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.
Klebsiella aerogenes urease uses a dinuclear nickel active site to catalyze urea hydrolysis at >10(14)-fold the spontaneous rate. To better define the enzyme mechanism, we examined the kinetics and structures for a suite of site-directed variants involving four residues at the active site: His320, His219, Asp221, and Arg336. Compared to wild-type urease, the H320A, H320N, and H320Q variants exhibit similar approximately 10(-)(5)-fold deficiencies in rates, modest K(m) changes, and disorders in the peptide flap covering their active sites. The pH profiles for these mutant enzymes are anomalous with optima near 6 and shoulders that extend to pH 9. H219A urease exhibits 10(3)-fold increased K(m) over that of native enzyme, whereas the increase is less marked ( approximately 10(2)-fold) in the H219N and H219Q variants that retain hydrogen bonding capability. Structures for these variants show clearly resolved active site water molecules covered by well-ordered peptide flaps. Whereas the D221N variant is only moderately affected compared to wild-type enzyme, D221A urease possesses low activity ( approximately 10(-)(3) that of native enzyme), a small increase in K(m), and a pH 5 optimum. The crystal structure for D221A urease is reminiscent of the His320 variants. The R336Q enzyme has a approximately 10(-)(4)-fold decreased catalytic rate with near-normal pH dependence and an unaffected K(m). Phenylglyoxal inactivates the R336Q variant at over half the rate observed for native enzyme, demonstrating that modification of non-active-site arginines can eliminate activity, perhaps by affecting the peptide flap. Our data favor a mechanism in which His219 helps to polarize the substrate carbonyl group, a metal-bound terminal hydroxide or bridging oxo-dianion attacks urea to form a tetrahedral intermediate, and protonation occurs via the general acid His320 with Asp221 and Arg336 orienting and influencing the acidity of this residue. Furthermore, we conclude that the simple bell-shaped pH dependence of k(cat) and k(cat)/K(m) for the native enzyme masks a more complex underlying pH dependence involving at least four pK(a)s.  相似文献   

2.
Xanthine/α-ketoglutarate (αKG) dioxygenase (XanA) is a non-heme mononuclear FeII enzyme that decarboxylates αKG to succinate and CO2 while hydroxylating xanthine to generate uric acid. In the absence of a XanA crystal structure, a homology model was used to target several putative active site residues for mutagenesis. Wild-type XanA and ten enzyme variants were purified from recombinant Escherichia coli cells and characterized. The H149A and D151A variants were inactive and the H340A variant exhibited only 0.17% of the wild-type enzyme activity, consistent with the proposed role of His149, Asp151, and His340 as Fe ligands. The K122A variant led to a 2-fold increase in the Kd of αKG as measured by fluorescence quenching analysis, in agreement with Lys122 acting to stabilize the binding of αKG. The N358A variant exhibited a 23-fold decrease in kcat/Km compared to wild-type XanA, pointing to a key role of Asn358 in catalysis. 9-Methylxanthine was exploited as an alternate substrate, and the C357A, E137A, and D138A variants were found to exhibit relatively enhanced activity consistent with Cys357, Glu137, and Asp138 being proximal to N-9 or involved in its proper positioning. 6,8-Dihydroxypurine was identified as a slow-binding competitive inhibitor of XanA, and significant decreases (E137A and D138A) or increases (Q356A and N358A) in of the variants were interpreted in terms of distinct interactions between this compound and the corresponding active site side chains. Further support for Cys357 residing at the active site was obtained using thiol-specific reagents that inactivated wild-type enzyme (with partial protection by substrate), whereas the C357A variant was resistant to these reagents. The Q101A, Q356A, and C357A variants showed elevated ferroxidase activity in the absence of substrates, pointing to the presence of the corresponding side chains at the active site. These results confirm most aspects of the homology model and provide additional insight into the enzyme reactivity.  相似文献   

