首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Homoisocitrate dehydrogenase is involved in the alpha-aminoadipate pathway of L-lysine biosynthesis in higher fungi such as yeast and human pathogenic fungi. This enzyme catalyzes the oxidative decarboxylation of (2R,3S)-homoisocitrate into 2-ketoadipate using NAD(+) as a coenzyme. A series of aza-, oxa-, and thia-analogues of homoisocitrate was designed and synthesized as an inhibitor for homoisocitrate dehydrogenase. Among them, thia-analogue showed strong competitive inhibitory activity as K(i)=97 nM toward homoisocitrate dehydrogenase derived from Saccharomyces cerevisiae. Kinetic studies suggested that the formation of the enolate intermediate played an important role in inhibition.  相似文献   

2.
Although the presence of an enzyme that catalyzes beta-decarboxylating dehydrogenation of homoisocitrate to synthesize 2-oxoadipate has been postulated in the lysine biosynthesis pathway through alpha-aminoadipate (AAA), the enzyme has not yet been analyzed at all, because no gene encoding the enzyme has been identified until recently. A gene encoding a protein with a significant amino acid sequence identity to both isocitrate dehydrogenase and 3-isopropylmalate dehydrogenase was cloned from Thermus thermophilus HB27. The gene product produced in recombinant Escherichia coli cells demonstrated homoisocitrate dehydrogenase (HICDH) activity. A knockout mutant of the gene showed an AAA-auxotrophic phenotype, indicating that the gene product is involved in lysine biosynthesis through AAA. We therefore named this gene hicdh. HICDH, the gene product, did not catalyze the conversion of 3-isopropylmalate to 2-oxoisocaproate, a leucine biosynthetic reaction, but it did recognize isocitrate, a related compound in the tricarboxylic acid cycle, as well as homoisocitrate as a substrate. It is of interest that HICDH catalyzes the reaction with isocitrate about 20 times more efficiently than the reaction with the putative native substrate, homoisocitrate. The broad specificity and possible dual function suggest that this enzyme represents a key link in the evolution of the pathways utilizing citrate derivatives. Site-directed mutagenesis study reveals that replacement of Arg(85) with Val in HICDH causes complete loss of activity with isocitrate but significant activity with 3-isopropylmalate and retains activity with homoisocitrate. These results indicate that Arg(85) is a key residue for both substrate specificity and evolution of beta-decarboxylating dehydrogenases.  相似文献   

3.
Homoaconitase enzymes catalyze hydrolyase reactions in the alpha-aminoadipate pathway for lysine biosynthesis or the 2-oxosuberate pathway for methanogenic coenzyme B biosynthesis. Despite the homology of this iron-sulfur protein to aconitase, previously studied homoaconitases catalyze only the hydration of cis-homoaconitate to form homoisocitrate rather than the complete isomerization of homocitrate to homoisocitrate. The MJ1003 and MJ1271 proteins from the methanogen Methanocaldococcus jannaschii formed the first homoaconitase shown to catalyze both the dehydration of (R)-homocitrate to form cis-homoaconitate, and its hydration is shown to produce homoisocitrate. This heterotetrameric enzyme also used the analogous longer chain substrates cis-(homo)(2)aconitate, cis-(homo)(3)aconitate, and cis-(homo)(4)aconitate, all with similar specificities. A combination of the homoaconitase with the M. jannaschii homoisocitrate dehydrogenase catalyzed all of the isomerization and oxidative decarboxylation reactions required to form 2-oxoadipate, 2-oxopimelate, and 2-oxosuberate, completing three iterations of the 2-oxoacid elongation pathway. Methanogenic archaeal homoaconitases and fungal homoaconitases evolved in parallel in the aconitase superfamily. The archaeal homoaconitases share a common ancestor with isopropylmalate isomerases, and both enzymes catalyzed the hydration of the minimal substrate maleate to form d-malate. The variation in substrate specificity among these enzymes correlated with the amino acid sequences of a flexible loop in the small subunits.  相似文献   

