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.
Substitution of a lysyl residue for Arg-386 of Escherichia coli aspartate aminotransferase resulted in an extensive decrease in Vmax values (0.8% with the aspartate-2-oxoglutarate pair and 0.2% with the glutamate-oxalacetate pair, compared with the corresponding values for the wild-type enzyme). Kinetic analysis of the four sets of half-reactions, the pyridoxal form of the enzyme with aspartate or glutamate and the pyridoxamine form with 2-oxoglutarate or oxalacetate, allowed us to define the independent effect of the mutation on the reactivity of each substrate. Decrease in the first order rate constant (kmax) was more pronounced in the reactions with five-carbon substrates (glutamate and 2-oxoglutarate) than in those with four-carbon substrates (aspartate and oxalacetate), while the increase in the apparent dissociation constant (Kd) was greater for four-carbon substrates than for five-carbon substrates. The decrease of overall catalytic efficiency as judged by the values, kmax/Kd, was more pronounced in the reactions with five-carbon substrates than in those with four-carbon substrates. Affinities for substrate analogs such as succinate, glutarate, 2-methylaspartate, and erythro-3-hydroxyaspartate, were also considerably decreased by the mutation of the enzyme. These findings indicate that the side chain of the lysyl residue, although it bears a positive charge similar to that of the arginyl residue, is not structurally adequate for the productive binding of a substrate during catalysis.  相似文献   

2.
Tyr225 in the active site of Escherichia coli aspartate aminotransferase (AspAT) was replaced by phenylalanine or arginine by site-directed mutagenesis. X-ray crystallographic analysis of Y225F AspAT showed that the benzene ring of Phe225 was situated at the same position as the phenol ring of Tyr225 in wild-type AspAT. The mutations resulted in a great decrease in the rate of the transamination reaction, suggesting that Tyr225 is important for efficient catalysis. The kinetic analysis of half-transamination reactions of Y225F AspAT with four substrates (aspartate, glutamate, oxalacetate, and 2-oxoglutarate) and some analogues (2-methylaspartate, succinate, and glutarate) revealed a considerable increase in the affinities for all these compounds. In contrast, affinity for the amino acid substrates was decreased by mutation to arginine, but affinities for the keto acid substrates and the two dicarboxylates (succinate and glutarate) were increased. The electrostatic interaction between O(3') of the coenzyme [pyridoxal 5'-phosphate (PLP)] and the residue at position 225 affected the pKa value of the Schiff base, which is formed between the epsilon-amino group of Lys258 and the aldehyde group of PLP; based on the spectrophotometric titration the pKa values were determined to be 6.8 for wild-type AspAT, 8.5 for Y225F AspAT, and 6.1 for Y225R AspAT in the absence of substrate. The absorption spectra of the three AspATs were almost identical in the acidic pH region, but the spectrum of Y225F AspAT differed from that of wild-type or Y225R AspAT in the alkaline pH region.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
1. The steady residual activity of ox liver glutamate dehydrogenase at equilibrium with the reversible inactivator pyridoxal 5'-phosphate was measured in the presence and absence of various protecting agents. 2. NAD(+) (up to 15mm) and its 3-acetylpyridine analogue (up to 5mm) both failed to protect, in contrast with NADH. 3. Partial protection was given by glutarate and by succinate. Adipate and pentanoate were much less effective. 4. Correspondingly, whereas succinate and glutarate were both shown to be strong inhibitors of the catalytic reaction, competitive with glutamate, adipate was only weakly competitive, and the still weaker inhibition by pentanoate was non-competitive. 5. When the enzyme was saturated with glutarate, NAD(+) became a good, although still partial, protecting agent. In the absence of protection, 1.8mm-pyridoxal 5'-phosphate decreased enzyme activity to 9%, in the presence of 150mm-glutarate to 29%, and with glutarate and 1mm-NAD(+) only to 73%. 6. 2-Oxoglutarate also promoted protection by NAD(+), but neither pentanoate nor succinate did so. The finding with succinate is remarkable in view of findings 3 and 4 above. 7. It seems possible that substrates or analogues possessing the glutarate structure promote a conformational change that alters the mode of NAD(+) binding. This may explain why glutamate is a much better substrate than norvaline or aspartate and why negative interactions in coenzyme binding occur only in the formation of ternary complexes with glutamate or its analogues.  相似文献   

4.
In aspartate aminotransferase (AspAT), His143 is located within a hydrogen-bonding distance to Asp222 that forms a strong ion pair with the ring nitrogen of the coenzyme, pyridoxal 5'-phosphate (PLP) or pyridoxamine 5'-phosphate (PMP). His143 of Escherichia coli AspAT was replaced by Ala or Asn. The mutant enzyme H143A showed a slight increase in the maximum velocity of the overall transamination reaction between aspartate and 2-oxoglutarate, while H143N AspAT showed a decrease to 60% in the maximum rate of the overall reactions in both directions. In all of the half-transamination reactions with four substrates, aspartate, glutamate, oxalacetate, and 2-oxoglutarate, the catalytic competence as defined by kmax/Kd decreased by 3-18-fold upon replacing His143 by either Ala or Asn. The extent of the decrease varied from one substrate to another; it was largely contributed to by the decrease in affinities for all substrates. The equilibrium constants, [PMP-form] [keto acid]/[( PLP-form] [amino acid]), decreased by over 10-fold upon the mutations at position 143. Both H143A and H143N AspATs exhibited a considerably decreased affinity for 2-methylaspartate, an external-aldimine-forming substrate analogue, yet without appreciable alteration in the affinity for succinate and glutarate, which are non-aldimine-forming analogues. All these findings suggest that, although His143 is not essential for catalysis, it might assist the formation of enzyme-substrate complex.  相似文献   

