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 15 毫秒
1.
The anaerobic dehalogenation of organohalides is catalyzed by the reductive dehalogenase (RdhA) enzymes produced in phylogenetically diverse bacteria. These enzymes contain a cobamide cofactor at the active site and two iron-sulfur clusters. In this study, the tetrachloroethene (PCE) reductive dehalogenase (PceA) of the Gram-positive Desulfitobacterium hafniense strain Y51 was produced in a catalytically active form in the nondechlorinating, cobamide-producing bacterium Shimwellia blattae (ATCC 33430), a Gram-negative gammaproteobacterium. The formation of recombinant catalytically active PceA enzyme was significantly enhanced when its dedicated PceT chaperone was coproduced and when 5,6-dimethylbenzimidazole and hydroxocobalamin were added to the S. blattae cultures. The experiments were extended to D. hafniense DCB-2, a reductively dehalogenating bacterium harboring multiple rdhA genes. To elucidate the substrate spectrum of the rdhA3 gene product of this organism, the recombinant enzyme was tested for the conversion of different dichlorophenols (DCP) in crude extracts of an RdhA3-producing S. blattae strain. 3,5-DCP, 2,3-DCP, and 2,4-DCP, but not 2,6-DCP and 3,4-DCP, were reductively dechlorinated by the recombinant RdhA3. In addition, this enzyme dechlorinated PCE to trichloroethene at low rates.  相似文献   

2.
The structures of two alkylurea inhibitors complexed with murine soluble epoxide hydrolase have been determined by x-ray crystallographic methods. The alkyl substituents of each inhibitor make extensive hydrophobic contacts in the soluble epoxide hydrolase active site, and each urea carbonyl oxygen accepts hydrogen bonds from the phenolic hydroxyl groups of Tyr(381) and Tyr(465). These hydrogen bond interactions suggest that Tyr(381) and/or Tyr(465) are general acid catalysts that facilitate epoxide ring opening in the first step of the hydrolysis reaction; Tyr(465) is highly conserved among all epoxide hydrolases, and Tyr(381) is conserved among the soluble epoxide hydrolases. In one enzyme-inhibitor complex, the urea carbonyl oxygen additionally interacts with Gln(382). If a comparable interaction occurs in catalysis, then Gln(382) may provide electrostatic stabilization of partial negative charge on the epoxide oxygen. The carboxylate side chain of Asp(333) accepts a hydrogen bond from one of the urea NH groups in each enzyme-inhibitor complex. Because Asp(333) is the catalytic nucleophile, its interaction with the partial positive charge on the urea NH group mimics its approach toward the partial positive charge on the electrophilic carbon of an epoxide substrate. Accordingly, alkylurea inhibitors mimic features encountered in the reaction coordinate of epoxide ring opening, and a structure-based mechanism is proposed for leukotoxin epoxide hydrolysis.  相似文献   

3.
Matthias Boll 《BBA》2005,1707(1):34-50
Several novel enzyme reactions have recently been discovered in the aromatic metabolism of anaerobic bacteria. Many of these reactions appear to be catalyzed by oxygen-sensitive enzymes by means of highly reactive radical intermediates. This contribution deals with two key reactions in this metabolism: the ATP-driven reductive dearomatisation of the benzene ring and the reductive removal of a phenolic hydroxyl group. The two reactions catalyzed by benzoyl-CoA reductase (BCR) and 4-hydroxybenzoyl-CoA reductase (4-HBCR) are both mechanistically difficult to achieve; both are considered to proceed in ‘Birch-like’ reductions involving single electron and proton transfer steps to the aromatic ring. The problem of both reactions is the extremely high redox barrier for the first electron transfer to the substrate (e.g., −1.9 V in case of a benzoyl-CoA (BCoA) analogue), which is solved in the two enzymes in different manners. Studying these enzymatic reactions provides insights into general principles of how oxygen-dependent reactions are replaced by alternative processes under anoxic conditions.  相似文献   

