drennan C.L. , Huang S. , Drummond J.T. , Matthews R.G. , Ludwig M.L. : 1994,
    How a protein binds B12: a 3.0 Å X-ray structure of B12-binding domains of methionine synthase.
    Science 266: 1669–1674. [Medline]
    Return to citation reference [1] [2] [3] [4]

  19. Marsh E.N.G. , Holloway D.E. : 1992,
    Cloning and sequencing of glutamate mutase component S from Clostridium tetanomorphum. Homologies with other cobalamin-dependent enzymes.
    FEBS Lett. 310: 167–170. [Medline]
    Return to citation reference [1] [2]

  20. Padmakumar R. , Taoka S. , Padmakumar R. , Banerjee R. : 1995,
    Coenzyme B12 is coordinated by histidine and not by dimethylbenzimidazole on methylmalonyl-CoA mutase.
    J. Am. Chem. Soc. 117: 7033–7034.
    Return to citation reference [1]

  21. Brünger A.T. : 1993, X-PLOR: Version 3.1. A System for Protein Crystallography and NMR. Yale University Press, New Haven, CT:
    Return to citation reference [1]

  22. Brändén C. , Tooze J. : 1991, Introduction to Protein Structure. pp. 43–48. Garland Publishing, Inc., New York and London:
    Return to citation reference [1]

  23. Ito N. et al , Knowles P.F. : 1991,
    Novel thioether bond revealed by a 1.7 Å crystal structure of galactose oxidase.
    Nature 350: 87–90. [Medline]
    Return to citation reference [1]

  24. Parsons M.R. et al , Knowles P.F. : 1995,
    Crystal structure of a quinoenzyme: copper amine oxidase of Escherichia coli at 2 Å resolution.
    Structure 3: 1171–1184.
    Return to citation reference [1]

  25. Picot D. , Loll P.J. , Garavito R.M. : 1994,
    The crystal structure of the membrane protein prostaglandin H2 synthase-1.
    Nature 367: 243–249.
    Return to citation reference [1]

  26. Halpern J. , Kim S.H. , Leung T.W. : 1984,
    Cobalt–carbon bond dissociation energy of coenzyme B12.
    J. Am. Chem. Soc. 106: 8317–8319.
    Return to citation reference [1]

  27. Finke R.G. , Hay B.P. : 1984,
    Thermolysis of adenosylcobalamin: a product, kinetic and Co–C5' bond dissociation study.
    Inorg. Chem. 23: 3041–3043.
    Return to citation reference [1]

  28. Dolphin D. : 1982, B12 John Wiley & Sons, Inc., New York:
    Return to citation reference [1]

  29. Vallee B.L. , Williams R.J.P. : 1968,
    Metalloenzymes: the entatic nature of their active sites.
    Proc. Natl. Acad. Sci. USA 59: 498–505.
    Return to citation reference [1]

  30. Perutz M.F. , Fermi G. , Luisi B. , Shaanan B. , Liddington R.C. : 1987,
    Stereochemistry of cooperative mechanisms in hemoglobin.
    Accounts Chem. Res. 20: 309–321.
    Return to citation reference [1]

  31. McKie N. , Keep N.H. , Patchett M.L. , Leadlay P.F. : 1990,
    Adenosylcobalamin-dependent methylmalonyl-CoA mutase from Propionibacterium shermanii. Active holoenzyme produced from Escherichia coli..
    Biochem J. 269: 293–298. [Medline]
    Return to citation reference [1]

  32. Marsh N. , Leadlay P.F. , Evans P.R. : 1988,
    Crystallisation and preliminary diffraction data for adenosylcobalamin-dependent methylmalonyl-CoA mutase from Propionibacterium shermanii.
    J. Mol. Biol. 200: 421–422. [Medline]
    Return to citation reference [1]

  33. Teng T. : 1990,
    Mounting of crystals for macromolecular crystallography in a free-standing thin film.
    J. Appl. Cryst. 23: 387–391.
    Return to citation reference [1]

  34. Cosier J. , Glazer A.M. : 1986,
    A nitrogen gas stream cryostat for general X-ray diffraction studies.
    J. Appl. Cryst. 19: 105–107.
    Return to citation reference [1]

  35. Leslie A.G.W. : 1992, Recent changes to the MOSFLM package for processing film and image plate data. In Joint CCP4 and ESF-EACMB Newsletter on Protein Crystallography No. 26. SERC Daresbury Laboratory, Warrington, UK:
    Return to citation reference [1]

  36. Collaborative Computational Project No.4.: 1994,
    The CCP4 suite: programs for protein crystallography.
    Acta Cryst. D 50: 760–763.
    Return to citation reference [1]

  37. Sheldrick G.M. : 1991, Heavy atom location using SHELXS-90. In Isomorphous Replacement and Anomalous Scattering: Proceedings of the CCP4 Study Weekend, 25–26 January 1991. Edited by Wolf, W., Evans, P.R. & Leslie, A.G.W.. 23–38. SERC Daresbury Laboratory, Warrington, UK:
    Return to citation reference [1]

  38. Otwinowski Z. : 1991, Maximum likelihood refinement of heavy atom parameters. In Isomorphous Replacement and Anomalous Scattering: Proceedings of the CCP4 study weekend, 25–26 January 1991. Edited by Wolf, W., Evans, P.R. & Leslie, A.G.W.. 80–86. SERC Daresbury Laboratory, Warrington, UK:
    Return to citation reference [1]

  39. Vellieux F.M.D.A.P. , Hunt J.F. , Roy S. , Read R.J. : 1995,
    DEMON/ANGEL: a suite of programs to carry out density modification.
    J. Appl. Cryst. 28: 347–351.
    Return to citation reference [1]

