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 187 毫秒
1.
Histidine decarboxylase (HisDCase) from Lactobacillus buchneri was purified to homogeneity. Its subunit structure, (alpha beta)6, and enzymatic properties resemble closely those of the immunologically cross-reactive HisDCase of Lactobacillus 30a (Recsei, P. A., and Snell, E. E. (1984) Annu. Rev. Biochem. 53, 357-387). The complete amino acid sequences of the beta chains of the HisDCase from L. buchneri (81 residues) and Clostridium perfringens (86 residues) were then determined to be a and b, respectively. (a) SEFDKKLNTLGVDRISVSPYKKWSRGYMEPGNIGNGYVSGLKVDAG VVDKTDDMVLDGIGSYDRAETKNAYIGQINMTTAS. (b) TLSEGIHKNIKNIKVRAP KIDKTAISPYDRYCDGYGMPGAYGDGYVSVLKVSVGTVKK TDDILLDGIVSYDRAEINDAYVGQINMLTAS. SEFDKKLNTLGVDRISVSPYKKWSRGYMEPGNIGNGYVSGLKVDAGVV. Although these sequences differ substantially near the NH2-terminal ends, there is striking homology near the COOH termini and also near the NH2 terminus of the two alpha chains (pyruvoyl-Phe-X-Gly-Val-, where X is Ser or Cys). If the four known pyruvoyl-dependent HisDCases arise from inactive proenzymes by the mechanism previously demonstrated for the HisDCase of Lactobacillus 30a (Recsei, P. A., Huynh, Q. K. and Snell, E. E. (1983) Proc. Natl. Acad. Sci. U.S.A. 80, 973-977), then each of these proenzymes has the sequence -Thr-Ala-Ser-Ser-Phe- at the activation site (where -Ser- becomes the COOH terminus of the beta chain and -Ser- becomes the pyruvoyl group blocking the NH2 terminus of the alpha chain), and the sequences around this activation site are highly conserved in all four enzymes. These facts support the assumptions that the four enzymes have evolved from a common ancestral protein, are formed from inactive pyruvate-free proenzymes by similar mechanisms, and have similar catalytic mechanisms.  相似文献   

2.
The two cysteinyl residues present in histidine decarboxylase from Lactobacillus 30a differ greatly in reactivity. One (class 1) reacts readily in the native state with dithiobis-(2-nitrobenzoate) with complete loss of enzyme activity; the other (class 2) reacts only after denaturation of the enzyme (Lane, R. S., and Snell, E. E. (1976) Biochemistry 15, 4175-4179). These differences in reactivity permitted use of covalent (disulfide) chromatography to isolate separate peptides that contain these two residues. Sequence analysis showed that the class 1 cysteinyl residue is at position 147 in a hydrophilic portion of the alpha chain (Huynh, Q. K., Recsei, P. A., Vaaler, G. L., and Snell, E. E. (1984) J. Biol. Chem. 259, 2833-2839), while the class 2 cysteinyl residue is present at position 71, adjacent to a hydrophobic portion of the same chain. Cysteinyl peptides identical with or homologous to the class 2 cysteinyl peptide of the Lactobacillus 30a enzyme were isolated from the alpha subunits of histidine decarboxylases from Lactobacillus buchneri and Clostridium perfringens, respectively. The L. buchneri enzyme also contained a peptide homologous to the class 1 cysteinyl peptide from Lactobacillus 30a. However, no corresponding peptide was present in the enzyme from C. perfringens, in which the second cysteinyl residue of the alpha chain occupies position 3, very near the essential pyruvoyl residue. This enzyme, unlike those from Lactobacillus 30a or L. buchneri, also contains one cysteinyl residue in its beta chain. Although Cys 147 is an active site residue in histidine decarboxylase from Lactobacillus 30a, the absence of a corresponding residue in the C. perfringens enzyme confirms previous indications (Recsei, P. A., and Snell, E. E. (1982) J. Biol. Chem. 257, 7196-7202) that this SH group is not essential for decarboxylase action.  相似文献   