4.
Cobalamin-dependent enzymes enhance the rate of C–Co bond cleavage by up to ∼1012-fold to generate cob(II)alamin and a transient adenosyl radical. In the case of the pyridoxal 5′-phosphate (PLP) and cobalamin-dependent enzymes lysine 5,6-aminomutase and ornithine 4,5 aminomutase (OAM), it has been proposed that a large scale domain reorientation of the cobalamin-binding domain is linked to radical catalysis. Here, OAM variants were designed to perturb the interface between the cobalamin-binding domain and the PLP-binding TIM barrel domain. Steady-state and single turnover kinetic studies of these variants, combined with pulsed electron-electron double resonance measurements of spin-labeled OAM were used to provide direct evidence for a dynamic interface between the cobalamin and PLP-binding domains. Our data suggest that following ligand binding-induced cleavage of the Lys629-PLP covalent bond, dynamic motion of the cobalamin-binding domain leads to conformational sampling of the available space. This supports radical catalysis through transient formation of a catalytically competent active state. Crucially, it appears that the formation of the state containing both a substrate/product radical and Co(II) does not restrict cobalamin domain motion. A similar conformational sampling mechanism has been proposed to support rapid electron transfer in a number of dynamic redox systems.  相似文献   

5.
The three-dimensional structures of human and rabbit liver cytosolic recombinant serine hydroxymethyltransferases (hcSHMT and rcSHMT) revealed that E75 and Y83 (numbering according to hcSHMT) are probable candidates for proton abstraction and Calpha-Cbeta bond cleavage in the reaction catalyzed by serine hydroxymethyltransferase. Both these residues are completely conserved in all serine hydroxymethyltransferases sequenced to date. In an attempt to decipher the role of these residues in sheep liver cytosolic recombinant serine hydroxymethyltransferase (scSHMT), E74 (corresponding residue is E75 in hcSHMT) was mutated to Q and K, and Y82 (corresponding residue is Y83 in hcSHMT) was mutated to F. The specific activities using serine as the substrate for the E74Q and E74K mutant enzymes were drastically reduced. These mutant enzymes catalyzed the transamination of D-alanine and 5,6,7, 8-tetrahydrofolate independent retroaldol cleavage of Lallo threonine at rates comparable with wild-type enzyme, suggesting that E74 was not involved directly in the proton abstraction step of catalysis, as predicted earlier from crystal structures of hcSHMT and rcSHMT. There was no change in the apparent Tm value of E74Q upon the addition of L-serine, whereas the apparent Tm value of scSHMT was enhanced by 10 degrees C. Differential scanning calorimetric data and proteolytic digestion patterns in the presence of L-serine showed that E74Q was different to scSHMT. These results indicated that E74 might be required for the conformational change involved in reaction specificity. It was predicted from the crystal structures of hcSHMT and rcSHMT that Y82 was involved in hemiacetal formation following Calpha-Cbeta bond cleavage of L-serine and mutation of this residue to F could lead to a rapid release of HCHO. However, the Y82F mutant had only 5% of the activity and failed to form a quinonoid intermediate, suggesting that this residue is not involved in the formation of the hemiacetal intermediate, but might be involved indirectly in the abstraction of the proton and in stabilizing the quinonoid intermediate.  相似文献   

6.
Thibodeaux CJ  Liu HW 《Biochemistry》2011,50(11):1950-1962
1-Aminocyclopropane-1-carboxylic acid (ACC) deaminase (ACCD) is a pyridoxal 5'-phosphate (PLP)-dependent enzyme that cleaves the cyclopropane ring of ACC, to give α-ketobutyric acid and ammonia as products. The cleavage of the C(α)-C(β) bond of an amino acid substrate is a rare event in PLP-dependent enzyme catalysis. Potential chemical mechanisms involving nucleophile- or acid-catalyzed cyclopropane ring opening have been proposed for the unusual transformation catalyzed by ACCD, but the actual mode of cyclopropane ring cleavage remains obscure. In this report, we aim to elucidate the mechanistic features of ACCD catalysis by investigating the kinetic properties of ACCD from Pseudomonas sp. ACP and several of its mutant enzymes. Our studies suggest that the pK(a) of the conserved active site residue, Tyr294, is lowered by a hydrogen bonding interaction with a second conserved residue, Tyr268. This allows Tyr294 to deprotonate the incoming amino group of ACC to initiate the aldimine exchange reaction between ACC and the PLP coenzyme and also likely helps to activate Tyr294 for a role as a nucleophile to attack and cleave the cyclopropane ring of the substrate. In addition, solvent kinetic isotope effect (KIE), proton inventory, and (13)C KIE studies of the wild type enzyme suggest that the C(α)-C(β) bond cleavage step in the chemical mechanism is at least partially rate-limiting under k(cat)/K(m) conditions and is likely preceded in the mechanism by a partially rate-limiting step involving the conversion of a stable gem-diamine intermediate into a reactive external aldimine intermediate that is poised for cyclopropane ring cleavage. When viewed within the context of previous mechanistic and structural studies of ACCD enzymes, our studies are most consistent with a mode of cyclopropane ring cleavage involving nucleophilic catalysis by Tyr294.  相似文献   

