首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Coenzyme B(12) serves as a cofactor for enzymatic radical reactions. The essential steps in all the coenzyme B(12)-dependent rearrangements are two hydrogen abstraction steps: hydrogen abstraction of the adenosyl radical from substrates, and hydrogen back-abstraction (recombination) of a product-derived radical from 5'-deoxyadenosine. The energetic feasibility of these hydrogen abstraction steps in the diol dehyratase reaction was examined by theoretical calculations with a protein-free, simplified model at the B3LYP/6-311G* level of density functional theory. Activation energies for the hydrogen abstraction and recombination with 1,2-propanediol as substrate are 9.0 and 15.1 kcal/mol, respectively, and essentially not affected by coordination of the substrate and the radical intermediate to K+. Since these energies can be considered to be supplied by the substrate-binding energy, the computational results with this simplified model indicate that the hydrogen abstraction and recombination in the coenzyme B(12)-dependent diol dehydratase reaction are energetically feasible.  相似文献   

2.
Crystal structure of substrate free form of glycerol dehydratase   总被引:13,自引:0,他引:13  
Glycerol dehydratase (GDH) and diol dehydratase (DDH) are highly homologous isofunctional enzymes that catalyze the elimination of water from glycerol and 1,2-propanediol (1,2-PD) to the corresponding aldehyde via a coenzyme B(12)-dependent radical mechanism. The crystal structure of substrate free form of GDH in complex with cobalamin and K(+) has been determined at 2.5 A resolution. Its overall fold and the subunit assembly closely resemble those of DDH. Comparison of this structure and the DDH structure, available only in substrate bound form, shows the expected change of the coordination of the essential K(+) from hexacoordinate to heptacoordinate with the displacement of a single coordinated water by the substrate diol. In addition, there appears to be an increase in the rigidity of the K(+) coordination (as measured by lower B values) upon the binding of the substrate. Structural analysis of the locations of conserved residues among various GDH and DDH sequences has aided in identification of residues potentially important for substrate preference or specificity of protein-protein interactions.  相似文献   

3.
Alternative substrates and site-directed mutations of active-site residues are used to probe factors controlling the catalytic efficacy of scytalone dehydratase. In the E1cb-like, syn-elimination reactions catalyzed, efficient catalysis requires distortion of the substrate ring system to facilitate proton abstraction from its C2 methylene and elimination of its C3 hydroxyl group. Theoretical calculations indicate that such distortions are more readily achieved in the substrate 2,3-dihydro-2,5-dihydroxy-4H-benzopyran-4-one (DDBO) than in the physiological substrates vermelone and scytalone by approximately 2 kcal/mol. A survey of 12 active-site amino acid residues reveals 4 site-directed mutants (H110N, N131A, F53A, and F53L) have higher relative values of k(cat) and k(cat)/K(m) for DDBO over scytalone and for DDBO over vermelone than the wild-type enzyme, thus suggesting substrate-distortion roles for the native residues in catalysis. A structural link for this function is in the modeled enzyme-substrate complex where F53 and H110 are positioned above and below the substrate's C3 hydroxyl group, respectively, for pushing and pulling the leaving group into the axial orientation of a pseudo-boat conformation; N131 hydrogen-bonds to the C8 hydroxyl group at the opposite end of the substrate, serving as a pivot for the actions of F53 and H110. Deshydroxyvermelone lacks the phenolic hydroxyl group and the intramolecular hydrogen bond of vermelone. The relative values of k(cat) (95) and k(cat)/K(m) (1800) for vermelone over deshydroxyvermelone for the wild-type enzyme indicate the importance of the hydroxyl group for substrate recognition and catalysis. Off the enzyme, the much slower rates for the solvolytic dehydration of deshydroxyvermelone and vermelone are similar, thus specifying the importance of the hydroxyl group of vermelone for enzyme catalysis.  相似文献   