4.
Homoisocitrate dehydrogenase (HIcDH, 3-carboxy-2-hydroxyadipate dehydrogenase) catalyzes the fourth reaction of the alpha-aminoadipate pathway for lysine biosynthesis, the conversion of homoisocitrate to alpha-ketoadipate using NAD as an oxidizing agent. A chemical mechanism for HIcDH is proposed on the basis of the pH dependence of kinetic parameters, dissociation constants for competitive inhibitors, and isotope effects. According to the pH-rate profiles, two enzyme groups act as acid-base catalysts in the reaction. A group with a p K a of approximately 6.5-7 acts as a general base accepting a proton as the beta-hydroxy acid is oxidized to the beta-keto acid, and this residue participates in all three of the chemical steps, acting to shuttle a proton between the C2 hydroxyl and itself. The second group acts as a general acid with a p K a of 9.5 and likely catalyzes the tautomerization step by donating a proton to the enol to give the final product. The general acid is observed in only the V pH-rate profile with homoisocitrate as a substrate, but not with isocitrate as a substrate, because the oxidative decarboxylation portion of the isocitrate reaction is limiting overall. With isocitrate as the substrate, the observed primary deuterium and (13)C isotope effects indicate that hydride transfer and decarboxylation steps contribute to rate limitation, and that the decarboxylation step is the more rate-limiting of the two. The multiple-substrate deuterium/ (13)C isotope effects suggest a stepwise mechanism with hydride transfer preceding decarboxylation. With homoisocitrate as the substrate, no primary deuterium isotope effect was observed, and a small (13)C kinetic isotope effect (1.0057) indicates that the decarboxylation step contributes only slightly to rate limitation. Thus, the chemical steps do not contribute significantly to rate limitation with the native substrate. On the basis of data from solvent deuterium kinetic isotope effects, viscosity effects, and multiple-solvent deuterium/ (13)C kinetic isotope effects, the proton transfer step(s) is slow and likely reflects a conformational change prior to catalysis.  相似文献   

5.
Beta-decarboxylating dehydrogenases comprise 3-isopropylmalate dehydrogenase, isocitrate dehydrogenase, and homoisocitrate dehydrogenase. They share a high degree of amino acid sequence identity and occupy equivalent positions in the amino acid biosynthetic pathways for leucine, glutamate, and lysine, respectively. Therefore, not only the enzymes but also the whole pathways should have evolved from a common ancestral pathway. In Pyrococcus horikoshii, only one pathway of the three has been identified in the genomic sequence, and PH1722 is the sole beta-decarboxylating dehydrogenase gene. The organism does not require leucine, glutamate, or lysine for growth; the single pathway might play multiple (i.e., ancestral) roles in amino acid biosynthesis. The PH1722 gene was cloned and expressed in Escherichia coli and the substrate specificity of the recombinant enzyme was investigated. It exhibited activities on isocitrate and homoisocitrate at near equal efficiency, but not on 3-isopropylmalate. PH1722 is thus a novel, bifunctional beta-decarboxylating dehydrogenase, which likely plays a dual role in glutamate and lysine biosynthesis in vivo.  相似文献   

6.
Sulfolobus acidocaldarius, a hyperthermoacidophilic archaeon, possesses two β-decarboxylating dehydrogenase genes, saci_0600 and saci_2375, in its genome, which suggests that it uses these enzymes for three similar reactions in lysine biosynthesis through 2-aminoadipate, leucine biosynthesis, and the tricarboxylic acid cycle. To elucidate their roles, these two genes were expressed in Escherichia coli in the present study and their gene products were characterized. Saci_0600 recognized 3-isopropylmalate as a substrate, but exhibited slight and no activity for homoisocitrate and isocitrate, respectively. Saci_2375 exhibited distinct and similar activities for isocitrate and homoisocitrate, but no detectable activity for 3-isopropylmalate. These results suggest that Saci_0600 is a 3-isopropylmalate dehydrogenase for leucine biosynthesis and Saci_2375 is a dual function enzyme serving as isocitrate-homoisocitrate dehydrogenase. The crystal structure of Saci_0600 was determined as a closed-form complex that binds 3-isopropylmalate and Mg2+, thereby revealing the structural basis for the extreme thermostability and novel-type recognition of the 3-isopropyl moiety of the substrate.  相似文献   