7.
The bacterial enzyme designated QhpD belongs to the radical S-adenosyl-l-methionine (SAM) superfamily of enzymes and participates in the post-translational processing of quinohemoprotein amine dehydrogenase. QhpD is essential for the formation of intra-protein thioether bonds within the small subunit (maturated QhpC) of quinohemoprotein amine dehydrogenase. We overproduced QhpD from Paracoccus denitrificans as a stable complex with its substrate QhpC, carrying the 28-residue leader peptide that is essential for the complex formation. Absorption and electron paramagnetic resonance spectra together with the analyses of iron and sulfur contents suggested the presence of multiple (likely three) [4Fe-4S] clusters in the purified and reconstituted QhpD. In the presence of a reducing agent (sodium dithionite), QhpD catalyzed the multiple-turnover reaction of reductive cleavage of SAM into methionine and 5′-deoxyadenosine and also the single-turnover reaction of intra-protein sulfur-to-methylene carbon thioether bond formation in QhpC bound to QhpD, producing a multiknotted structure of the polypeptide chain. Homology modeling and mutagenic analysis revealed several conserved residues indispensable for both in vivo and in vitro activities of QhpD. Our findings uncover another challenging reaction catalyzed by a radical SAM enzyme acting on a ribosomally translated protein substrate.  相似文献   

8.
The catalytic mechanism of epoxide hydrolase (EC 3.3.2.3) involves acid-assisted ring opening of the oxirane during the alkylation half-reaction of hydrolysis. Two tyrosyl residues in the active site of epoxide hydrolases have been shown to contribute to the catalysis of enzyme alkylation, but their mechanism of action has not been fully described. We have investigated the involvement of the active site Tyr154 and Tyr235 during S,S-trans-stilbene oxide hydrolysis catalyzed by potato epoxide hydrolase StEH1. Tyr phenol ionizations of unliganded enzyme as well as under pre-steady-state conditions during catalysis were studied by direct absorption spectroscopy. A transient UV absorption, indicative of tyrosinate formation, was detected during the lifetime of the alkyl-enzyme intermediate. The apparent pKa of Tyr ionization was 7.3, a value more than 3 pH units below the estimated pKa of protein Tyr residues in the unliganded enzyme. In addition, the pH dependencies of microscopic kinetic rates of catalyzed S,S-trans-stilbene oxide hydrolysis were determined. The alkylation rate increased with pH and displayed a pKa value identical to that of Tyr ionization (7.3), whereas the reverse (epoxidation) reaction did not display any pH dependence. The rate of alkyl-enzyme hydrolysis was inversely dependent on tyrosinate formation, decreasing with its buildup in the active site. Since alkyl-enzyme hydrolysis is the rate-limiting step of the overall reaction, kcat displayed the same decrease with pH as the hydrolysis rate. The compiled results suggested that the role of the Tyr154/Tyr235 pair was not as ultimate proton donor to the alkoxide anion but to stabilize the negatively charged alkyl-enzyme through electrophilic catalysis via hydrogen bonding.  相似文献   