3.
15N isotope effects and solvent deuterium isotope effects have been measured for the hydrolytic deamination of cytidine catalyzed by Escherichia coli cytidine deaminase and for the uncatalyzed reaction proceeding spontaneously in neutral solution at elevated temperatures. The primary (15)(V/K) arising from the exocyclic amino group for wild-type cytidine deaminase acting on its natural substrate, cytidine, is 1.0109 (in H(2)O, pH 7.3), 1.0123 (in H(2)O, pH 4.2), and 1.0086 (in D(2)O, pD 7.3). Increasing solvent D(2)O content has no substantial effect on k(cat) but enhances k(cat)/K(m), with a proton inventory showing that the fractionation factors of at least two protons increase markedly during the reaction. Mutant cytidine deaminases with reduced catalytic activity show more pronounced (15)N isotope effects of 1.0124 (Glu91Ala), 1.0134 (His102Ala), and 1.0158 (His102Asn) at pH 7.3 in H(2)O, as expected for processes in which the chemical transformation of the substrate becomes more rate determining. The isotope effect of mutant His102Asn is 1.033 after correcting for protonation of the -NH(2) group, and represents the intrinsic isotope effect on C-N bond cleavage. This result allows an estimation of the forward commitment of the reaction with the wild-type enzyme. The observed (15)N kinetic isotope effect of the pyrimidine N-3, for wild-type cytidine deaminase acting on cytidine, is 0.9879, which is consistent with protonation and rehybidization of N-3 with hydroxide ion attack on the adjacent carbon to create a tetrahedral intermediate. These results show that enzymatic deamination of cytidine proceeds stepwise through a tetrahedral intermediate with ammonia elimination as the major rate-determining step. The primary (15)N isotope effects observed for the uncatalyzed reaction at pH 7 (1.0021) and pH 12.5 (1.0034) were found to be insensitive to changing temperatures between 100 and 185 degrees C. These results show that the uncatalyzed and the enzymatic deaminations of cytidine proceed by similar mechanisms, although the commitment to C-N bond breaking is greater for the spontaneous reaction.  相似文献   

4.
A detailed kinetic analysis of the catalytic trimer of aspartate transcarbamoylase containing the active site substitution H134A was performed to investigate the role of His 134 in the catalytic mechanism. Replacement of histidine by alanine resulted in decreases in the affinities for the two substrates, carbamoyl phosphate and aspartate, and the inhibitor succinate, by factors of 50, 10, and 6, respectively, and yielded a maximum velocity that was 5% that of the wild-type enzyme. However, the pK values determined from the pH dependence of the kinetic parameters, log V and log (V/K) for aspartate, the pK(i) for succinate, and the pK(ia) for carbamoyl phosphate, were similar for both the mutant and the wild-type enzymes, indicating that the protonated form of His 134 does not participate in binding and catalysis between pH 6.2 and 9.2. 13C and 15N isotope effects were studied to determine which steps in the catalytic mechanism were altered by the amino acid substitutions. The 13(V/K) for carbamoyl phosphate exhibited by the catalytic trimer containing alanine at position 134 revealed an isotope effect of 4.1%, probably equal to the intrinsic value and, together with quantitative analysis of the 15N isotope effects, showed that formation of the tetrahedral intermediate is rate-determining for the mutant enzyme. Thus, His 134 plays a role in the chemistry of the reaction in addition to substrate binding. The initial velocity pattern for the reaction catalyzed by the H134A mutant intersected to the left of the vertical axis, negating an equilibrium ordered kinetic mechanism.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
Brosius JL  Colman RF 《Biochemistry》2000,39(44):13336-13343
Adenylosuccinate lyase of Bacillus subtilis is a tetrameric enzyme which catalyzes the cleavage of adenylosuccinate to AMP and fumarate. We have mutated His(89), one of three conserved histidines, to Gln, Ala, Glu, and Arg. The enzymes were expressed in Escherichia coli and purified to homogeneity. As compared to a specific activity of 1. 56 micromol of adenylosuccinate converted/min/mg protein for wild-type enzyme, the mutant enzymes exhibit specific activities of 0.0225, 0.0036, 0.0036, and 0.0009 for H89Q, H89A, H89E, and H89R, respectively. Circular dichroism and FPLC gel filtration reveal that mutant enzymes have a similar conformation and oligomeric state to that of wild-type enzyme. In H89Q, the K(M) for adenylosuccinate increases slightly to 2.5-fold that of wild-type, the K(M) for fumarate is elevated 3.3-fold, and the K(M) for AMP is 13 times higher than that observed in wild-type enzyme. The catalytic efficiency of the H89Q enzyme is compromised, with k(cat)/K(M) reduced 174-fold in the direction of AMP formation. These data suggest that His(89) plays a role in both the binding of the AMP portion of the substrate and in correctly orienting the substrate for catalysis. Incubation of H89Q with inactive H141Q enzyme [Lee, T. T., Worby, C., Bao, Z.-Q., Dixon, J. E., and Colman, R. F. (1999) Biochemistry 38, 22-32] leads to a 30-fold increase in activity. This intersubunit complementation indicates that His(89) and His(141) from different subunits participate in the active site and that both are required for catalysis.  相似文献   