7.
G W Plaut  R L Beach  T Aogaichi 《Biochemistry》1975,14(12):2581-2588
D-Garcinia acid (D-threo-1,2-dihydroxy-1,2,3-propanetricarboxylate), like D-isocitrate, has an alpha-DS-hydroxyl group and a beta-LS configuration of the second carboxyl group. The maximal velocity of pyridine nucleotide reduction with D-garcinia acid is 8 and 21% of D-threo-isocitrate with the DPN-linked and TPN-linked isocitrate dehydrogenase from bovine heart, respectively. The other stereoisomers of hydroxycitrate [L-garcinia acid, D- and L-hibiscus acid (D- and L-erythro-1,2-dihydroxy-1,2,3-propanetricarboxylate)] are inactive. DL-threo-Homoisocitrate (DL-threo-1-hydroxy-1,2,4-butanetricarboxylate) supports DPN+ reduction at 10-15% of the rate observed for isocitrate with the DPN-specific enzyme, but is not a substrate for TPN-linked isocitrate dehydrogenase. The values of apparent S0.5 for total isocitrate and total garcinia acid are similar with both enzymes; the apparent S0.5 of total homoisocitrate is two- to threefold higher than that of total isocitrate with the DPN-linked enzyme. Enzymatic oxidative decarboxylation of garcinia acid and homoisocitrate leads to formation of alpha-keto-beta-hydroxyglutarate and alpha-ketoadipate, respectively. DL-Methylmalate (DL-1-hydroxy-2-methylsuccinate) is inactive as a substrate for either dehydrogenase as are the newly synthesized compounds: DL-threo-gamma-isocitrate amide (DL-threo-1-hydroxy-3-carbamy01,2-propanedicarboxylate), beta-methyl-DL-isocitrate (DL-1-hydroxy-2-methyl-1,2,3-propanetricarboxylate), beta-methyl-DL-garcinia acid (DL-threo-1-hydroxyl-2-methoxy-1,2,3-propanetricarboxylate), DL-1-hydroxyl-1,2,2-ethanetricarboxylate, and DL-1,4-dihydroxy-1,2-butanedicarboxylate.  相似文献   

8.
(3R)-hydroxyacyl-CoA dehydrogenase is part of multifunctional enzyme type 2 (MFE-2) of peroxisomal fatty acid beta-oxidation. The MFE-2 protein from yeasts contains in the same polypeptide chain two dehydrogenases (A and B), which possess difference in substrate specificity. The crystal structure of Candida tropicalis (3R)-hydroxyacyl-CoA dehydrogenase AB heterodimer, consisting of dehydrogenase A and B, determined at the resolution of 2.2A, shows overall similarity with the prototypic counterpart from rat, but also important differences that explain the substrate specificity differences observed. Docking studies suggest that dehydrogenase A binds the hydrophobic fatty acyl chain of a medium-chain-length ((3R)-OH-C10) substrate as bent into the binding pocket, whereas the short-chain substrates are dislocated by two mechanisms: (i) a short-chain-length 3-hydroxyacyl group ((3R)-OH-C4) does not reach the hydrophobic contacts needed for anchoring the substrate into the active site; and (ii) Leu44 in the loop above the NAD(+) cofactor attracts short-chain-length substrates away from the active site. Dehydrogenase B, which can use a (3R)-OH-C4 substrate, has a more shallow binding pocket and the substrate is correctly placed for catalysis. Based on the current structure, and together with the structure of the 2-enoyl-CoA hydratase 2 unit of yeast MFE-2 it becomes obvious that in yeast and mammalian MFE-2s, despite basically identical functional domains, the assembly of these domains into a mature, dimeric multifunctional enzyme is very different.  相似文献   

9.
Antiinflammatory agents and estrogens have been tested as inhibitors of two isozymes of guinea pig liver testosterone 17 beta-dehydrogenase (NADP) 1.1.1.64) and rat liver 3 alpha-hydroxysteroid dehydrogenase (EC 1.1.1.50). Antiinflammatory steroids and estradiols were highly inhibitory to 3 alpha-hydroxysteroid dehydrogenase and one isozyme of testosterone 17 beta-dehydrogenase, respectively, but nonsteroidal antiinflammatory agents and nonsteroidal estrogens such as hexestrol, dienstrol, diethylstilbestrol and zearalenone showed potent inhibitions on all the enzymes. Although the inhibitory potency of indomethacin for one isozymes of testosterone 17 beta-dehydrogenase and 3 alpha-hydroxysteroid dehydrogenase decreased with changing pH from 9.7 to 7.0, that of the nonsteroidal estrogens for all the enzymes was little affected by pH. No additive effect in double inhibitor experiments with indomethacin and the nonsteroidal estrogens was observed, and the compounds were all competitive inhibitors with respect to steroidal substrate. The results suggest that there is a very similar region in substrate binding sites of the enzymes.  相似文献   