3.
The complete amino acid sequence of the alpha chain of histidine decarboxylase of Lactobacillus 30a has been established by isolation and analysis of the eight methionine-containing tryptic peptides of this chain. These peptides provide the overlaps required to order all nine peptides derived by complete cyanogen bromide cleavage of the alpha chain (Huynh, Q.K., Vaaler, G.L., Recsei, P.A., and Snell, E.E. (1984) J. Biol. Chem. 259, 2826-2832). Ordering of six of the latter peptides was confirmed by isolation and analysis of four peptides derived by incomplete cyanogen bromide cleavage. The alpha chain is composed of 226 residues and has a molecular weight of 24,892 calculated from the sequence. These results and the previously determined sequence of the beta chain (Vaaler, G.L., Recsei, P.A., Fox, J.L., and Snell, E.E. (1982) J. Biol. Chem. 257, 12770-12774) establish the complete amino acid sequence of the enzyme and of the pi chain of prohistidine decarboxylase. The latter is composed of 307 amino acids and has a calculated molecular weight of 33,731. Four segments of the pi chain sequence are repeated. The bond between Ser-81 and Ser-82 that is cleaved during proenzyme activation is in an uncharged portion of the sequence that is rich in serine and threonine residues and is predicted to be part of a beta sheet structure.  相似文献   

4.
Conversion of the pi subunit of prohistidine decarboxylase to the alpha beta subunits of the active enzyme proceeds by a nonhydrolytic, monovalent cation-dependent, serinolysis reaction in which the hydroxyl oxygen of serine 82 of the pi chain is incorporated into the carboxyl group at the COOH terminus (serine 81) of the beta chain. Serine-82 becomes the pyruvate residue at the NH2 terminus of the alpha chain (Recsei, P.A., Huynh, Q. K., and Snell, E.E. (1983) Proc. Natl. Acad. Sci. U.S.A. 80, 973-977). The unusual reactivity of this particular -Ser-Ser- bond is demonstrated by its sensitivity to 1 M hydroxylamine, which cleaves the native proenzyme under mild conditions (pH 8.0, 37 degrees C) to yield a modified beta chain with serine hydroxamate at the COOH terminus (Ser-81) and a modified alpha chain containing serine (Ser-82 of the proenzyme) rather than pyruvate at the NH2 terminus. Neither an -Asn-Gly- bond nor other -Ser-Ser- bonds in the proenzyme were cleaved under these conditions. The reaction also did not occur with the denatured enzyme or with model peptides, indicating that the enhanced reactivity is a result of the particular conformation at this position in the native protein. The reaction with the native proenzyme proceeded optimally at pH 7.5-8.0 with a half-time (30 min) substantially less than that (3.5-4.5 h) required for the activation reaction and was not increased in rate by addition of K+. Correspondingly, preincubation of the proenzyme at pH 8.0 in the absence of both hydroxylamine and K+ modestly increased the rate of activation when K+ was subsequently added. Although these findings do not exclude other mechanisms, they are all consistent with and most easily explained by rearrangement of the pi chain to form an internal ester intermediate prior to the beta-elimination that occurs during activation to yield the alpha and beta chains of the mature enzyme.  相似文献   