9.
10.
Lévêque VJ  Vance CK  Nick HS  Silverman DN 《Biochemistry》2001,40(35):10586-10591
The redox potential of human manganese superoxide dismutase (MnSOD) has been difficult to determine because of the problem of finding suitable electron mediators. We have found that ferricyanide and pentacyanoaminoferrate can be used as electron mediators, although equilibration is very slow with a half-time near 6 h. Values of the midpoint potential were determined both by allowing enzyme and mediators to equilibrate up to 38 h and by reductive titration adding dithionite to enzyme and mediator. An overall value of the midpoint potential was found to be 393 +/- 29 mV. To elucidate the role of His30 and Tyr34 in the active site of human MnSOD, we have also measured the redox properties of the site-specific mutants His30Asn (H30N) and Tyr34Phe (Y34F) and compared them with the wild-type enzyme. Crystal structures have shown that each mutation interrupts a hydrogen bond network in the active site, and each causes a 10-fold decrease in the maximal velocity of catalysis of superoxide dismutation as compared with wild type. The present study shows that H30N and Y34F human MnSOD have very little effect, within experimental uncertainty, on the redox potential of the active-site metal. The redox potentials determined electrochemically were 365 +/- 28 mV for H30N and 435 +/- 30 mV for Y34F MnSOD. These results suggest that the role of His30 and Tyr34 is more in support of catalysis, probably proton transport, and not in the tuning of the redox potential.  相似文献   

11.
Haloalcohol dehalogenases are bacterial enzymes that catalyze the cofactor-independent dehalogenation of vicinal haloalcohols such as the genotoxic environmental pollutant 1,3-dichloro-2-propanol, thereby producing an epoxide, a chloride ion and a proton. Here we present X-ray structures of the haloalcohol dehalogenase HheC from Agrobacterium radiobacter AD1, and complexes of the enzyme with an epoxide product and chloride ion, and with a bound haloalcohol substrate mimic. These structures support a catalytic mechanism in which Tyr145 of a Ser-Tyr-Arg catalytic triad deprotonates the haloalcohol hydroxyl function to generate an intramolecular nucleophile that substitutes the vicinal halogen. Haloalcohol dehalogenases are related to the widespread family of NAD(P)H-dependent short-chain dehydrogenases/reductases (SDR family), which use a similar Ser-Tyr-Lys/Arg catalytic triad to catalyze reductive or oxidative conversions of various secondary alcohols and ketones. Our results reveal the first structural details of an SDR-related enzyme that catalyzes a substitutive dehalogenation reaction rather than a redox reaction, in which a halide-binding site is found at the location of the NAD(P)H binding site. Structure-based sequence analysis reveals that the various haloalcohol dehalogenases have likely originated from at least two different NAD-binding SDR precursors.  相似文献   

12.
In reaction center proteins of photosynthetic bacteria, the amplitude of proton uptake induced by the one-electron reduction of either of the two quinone electron acceptors (Q(A) and Q(B)) is an intrinsic observable of the electrostatic interactions associated with the redox function of the complex. We report here that, in Rhodobacter capsulatus, complete restoration of proton uptake (upon formation of Q(A)(-) and Q(B)(-)) to the level found in the wild type is observed in a mutant reaction center in which a tyrosine substitution in the Q(A) environment (Ala(M274) --> Tyr) is coupled with mutations of acidic residues near Q(B) (Glu(L212) --> Ala/Asp(L213) --> Ala) that initially cancel the proton uptake above pH 8. This result demonstrates that proton uptake occurs by strong cooperation between structural motifs, such as hydrogen-bonded networks, that span the 18 A distance between the two quinone acceptors.  相似文献   