10.
The crystal structure of homoisocitrate dehydrogenase involved in lysine biosynthesis from Thermus thermophilus (TtHICDH) was determined at 1.85-A resolution. Arg85, which was shown to be a determinant for substrate specificity in our previous study, is positioned close to the putative substrate binding site and interacts with Glu122. Glu122 is highly conserved in the equivalent position in the primary sequence of ICDH and archaeal 3-isopropylmalate dehydrogenase (IPMDH) but interacts with main- and side-chain atoms in the same domain in those paralogs. In addition, a conserved Tyr residue (Tyr125 in TtHICDH) which extends its side chain toward a substrate and thus has a catalytic function in the related beta-decarboxylating dehydrogenases, is flipped out of the substrate-binding site. These results suggest the possibility that the conformation of the region containing Glu122-Tyr125 is changed upon substrate binding in TtHICDH. The crystal structure of TtHICDH also reveals that the arm region is involved in tetramer formation via hydrophobic interactions and might be responsible for the high thermotolerance. Mutation of Val135, located in the dimer-dimer interface and involved in the hydrophobic interaction, to Met alters the enzyme to a dimer (probably due to steric perturbation) and markedly decreases the thermal inactivation temperature. Both the crystal structure and the mutation analysis indicate that tetramer formation is involved in the extremely high thermotolerance of TtHICDH.  相似文献   

11.
NAD(+)-dependent (R)-2-hydroxyglutarate dehydrogenase (HGDH) catalyses the reduction of 2-oxoglutarate to (R)-2-hydroxyglutarate and belongs to the d-2-hydroxyacid NAD(+)-dependent dehydrogenase (d-2-hydroxyacid dehydrogenase) protein family. Its crystal structure was determined by phase combination to 1.98 A resolution. Structure-function relationships obtained by the comparison of HGDH with other members of the d-2-hydroxyacid dehydrogenase family give a chemically satisfying view of the substrate stereoselectivity and catalytic requirements for the hydride transfer reaction. A model for substrate recognition and turnover is discussed. The HGDH active site architecture is structurally optimized to recognize and bind the negatively charged substrate 2-oxoglutarate. The structural position of the side chain of Arg52, and its counterparts in other family members, strongly correlates with substrate specificity towards substitutions at the C3 atom (linear or branched substrates). Arg235 interacts with the substrate's alpha-carboxylate and carbonyl groups, having a dual role in both substrate binding and activation, and the gamma-carboxylate group can dock at an arginine cluster. The proton-relay system built up by Glu264 and His297 permits His297 to act as acid-base catalyst and the 4Re-hydrogen from NADH is transferred as hydride to the carbonyl group Si-face leading to the formation of the correct enantiomer (R)-2-hydroxyglutarate.  相似文献   

12.
Trypsin-like protease from soybean seeds. Purification and some properties   总被引:2,自引:0,他引:2  
An enzyme was purified from soybean seeds mainly by repeated ion-exchange chromatography using benzoyl-L-arginine p-nitroanilide (BAPA) as a substrate. The purified enzyme was homogeneous as judged by disc gel electrophoresis. The molecular weight was estimated as 59,000 by gel filtration. The enzyme was most active toward BAPA between pH 8 and 10. The enzyme was inactive toward protein substrates but hydrolyzed synthetic substrates and oligopeptides exclusively at the carboxyl side of L-arginine and L-lysine. Kinetic studies using synthetic substrates showed that, on the basis of Vmax/Km, the enzyme preferentially hydrolyzed amide substrates over ester substrates. Benzoyl-L-arginine 4-methylcoumaryl-7-amide (Bz-Arg-MCA) was the best substrate. The enzyme was strongly inhibited by diisopropylfluorophosphate (DFP), tosyl-L-lysine chloromethyl ketone (Tos-Lys-CH2Cl), leupeptin, and antipain. p-Chloromercuribenzoate (PCMB) was only partially inhibitory. Various protein inhibitors of trypsin such as soybean trypsin inhibitor were ineffective. From the primary specificity and susceptibility to chemicals, the enzyme can be said to be a trypsin-like serine protease. Although the physiological role of the enzyme is unclear, it seems likely that it is involved in limited hydrolysis of certain physiological peptides during processing.  相似文献   