4.
BACKGROUND: Adenosylcobalamin (coenzyme B(12)) serves as a cofactor for enzymatic radical reactions. The adenosyl radical, a catalytic radical in these reactions, is formed by homolysis of the cobalt-carbon bond of the coenzyme, although the mechanism of cleavage of its organometallic bond remains unsolved. RESULTS: We determined the three-dimensional structures of diol dehydratase complexed with adeninylpentylcobalamin and with cyanocobalamin at 1.7 A and 1.9 A resolution, respectively, at cryogenic temperatures. In the adeninylpentylcobalamin complex, the adenine ring is bound parallel to the corrin ring as in the free form and methylmalonyl-CoA-mutase-bound coenzyme, but with the other side facing pyrrole ring C. All of its nitrogen atoms except for N(9) are hydrogen-bonded to mainchain amide oxygen and amide nitrogen atoms, a sidechain hydroxyl group, and a water molecule. As compared with the cyanocobalamin complex, the sidechain of Seralpha224 rotates by 120 degrees to hydrogen bond with N(3) of the adenine ring. CONCLUSIONS: The structure of the adenine-ring-binding site provides a molecular basis for the strict specificity of diol dehydratase for the coenzyme adenosyl group. The superimposition of the structure of the free coenzyme on that of enzyme-bound adeninylpentylcobalamin demonstrated that the tight enzyme-coenzyme interactions at both the cobalamin moiety and adenine ring of the adenosyl group would inevitably lead to cleavage of the cobalt-carbon bond. Rotation of the ribose moiety around the glycosidic linkage makes the 5'-carbon radical accessible to the hydrogen atom of the substrate to be abstracted.  相似文献   

5.
Background: Diol dehydratase is an enzyme that catalyzes the adenosylcobalamin (coenzyme B12) dependent conversion of 1,2-diols to the corresponding aldehydes. The reaction initiated by homolytic cleavage of the cobalt–carbon bond of the coenzyme proceeds by a radical mechanism. The enzyme is an α2β2γ2 heterooligomer and has an absolute requirement for a potassium ion for catalytic activity. The crystal structure analysis of a diol dehydratase–cyanocobalamin complex was carried out in order to help understand the mechanism of action of this enzyme.Results: The three-dimensional structure of diol dehydratase in complex with cyanocobalamin was determined at 2.2 Å resolution. The enzyme exists as a dimer of heterotrimers (α β γ)2. The cobalamin molecule is bound between the α and β subunits in the ‘base-on’ mode, that is, 5,6-dimethylbenzimidazole of the nucleotide moiety coordinates to the cobalt atom in the lower axial position. The α subunit includes a (β/α)8 barrel. The substrate, 1,2-propanediol, and an essential potassium ion are deeply buried inside the barrel. The two hydroxyl groups of the substrate coordinate directly to the potassium ion.Conclusions: This is the first crystallographic indication of the ‘base-on’ mode of cobalamin binding. An unusually long cobalt–base bond seems to favor homolytic cleavage of the cobalt–carbon bond and therefore to favor radical enzyme catalysis. Reactive radical intermediates can be protected from side reactions by spatial isolation inside the barrel. On the basis of unique direct interactions between the potassium ion and the two hydroxyl groups of the substrate, direct participation of a potassium ion in enzyme catalysis is strongly suggested.  相似文献   

