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.
In this study, rate equations that predict the regulatory kinetic behavior of homocitrate synthase were derived, and simulation of the predicted behavior was carried out over a range of values for the kinetic parameters. The data obtained allow application of the resulting expressions to enzyme systems that exhibit activation and inhibition as a result of the interaction of effectors at multiple sites in the free enzyme. Homocitrate synthase was used as an example in terms of its activation by Na+ binding to the active enzyme conformer at an allosteric site, inhibition by binding to the active site, and inhibition by lysine binding to the less active enzyme conformer.  相似文献   

2.
Homocitrate synthase (HCS) catalyzes one of the regulated steps of the alpha-aminoadipate pathway for lysine biosynthesis in fungi. The kinetic mechanism of regulation of HCS from Saccharomyces cerevisiae by Na+ and the feedback inhibitor lysine was studied by measuring the initial rate in the absence and presence of the effectors. The data suggest that Na+ is an activator at low concentrations and an inhibitor at high concentrations and that these effects occur as a result of the monovalent ion binding to two different sites in the free enzyme. Inhibition and activation by Na+ can occur simultaneously, with the net rate of the enzyme determined by Na+/K(iNa+) and Na+/K(act), where K(iNa+) and K(act) are the inhibition and activation constants, respectively. The inhibition by Na+ was eliminated at high concentrations of acetyl-CoA, the second substrate bound, but the activation remained. Fluorescence binding studies indicated that lysine bound with high affinity to its binding site as an inhibitor. The inhibition by lysine was competitive versus alpha-ketoglutarate and linear in the physiological range of lysine concentrations up to 5 mm. The effects of Na+ and lysine were independent of one another. A model is developed for regulation of HCS that takes into account all of the effects discussed above.  相似文献   

3.
Histidine-tagged homocitrate synthase from Saccharomyces cerevisiae was purified to about 98% using a Ni-NTA resin and stabilized using a combination of 100 mM guanidine hydrochloride, 100 mM alpha-cyclodextrin, and 600 mM ammonium sulfate. The enzyme was assayed using dichlorophenol indophenol (DCPIP) as an oxidant to oxidize the CoASH produced in the reaction. A stoichiometry of 1:1 was obtained between DCPIP and CoASH. Kinetic parameters for the stable enzyme at pH 7.5 are: Km (AcCoA), 24 microM: Km (alpha-kg), 1.3 mM; and kcat, 37 min(-1). The enzyme, in the absence of reactants, self-associates, as suggested by size exclusion chromatography. Fluorescence and circular dichroic spectra suggested a partially exposed tryptophan residue and a mixed (alpha/beta) secondary structure for the enzyme. Fluorescence quenching studies with KI, CsCl, and acrylamide suggest that the microenvironment around the single tryptophan residue of the enzyme has some positive charge.  相似文献   

4.
Qian J  Khandogin J  West AH  Cook PF 《Biochemistry》2008,47(26):6851-6858
Homocitrate synthase (acetyl-coenzyme A: 2-ketoglutarate C-transferase; E.C. 2.3.3.14) (HCS) catalyzes the condensation of acetyl-CoA (AcCoA) and alpha-ketoglutarate (alpha-KG) to give homocitrate and CoA. Although the structure of an HCS has not been solved, the structure of isopropylmalate synthase (IPMS), a homologue, has been solved (Koon, N., Squire, C. J., and Baker, E. N. (2004) Proc. Natl. Acad. Sci. U.S.A. 101, 8295-8300). Three active site residues in IPMS, Glu-218, His-379, and Tyr-410, were proposed as candidates for catalytic residues involved in deprotonation of the methyl group of AcCoA prior to the Claisen condensation to give homocitrylCoA. All three of the active site residues in IPMS are conserved in the HCS from Saccharomyces cerevisiae. Site-directed mutagenesis has been carried out to probe the role of the homologous residues, Glu-155, His-309, and Tyr-320, in the S. cerevisiae HCS. No detectable activity was observed for the H309A and H309N mutant enzyme, but a slight increase in activity was observed for H309A in the presence of 300 mM imidazole, which is still 1000-fold lower than that of wild type (wt). The E155Q and E155A mutant enzymes exhibited 1000-fold lower activity than wt. The activity of E155A, but not of E155Q, could be partially rescued by formate; a K act of 60 mM with a modest 4-fold maximum activation was observed. In the presence of formate, E155A gives k cat, K AcCoA, and K alpha-KG values of 0.0031 s (-1), 13 muM, and 39 microM, respectively, while a primary kinetic deuterium isotope effect of about 1.4 was obtained on V, with deuterium in the methyl of AcCoA. The pH dependence of k cat for E155A in the presence of formate gave a p K a of 7.9 for a group that must be protonated for optimum activity, similar to that observed for the wt enzyme. However, a partial change was observed on the acid side of the profile, compared to the all or none change observed for wt giving a p K a of about 6.7. The k cat for E155Q decreased at high pH, similar to the wt enzyme, but was pH independent at low pH. The Y320F mutant enzyme only lost 25-fold activity compared to that of the wt, giving k cat, K AcCoA, and K alpha-KG values of 0.039 s (-1), 33 microM, and 140 microM, respectively, and a primary kinetic deuterium isotope effect of 1.3 and 1.8 on V/ K AcCoA and V, respectively; the pH dependence of k cat was similar to that of the wt. These data, combined with a constant pH molecular dynamics simulation study, suggest that a catalytic dyad comprising Glu-155 and His-309 acts to deprotonate the methyl group of AcCoA, while Tyr320 is likely not directly involved in catalysis, but may aid in orienting the reactant and/or the catalytic dyad.  相似文献   