  40. Kleywegt G.J. , Jones T.A. : 1994, Halloween... masks and bones. In From First Map to Final Model: Proceedings of the CCP4 Study Weekend, 6–7 January 1994. Edited by Bailey, S., Hubbard, R. & Waller, D.. 59–66. SERC Daresbury Laboratory, Warrington, UK:
    Return to citation reference [1]

  41. Navaza J. : 1994,
    AMoRe: an automated package for molecular replacement.
    Acta Cryst. A 50: 157–163.
    Return to citation reference [1]

  42. Tickle I.J. : 1992, Fast Fourier translation functions. In Molecular Replacement: proceedings of the CCP4 study weekend, 31 January–1 February 1992. Edited by Dodson, E.J., Gover, S. & Wolf, W.. 20–32. SERC Daresbury Laboratory, Warrington, UK:
    Return to citation reference [1]

  43. Abrahams J.P. , Leslie A.G.W. : 1996,
    Methods used in the structure determination of bovine mitochondrial F1 ATPase.
    Acta Cryst. D 52: 30–42.
    Return to citation reference [1]

  44. Read R.J. : 1986,
    Improved Fourier coefficients for maps using phases from partial structures with errors.
    Acta Cryst. A 42: 140–149.
    Return to citation reference [1] [2]

  45. Jones T.A. , Zou J.Y. , Cowan S.W. , Kjeldgaard M. : 1991,
    Improved methods for building protein models in electron density maps and the location of errors in these models.
    Acta Cryst. A 47: 110–119. [Medline]
    Return to citation reference [1]

  46. Tronrud D.E. , TenEyck L.F. , Matthews B.W. : 1987,
    An efficient general-purpose least-squares refinement program for macromolecular structures.
    Acta Cryst. A 43: 489–501.
    Return to citation reference [1]

  47. Konnert J.H. , Hendrickson W.A. : 1980,
    A restrained-parameter thermal-factor refinement procedure.
    Acta Cryst. A 36: 344–350.
    Return to citation reference [1]

  48. Kräutler B. , Konrat R. , Stupperich E. , Färber G. , Gruber K. , Kratky C. : 1994,
    Direct evidence for the conformational distortion of the corrin ring by the nucleotide base in vitamin B12: synthesis and solution spectroscopic and crystal structure analysis of cob-cyanoimidazolylcobamide.
    Inorg. Chem. 33: 4128–4139.
    Return to citation reference [1] [2]

  49. Kratky C. et al , Kräutler B. : 1995,
    Accurate structural data demystify B12: high resolution solid-state structure of aquocobalamin perchlorate and structural analysis of the aquocobalamin ion in solution.
    J. Am. Chem. Soc. 117: 456–4670.
    Return to citation reference [1]

  50. Kräutler B. , Keller W. , Kratky C. : 1989,
    Coenzyme B12 chemistry: the crystal and molecular structure of cob(II)alamin.
    J. Am. Chem. Soc. 111: 8936–8939.
    Return to citation reference [1]

  51. Rossi M. et al , Marzilli L.G. : 1985,
    The structure of a B12 coenzyme: methylcobalamin studies by X-ray and NMR methods.
    J. Am. Chem. Soc. 107: 1729–738.
    Return to citation reference [1]

  52. Kraulis P.J. : 1991,
    MOLSCRIPT: a program to produce both detailed and schematic structures.
    J. Appl. Cryst. 24: 946–950.
    Return to citation reference [1]

  53. Nicholls A. , Sharp K.A. , Honig B. : 1991,
    Protein folding and association: insights from the interfacial and thermodynamic properties of hydrocarbons.
    Proteins 11: 281–296. [Medline]
    Return to citation reference [1]



Received / Accepted
Received: 15 Jan 1996
Accepted: 1 Feb 1996
Electronic revisions: PDB links - Created: 1REQ (misc).
Electronic revisions: PDB links - Updated: 1REQ (extlink).

Author Contacts
Filippo Mancia, Atsushi Nakagawa and Philip R Evans (corresponding author), MRC Laboratory of Molecular Biology, Hills Road, Cambridge CB2 2QH, UK. Nicholas H Keep, MRC Laboratory of Molecular Biology, Hills Road, Cambridge CB2 2QH, UK, and Department of Biochemistry, University of Cambridge, Tennis Court Road, Cambridge CB2 1QW, UK. Peter F Leadlay, Department of Biochemistry, University of Cambridge, Tennis Court Road, Cambridge CB2 1QW, UK. Sean McSweeney, SERC Daresbury Laboratory, Warrington, Cheshire WA4 4AD, UK. Bjarne Rasmussen, EMBL, c/o ILL, 156X, F-38042 Grenoble Cedex, France. Peter Bösecke and Olivier Diat, European Synchrotron Radiation Facility, BP 220, F-38043 Grenoble Cedex, France. Present address for Nicholas H Keep: Department of Medicine, Rayne Institute, University College, London WC1E 6JJ, UK. Present address for Atsushi Nakagawa: Division of Biological Sciences, Graduate School of Science, Hokkaido University, Sapporo, Hokkaido 060, Japan. Present address for Sean McSweeney: EMBL, c/o ILL, 156X, F-38042 Grenoble Cedex, France.
Return to author list


Copyright

Copyright © 1996 Current Biology Publishing
Help   Feedback      Search   Map
Black Line
© 1998 BioMedNet Limited. All rights reserved.
email: info@biomednet.com.
How coenzyme B12 radicals are generated: the crystal structure of methylmalonyl-coenzyme A mutase at 2 A resolution