5.
P D van Poelje  A V Kamath  E E Snell 《Biochemistry》1990,29(45):10413-10418
To clarify the mechanism of biogenesis and catalysis by the pyruvoyl-dependent histidine decarboxylase (HisDCase) from Clostridium perfringens, 12 mutant genes encoding amino acid substitutions at the active site of this enzyme were constructed and expressed in Escherichia coli. The resulting mutant proteins were purified to homogeneity, characterized, and subjected to kinetic analysis. The results (a) exclude all polar amino acid residues in the active site except Glu-214 as donor of the proton that replaces the carboxyl group of histidine during decarboxylation and, since E214I and E214H are nearly inactive, indicate that Glu-214 is the essential proton donor; (b) demonstrate the importance to substrate binding of hydrophobic interactions between Phe-98, Ile-74, and the imidazole ring of histidine, and of hydrogen bonding between Asp-78 and N2 of the substrate; and (c) demonstrate a significant unidentified role for Glu-81 in the maintenance of the active-site structure. The proposed roles of these amino acid residues are consistent with those assigned on the basis of crystallographic evidence to the corresponding residues at the active site of the related HisDCase from Lactobacillus 30a [Gallagher, T., Snell, E. E., & Hackert, M. L. (1989) J. Biol. Chem. 264, 12737-12743]. Of the residues altered, only Ser-97 was essential for the autocatalytic serinolysis reaction by which this HisDCase, (alpha beta)6, is derived from its inactive, pyruvate-free precursor, proHisDCase, pi 6.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
S-Adenosylmethionine decarboxylase is one of a small group of enzymes that use a pyruvoyl residue as a cofactor. Histidine decarboxylase from Lactobacillus 30a, the best studied pyruvoyl-containing enzyme, has an (alpha beta)6 subunit structure with the pyruvoyl moiety linked through an amide bond to the NH2-terminal of the larger alpha subunit (Recsei, P. A., Huynh, Q. K., and Snell, E. E. (1983) Proc. Natl. Acad. Sci. U. S. A. 80, 973-977). To examine potential structural analogies between the two enzymes, we have isolated and partially characterized S-adenosylmethionine decarboxylase. The purified enzyme comprises equimolar amounts of two subunits of Mr = 14,000 and 19,000 (by sodium dodecyl sulfate-polyacrylamide gel electrophoresis) and has a native molecular weight of 136,000 (by gel filtration). Approximately 4 mol of [methyl-3H] adenosylmethionine are incorporated per mol of enzyme (Mr = 136,000) when the enzyme is inactivated with this substrate and NaCNBH3. These data suggest an (alpha beta)4 structure with 1 pyruvoyl residue for each alpha beta pair. The two subunits have been separated by reversed-phase high performance liquid chromatography after reduction and carboxymethylation. The smaller subunit (beta) has a free amino terminus. The amino terminus of the larger subunit (alpha) appears to be blocked by a pyruvoyl group; this subunit can be sequenced only after this group is converted to an alanyl residue by reduction with sodium cyanoborohydride in the presence of ammonium acetate. This work suggests that S-adenosylmethionine decarboxylase is structurally much more similar to histidine decarboxylase than previously thought.  相似文献   

7.
The hdc genes encoding the inducible pyridoxal-P-dependent histidine decarboxylase (HisDCase) of Klebsiella planticola and Enterobacter aerogenes were isolated, sequenced, and expressed in Escherichia coli under control of the lac promoter, and the overproduced enzymes were purified to homogeneity from the recombinant host. Formation of inclusion bodies during synthesis of the E. aerogenes enzyme was avoided by cooling the culture and inducing at 25 degrees C. The cloned enzymes were produced in amounts three to four times those present in the fully induced native hosts and were identical in properties to those isolated earlier (Guirard, B. M., and Snell, E. E. (1987) J. Bacteriol. 169, 3963-3968). The two enzymes showed 85% sequence identity and also showed 80% sequence identity with the previously sequenced (Vaaler, G. L., Brasch, M. A., and Snell, E. E. (1986) J. Biol. Chem. 261, 11010-11014) HisDCase of Morganella morganii. Nevertheless, antibodies to the M. morganii HisDCase do not cross-react with these enzymes suggesting that the regions of amino acid variations are located on the outer surface of the proteins. All three HisDCases are the same length (377 amino acid residues); encoded N-terminal methionine was completely removed in each case. These closely related pyridoxal-P enzymes show no sequence homology with the pyruvoyl-dependent HisDCases of Gram-positive bacteria.  相似文献   