6.
Diol dehydratase of Klebsiella oxytoca contains an essential histidine residue. Its X-ray structure revealed that the migrating hydroxyl group on C2 of substrate is hydrogen-bonded to Hisalpha143. Mutant enzymes in which Hisalpha143 was mutated to another amino acid residue were expressed in Escherichia coli, purified, and examined for enzymatic activity. The Halpha143Q mutant was 34% as active as the wild-type enzyme. Halpha143A and Halpha143L showed only a trace of activity. Kinetic analyses indicated that the hydrogen bonding interaction between the hydroxyl group on C2 of substrate and the side chain of residue alpha143 is important not only for catalysis but also for protecting radical intermediates. Halpha143E and Halpha143K that did not exist as (alphabetagamma) 2 complexes were inactive. The deuterium kinetic isotope effect on the overall reaction suggested that a hydrogen abstraction step is fully rate-determining for the wild type and Halpha143Q and partially rate-determining for Halpha143A. The preference for substrate enantiomers was reversed by the Halpha143Q mutation in both substrate binding and catalysis. Upon the inactivation of the Halpha143A holoenzyme by 1,2-propanediol, cob(II)alamin without an organic radical coupling partner accumulated, 5'-deoxyadenosine was quantitatively formed from the coenzyme adenosyl group, and the apoenzyme itself was not damaged. This inactivation was thus concluded to be a mechanism-based inactivation. The holoenzyme of Halpha143Q underwent irreversible inactivation by O 2 in the absence of substrate at a much lower rate than the wild type.  相似文献   

7.
Jiang YL  Ichikawa Y  Song F  Stivers JT 《Biochemistry》2003,42(7):1922-1929
The reaction catalyzed by the DNA repair enzyme uracil DNA glycosylase (UDG) proceeds through an unprecedented stepwise mechanism involving a positively charged oxacarbenium ion sugar and uracil anion leaving group. Here we use a novel approach to evaluate the catalytic contribution of electrostatic interactions between four essential phosphodiester groups of the DNA substrate and the cationic transition state. Our strategy was to substitute each of these phosphate groups with an uncharged (R)- or (S)-methylphosphonate linkage (MeP). We then compared the damaging effects of these methylphosphonate substitutions on catalysis with their damaging effects on binding of a cationic 1-azadeoxyribose (1-aza-dR(+)) oxacarbenium ion analogue to the UDG-uracil anion binary complex. A plot of log k(cat)/K(m) for the series of MeP-substituted substrates against log K(D) for binding of the 1-aza-dR(+) inhibitors gives a linear correlation of unit slope, confirming that the electronic features of the transition state resemble that of the 1-aza-dR(+), and that the anionic backbone of DNA is used in transition state stabilization. We estimate that all of the combined phosphodiester interactions with the substrate contribute 6-8 kcal/mol toward lowering the activation barrier, a stabilization that is significant compared to the 16 kcal/mol catalytic power of UDG. However, unlike groups of the enzyme that selectively stabilize the charged transition state by an estimated 7 kcal/mol, these phosphodiester groups also interact strongly in the ground state. To our knowledge, these results provide the first experimental evidence for electrostatic stabilization of a charged enzymatic transition state and intermediate using the anionic backbone of DNA.  相似文献   

8.
Sun L  Warncke K 《Proteins》2006,64(2):308-319
The structure of the EutB protein from Salmonella typhimurium, which contains the active site of the coenzyme B12 (adenosylcobalamin)-dependent enzyme, ethanolamine ammonia-lyase, has been predicted by using structural proteomics techniques of comparative modelling. The 453-residue EutB protein displays no significant sequence identity with proteins of known structure. Therefore, secondary structure prediction and fold recognition algorithms were used to identify templates. Multiple three-dimensional template matching (threading) servers identified predominantly beta8alpha8, TIM-barrel proteins, and in particular, the large subunits of diol dehydratase (PDB: 1eex:A, 1dio:A) and glycerol dehydratase (PDB: 1mmf:A), as templates. Consistent with this identification, the dehydratases are, like ethanolamine ammonia-lyase, Class II coenzyme B12-dependent enzymes. Model building was performed by using MODELLER. Models were evaluated by using different programs, including PROCHECK and VERIFY3D. The results identify a beta8alpha8, TIM-barrel fold for EutB. The beta8alpha8, TIM-barrel fold is consistent with a central role of the alpha/beta-barrel structures in radical catalysis conducted by the coenzyme B12- and S-adenosylmethionine-dependent (radical SAM) enzyme superfamilies. The EutB model and multiple sequence alignment among ethanolamine ammonia-lyase, diol dehydratase, and glycerol dehydratase from different species reveal the following protein structural features: (1) a "cap" loop segment that closes the N-terminal region of the barrel, (2) a common cobalamin cofactor binding topography at the C-terminal region of the barrel, and (3) a beta-barrel-internal guanidinium group from EutB R160 that overlaps the position of the active-site potassium ion found in the dehydratases. R160 is proposed to have a role in substrate binding and radical catalysis.  相似文献   