5.
The homocitrate synthase from Thermus thermophilus (TtHCS) is a metal-activated enzyme with either Mg(2+) or Mn(2+) capable of serving as the divalent cation. The enzyme exhibits a sequential kinetic mechanism. The mechanism is steady state ordered with α-ketoglutarate (α-Kg) binding prior to acetyl-CoA (AcCoA) with Mn(2+), whereas it is steady state random with Mg(2+), suggesting a difference in the competence of the E·Mn·α-Kg·AcCoA and E·Mg·α-Kg·AcCoA complexes. The mechanism is supported by product and dead-end inhibition studies. The primary isotope effect obtained with deuterioacetylCoA (AcCoA-d(3)) in the presence of Mg(2+) is unity (value 1.0) at low concentrations of AcCoA, whereas it is 2 at high concentrations of AcCoA. Data suggest the presence of a slow conformational change induced by binding of AcCoA that accompanies deprotonation of the methyl group of AcCoA. The solvent kinetic deuterium isotope effect is also unity at low AcCoA, but is 1.7 at high AcCoA, consistent with the proposed slow conformational change. The maximum rate is pH independent with either Mg(2+) or Mn(2+) as the divalent metal ion, whereas V/K(α-Kg) (with Mn(2+)) decreases at low and high pH giving pK values of about 6.5 and 8.0. Lysine is a competitive inhibitor that binds to the active site of TtHCS, and shares some of the same binding determinants as α-Kg. Lysine binding exhibits negative cooperativity, indicating cross-talk between the two monomers of the TtHCS dimer. Data are discussed in terms of the overall mechanism of TtHCS.  相似文献   

6.
Membrane-associated cytidine 5'-diphospho-1,2-diacyl-sn-glycerol (CDP-diacylglycerol):L-serine O-phosphatidyltransferase (phosphatidylserine synthase, EC2.7.8.8.) and CDP-diacylglycerol:myo-inositol phosphatidyltransferase (phosphatidylinositol synthase, EC 2.7.8.11) were solubilized from the microsomal fraction of Saccharomyces cerevisiae. A variety of detergents were examined for their ability to release phosphatidylserine synthase and phosphatidylinositol synthase activities from the microsome fraction. Both enzymes were solubilized from the microsome fraction with Renex 690 in yield over 80% with increase to specific activity of 1.6-fold. Both solubilized enzymatic activities were dependent on manganese ions and Triton X-100 for maximum activity. The pH optimum for each reaction was 8.0. The apparent Km values for CDP-diacylglycerol and serine for the phosphatidylserine synthase reaction were 0.1 and 0.25 mM, respectively. The apparent Km values for CDP-diacylglycerol and inositol for the phosphatidylinositol synthase reaction were 70 microM and 0.1 mM, respectively. Thioreactive agents inhibited both enzymatic activities. Both solubilized enzymatic activities were thermally inactivated at temperatures above 30 degrees C.  相似文献   