13.
HACN (homoaconitase) is a member of a family of [4Fe-4S] cluster-dependent enzymes that catalyse hydration/dehydration reactions. The best characterized example of this family is the ubiquitous ACN (aconitase), which catalyses the dehydration of citrate to cis-aconitate, and the subsequent hydration of cis-aconitate to isocitrate. HACN is an enzyme from the alpha-aminoadipate pathway of lysine biosynthesis, and has been identified in higher fungi and several archaea and one thermophilic species of bacteria, Thermus thermophilus. HACN catalyses the hydration of cis-homoaconitate to (2R,3S)-homoisocitrate, but the HACN-catalysed dehydration of (R)-homocitrate to cis-homoaconitate has not been observed in vitro. We have synthesized the substrates and putative substrates for this enzyme, and in the present study report the first steady-state kinetic data for recombinant HACN from T. thermophilus using a (2R,3S)-homoisocitrate dehydrogenase-coupled assay. We have also examined the products of the reaction using HPLC. We do not observe HACN-catalysed 'homocitrate dehydratase' activity; however, we have observed that ACN can catalyse the dehydration of (R)-homocitrate to cis-homoaconitate, but HACN is required for subsequent conversion of cis-homoaconitate into homoisocitrate. This suggests that the in vivo process for conversion of homocitrate into homoisocitrate requires two enzymes, in simile with the propionate utilization pathway from Escherichia coli. Surprisingly, HACN does not show any activity when cis-aconitate is substituted for the substrate, even though other enzymes from the alpha-aminoadipate pathway can accept analogous tricarboxylic acid-cycle substrates. The enzyme shows no apparent feedback inhibition by L-lysine.  相似文献   

14.
The substrate specificity of spermidine dehydrogenase from Serratia marcescens was studied using many kinds of naturally occurring and synthetic polyamines. Diamines inhibited the enzyme competitively and their inhibitor constants tended to decrease with increasing methylene chain length in the diamines. All of the triamines and tetramines examined were active as substrates, and the amines containing a 4-aminobutylimino moiety (NH2(CH2)4NH-) in their structures were more active. N-Alkylputrescine was also oxidized by the enzyme. All of the amines containing a 4-aminobutylimino group were oxidized to form 1-pyrroline stoichiometrically as one of the products. Tetramines containing a 3-aminopropylimino group (NH2(CH2)3NH-) were oxidized to form 1,3-diaminopropane. However, in the case of an amine containing both 4-aminobutylimino and 3-aminopropylimino groups, the imino moiety of the former was preferentially oxidized by the enzyme. On the basis of the substrate specificity, the binding characteristics of the enzyme are discussed and a subsite model for the binding site is proposed.  相似文献   

15.
1-Acyl-2(R)-acylamino phospholipids are effective competitive inhibitors of porcine pancreatic phospholipase A2 (EC 3.1.1.4, Bonsen et al. (1972) Biochim. Biophys. Acta 270, 364-382). By systematically varying the substituent at C-1 and the acyl chain length at C-2, a series of phospholipid analogues was obtained for which the inhibitory power was determined in a detergent-containing and occasionally also in a detergent-free micellar substrate system. The recently proposed kinetic model applicable to water-insoluble inhibitors (Ransac et al. (1990) Biochim. Biophys. Acta 1043, 57-66) allowed a quantitative comparison of the inhibitory power Z of the various substrate analogues. The most powerful inhibitors of the enzyme were found to possess the general 2-(R)-structure: (formula; see text) Using as substrate (R)-1,2-didodecanoylglycero-3-phosphocholine in mixed micelles with sodium taurodeoxycholate, the inhibitor molecule with m = 4 and n = 11 showed a Z-value of 15,000. This implies an affinity of the inhibitor for the active site of the enzyme higher than 4 orders of magnitude stronger as compared with the substrate molecule. Slightly higher and lower m-values resulted in a sharp drop of the inhibitory power, which suggests that the enzyme must possess a rather short, but well-defined hydrophobic binding pocket for the C-1 alkyl chain. Variation of n (keeping m = 2 constant) resulted in inhibitors with nearly equal Z-values for n = 11, 13 and 15. Most probably the binding cleft on the enzyme for the C-2 acylamino chain is longer, more losely constructed and contributing less to the overall binding energy. Several members of the 2-acylamino phospholipids are water-soluble and possess relatively high critical micelle concentrations. Their inhibitory power could be tested not only in micellar substrate dispersions but also in assay systems where both the inhibitor and substrate are molecularly dispersed. It appeared that these water-soluble phospholipid analogues are effective inhibitors of the enzyme only after incorporation into an organized substrate/water interface. In contrast, in molecularly dispersed substrate solutions the same molecules have completely lost their inhibitory power. These observations support our kinetic model of lipolysis and interfacial inhibition.  相似文献   