9.
Klebsiella pneumoniae ATCC 25955 (formerly named Aerobacter aerogenes PZH 572, Warsaw), which is known to produce coenzyme-B12-dependent glycerol dehydratase when grown anaerobically in a glycerol medium, formed coenzyme-B12-dependent diol dehydratase in a 1,2-propanediol-containing medium. Both the diol dehydratase and the glycerol dehydratase produced by the organism catalyzed the conversion of glycerol, 1,2-propanediol and 1,2-ethanediol to the corresponding aldehydes and underwent concomitant inactivation during the catalysis of glycerol dehydration, as does the diol dehydratase of K. pneumoniae (A. aerogenes) ATCC 8724. However, the two enzymes were distinguishable from each other by the monovalent-cation-selectivity pattern and by substrate specificity; that is, glycerol dehydratase preferred glycerol to 1,2-propanediol as a substrate, whereas diol dehydratase preferred 1,2-propanediol to glycerol, as judged from initial velocity studies. Ouchterlony double-diffusion analysis and immunochemical titration with rabbit antiserum against diol dehydratase of K. pneumoniae ATCC 8724 established clearly that the diol dehydratase of K. pneumoniae ATCC 25955 is immunologically similar to that of K. pneumoniae ATCC 8724, while the glycerol dehydratase of the former is different from the diol dehydratase of both strains. Both the enzymes were found to be distributed in several bacteria of the family Enterobacteriaceae.  相似文献   

10.
The reactions of diol dehydratase with 3-unsaturated 1,2-diols and thioglycerol were investigated. Holodiol dehydratase underwent rapid and irreversible inactivation by either 3-butene-1,2-diol, 3-butyne-1,2-diol or thioglycerol without catalytic turnovers. In the inactivation, the Co-C bond of adenosylcobalamin underwent irreversible cleavage forming unidentified radicals and cob(II)alamin that resisted oxidation even in the presence of oxygen. Two moles of 5'-deoxyadenosine per mol of enzyme was formed as an inactivation product from the coenzyme adenosyl group. Inactivated holoenzymes underwent reactivation by diol dehydratase-reactivating factor in the presence of ATP, Mg(2+) and adenosylcobalamin. It was thus concluded that these substrate analogues served as mechanism-based inactivators or pseudosubstrates, and that the coenzyme was damaged in the inactivation, whereas apoenzyme was not damaged. In the inactivation by 3-unsaturated 1,2-diols, product radicals stabilized by neighbouring unsaturated bonds might be unable to back-abstract the hydrogen atom from 5'-deoxyadenosine and then converted to unidentified products. In the inactivation by thioglycerol, a product radical may be lost by the elimination of sulphydryl group producing acrolein and unidentified sulphur compound(s). H(2)S or sulphide ion was not formed. The loss or stabilization of product radicals would result in the inactivation of holoenzyme, because the regeneration of the coenzyme becomes impossible.  相似文献   