6.
Huang YC  Colman RF 《Biochemistry》2002,41(17):5637-5643
Sequence alignment predicts that His(309) of pig heart NADP-dependent isocitrate dehydrogenase is equivalent to His(339) of the Escherichia coli enzyme, which interacts with the coenzyme in the crystal structure [Hurley et al. (1991) Biochemistry 30, 8671-8688], and porcine His(315) and His(319) are close to that site. The mutant porcine enzymes H309Q, H309F, H315Q, and H319Q were prepared by site-directed mutagenesis, expressed in E. coli, and purified. The H319Q mutant has K(m) values for NADP, isocitrate, and Mn(2+) similar to those of wild-type enzyme, and V(max) = 20.1, as compared to 37.8 micromol of NADPH min(-1) (mg of protein)(-1) for wild type. Thus, His(319) is not involved in coenzyme binding and has a minimal effect on catalysis. In contrast, H315Q exhibits a K(m) for NADP 40 times that of wild type and V(max) = 16.2 units/mg of protein, with K(m) values for isocitrate and Mn(2+) similar to those of wild type. These results implicate His(315) in the region of the NADP site. Replacement of His(309) by Q or F yields enzyme with no detectable activity. The His(309) mutants bind NADPH poorly, under conditions in which wild type and H319Q bind 1.0 mol of NADPH/mol of subunit, indicating that His(309) is important for the binding of coenzyme. The His(309) mutants bind isocitrate stoichiometrically, as do wild-type and the other mutant enzymes. However, as distinguished from the wild-type enzyme, the His(309) mutants are not oxidatively cleaved by metal isocitrate, implying that the metal ion is not bound normally. Since circular dichroism spectra are similar for wild type, H315Q, and H319Q, these amino acid substitutions do not cause major conformational changes. In contrast, replacement of His(309) results in detectable change in the enzyme's CD spectrum and therefore in its secondary structure. We propose that His(309) plays a significant role in the binding of coenzyme, contributes to the proper coordination of divalent metal ion in the presence of isocitrate, and maintains the normal conformation of the enzyme.  相似文献   

7.
1-Deoxy-d-xylulose 5-phosphate (DXP) reductoisomerase, which simultaneously catalyzes the intramolecular rearrangement and reduction of DXP to form 2-C-methyl-d-erythritol 4-phosphate, constitutes a key enzyme of an alternative mevalonate-independent pathway for isopentenyl diphosphate biosynthesis. The dxr gene encoding this enzyme from Escherichia coli was overexpressed as a histidine-tagged protein and characterized in detail. DNA sequencing analysis of the dxr genes from 10 E. coli dxr-deficient mutants revealed base substitution mutations at four points: two nonsense mutations and two amino acid substitutions (Gly(14) to Asp(14) and Glu(231) to Lys(231)). Diethyl pyrocarbonate treatment inactivated DXP reductoisomerase, and subsequent hydroxylamine treatment restored the activity of the diethyl pyrocarbonate-treated enzyme. To characterize these defects, we overexpressed the mutant enzymes G14D, E231K, H153Q, H209Q, and H257Q. All of these mutant enzymes except for G14D were obtained as soluble proteins. Although the purified enzyme E231K had wild-type K(m) values for DXP and NADPH, the mutant enzyme had less than a 0.24% wild-type k(cat) value. K(m) values of H153Q, H209Q, and H257Q for DXP increased to 3.5-, 7.6-, and 19-fold the wild-type value, respectively. These results indicate that Glu(231) of E. coli DXP reductoisomerase plays an important role(s) in the conversion of DXP to 2-C-methyl-d-erythritol 4-phosphate, and that His(153), His(209), and His(257), in part, associate with DXP binding in the enzyme molecule.  相似文献   