16.
Human acid beta-glucosidase (D-glucosyl-N-acylsphingosine glucohydrolase, EC 3.2.1.45) cleaves the beta-glucosidic bonds of glucosylceramide and synthetic beta-glucosides. The specificity of binding to the active site of this enzyme was evaluated using series of inhibitors including synthetic sphingosines, N-alkyl(Cn)-deoxynojirimycins (1,5-dideoxy-5-iminoglucose) and N-Cn-glucosylamines. The sphingosines were rapidly reversible inhibitors with maximal potency (IC50 approximately 78-150 micro M) at chain lengths of 14-18 carbons. The presence of unsaturation between C4 and C5 was required for inhibition of enzyme activity. Neither the nature of this bond (double or triple bond) nor the presence of erythro or threo configurations at C2 influenced inhibitory potency. The N-C10- to N-C14-deoxynojirimycins were rapidly reversible inhibitors with Ki approximately 8.5 nM. In comparison, the 1-amino glucose derivatives, i.e., N-Cn-glucosylamines (n = 12-18), were more potent (IC50 approximatley 0.3-3 nM) and their maximal inhibitory potencies were dependent on time as well as enzyme and substrate concentrations: i.e., the N-C12- to N-C18-glucosylamines were competitive, slow-tight binding inhibitors. Analyses of progress curves at various N-Cn-glucosylamine (n = 14-18) concentrations indicated the formation of rapidly dissociating initial EI collison complex which then undergoes a conformational change to a slowly reversible EI complex. These results were consistent with the long chain N-Cn-glucosylamines being reaction intermediate analogues and with this enzyme's hydrolytic mechanism requiring a conformational change during the transition state.  相似文献   

17.
Purification and characterization of human liver sorbitol dehydrogenase   总被引:1,自引:0,他引:1  
W Maret  D S Auld 《Biochemistry》1988,27(5):1622-1628
Sorbitol dehydrogenase from human liver was purified to homogeneity by affinity chromatography on immobilized triazine dyes, conventional cation-exchange chromatography, and high-performance liquid chromatography. The major form is a tetrameric, NAD-specific enzyme containing one zinc atom per subunit. Human liver sorbitol dehydrogenase oxidizes neither ethanol nor other primary alcohols. It catalyzes the oxidation of a secondary alcohol group of polyol substrates such as sorbitol, xylitol, or L-threitol. However, the substrate specificity of human liver sorbitol dehydrogenase is broader than that of the liver enzymes of other sources. The present report describes the stereospecific oxidation of (2R,3R)-2,3-butanediol, indicating a more general function of sorbitol dehydrogenase in the metabolism of secondary alcohols. Thus, the enzyme complements the substrate specificities covered by the three classes of human liver alcohol dehydrogenase.  相似文献   