11.
Adenosylcobalamin-dependent glycerol and diol dehydratases undergo inactivation by the physiological substrate glycerol during catalysis. In the permeabilized cells of Klebsiella pneumoniae, Klebsiella oxytoca, and recombinant Escherichia coli, glycerol-inactivated glycerol dehydratase and diol dehydratase are reactivated by their respective reactivating factors in the presence of ATP, Mg2+, and adenosylcobalamin. Both of the reactivating factors consist of two subunits. To examine the specificities of the reactivating factors, their genes or their hybrid genes were co-expressed with dehydratase genes in E. coli cells in various combinations. The reactivating factor of K. oxytoca for diol dehydratase efficiently cross-reactivated the inactivated glycerol dehydratase, whereas the reactivating factor of K. pneumoniae for glycerol dehydratase hardly cross-reactivated the inactivated diol dehydratase. Both of the two hybrid reactivating factors rapidly reactivated the inactivated glycerol dehydratase. In contrast, the hybrid reactivating factor containing the large subunit of the glycerol dehydratase reactivating factor hardly reactivated the inactivated diol dehydratase. These results indicate that the glycerol dehydratase reactivating factor is much more specific for the dehydratase partner than the diol dehydratase reactivating factor and that a large subunit of the reactivating factors principally determines the specificity for a dehydratase.  相似文献   

12.
Yamanishi M  Ide H  Murakami Y  Toraya T 《Biochemistry》2005,44(6):2113-2118
The reaction catalyzed by adenosylcobalamin-dependent diol dehydratase proceeds by a radical mechanism. A radical pair consisting of the Co(II) of cob(II)alamin and an organic radical intermediate formed during catalysis gives EPR spectra. The high-field doublet and the low-field broad signals arise from the weak interaction of an organic radical with the low-spin Co(II) of cob(II)alamin. To characterize the organic radical intermediate in the diol dehydratase reaction, several deuterated and (13)C-labeled 1,2-propanediols were synthesized, and the EPR spectra observed in the catalysis were measured using them as substrate. The EPR spectra with the substrates deuterated on C1 showed significant line width narrowing of the doublet signal. A distinct change in the hyperfine coupling was seen with [1-(13)C]-1,2-propanediol, but not with the [2-(13)C]-counterpart. Thus, the organic radical intermediate observed by EPR spectroscopy was identified as the 1,2-propanediol-1-yl radical, a C1-centered substrate-derived radical.  相似文献   

13.
Adenosylcobalamin-dependent diol dehydratase of Klebsiella oxytoca is apparently not stereospecific and catalyzes the conversion of both (R)- and (S)-1,2-propanediol to propionaldehyde. To explain this unusual property of the enzyme, we analyzed the crystal structures of diol dehydratase in complexes with cyanocobalamin and (R)- or (S)-1,2-propanediol. (R)- and (S)-isomers are bound in a symmetrical manner, although the hydrogen-bonding interactions between the substrate and the active-site residues are the same. From the position of the adenosyl radical in the modeled "distal" conformation, it is reasonable for the radical to abstract the pro-R and pro-S hydrogens from (R)- and (S)-isomers, respectively. The hydroxyl groups in the substrate radicals would migrates from C(2) to C(1) by a suprafacial shift, resulting in the stereochemical inversion at C(1). This causes 60 degrees clockwise and 70 degrees counterclockwise rotations of the C(1)-C(2) bond of the (R)- and (S)-isomers, respectively, if viewed from K+. A modeling study of 1,1-gem-diol intermediates indicated that new radical center C(2) becomes close to the methyl group of 5'-deoxyadenosine. Thus, the hydrogen back-abstraction (recombination) from 5'-deoxyadenosine by the product radical is structurally feasible. It was also predictable that the substitution of the migrating hydroxyl group by a hydrogen atom from 5'-deoxyadenosine takes place with the inversion of the configuration at C(2) of the substrate. Stereospecific dehydration of the 1,1-gem-diol intermediates can also be rationalized by assuming that Asp-alpha335 and Glu-alpha170 function as base catalysts in the dehydration of the (R)- and (S)-isomers, respectively. The structure-based mechanism and stereochemical courses of the reaction are proposed.  相似文献   