13.
Hydroxymethylglutaryl-CoA synthase-catalyzed condensation of acetyl-CoA with acetoacetyl-CoA requires enolization/carbanion formation from the acetyl C-2 methyl group prior to formation of a new carbon-carbon bond. Acetyldithio-CoA, a readily enolizable analog of acetyl-CoA, was an effective competitive inhibitor of avian hydroxymethylglutaryl-CoA synthase (Ki = 28 microm). In the absence of cosubstrate, enzyme catalyzed the enolization/proton exchange from the C-2 methyl group of acetyldithio-CoA. Mutant enzymes that exhibited impaired formation of the covalent acetyl-S-enzyme reaction intermediate exhibited diminished (D159A and D203A) or undetectable (C129S) rates of enolization of acetyldithio-CoA. The results suggest that covalent thioacetylation of protein, which has not been detected previously for other enzymes that enolize this analog, occurs with hydroxymethylglutaryl-CoA synthase. Enzyme catalyzed the transfer of the thioacetyl group of this analog to 3'-dephospho-CoA suggesting the intermediacy of a covalent thioacetyl-S-enzyme species, which appears to be important for proton abstraction from C-2 of the thioacetyl group. Avian enzyme glutamate 95 is crucial to substrate condensation to form a new carboncarbon bond. Mutations of this invariant residue (avian enzyme E95A and E95Q; Staphylococcus aureus enzyme E79Q) correlated with diminished ability to catalyze enolization of acetyldithio-CoA. Enolization by E95Q was not stimulated in the presence of acetoacetyl-CoA. These observations suggest either a direct (proton abstraction) or indirect (solvent polarization) role for this active site glutamate.  相似文献   

14.
Protein dynamics are crucial for realizing the catalytic power of enzymes, but how enzymes have evolved to achieve catalysis is unknown. The light-activated enzyme protochlorophyllide oxidoreductase (POR) catalyzes sequential hydride and proton transfers in the photoexcited and ground states, respectively, and is an excellent system for relating the effects of motions to catalysis. Here, we have used the temperature dependence of isotope effects and solvent viscosity measurements to analyze the dynamics coupled to the hydride and proton transfer steps in three cyanobacterial PORs and a related plant enzyme. We have related the dynamic profiles of each enzyme to their evolutionary origin. Motions coupled to light-driven hydride transfer are conserved across all POR enzymes, but those linked to thermally activated proton transfer are variable. Cyanobacterial PORs require complex and solvent-coupled dynamic networks to optimize the proton donor-acceptor distance, but evolutionary pressures appear to have minimized such networks in plant PORs. POR from Gloeobacter violaceus has features of both the cyanobacterial and plant enzymes, suggesting that the dynamic properties have been optimized during the evolution of POR. We infer that the differing trajectories in optimizing a catalytic structure are related to the stringency of the chemistry catalyzed and define a functional adaptation in which active site chemistry is protected from the dynamic effects of distal mutations that might otherwise impact negatively on enzyme catalysis.  相似文献   

15.
The bacterial metabolism of epoxypropane formed from propylene oxidation uses the atypical cofactor coenzyme M (CoM, 2-mercaptoethanesulfonate) as the nucleophile for epoxide ring opening and as a carrier of intermediates that undergo dehydrogenation, reductive cleavage, and carboxylation to form acetoacetate in a three-step metabolic pathway. 2-Ketopropyl-CoM carboxylase/oxidoreductase (2-KPCC), the terminal enzyme of this pathway, is the only known member of the disulfide oxidoreductase family of enzymes that is a carboxylase. In the present work, the CoM analog 2-bromoethanesulfonate (BES) is shown to be a reversible inhibitor of 2-KPCC and hydroxypropyl-CoM dehydrogenase but not of epoxyalkane:CoM transferase. Further investigations revealed that BES is a time-dependent inactivator of dithiothreitol-reduced 2-KPCC, where the redox active cysteines are in the free thiol forms. BES did not inactivate air-oxidized 2-KPCC, where the redox active cysteine pair is in the disulfide form. The inactivation of 2-KPCC exhibited saturation kinetics, and CoM slowed the rate of inactivation. Mass spectral analysis demonstrated that BES inactivation of reduced 2-KPCC occurs with covalent modification of the interchange thiol (Cys82) by a group with a molecular mass identical to that of ethylsulfonate. The flavin thiol Cys87 was not alkylated by BES under reducing conditions, and no amino acid residues were modified by BES in the oxidized enzyme. The UV-visible spectrum of BES-modifed 2-KPCC showed the characteristic charge transfer absorbance expected with alkylation at Cys82. These results identify BES as a reactive CoM analog that specifically alkylates the interchange thiol that facilitates thioether bond cleavage and enolacetone formation during catalysis.  相似文献   