8.
T A Alston  R H Abeles 《Biochemistry》1987,26(13):4082-4085
L-Histidine methyl ester inactivates histidine decarboxylase in a time-dependent manner. The possibility was considered that an irreversible reaction between enzyme and inhibitor occurs [Recsei, P. A., & Snell, E. E. (1970) Biochemistry 9, 1492-1497]. We have confirmed time-dependent inactivation by histidine methyl ester and have investigated the structure of the enzyme-inhibitor complex. Upon exposure to either 8 M guanidinium chloride or 6% trichloroacetic acid, unchanged histidine methyl ester is recovered. Formation of the complex involves Schiff base formation, most likely with the active site pyruvyl residue [Huynh, Q. K., & Snell, E. E. (1986) J. Biol. Chem. 261, 4389-4394], but does not involve additional irreversible covalent interaction between inhibitor and enzyme. Complex formation is a two-step process involving rapidly reversible formation of a loose complex and essentially irreversible formation of a tight complex. For the formation of the tight complex, Ki = 80 nM and koff = 2.5 X 10(-4) min-1. Time-dependent inhibition was also observed with L-histidine ethyl ester, L-histidinamide, and DL-3-amino-4-(4-imidazolyl)-2-butanone. No inactivation was observed with glycine methyl ester or histamine. We propose that in the catalytic reaction the carboxyl group of the substrate is in a hydrophobic region. The unfavorable interaction between the carboxylate group and the hydrophobic region facilitates decarboxylation [Crosby, J., Stone, R., & Liehard, G. E. (1970) J. Am. Chem. Soc. 92, 2891-2900]. With histidine methyl ester this unfavorable interaction is no longer present; hence, there is tight binding.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
The alpha chain of histidine decarboxylase contains eight internal methionine residues. After reductive amination to convert the NH2-terminal pyruvoyl residue to an alanyl residue and cyanogen bromide treatment, 13 pure peptides were isolated. Four of these are incomplete cleavage products. The sequence of each of the remaining nine peptides was established by automated and manual degradation of the intact peptides and of smaller peptides obtained from tryptic, chymotryptic, and staphylococcal protease digests of the cyanogen bromide peptides. These results, together with the data on overlapping peptides reported in the accompanying paper (Huynh, Q. K., Recsei, P. A., Vaaler, G. L., and Snell, E. E. (1984) J. Biol. Chem. 259, 2833-2839), establish the complete amino acid sequence of the alpha chain.  相似文献   

10.
A 2.4 kilobase cDNA for rat mitochondrial aspartate aminotransferase (E.C. 2.6.1.1.) was isolated and sequenced. The predicted presequence is 93% homologous to the presequences of the enzyme from pig and mouse. The predicted amino acid sequence of the mature enzyme differs from that determined directly by amino acid sequencing (Huynh, Q.K., Sakakibara, R., Watanabe, T., and Wada, H. (1981) J. Biochem. (Tokyo) 90, 863-875) at 13 amino acids residues. The most important difference is at position 140 where the cDNA encodes a tryptophanyl residue rather than the previously reported glycine. This critical residue is now seen to be conserved in all aspartate aminotransferases. The coding region of this cDNA was inserted into the plasmid cloning vector pKK233-2 and used to stably express an unfused precursor in Escherichia coli JM105.  相似文献   

11.
M Tagaya  K Yamano  T Fukui 《Biochemistry》1989,28(11):4670-4675
Pyridoxal kinase from pig liver has been purified 10,000-fold to apparent homogeneity. The enzyme is a dimer of subunits of Mr 32,000. The enzyme is strongly inhibited by the product pyridoxal 5'-phosphate. Liver pyridoxamine phosphate oxidase, another enzyme involved in the biosynthesis of pyridoxal 5'-phosphate, is also strongly inhibited by this compound [Wada, H., & Snell, E. E. (1961) J. Biol. Chem. 236, 2089-2095]. Thus, the biosynthesis of pyridoxal 5'-phosphate in the liver might be regulated by the product inhibition of both pyridoxamine phosphate oxidase and pyridoxal kinase. Kinetic studies revealed that the catalytic reaction of liver pyridoxal kinase follows an ordered mechanism in which pyridoxal and ATP bind to the enzyme and ADP and pyridoxal 5'-phosphate are released from the enzyme, in this order. Adenosine tetraphosphopyridoxal was found to be a slow-binding inhibitor of pyridoxal kinase. Pre-steady-state kinetics of the inhibition revealed that the inhibitor and the enzyme form an initial weak complex prior to the formation of a tighter and slowly reversing complex. The overall inhibition constant was 2.4 microM. ATP markedly protects the enzyme against time-dependent inhibition by the inhibitor, whereas another substrate pyridoxal affords no protection. By contrast, adenosine triphosphopyridoxal is not a slow-binding inhibitor of this enzyme.  相似文献   