14.
The roles of the D-ribosyl moiety and the bulky axial ligand of the nucleotide loop of adenosylcobalamin in coenzymic function have been investigated using two series of coenzyme analogs bearing various artificial bases. The 2-methylbenzimidazolyl trimethylene analog that exists exclusively in the base-off form was a totally inactive coenzyme for diol dehydratase and served as a competitive inhibitor. The benzimidazolyl trimethylene analog and the benzimidazolylcobamide coenzyme were highly active for diol dehydratase and ethanolamine ammonia-lyase. The imidazolylcobamide coenzyme was 59 and 9% as active as the normal coenzyme for diol dehydratase and ethanolamine ammonia-lyase, respectively. The latter analog served as an effective suicide coenzyme for both enzymes, although the partition ratio (k(cat)/k(inact)) of 630 for ethanolamine ammonia-lyase is much lower than that for diol dehydratase. Suicide inactivation was accompanied by the accumulation of a cob(II)amide species, indicating irreversible cleavage of the coenzyme Co-C bond during the inactivation. It was thus concluded that the bulkiness of a Co-coordinating base of the nucleotide loop is essential for both the initial activity and continuous catalytic turnovers. Since the k(cat)/k(inact) value for the imidazolylcobamide in diol dehydratase was 27-times higher than that for the imidazolyl trimethylene analog, it is clear that the ribosyl moiety protects the reaction intermediates from suicide inactivation. Stopped-flow measurements indicated that the rate of Co-C bond homolysis is essentially unaffected by the bulkiness of the Co-coordinating base for diol dehydratase. Thus, it seems unlikely that the Co-C bond is labilized through a ground state mechanochemical triggering mechanism in diol dehydratase.  相似文献   

15.
The roles of particular amino acids in substrate and coenzyme binding and catalysis of glucose-6-phosphate dehydrogenase of Leuconostoc mesenteroides have been investigated by site-directed mutagenesis, kinetic analysis, and determination of binding constants. The enzyme from this species has functional dual NADP(+)/NAD(+) specificity. Previous investigations in our laboratories determined the three-dimensional structure. Kinetic studies showed an ordered mechanism for the NADP-linked reaction while the NAD-linked reaction is random. His-240 was identified as the catalytic base, and Arg-46 was identified as important for NADP(+) but not NAD(+) binding. Mutations have been selected on the basis of the three-dimensional structure. Kinetic studies of 14 mutant enzymes are reported and kinetic mechanisms are reported for 5 mutant enzymes. Fourteen substrate or coenzyme dissociation constants have been measured for 11 mutant enzymes. Roles of particular residues are inferred from k(cat), K(m), k(cat)/K(m), K(d), and changes in kinetic mechanism. Results for enzymes K182R, K182Q, K343R, and K343Q establish Lys-182 and Lys-343 as important in binding substrate both to free enzyme and during catalysis. Studies of mutant enzymes Y415F and Y179F showed no significant contribution for Tyr-415 to substrate binding and only a small contribution for Tyr-179. Changes in kinetics for T14A, Q47E, and R46A enzymes implicate these residues, to differing extents, in coenzyme binding and discrimination between NADP(+) and NAD(+). By the same measure, Lys-343 is also involved in defining coenzyme specificity. Decrease in k(cat) and k(cat)/K(m) for the D374Q mutant enzyme defines the way Asp-374, unique to L. mesenteroides G6PD, modulates stabilization of the enzyme during catalysis by its interaction with Lys-182. The greatly reduced k(cat) values of enzymes P149V and P149G indicate the importance of the cis conformation of Pro-149 in accessing the correct transition state.  相似文献   