18.
It has been shown that the P1 site (the center of the reactive site) of protease inhibitors corresponds to the specificity of the cognate protease, and consequently specificity of Streptomyces subtilisin inhibitor (SSI) can be altered by substitution of a single amino acid at the P1 site. In this paper, to investigate whether similar correlation between inhibitory activity of mutated SSI and substrate preference of protease is observed for subtilisin BPN', which has broad substrate specificity, a complete set of mutants of SSI at the reaction site P1 (position 73) was constructed by cassette and site-directed mutagenesis and their inhibitory activities toward subtilisin BPN' were measured. Mutated SSIs which have a polar (Ser, Thr, Gln, Asn), basic (Lys, Arg), or aromatic amino acid (Tyr, Phe, Trp, His), or Ala or Leu, at the P1 site showed almost the same strong inhibitory activity toward subtilisin as the wild type (Met) SSI. However, the inhibitory activity of SSI variants with an acidic (Glu, Asp), or a beta-branched aliphatic amino acid (Val, Ile), or Gly or Pro, at P1 was decreased. The values of the inhibitor constant (Ki) of mutated SSIs toward subtilisin BPN' were consistent with the substrate preference of subtilisin BPN'. A linear correlation was observed between log(1/Ki) of mutated SSIs and log(1/Km) of synthetic substrates. These results demonstrate that the inhibitory activities of P1 site mutants of SSI are linearly related to the substrate preference of subtilisin BPN', and indicate that the binding mode of the inhibitors with the protease may be similar to that of substrates, as in the case of trypsin and chymotrypsin.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
The completion of the Saccharomyces cerevisiae genome project in 1996 showed that almost 60% of the potential open reading frames of the genome had no experimentally determined function. Using a conserved sequence motif present in the zinc-containing medium-chain alcohol dehydrogenases, we found several potential alcohol dehydrogenase genes with no defined function. One of these, YAL060W, was overexpressed using a multicopy inducible vector, and its protein product was purified to homogeneity. The enzyme was found to be a homodimer that, in the presence of NAD(+), but not of NADP, could catalyze the stereospecific oxidation of (2R,3R)-2, 3-butanediol (K(m) = 14 mm, k(cat) = 78,000 min(-)(1)) and meso-butanediol (K(m) = 65 mm, k(cat) = 46,000 min(-)(1)) to (3R)-acetoin and (3S)-acetoin, respectively. It was unable, however, to further oxidize these acetoins to diacetyl. In the presence of NADH, it could catalyze the stereospecific reduction of racemic acetoin ((3R/3S)- acetoin; K(m) = 4.5 mm, k(cat) = 98,000 min(-)(1)) to (2R,3R)-2,3-butanediol and meso-butanediol, respectively. The substrate stereospecificity was determined by analysis of products by gas-liquid chromatography. The YAL060W gene product can therefore be classified as an NAD-dependent (2R,3R)-2,3-butanediol dehydrogenase (BDH). S. cerevisiae could grow on 2,3-butanediol as the sole carbon and energy source. Under these conditions, a 3. 5-fold increase in (2R,3R)-2,3-butanediol dehydrogenase activity was observed in the total cell extracts. The isoelectric focusing pattern of the induced enzyme coincided with that of the pure BDH (pI 6.9). The disruption of the YAL060W gene was not lethal for the yeast under laboratory conditions. The disrupted strain could also grow on 2,3-butanediol, although attaining a lesser cell density than the wild-type strain. Taking into consideration the substrate specificity of the YAL060W gene product, we propose the name of BDH for this gene. The corresponding enzyme is the first eukaryotic (2R, 3R)-2,3-butanediol dehydrogenase characterized of the medium-chain dehydrogenase/reductase family.  相似文献   

20.
The kinetic mechanism of homoisocitrate dehydrogenase from Saccharomyces cerevisiae was determined using initial velocity studies in the absence and presence of product and dead end inhibitors in both reaction directions. Data suggest a steady state random kinetic mechanism. The dissociation constant of the Mg-homoisocitrate complex (MgHIc) was estimated to be 11 +/- 2 mM as measured using Mg2+ as a shift reagent. Initial velocity data indicate the MgHIc complex is the reactant in the direction of oxidative decarboxylation, while in the reverse reaction direction, the enzyme likely binds uncomplexed Mg2+ and alpha-ketoadipate. Curvature is observed in the double-reciprocal plots for product inhibition by NADH and the dead-end inhibition by 3-acetylpyridine adenine dinucleotide phosphate when MgHIc is the varied substrate. At low concentrations of MgHIc, the inhibition by both nucleotides is competitive, but as the MgHIc concentration increases, the inhibition changes to uncompetitive, consistent with a steady state random mechanism with preferred binding of MgHIc before NAD. Release of product is preferred and ordered with respect to CO2, alpha-ketoadipate, and NADH. Isocitrate is a slow substrate with a rate (V/E(t)) 216-fold slower than that measured with HIc. In contrast to HIc, the uncomplexed form of isocitrate and Mg2+ bind to the enzyme. The kinetic mechanism in the direction of oxidative decarboxylation of isocitrate, on the basis of initial velocity studies in the absence and presence of dead-end inhibitors, suggests random addition of NAD and isocitrate with Mg2+ binding before isocitrate in rapid equilibrium, and the mechanism approximates rapid equilibrium random. The Keq for the overall reaction measured directly using the change in NADH as a probe is 0.45 M.  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