16.
Using DTT(red) as the reducing agent, the kinetics of the reductive unfolding of onconase, a frog ribonuclease, has been examined. An intermediate containing three disulfides, Ir, that is formed rapidly in the reductive pathway, is more resistant to further reduction than the parent molecule, indicating that the remaining disulfides in onconase are less accessible to DTT(red). Disulfide-bond mapping of Ir indicated that it is a single species lacking the (30-75) disulfide bond. The reductive unfolding pattern of onconase is consistent with an analysis of the exposed surface area of the cysteine sulfur atoms in the (30-75) disulfide bond, which reveals that these atoms are about four- and sevenfold, respectively, more exposed than those in the next two maximally exposed disulfides. By contrast, in the reductive unfolding of the homologue, RNase A, there are two intermediates, arising from the reduction of the (40-95) and (65-72) disulfide bonds, which takes place in parallel, and on a much longer time-scale, compared to the initial reduction of onconase; this behavior is consistent with the almost equally exposed surface areas of the cysteine sulfur atoms that form the (40-95) and (65-72) disulfide bonds in RNase A and the fourfold more exposed cysteine sulfur atoms of the (30-75) disulfide bond in onconase. Analysis and in silico mutation of the residues around the (40-95) disulfide bond in RNase A, which is analogous to the (30-75) disulfide bond of onconase, reveal that the side-chain of tyrosine 92 of RNase A, a highly conserved residue among mammalian pancreatic ribonucleases, lies atop the (40-95) disulfide bond, resulting in a shielding of the corresponding sulfur atoms from the solvent; such burial of the (30-75) sulfur atoms is absent from onconase, due to the replacement of Tyr92 by Arg73, which is situated away from the (30-75) disulfide bond and into the solvent, resulting in the large exposed surface-area of the cysteine sulfur atoms forming this bond. Removal of Tyr92 from RNase A resulted in the relatively rapid reduction of the mutant to form a single intermediate (des [40-95] Y92A), i.e. it resulted in an onconase-like reductive unfolding behavior. The reduction of the P93A mutant of RNase A proceeds through a single intermediate, the des [40-95] P93A species, as in onconase. Although mutation of Pro93 to Ala does not increase the exposed surface area of the (40-95) cysteine sulfur atoms, structural analysis of the mutant reveals that there is greater flexibility in the (40-95) disulfide bond compared to the (65-72) disulfide bond that may make the (40-95) disulfide bond much easier to expose, consistent with the reductive unfolding pathway and kinetics of P93A. Mutation of Tyr92 to Phe92 in RNase A has no effect on its reductive unfolding pathway, suggesting that the hydrogen bond between the hydroxyl group of Tyr92 and the carbonyl group of Lys37 has no impact on the local unfolding free energy required to expose the (40-95) disulfide bond. Thus, these data shed light on the differences between the reductive unfolding pathways of the two homologous proteins and provide a structural basis for the origin of this difference.  相似文献   

17.
The cytochrome bc(1) complex catalyzes electron transfer from ubiquinol to cytochrome c by a protonmotive Q cycle mechanism in which electron transfer is linked to proton translocation across the inner mitochondrial membrane. In the Q cycle mechanism proton translocation is the net result of topographically segregated reduction of quinone and reoxidation of quinol on opposite sides of the membrane, with protons being carried across the membrane as hydrogens on the quinol. The linkage of proton chemistry to electron transfer during quinol oxidation and quinone reduction requires pathways for moving protons to and from the aqueous phase and the hydrophobic environment in which the quinol and quinone redox reactions occur. Crystal structures of the mitochondrial cytochrome bc(1) complexes in various conformations allow insight into possible proton conduction pathways. In this review we discuss pathways for proton conduction linked to ubiquinone redox reactions with particular reference to recently determined structures of the yeast bc(1) complex.  相似文献   