7.
The membrane-associated phospholipid biosynthetic enzyme CDP-diacylglycerol synthase (CTP:phosphatidate cytidylyltransferase, EC 2.7.7.41) was purified 2,300-fold from Saccharomyces cerevisiae. The purification procedure included Triton X-100 solubilization of mitochondrial membranes, CDP-diacylglycerol-Sepharose affinity chromatography, and hydroxylapatite chromatography. The procedure resulted in a nearly homogeneous enzyme preparation as determined by native and sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Radiation inactivation of mitochondrial associated and purified CDP-diacylglycerol synthase suggested that the molecular weight of the native enzyme was 114,000. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis of the purified enzyme preparation yielded two subunits with molecular weights of 56,000 and 54,000. Antibodies prepared against the purified enzyme immunoprecipitated CDP-diacylglycerol synthase activity and subunits. CDP-diacylglycerol synthase activity was dependent on magnesium ions and Triton X-100 at pH 6.5. Thio-reactive agents inhibited activity. The activation energy for the reaction was 9 kcal/mol, and the enzyme was thermally labile above 30 degrees C. The Km values for CTP and phosphatidate were 1 and 0.5 mM, respectively, and the Vmax was 4,700 nmol/min/mg. Results of kinetic and isotopic exchange reactions suggested that the enzyme catalyzes a sequential Bi Bi reaction mechanism.  相似文献   

8.
1. An improved filtration method is introduced to perform kinetic investigations on the chitin synthase reaction. This method is especially suited for the assay of a large number of samples necessary in enzyme kinetic studies. 2. From initial rate data the possibility could be excluded that the two-substrate reactions occurs by a random or a ping-pong mechanism. 3. Investigations of product inhibition exclude a rapid random mechanism but favour an ordered mechanism with UDP-N-acetylglucosamine as the first substrate. 4. This result was confirmed by isotope exchange studies.  相似文献   

9.
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   

10.
Chorismate synthase (EC 4.2.3.5), the seventh enzyme in the shikimate pathway, catalyzes the transformation of 5-enolpyruvylshikimate 3-phosphate (EPSP) to chorismate, which is the last common precursor in the biosynthesis of numerous aromatic compounds in bacteria, fungi, and plants. The chorismate synthase reaction involves a 1,4-trans-elimination of phosphoric acid from EPSP and has an absolute requirement for reduced FMN as a cofactor. We have determined the three-dimensional x-ray structure of the yeast chorismate synthase from selenomethionine-labeled crystals at 2.2-A resolution. The structure shows a novel betaalphabetaalpha fold consisting of an alternate tight packing of two alpha-helical and two beta-sheet layers, showing no resemblance to any documented protein structure. The molecule is arranged as a tight tetramer with D2 symmetry, in accordance with its quaternary structure in solution. Electron density is missing for 23% of the amino acids, spread over sequence regions that in the three-dimensional structure converge on the surface of the protein. Many totally conserved residues are contained within these regions, and they probably form a structured but mobile domain that closes over a cleft upon substrate binding and catalysis. This hypothesis is supported by previously published spectroscopic measurements implying that the enzyme undergoes considerable structural changes upon binding of both FMN and EPSP.  相似文献   

11.
Phosphatidylglycerophosphate synthase activity in Saccharomyces cerevisiae   总被引:4,自引:0,他引:4  
Cytidine 5'-diphospho-1,2-diacyl-sn-glycerol (CDP-diacylglycerol): sn-glycerol-3-phosphate phosphatidyltransferase (phosphatidylglycerophosphate synthase, EC 2.7.8.5) activity was characterized from the mitochondrial fraction of Saccharomyces cerevisiae. The pH optimum for the reaction was 7.0. Maximum activity was dependent on manganese (0.1 mM), magnesium (0.3 mM), or cobalt (1 mM) ions and the nonionic detergent Triton X-100 (1 mM). The apparent Km values for CDP-diacylglycerol and glycerol-3-phosphate were 33 and 27 microM, respectively. Optimal activity was at 30 degrees C with an energy of activation of 5.4 kcal/mol (1 cal = 4.1868 J). Phosphatidylglycerophosphate synthase activity was thermally labile above 40 degrees C. p-Chloromecuriphenylsulfonic acid, N-ethylmaleimide, and mercurous ions inhibited activity. Phosphatidylglycerophosphate synthase activity was partially solubilized from the mitochondrial fraction with 1% Triton X-100.  相似文献   