8.
Flavocytochrome b(2) catalyzes the oxidation of lactate to pyruvate. Primary deuterium and solvent kinetic isotope effects have been used to determine the relative timing of cleavage of the lactate O-H and C-H bonds by the wild-type enzyme, a mutant protein lacking the heme domain, and the D282N enzyme. The (D)V(max) and (D)(V/K(lactate)) values are both 3.0 with the wild-type enzyme at pH 7.5 and 25 degrees C, increasing to about 3.6 with the flavin domain and increasing further to about 4.5 with the D282N enzyme. Under these conditions, the (D20)V(max) values are 1.38, 1.18, and 0.98 for the wild-type enzyme, the flavin domain, and the D282N enzyme, respectively; the (D20)(V/K(lactate)) values are 0.9, 0.44, and 1.0, respectively. The (D)k(red) value is 5.4 for the wild-type enzyme and 3.5 for the flavin domain, whereas the solvent isotope effect on this kinetic parameter is 1.0 for both enzymes. The V(max) values for the wild-type enzyme and the flavin domain are 32 and 15% limited by viscosity, respectively. In contrast, the V/K(lactate) value for the flavin domain increases about 2-fold at moderate concentrations of glycerol. The data are consistent with a minimal chemical mechanism in which the lactate hydroxyl proton is not in flight in the transition state for C-H bond cleavage and there is an internal equilibrium involving the lactate-bound enzyme prior to C-H bond cleavage which is sensitive to solution conditions. Removal of the hydroxyl proton may occur in this pre-equilibrium.  相似文献   

9.
The flavoprotein oxidase Fms1 from Saccharomyces cerevisiae catalyzes the oxidation of spermine and N(1)-acetylspermine to spermidine and 3-aminopropanal or N-acetyl-3-aminopropanal. Within the active site of Fms1, His67 is positioned to form hydrogen bonds with the polyamine substrate. This residue is also conserved in other polyamine oxidases. The catalytic properties of H67Q, H67N, and H67A Fms1 have been characterized to evaluate the role of this residue in catalysis. With both spermine and N(1)-acetylspermine as the amine substrate, the value of the first-order rate constant for flavin reduction decreases 2-3 orders of magnitude, with the H67Q mutation having the smallest effect and H67N the largest. The k(cat)/K(O2) value changes very little upon mutation with N(1)-acetylspermine as the amine substrate and decreases only an order of magnitude with spermine. The k(cat)/K(M)-pH profiles with N(1)-acetylspermine are bell-shaped for all the mutants; the similarity to the profile of the wild-type enzyme rules out His67 as being responsible for either of the pK(a) values. The pH profiles for the rate constant for flavin reduction for all the mutant enzymes similarly show the same pK(a) as wild-type Fms1, about ~7.4; this pK(a) is assigned to the substrate N4. The k(cat)/K(O2)-pH profiles for wild-type Fms1 and the H67A enzyme both show a pK(a) of about ~6.9; this suggests His67 is not responsible for this pH behavior. With the H67Q, H67N, and H67A enzymes the k(cat) value decreases when a single residue is protonated, as is the case with the wild-type enzyme. The structure of H67Q Fms1 has been determined at a resolution of 2.4 ?. The structure shows that the mutation disrupts a hydrogen bond network in the active site, suggesting that His67 is important both for direct interactions with the substrate and to maintain the overall active site structure.  相似文献   