12.
Chemical modification studies with pyridoxal 5'-phosphate have indicated that lysine(s) appear to be at or near the active site of Escherichia coli glutamine synthetase (Colanduoni, J., and Villafranca, J. J. (1985) J. Biol. Chem. 260, 15042-15050; Whitley, E. J., Jr., and Ginsburg, A. (1978) J. Biol. Chem. 253, 7017-7025). Enzyme samples were prepared that contained approximately 1, approximately 2, and approximately 3 pyridoxamine 5'-phosphate residues/50,000-Da monomer; the activity of each sample was 100, 25, and 14% of the activity of unmodified enzyme, respectively. Cyanogen bromide cleavage of each enzyme sample was performed, the peptides were separated by high performance liquid chromatography, and the peptides containing pyridoxamine 5'-phosphate were identified by their absorbance at 320 nm. These isolated peptides were analyzed for amino acid composition and sequenced. The N terminus of the protein (a serine residue) was modified by pyridoxal 5'-phosphate at a stoichiometry of approximately 1/50,000 Da and this modified enzyme had full catalytic activity. Beyond a stoichiometry of approximately 1, lysines 383 and 352 reacted with pyridoxal 5'-phosphate and each modification results in a partial loss of activity. When various combinations of substrates and substrate analogs (ADP/Pi or L-methionine-SR-sulfoximine phosphate/ADP) were used to protect the enzyme from modification, Lys-352 was protected from modification indicating that this residue is at the active site. Under all experimental conditions employed, Lys-47, which reacts with the ATP analog 5'-p-fluorosulfonylbenzoyl-adenosine does not react with pyridoxal 5'-phosphate.  相似文献   

13.
The nucleotide sequence of a 1.3-kilobase NaeI fragment from Morganella morganii AM-15 that contains the gene for histidine decarboxylase has been determined. The gene was initially identified among total chromosomal digests using a mixed sequence oligonucleotide probe corresponding to amino acids 11-16 of histidine decarboxylase and then cloned on a 5.5-kilobase PstI fragment. The structural gene contains 1131 nucleotides and encodes 377 amino acids with the sequence: (sequence: in text). The independently determined NH2-terminal sequence of this enzyme (Tanase, S., Guirard, B. M., and Snell, E. E. (1985) J. Biol. Chem. 260, 6738-6746) and the amino acid sequences of two tryptic peptides reported in the accompanying paper (Hayashi, H., Tanase, S., and Snell, E. E. (1986) J. Biol. Chem. 261, 11003-11009) are localized in the sequence presented here; the lysine that binds pyridoxal phosphate is situated at residue 232, whereas the serine that binds the adduct formed between pyridoxal phosphate and the inhibitor alpha-fluoromethylhistidine is positioned at residue 322.  相似文献   