16.
The importance of each active-site residue in adenosylcobalamin-dependent diol dehydratase of Klebsiella oxytoca was estimated using mutant enzymes in which one of the residues interacting with substrate and/or K(+) was mutated to Ala or another amino acid residue. The Ealpha170A and Dalpha335A mutants were totally inactive, and the Halpha143A mutant showed only a trace of activity, indicating that Glu-alpha170, Asp-alpha335, and His-alpha143 are catalytic residues. The Qalpha141A, Qalpha296A, and Salpha362A mutants showed partial activity. It was suggested from kinetic parameters that Gln-alpha296 is important for substrate binding and Gln-alpha296 and Gln-alpha141 for preventing the enzyme from mechanism-based inactivation. The Ealpha221A, Ealpha170H, and Dalpha335A did not form the (alphabetagamma)(2) complex, suggesting that these mutations indirectly disrupt subunit contacts. Among other Glu-alpha170 and Asp-alpha335 mutants, Ealpha170D and Ealpha170Q were 2.2 +/- 0.3% and 0.02% as active as the wild-type enzyme, respectively, whereas Dalpha335N was totally inactive. Kinetic analysis indicated that the presence and the position of a carboxyl group in the residue alpha170 are essential for catalysis as well as for the continuous progress of catalytic cycles. It was suggested that the roles of Glu-alpha170 and Asp-alpha335 are to participate in the binding of substrate and intermediates and keep them appropriately oriented and to function as a base in the dehydration of the 1,1-diol intermediate. In addition, Glu-alpha170 seems to stabilize the transition state for the hydroxyl group migration from C2 to C1 by accepting the proton of the spectator hydroxyl group on C1.  相似文献   

17.
[Omega-(Adenosyl)alkyl]cobalamins (homoadenosylcobalamins) are useful analogues of adenosylcobalamin to get information about the distance between Co and C5', which is critical for Co-C bond activation. In order to use them as probes for exploring the active sites of enzymes, the coenzymic properties of homoadenosylcobalamins for diol dehydratase and ethanolamine ammonia-lyase were investigated. The kcat and kcat/Km values for adenosylmethylcobalamin were about 0.27% and 0.15% that for the regular coenzyme with diol dehydratase, respectively. The kcat/kinact value showed that the holoenzyme with this analogue becomes inactivated on average after about 3000 catalytic turnovers, indicating that the probability of inactivation during catalysis is almost 500 times higher than that for the regular holoenzyme. The kcat value for adenosylmethylcobalamin was about 0.13% that of the regular coenzyme for ethanolamine ammonia-lyase, as judged from the initial velocity, but the holoenzyme with this analogue underwent inactivation after on average about 50 catalytic turnovers. This probability of inactivation is 3800 times higher than that for the regular holoenzyme. When estimated from the spectra of reacting holoenzymes, the steady state concentration of cob(II)alamin intermediate from adenosylmethylcobalamin was very low with either diol dehydratase or ethanolamine ammonia-lyase, which is consistent with its extremely low coenzymic activity. In contrast, neither adenosylethylcobalamin nor adeninylpentylcobalamin served as active coenzyme for either enzyme and did not undergo Co-C bond cleavage upon binding to apoenzymes.  相似文献   

18.
Leucoanthocyanidin reductase (LAR) catalyzes the NADPH-dependent reduction of 2R,3S,4S-flavan-3,4-diols into 2R,3S-flavan-3-ols, a subfamily of flavonoids that is important for plant survival and for human nutrition. LAR1 from Vitis vinifera has been co-crystallized with or without NADPH and one of its natural products, (+)-catechin. Crystals diffract to a resolution between 1.75 and 2.72 Å. The coenzyme and substrate binding pocket is preformed in the apoprotein and not markedly altered upon NADPH binding. The structure of the abortive ternary complex, determined at a resolution of 2.28 Å, indicates the ordering of a short 310 helix associated with substrate binding and suggests that His122 and Lys140 act as acid-base catalysts. Based on our 3D structures, a two-step catalytic mechanism is proposed, in which a concerted dehydration precedes an NADPH-mediated hydride transfer at C4. The dehydration step involves a Lys-catalyzed deprotonation of the phenolic OH7 through a bridging water molecule and a His-catalyzed protonation of the benzylic hydroxyl at C4. The resulting quinone methide serves as an electrophilic target for hydride transfer at C4. LAR belongs to the short-chain dehydrogenase/reductase superfamily and to the PIP (pinoresinol-lariciresinol reductase, isoflavone reductase, and phenylcoumaran benzylic ether reductase) family. Our data support the concept that all PIP enzymes reduce a quinone methide intermediate and that the major role of the only residue that has been conserved from the short-chain dehydrogenase/reductase catalytic triad (Ser…TyrXXXLys), that is, lysine, is to promote the formation of this intermediate by catalyzing the deprotonation of a phenolic hydroxyl. For some PIP enzymes, this lysine-catalyzed proton abstraction may be sufficient to trigger the extrusion of the leaving group, whereas in LAR, the extrusion of a hydroxide group requires a more sophisticated mechanism of concerted acid-base catalysis that involves histidine and takes advantage of the OH4, OH5, and OH7 substituents of leucoanthocyanidins.  相似文献   