10.
Adenylosuccinate lyase (ASL) of Bacillus subtilis contains three conserved histidines, His(68), His(89), and His(141), identified by affinity labeling and site-directed mutagenesis as critical to the intersubunit catalytic site. The pH-V(max) profile for wild-type ASL is bell-shaped (pK (1) = 6.74 and pK (2) = 8.28). Only the alkaline side changes with temperature, characteristic of histidine pKs. To identify determinants of pK (2) in the enzyme-substrate complex, we replaced residues at two positions close to His(68) (but not to His(89) or His(141)) in the structure. Compared with the specific activity of 1.75 mumol adenylosuccinate reacting/min/mg of wild-type enzyme at pH 7.0, mutant enzymes D69E, D69N, R310Q, and R310K exhibit specific activities of 0.40, 0.04, 0.00083, and 0.10, respectively. While D69E has a K (m) for adenylosuccinate similar to that of wild-type ASL, D69N and R310K exhibit modest increases in K (m), and R310Q has an 11-fold increase in K (m). The mutant enzymes show no significant change in molecular weight or secondary structure. The major change is in the pH-V(max) profile: pK (2) is 8.48 for the D69E mutant and is decreased to 7.83 in D69N, suggesting a proximal negative charge is needed to maintain the high pK of 8.28 observed for wild-type enzyme and attributed to His(68). Similarly, R310Q exhibits a decrease in its pK (2) (7.33), whereas R310K shows little change in pK (2) (8.24). These results suggest that Asp(69) interacts with His(68), that Arg(310) interacts with and orients the beta-carboxylate of Asp(69), and that His(68) must be protonated for ASL to be active.  相似文献   

11.
Human glycine N-acyltransferase (human GLYAT) detoxifies a wide range of endogenous and xenobiotic metabolites, including benzoate and salicylate. Significant inter-individual variation exists in glycine conjugation capacity. The molecular basis for this variability is not known. To investigate the influence of single nucleotide polymorphisms (SNPs) in the GLYAT coding sequence on enzyme activity, we expressed and characterised a recombinant human GLYAT. Site-directed mutagenesis was used to generate six non-synonymous SNP variants of the enzyme (K16N; S17T; R131H; N156S; F168L; R199C). The variants were expressed, purified, and enzymatically characterised. The enzyme activities of the K16N, S17T and R131H variants were similar to that of the wild-type, whereas the N156S variant was more active, the F168L variant less active, and the R199C variant was inactive. We also generated an E227Q mutant, which lacks the catalytic residue proposed by Badenhorst et al. (2012). This mutant was inactive compared to the wild-type recombinant human GLYAT. A molecular model of human GLYAT containing coenzyme A (CoA) was generated which revealed that the inactivity of the R199C variant could be due to the substitution of the highly conserved Arg199 and destabilisation of an α-loop-α motif which is important for substrate binding in the GNAT superfamily. The finding that SNP variations in the human GLYAT gene influence the kinetic properties of the enzyme may explain some of the inter-individual variation in glycine conjugation capacity, which is relevant to the metabolism of xenobiotics such as aspirin and the industrial solvent xylene, and to the treatment of some metabolic disorders.  相似文献   

12.
The catalytic contribution of His48 in the active site of porcine pancreatic phospholipase A2 was examined using site-directed mutagenesis. Replacement of His48 by lysine (H48K) gives rise to a protein having a distorted lipid binding pocket. Activity of this variant drops below the detection limit which is 10(7)-fold lower than that of the wild-type enzyme. On the other hand, the presence of glutamine (H48Q) or asparagine (H48N) at this position does not affect the structural integrity of the enzyme as can be derived from the preserved lipid binding properties of these variants. However, the substitutions H48Q and H48N strongly reduce the turnover number, i.e. by a factor of 10(5). Residual activity is totally lost after addition of a competitive inhibitor. We conclude that proper lipid binding on its own accelerates ester bond hydrolysis by a factor of 10(2). With the selected variants, we were also able to dissect the contribution of the hydrogen bond between Asp99 and His48 on conformational stability, being 5.2 kJ/mol. Another hydrogen bond with His48 is formed when the competitive inhibitor (R)-2-dodecanoylamino-hexanol-1-phosphoglycol interacts with the enzyme. Its contribution to binding of the inhibitor in the presence of an interface was found to be 5.7 kJ/mol.  相似文献   

13.
To engineer dehairing alkaline protease (DHAP) variants to improve cold activity and increase thermostability so these variants are suitable for the leather processing industry. Based on previous studies with bacterial alkaline proteases, double-site mutations (W106K/V149I and W106K/M124L) were introduced into the DHAP from Bacillus pumilus. Compared with the wild-type DHAP hydrolytic activity, the double-site variant W106K/V149I showed an increase in specific hydrolytic activity at 15 °C by 2.3-fold toward casein in terms of hydrolytic rate and 2.7-fold toward the synthetic peptide AAPF-pN by means of kcat/Km value. The thermostability of the variant (W106K/V149I) was improved with the half-life at 60 and 70 °C increased by 2.7- and 5.0-fold, respectively, when compared with the thermostability of the wild-type DHAP. Conclusively, an increase in the cold activity and thermostability of a bacterial alkaline protease was achieved by protein engineering.  相似文献   