12.
Inactivation of chitin synthase in Saccharomyces cerevisiae   总被引:3,自引:0,他引:3  
  相似文献   

13.
The activity of chitin synthase extracted from whole cells of Saccharomyces cerevisiae shows reproducible changes during the course of batch cultivation. During exponential growth 5–10% of the enzyme occurs in the active form, whereas in the stationary phase no active enzyme can be detected. Of three yeast proteinases, A, B and C, only B is able to activate pre-chitin synthase and inactivate chitin synthase. A new model of the regulation is presented which accounts for the specific location as well as for termination of chitin synthesis during the budding cycle.These results were reported at the 4th International Symposium on Yeasts in Vienna, July 1974, and are part of doctoral thesis by A.H., University Freiburg (1974).  相似文献   

14.
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   

15.
Xu H  West AH  Cook PF 《Biochemistry》2006,45(39):12156-12166
Kinetic data have been measured for the histidine-tagged saccharopine dehydrogenase from Saccharomyces cerevisiae, suggesting the ordered addition of nicotinamide adenine dinucleotide (NAD) followed by saccharopine in the physiologic reaction direction. In the opposite direction, the reduced nicotinamide adenine dinucleotide (NADH) adds to the enzyme first, while there is no preference for the order of binding of alpha-ketoglutarate (alpha-Kg) and lysine. In the direction of saccharopine formation, data also suggest that, at high concentrations, lysine inhibits the reaction by binding to free enzyme. In addition, uncompetitive substrate inhibition by alpha-Kg and double inhibition by NAD and alpha-Kg suggest the existence of an abortive E:NAD:alpha-Kg complex. Product inhibition by saccharopine is uncompetitive versus NADH, suggesting a practical irreversibility of the reaction at pH 7.0 in agreement with the overall K(eq). Saccharopine is noncompetitive versus lysine or alpha-Kg, suggesting the existence of both E:NADH:saccharopine and E:NAD:saccharopine complexes. NAD is competitive versus NADH, and noncompetitive versus lysine and alpha-Kg, indicating the combination of the dinucleotides with free enzyme. Dead-end inhibition studies are also consistent with the random addition of alpha-Kg and lysine. Leucine and oxalylglycine serve as lysine and alpha-Kg dead-end analogues, respectively, and are uncompetitive against NADH and noncompetitive against alpha-Kg and lysine, respectively. Oxaloacetate (OAA), pyruvate, and glutarate behave as dead-end analogues of lysine, which suggests that the lysine-binding site has a higher affinity for keto acid analogues than does the alpha-Kg site or that dicarboxylic acids have more than one binding mode on the enzyme. In addition, OAA and glutarate also bind to free enzyme as does lysine at high concentrations. Glutarate gives S-parabolic noncompetitive inhibition versus NADH, indicating the formation of a E:(glutarate)2 complex as a result of occupying both the lysine- and alpha-Kg-binding sites. Pyruvate, a slow alternative keto acid substrate, exhibits competitive inhibition versus both lysine and alpha-Kg, suggesting the combination to the E:NADH:alpha-Kg and E:NADH:lysine enzyme forms. The equilibrium constant for the reaction has been measured at pH 7.0 as 3.9 x 10(-7) M by monitoring the change in NADH upon the addition of the enzyme. The Haldane relationship is in very good agreement with the directly measured value.  相似文献   

16.
The addition of ethanolamine or choline to inositol-containing growth medium of Saccharomyces cerevisiae wild-type cells resulted in a reduction of membrane-associated phosphatidylserine synthase (CDPdiacylglycerol:L-serine O-phosphatidyltransferase, EC 2.7.8.8) activity in cell extracts. The reduction of activity did not occur when inositol was absent from the growth medium. Under the growth conditions where a reduction of enzyme activity occurred, there was a corresponding qualitative reduction of enzyme subunit as determined by immunoblotting with antiserum raised against purified phosphatidylserine synthase. Water-soluble phospholipid precursors did not effect purified phosphatidylserine synthase activity. Phosphatidylserine synthase (activity and enzyme subunit) was not regulated by the availability of water-soluble phospholipid precursors in S. cerevisiae VAL2C(YEp CHO1) and the opi1 mutant. VAL2C(YEp CHO1) is a plasmid-bearing strain that over produces phosphatidylserine synthase activity, and the opi1 mutant is an inositol biosynthesis regulatory mutant. The results of this study suggest that the regulation of phosphatidylserine synthase by the availability of phospholipid precursors occurs at the level of enzyme formation and not at the enzyme activity level. Furthermore, the regulation of phosphatidylserine synthase is coupled to inositol synthesis.  相似文献   