14.
Serine hydroxymethyltransferase (SHMT), a pyridoxal-5'-phosphate (PLP)-dependent enzyme catalyzes the tetrahydrofolate (H(4)-folate)-dependent retro-aldol cleavage of serine to form 5,10-methylene H(4)-folate and glycine. The structure-function relationship of SHMT was studied in our laboratory initially by mutation of residues that are conserved in all SHMTs and later by structure-based mutagenesis of residues located in the active site. The analysis of mutants showed that K71, Y72, R80, D89, W110, S202, C203, H304, H306 and H356 residues are involved in maintenance of the oligomeric structure. The mutation of D227, a residue involved in charge relay system, led to the formation of inactive dimers, indicating that this residue has a role in maintaining the tetrameric structure and catalysis. E74, a residue appropriately positioned in the structure of the enzyme to carry out proton abstraction, was shown by characterization of E74Q and E74K mutants to be involved in conversion of the enzyme from an 'open' to 'closed' conformation rather than proton abstraction from the hydroxyl group of serine. K256, the residue involved in the formation of Schiffs base with PLP, also plays a crucial role in the maintenance of the tetrameric structure. Mutation of R262 residue established the importance of distal interactions in facilitating catalysis and Y82 is not involved in the formaldehyde transfer via the postulated hemiacetal intermediate but plays a role in stabilizing the quinonoid intermediate. The mutational analysis of scSHMT along with the structure of recombinant Bacillus stearothermophilus SHMT and its substrate(s) complexes was used to provide evidence for a direct transfer mechanism rather than retro-aldol cleavage for the reaction catalyzed by SHMT.  相似文献   

15.
Phosphatidylserine decarboxylase from Escherichia coli uses a pyruvate group as the enzyme cofactor (Satre, M., and Kennedy, E. P. (1978) J. Biol. Chem. 253, 479-483). Comparison of the DNA sequence of the psd gene with the partial amino acid sequence of the mature gene product suggests that the two nonidentical subunits of the mature enzyme are formed by cleavage of a proenzyme resulting in the conversion of Ser-254 to an amino-terminal pyruvate residue (Li, Q.-X., and Dowhan, W. (1988) J. Biol. Chem. 263, 11516-11522). The cleavage of the wild-type proenzyme occurs rapidly with a half-time on the order of 2 min. When Ser-254 is changed to cysteine (S254C), threonine (S254T), or alanine (S254A) by site-directed mutagenesis, the rate of processing of the proenzyme and the production of the functional enzyme are drastically affected. Proenzymes with S254C or S254T are cleaved with a half-time of around 2-4 h while the S254A proenzyme does not undergo processing. The reduced processing rate for the mutant proenzymes is consistent with less of the functional enzyme being made. Mutants encoding the S254C and S254T protein produce 16 and 2%, respectively, of the activity of the wild-type allele but can still complement a temperature-sensitive mutant in the psd locus. There is no detectable activity or complementation observed with the S254A protein. These results are consistent with the hydroxyl group of Ser-254 playing a critical role in the cleavage of the peptide bond between Gly-253 and Ser-254 of the prophosphatidylserine decarboxylase and support the mechanism proposed by Snell and coworkers (Recsei and Snell (1984) Annul Rev. Biochem. 53, 357-387) for the formation of the prosthetic group of pyruvate-dependent decarboxylases.  相似文献   

16.
The role of two glutamate residues (E164 and E144) in the active site of enoyl-CoA hydratase has been probed by site-directed mutagenesis. The catalytic activity of the E164Q and E144Q mutants has been determined using 3'-dephosphocrotonyl-CoA. Removal of the 3'-phosphate group reduces the affinity of the substrate for the enzyme, thereby facilitating the determination of K(m) and simplifying the analysis of the enzymes' pH dependence. k(cat) for the hydration of 3'-dephosphocrotonyl-CoA is reduced 7700-fold for the E144Q mutant and 630000-fold for the E164Q mutant, while K(m) is unaffected. These results indicate that both glutamate residues play crucial roles in the hydration chemistry catalyzed by the enzyme. Previously, we reported that, in contrast to the wild-type enzyme, the E164Q mutant was unable to exchange the alpha-proton of butyryl-CoA with D(2)O [D'Ordine, R. L., Bahnson, B. J., Tonge, P. J. , and Anderson, V. E. (1994) Biochemistry 33, 14733-14742]. Here we demonstrate that E144Q is also unable to catalyze alpha-proton exchange even though E164, the glutamate that is positioned to abstract the alpha-proton, is intact in the active site. The catalytic function of each residue has been further investigated by exploring the ability of the wild-type and mutant enzymes to eliminate 2-mercaptobenzothiazole from 4-(2-benzothiazole)-4-thiabutanoyl-CoA (BTTB-CoA). As expected, reactivity toward BTTB-CoA is substantially reduced (690-fold) for the E164Q enzyme compared to wild-type. However, E144Q is also less active than wild-type (180-fold) even though elimination of 2-mercaptobenzothiazole (pK(a) 6.8) should require no assistance from an acid catalyst. Clearly, the ability of E164 to function as an acid-base in the active site is affected by mutation of E144 and it is concluded that the two glutamates act in concert to effect catalysis.  相似文献   