18.
Sequence alignment of pig mitochondrial NADP-dependent isocitrate dehydrogenase with eukaryotic (human, rat, and yeast) and Escherichia coli isocitrate dehydrogenases reveals that Tyr316 is completely conserved and is equivalent to the E. coli Tyr345, which interacts with the 2'-phosphate of NADP in the crystal structure [Hurley et al., Biochemistry 30 (1991) 8671-8678]. Lys321 is also completely conserved in the five isocitrate dehydrogenases. Either an arginine or lysine residue is found among the enzymes from other species at the position corresponding to the pig enzyme Arg314. While Arg323 is not conserved among all species, its proximity to the coenzyme site makes it a good candidate for investigation. The importance of these four amino acids to the function of pig mitochondrial NADP-isocitrate dehydrogenase was studied by site-directed mutagenesis. Mutants (R314Q, Y316F, Y316L, K321Q, and R323Q) were generated by a megaprimer polymerase chain reaction method. Wild-type and mutant enzymes were expressed in E. coli and purified to homogeneity. All mutant and wild-type enzymes exhibited comparable molecular weights indicative of the dimeric enzyme. Mutations do not cause an appreciable change in enzyme secondary structure as revealed by circular dichroism measurements. The kinetic parameters (V(max) and K(M) values) of K321Q and R323Q are similar to those of wild-type, indicating that Lys321 and Arg323 are not involved in enzyme function. R314Q exhibits a 10-fold increase in K(M) for NADP as compared to that of wild-type, while they have comparable V(max) values. These results suggest that Arg314 contributes to the affinity between the enzyme and NADP. The hydroxyl group of Tyr316 is not required for enzyme function since Y316F exhibits similar kinetic parameters to those of wild-type. Y316L shows a 4-fold increase in K(M) for NADP and a decrease in V(max) as compared to wild-type, suggesting that the aromatic ring of the Tyr of isocitrate dehydrogenase contributes to the affinity for coenzyme, as well as to catalysis. The K(i) for NAD of R314Q, Y316F, and Y316L is comparable to that of wild-type, indicating that the Arg314 and Tyr316 may be located near the 2'-phosphate of enzyme-bound NADP.  相似文献   

19.
Several novel enzyme reactions have recently been discovered in the aromatic metabolism of anaerobic bacteria. Many of these reactions appear to be catalyzed by oxygen-sensitive enzymes by means of highly reactive radical intermediates. This contribution deals with two key reactions in this metabolism: the ATP-driven reductive dearomatisation of the benzene ring and the reductive removal of a phenolic hydroxyl group. The two reactions catalyzed by benzoyl-CoA reductase (BCR) and 4-hydroxybenzoyl-CoA reductase (4-HBCR) are both mechanistically difficult to achieve; both are considered to proceed in 'Birch-like' reductions involving single electron and proton transfer steps to the aromatic ring. The problem of both reactions is the extremely high redox barrier for the first electron transfer to the substrate (e.g., -1.9 V in case of a benzoyl-CoA (BCoA) analogue), which is solved in the two enzymes in different manners. Studying these enzymatic reactions provides insights into general principles of how oxygen-dependent reactions are replaced by alternative processes under anoxic conditions.  相似文献   

20.
Hydrogen bonds formed between photosynthetic reaction centers (RCs) and their cofactors were shown to affect the efficacy of electron transfer. The mechanism of such influence is determined by sensitivity of hydrogen bonds to electron density rearrangements, which alter hydrogen bonds potential energy surface. Quantum chemistry calculations were carried out on a system consisting of a primary quinone Q(A), non-heme Fe(2+) ion and neighboring residues(.) The primary quinone forms two hydrogen bonds with its environment, one of which was shown to be highly sensitive to the Q(A) state. In the case of the reduced primary quinone two stable hydrogen bond proton positions were shown to exist on [Q(A)-His(M219)] hydrogen bond line, while there is only one stable proton position in the case of the oxidized primary quinone. Taking into account this fact and also the ability of proton to transfer between potential energy wells along a hydrogen bond, theoretical study of temperature dependence of hydrogen bond polarization was carried out. Current theory was successfully applied to interpret dark P(+)/Q(A)(-) recombination rate temperature dependence.  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