14.
Saccharopine dehydrogenase (SDH) catalyzes the final reaction in the α-aminoadipate pathway, the conversion of l-saccharopine to l-lysine (Lys) and α-ketoglutarate (α-kg) using NAD? as an oxidant. The enzyme utilizes a general acid-base mechanism to conduct its reaction with a base proposed to accept a proton from the secondary amine of saccharopine in the oxidation step and a group proposed to activate water to hydrolyze the resulting imine. Crystal structures of an open apo form and a closed form of the enzyme with saccharopine and NADH bound have been determined at 2.0 and 2.2 ? resolution, respectively. In the ternary complex, a significant movement of domain I relative to domain II that closes the active site cleft between the two domains and brings H96 and K77 into the proximity of the substrate binding site is observed. The hydride transfer distance is 3.6 ?, and the side chains of H96 and K77 are properly positioned to act as acid-base catalysts. Preparation of the K77M and H96Q single-mutant and K77M/H96Q double-mutant enzymes provides data consistent with their role as the general acid-base catalysts in the SDH reaction. The side chain of K77 initially accepts a proton from the ε-amine of the substrate Lys and eventually donates it to the imino nitrogen as it is reduced to a secondary amine in the hydride transfer step, and H96 protonates the carbonyl oxygen as the carbinolamine is formed. The K77M, H976Q, and K77M/H96Q mutant enzymes give 145-, 28-, and 700-fold decreases in V/E(t) and >103-fold increases in V?/K(Lys)E(t) and V?/K(α-kg)E(t) (the double mutation gives >10?-fold decreases in the second-order rate constants). In addition, the K77M mutant enzyme exhibits a primary deuterium kinetic isotope effect of 2.0 and an inverse solvent deuterium isotope effect of 0.77 on V?/K(Lys). A value of 2.0 was also observed for (D)(V?/K(Lys))(D?O) when the primary deuterium kinetic isotope effect was repeated in D?O, consistent with a rate-limiting hydride transfer step. A viscosity effect of 0.8 was observed on V?/K(Lys), indicating the solvent deuterium isotope effect resulted from stabilization of an enzyme form prior to hydride transfer. A small normal solvent isotope effect is observed on V, which decreases slightly when repeated with NADD, consistent with a contribution from product release to rate limitation. In addition, V?/K(Lys)E(t) is pH-independent, which is consistent with the loss of an acid-base catalyst and perturbation of the pK(a) of the second catalytic group to a higher pH, likely a result of a change in the overall charge of the active site. The primary deuterium kinetic isotope effect for H96Q, measured in H?O or D?O, is within error equal to 1. A solvent deuterium isotope effect of 2.4 is observed with NADH or NADD as the dinucleotide substrate. Data suggest rate-limiting imine formation, consistent with the proposed role of H96 in protonating the leaving hydroxyl as the imine is formed. The pH-rate profile for V?/K(Lys)E(t) exhibits the pK(a) for K77, perturbed to a value of ~9, which must be unprotonated to accept a proton from the ε-amine of the substrate Lys so that it can act as a nucleophile. Overall, data are consistent with a role for K77 acting as the base that accepts a proton from the ε-amine of the substrate lysine prior to nucleophilic attack on the α-oxo group of α-ketoglutarate, and finally donating a proton to the imine nitrogen as it is reduced to give saccharopine. In addition, data indicate a role for H96 acting as a general acid-base catalyst in the formation of the imine between the ε-amine of lysine and the α-oxo group of α-ketoglutarate.  相似文献   