17.
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   

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.
T M Weigel  V P Miller  H W Liu 《Biochemistry》1992,31(7):2140-2147
CDP-4-keto-6-deoxy-D-glucose-3-dehydrase (E1) purified from Yersinia pseudotuberculosis is a pyridoxamine 5'-phosphate (PMP) dependent enzyme which catalyzes the C-O bond cleavage at C-3 of a CDP-4-keto-6-deoxy-D-glucose substrate, a key step in the formation of 3,6-dideoxyhexoses. Since enzyme E1 utilizes the PMP cofactor in a unique manner, it is essential to establish its role in E1 catalysis. When an incubation was conducted in [18O]H2O, incorporation of 18O into positions C-3 and C-4 of the recovered substrate was observed. This result not only provided the evidence necessary to reveal the reversibility of E1 catalysis but also lent credence to the formation of a delta 3,4-glucoseen intermediate. In view of E1 catalysis being initiated by a C-4' deprotonation of the PMP-substrate complex, the stereochemical course of this step was examined using chemically synthesized (4'S)- and (4'R)-[4'-3H]PMP as probes. Our results clearly demonstrated that the stereochemistry of this deprotonation is pro-S specific, which is in agreement with the stereochemical consistency found with other vitamin B6 phosphate dependent enzymes. The fact that reprotonation at C-4' of the PMP-delta 3,4-glucoseen complex in the reverse direction of E1 catalysis was also found to be pro-S stereospecific strongly suggested that enzyme E1, like most of its counterparts, has the si face of its cofactor-substrate complex exposed to solvent and accessible to active-site catalytic groups as well.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
Homology modeling of gastric H,K-ATPase based on the E2 model of sarcoplasmic reticulum Ca2+-ATPase (Toyoshima, C., and Nomura, H. (2002) Nature 392, 835-839) revealed the presence of a single high-affinity binding site for K+ and an E2 form-specific salt bridge between Glu820 (M6) and Lys791 (M5). In the E820Q mutant this salt bridge is no longer possible, and the head group of Lys791, together with a water molecule, fills the position of the K+ ion and apparently mimics the K+-filled cation binding pocket. This gives an explanation for the K+-independent ATPase activity and dephosphorylation step of the E820Q mutant (Swarts, H. G. P., Hermsen, H. P. H., Koenderink, J. B., Schuurmans Stekhoven, F. M. A. H., and De Pont, J. J. H. H. M. (1998) EMBO J. 17, 3029-3035) and, indirectly, for its E1 preference. The model is strongly supported by a series of reported mutagenesis studies on charged and polar amino acid residues in the membrane domain. To further test this model, Lys791 was mutated alone and in combination with other crucial residues. In the K791A mutant, the K+ affinity was markedly reduced without altering the E2 preference of the enzyme. The K791A mutation prevented, in contrast to the K791R mutation, the spontaneous dephosphorylation of the E820Q mutant as well as its conformational equilibrium change toward E1. This indicates that the salt bridge is essential for high-affinity K+ binding and the E2 preference of H,K-ATPase. Moreover, its breakage (E820Q) can generate a K+-insensitive activity and an E1 preference. In addition, the study gives a molecular explanation for the electroneutrality of H,K-ATPases.  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