17.
Geranylgeranyl pyrophosphate synthase (GGPPs) catalyzes a condensation reaction of farnesyl pyrophosphate with isopentenyl pyrophosphate to generate C(20) geranylgeranyl pyrophosphate, which is a precursor for carotenoids, chlorophylls, geranylgeranylated proteins, and archaeal ether-linked lipid. For short-chain trans-prenyltransferases that synthesize C(10)-C(25) products, bulky amino acid residues generally occupy the fourth or fifth position upstream from the first DDXXD motif to block further elongation of the final products. However, the short-chain type-III GGPPs in eukaryotes lack any large amino acid at these positions. In this study, the first structure of type-III GGPPs from Saccharomyces cerevisiae has been determined to 1.98 A resolution. The structure is composed entirely of 15 alpha-helices joined by connecting loops and is arranged with alpha-helices around a large central cavity. Distinct from other known structures of trans-prenyltransferases, the N-terminal 17 amino acids (9-amino acid helix A and the following loop) of this GGPPs protrude from the helix core into the other subunit and contribute to the tight dimer formation. Deletion of the first 9 or 17 amino acids caused the dissociation of dimer into monomer, and the Delta(1-17) mutant showed abolished enzyme activity. In each subunit, an elongated hydrophobic crevice surrounded by D, F, G, H, and I alpha-helices contains two DDXXD motifs at the top for substrate binding with one Mg(2+) coordinated by Asp(75), Asp(79), and four water molecules. It is sealed at the bottom with three large residues of Tyr(107), Phe(108), and His(139). Compared with the major product C(30) synthesized by mutant H139A, the products generated by mutant Y107A and F108A are predominantly C(40) and C(30), respectively, suggesting the most important role of Tyr(107) in determining the product chain length.  相似文献   

18.
The New York Structural Genomics Research Consortium has targeted highly conserved but uncharacterized enzyme families for structure determination. As part of this effort, the 2.65-Å crystal structure has been determined for Saccharomyces cerevisiae myo-inositol 1-phosphate synthase (MIP), an essential enzyme that catalyzes critical steps in inositol biosynthesis. The structure determination of four independent monomers in the asymmetric unit (240 kDa) reveals atomic details and residue composition for the partially closed NAD-containing active sites in apo-configuration. The structure further reveals extensive interactions involved in tetrameric assembly of the enzyme complex.  相似文献   

19.
Feedback inhibition of N-acetylgutamate synthase in a particulate fraction from Saccharomyces cerevisiae by L-arginine was synergistically enhanced by N-actylglutamate, whereas coenzyme A let to an additive enhancement of arginine inhibition. N-acetylglutamate synthase was not inhibited by polyamines, nor was the enzyme inactivated by incubation in the presence of coenzyme A and zinc ions. Evidence was obtained for the involvement of at least three different regulatory mechanisms in the expression of N-acetylglutamate synthase: arginine-specific repression, glucose repression and general amino acid control. The combined action of these control mechanisms led to a 90-fold variation in the specific activity of the enzyme.  相似文献   

20.
The yeast ilv2 gene, encoding acetolactate synthase, was subcloned in an Escherichia coli expression vector. Although a major part of the acetolactate synthase synthesized by E. coli cells harbouring this vector was packaged into protein inclusion bodies, we used these recombinant E. coli cells to produce large quantities of the yeast enzyme. The yeast acetolactate synthase was purified to homogeneity using first streptomycin and ammonium sulfate precipitations, followed by T-gel thiophilic interaction, Sephacryl S-300 gel filtration, Mono Q anion exchange, and Superose 12 gel filtration chromatography. SDS/PAGE and gel filtration of the purified enzyme showed that it is a dimer composed of two subunits, each with the molecular mass of 75 kDa. The purified yeast acetolactate synthase was further characterized with respect to pH optimum, dependence of the substrate, pyruvate, and requirements of the cofactors, thiamin diphosphate, Mg2+, and FAD.  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