15.
Selective (15)N isotope labeling of the cytochrome bo(3) ubiquinol oxidase from Escherichia coli with auxotrophs was used to characterize the hyperfine couplings with the side-chain nitrogens from residues R71, H98, and Q101 and peptide nitrogens from residues R71 and H98 around the semiquinone (SQ) at the high-affinity Q(H) site. The two-dimensional ESEEM (HYSCORE) data have directly identified N(ε) of R71 as an H-bond donor carrying the largest amount of unpaired spin density. In addition, weaker hyperfine couplings with the side-chain nitrogens from all residues around the SQ were determined. These hyperfine couplings reflect a distribution of the unpaired spin density over the protein in the SQ state of the Q(H) site and the strength of interaction with different residues. The approach was extended to the virtually inactive D75H mutant, where the intermediate SQ is also stabilized. We found that N(ε) of a histidine residue, presumably H75, carries most of the unpaired spin density instead of N(ε) of R71, as in wild-type bo(3). However, the detailed characterization of the weakly coupled (15)N atoms from selective labeling of R71 and Q101 in D75H was precluded by overlap of the (15)N lines with the much stronger ~1.6 MHz line from the quadrupole triplet of the strongly coupled (14)N(ε) atom of H75. Therefore, a reverse labeling approach, in which the enzyme was uniformly labeled except for selected amino acid types, was applied to probe the contribution of R71 and Q101 to the (15)N signals. Such labeling has shown only weak coupling with all nitrogens of R71 and Q101. We utilize density functional theory-based calculations to model the available information about (1)H, (15)N, and (13)C hyperfine couplings for the Q(H) site and to describe the protein-substrate interactions in both enzymes. In particular, we identify the factors responsible for the asymmetric distribution of the unpaired spin density and ponder the significance of this asymmetry to the quinone's electron transfer function.  相似文献   

16.
lac permease of Escherichia coli was modified by site-directed mutagenesis in order to investigate the effects of polarity, distance, and orientation between the components of a putative H+ relay system (Arg302/His322/Glu325) postulated to be involved in lactose-coupled H+ translocation. The importance of polarity between His322 and Glu325 was studied by interchanging the residues, and the modified permease--H322E/E325H--is inactive in all modes of translocation. The effect of distance and/or orientation between His322 and Glu325 was investigated by interchanging Glu325 with Val326, thereby moving the carboxylate one residue around putative helix X. The resulting permease molecule--E325V/V326E--is also completely inactive; control mutations, E325V [Carrasco, N., Püttner, I. B., Antes, L. M., Lee, J. A., Larigan, J. D., Lolkema, J. S., Roepe, P. D., & Kaback, H. R. (1989) Biochemistry (second paper of three in this issue)], and E325A/V326E, indicate that a Glu residue at position 326 inactivates the permease. The wild-type orientation between His and Glu was then restored by further mutation of E325V/V326E to introduce a His residue into position 323 or by interchanging Met323 with His322. The resulting permease molecules--M323H/E325V/V326E and H322M/M323H/E325V/V326E--contain the wild-type His/Glu orientation, but the His/Glu ion pair is rotated about the helical axis by 100 degrees relative to Arg302 in putative helix IX. Both mutants are inactive with respect to all modes of translocation. The results provide strong support for the contention that the polarity between His322 and Glu325 and the geometric relationship between Arg302, His322, and Glu325 are critical for permease activity.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
We have investigated the roles of highly conserved glycine (G175, G185), negatively charged (E188, D165) and histidine residues (H233, H237) in rat steroid 5alpha-reductase (isozyme-1), on NADPH, testosterone (T) binding and enzyme activity. The mutations G175R and G175S result in a two- to threefold increase in K(m)(NADPH) and an approximately fourfold decrease in the V(max) with no change in K(m)(T). The mutation G185W resulted in a fivefold decrease in K(m)(NADPH) and an eightfold decrease in V(max), with no change in K(m)(T), whereas the mutations E188Q and D165N both resulted in inactive enzyme. Steady-state kinetic measurements showed that the mutation H233R resulted in an approximately 40-fold decrease in V(max), an approximately 20-fold increase in K(m)(T) and no alteration in K(m)(NADPH), whereas the mutation H237R resulted in virtually inactive enzyme. The results suggest that the conserved glycines are not essential for cofactor binding and activity, and that the negatively charged residues may contribute to enzyme stability, whereas the C-terminal histidines appear to be involved in substrate binding and catalytic activity.  相似文献   