19.
Sohn J  Rudolph J 《Biophysical chemistry》2007,125(2-3):549-555
Using a combination of steady-state and single-turnover kinetics, we probe the temperature dependence of substrate association and chemistry for the reaction of Cdc25B phosphatase with its Cdk2-pTpY/CycA protein substrate. The transition state for substrate association is dominated by an enthalpic barrier (DeltaH(++) of 13 kcal/mol) and has a favorable entropic contribution of 4 kcal/mol at 298 K. Phosphate transfer from Cdk2-pTpY/CycA to enzyme (DeltaH(++) of 12 kcal/mol) is enthalpically more favorable than for the small molecule substrate p-nitrophenyl phosphate (DeltaH(++) of 18 kcal/mol), yet entropically less favorable (TDeltaS(++) of 2 vs. -6 kcal/mol at 298 K, respectively). By measuring the temperature dependence of binding and catalysis for several hotspot mutants involved in binding of protein substrate, we determine the enthalpy-entropy compensations for changes in rates of association and phosphate transfer compared to the wild type system. We conclude that the transition state for enzyme-substrate association involves tight and specific contacts at the remote docking site and that phospho-transfer from Cdk2-pTpY/CycA to the pre-organized active site of the enzyme is accompanied by unfavorable entropic rearrangements that promote rapid product dissociation.  相似文献   

20.
Previous studies have shown that the interaction of P450 reductase with bound NADP(H) is essential to ensure fast electron transfer through the two flavin cofactors. In this study we investigated in detail the interaction of the house fly flavoprotein with NADP(H) and a number of nucleotide analogues. 1,4,5,6-Tetrahydro-NADP, an analogue of NADPH, was used to characterize the interaction of P450 reductase with the reduced nucleotide. This analogue is inactive as electron donor, but its binding affinity and rate constant of release are very close to those for NADPH. The 2'-phosphate contributes about 5 kcal/mol of the binding energy of NADP(H). Oxidized nicotinamide does not interact with the oxidized flavoprotein, while reduced nicotinamide contributes 1.3 kcal/mol of the binding energy. Oxidized P450 reductase binds NADPH with a K(d) of 0.3 microM, while the affinity of the reduced enzyme is considerably lower, K(d) = 1.9 microM. P450 reductase catalyzes a transhydrogenase reaction between NADPH and oxidized nucleotides, such as thionicotinamide-NADP(+), acetylpyridine-NADP(+), or [(3)H]NADP(+). The reverse reaction, reduction of [(3)H]NADP(+) by the reduced analogues, is also catalyzed by P450 reductase. We define the mechanism of the transhydrogenase reaction as follows: NADPH binding, hydride ion transfer, and release of the NADP(+) formed. An NADP(+) or its analogue binds to the two-electron-reduced flavoprotein, and the electron-transfer steps reverse to transfer hydride ion to the oxidized nucleotide, which is released. Measurements of the flavin semiquinone content, rate constant for NADPH release, and transhydrogenase turnover rates allowed us to estimate the steady-state distribution of P450 reductase species during catalysis, and to calculate equilibrium constants for the interconversion of catalytic intermediates. Our results demonstrate that equilibrium redox potentials of the flavin cofactors are not the sole factor governing rapid electron transfer during catalysis, but conformational changes must be considered to understand P450 reductase catalysis.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号