18.
Hydrostatic pressure causes biphasic effects on the oxidation of alcohols by yeast alcohol dehydrogenase as expressed on the kinetic parameter V/K which measures substrate capture. Moderate pressure increases capture by activating hydride transfer, whose transition-state must therefore have a smaller volume than the free alcohol plus the capturing form of enzyme, with DeltaV(double dagger)=-30 mL mol(-1) for isopropanol. A comparison of these effects with those on the oxidation of deutero-isopropanol generates a monophasic decrease in the intrinsic isotope effect; therefore, the volume of activation for the transition-state of deuteride transfer must be even more negative, by 7.6 mL mol(-1). The pressure data extrapolate and factor the kinetic isotope effect into a semi-classical reactant-state component, with a null value of k(H)/k(D)=1, and a transition-state component of Q(H)/Q(D)=4, suggestive of hydrogen tunneling. Pressures above 1.5 kbar decrease capture by favoring a minor conformation of enzyme which binds nicotinamide adenine dinucleotide (NAD(+)) less tightly. This inactive conformation has a smaller volume than active E-NAD(+), with a difference of 74 mL mol(-1) and an equilibrium constant of 93 between them, at one atmosphere of pressure. These results are virtually identical to those obtained with benzyl alcohol and give credence to this method of analysis. Moreover, qualitatively similar results with greater pressure sensitivity but less precision are obtained using ethanol as a substrate, only with pressure driving the value of the isotope effect to a value less than (D)k=1.03 directly, without extrapolation. The ethanol data verify the most surprising finding of these studies, namely that the entire kinetic isotope effect arises from a transition-state phenomenon.  相似文献   

19.
alpha-Sarcin is a ribotoxin secreted by the mold Aspergillus giganteus that degrades the ribosomal RNA by acting as a cyclizing ribonuclease. Three residues potentially involved in the mechanism of catalysis--histidine-50, glutamic acid-96, and histidine-137--were changed to glutamine. Three different single mutation variants (H50Q, E96Q, H137Q) as well as a double variant (H50/137Q) and a triple variant (H50/137Q/E96Q) were prepared and isolated to homogeneity. These variants were spectroscopically (circular dichroism, fluorescence emission, and proton nuclear magnetic resonance) characterized. According to these results, the three-dimensional structure of these variants of alpha-sarcin was preserved; only very minor local changes were detected. All the variants were inactive when assayed against either intact ribosomes or poly(A). The effect of pH on the ribonucleolytic activity of alpha-sarcin was evaluated against the ApA dinucleotide. This assay revealed that only the H50Q variant still retained its ability to cleave a phosphodiester bond, but it did so to a lesser extent than did wild-type alpha-sarcin. The results obtained are interpreted in terms of His137 and Glu96 as essential residues for the catalytic activity of alpha-sarcin (His137 as the general acid and Glu96 as the general base) and His50 stabilizing the transition state of the reaction catalyzed by alpha-sarcin.  相似文献   

20.
This paper presents the complete amino acid sequence of the low molecular weight acid phosphatase from bovine liver. This isoenzyme of the acid phosphatase family is located in the cytosol, is not inhibited by L-(+)-tartrate and fluoride ions, but is inhibited by sulfhydryl reagents. The enzyme consists of 157 amino acid residues, has an acetylated NH2 terminus, and has arginine as the COOH-terminal residue. All 8 half-cystine residues are in the free thiol form. The molecular weight calculated from the sequence is 17,953. The sequence was determined by characterizing the peptides purified by reverse-phase high performance liquid chromatography from tryptic, thermolytic, peptic, Staphylococcus aureus protease, and chymotryptic digests of the carboxymethylated protein. No sequence homologies were found with the two known acylphosphatase isoenzymes or the metalloproteins porcine uteroferrin and purple acid phosphatase from bovine spleen (both of which have acid phosphatase activity). Two half-cystines at or near the active site were identified through the reaction of the enzyme with [14C] iodoacetate in the presence or in the absence of a competitive inhibitor (i.e. inorganic phosphate). Ac-A E Q V T K S V L F V C L G N I C R S P I A E A V F R K L V T D Q N I S D N W V I D S G A V S D W N V G R S P N P R A V S C L R N H G I N T A H K A R Q V T K E D F V T F D Y I L C M D E S N L R D L N R K S N Q V K N C R A K I E L L G S Y D P Q K Q L I I E D P Y Y G N D A D F E T V Y Q Q C V R C C R A F L E K V R-OH.  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