5.
Arg292 of E. coli aspartate aminotransferase was substituted with valine or leucine by site-directed mutagenesis. In comparison with the wild-type enzyme, either of the mutant enzymes showed a decrease by over 5 orders of magnitude of kcat/km values for aspartate and glutamate. This supports the contention that Arg292 is important for determining the specificity of this enzyme for dicarboxylic substrates. In contrast, mutant enzymes displayed a 5- to 10-fold increase in kcat/Km values for aromatic amino acids as substrates. Thus, introduction of an uncharged, hydrophobic side chain into position 292 leads to a striking alteration in substrate specificity of this enzyme, thereby improving catalytic efficiency toward aromatic amino acids.  相似文献   

6.
An Acinetobacter sp. genetic screen was used to probe structure-function relationships in vanillate demethylase, a two-component monooxygenase. Mutants with null, leaky, and heat-sensitive phenotypes were isolated. Missense mutations tended to be clustered in specific regions, most of which make known contributions to catalytic activity. The vanillate analogs m-anisate, m-toluate, and 4-hydroxy-3,5-dimethylbenzoate are substrates of the enzyme and weakly inhibit the metabolism of vanillate by wild-type Acinetobacter bacteria. PCR mutagenesis of vanAB, followed by selection for strains unable to metabolize vanillate, yielded mutant organisms in which vanillate metabolism is more strongly inhibited by the vanillate analogs. Thus, the procedure opens for investigation amino acid residues that may contribute to the binding of either vanillate or its chemical analogs to wild-type and mutant vanillate demethylases. Selection of phenotypic revertants following PCR mutagenesis gave an indication of the extent to which amino acid substitutions can be tolerated at specified positions. In some cases, only true reversion to the original amino acid was observed. In other examples, a range of amino acid substitutions was tolerated. In one instance, phenotypic reversion failed to produce a protein with the original wild-type sequence. In this example, constraints favoring certain nucleotide substitutions appear to be imposed at the DNA level.  相似文献   

7.
A procedure was developed for the selection of spontaneous mutants of Bacillus stearothermophilus NUB31 that are more efficient than the wild type in the restriction of phage at elevated temperatures. Inactivation studies revealed that two mutants contained a more thermostable restriction enzyme and one mutant contained three times more enzyme than the wild type. The restriction endonucleases from the wild type and one of the mutants were purified to apparent homogeneity. The mutant enzyme was more thermostable than the wild-type enzyme. The subunit molecular weight, amino acid composition, N-terminal and C-terminal amino acid residues, tryptic peptide map, and catalytic properties of the two enzymes were determined. The two enzymes have similar catalytic properties, but the molecular size of the mutant enzyme is approximately 6 to 7 kilodaltons larger than that of the wild-type enzyme. The mutant enzyme contains 54 additional amino acid residues, of which 26 to 28 are aspartate/asparagine, 8 to 15 are glutamate/glutamine, and 8 to 9 are tyrosine residues. The two enzymes contained similar amounts of the other amino acids, identical N-terminal residues, and different C-terminal residues. Tryptic peptide analyses revealed a high degree of homology between the two enzymes. The increased thermostability observed in the mutant enzyme appears to have been achieved by a mutation that resulted in the addition of amino acid residues to the wild-type enzyme. A number of mechanisms are discussed that could account for the observed difference between the mutant and wild-type enzymes.  相似文献   

8.
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   

9.
In earlier attempts to shift the substrate specificity of glutamate dehydrogenase (GDH) in favour of monocarboxylic amino-acid substrates, the active-site residues K89 and S380 were replaced by leucine and valine, respectively, which occupy corresponding positions in leucine dehydrogenase. In the GDH framework, however, the mutation S380V caused a steric clash. To avoid this, S380 has been replaced with alanine instead. The single mutant S380A and the combined double mutant K89L/S380A were satisfactorily overexpressed in soluble form and folded correctly as hexameric enzymes. Both were purified successfully by Remazol Red dye chromatography as routinely used for wild-type GDH. The S380A mutant shows much lower activity than wild-type GDH with glutamate. Activities towards monocarboxylic substrates were only marginally altered, and the pH profile of substrate specificity was not markedly altered. In the double mutant K89L/S380A, activity towards glutamate was undetectable. Activity towards L-methionine, L-norleucine and L-norvaline, however, was measurable at pH 7.0, 8.0 and 9.0, as for wild-type GDH. Ala163 is one of the residues that lines the binding pocket for the side chain of the amino-acid substrate. To explore its importance, the three mutants A163G, K89L/A163G and K89L/S380A/A163G were constructed. All three were abundantly overexpressed and showed chromatographic behaviour identical with that of wild-type GDH. With A163G, glutamate activity was lower at pH 7.0 and 8.0, but by contrast higher at pH 9.0 than with wild-type GDH. Activities towards five aliphatic amino acids were remarkably higher than those for the wild-type enzyme at pH 8.0 and 9.0. In addition, the mutant A163G used L-aspartate and L-leucine as substrates, neither of which gave any detectable activity with wild-type GDH. Compared with wild-type GDH, the A163 mutant showed lower catalytic efficiencies and higher K(m ) values for glutamate/2-oxoglutarate at pH 7.0, but a similar k(cat)/K(m) value and lower K(m) at pH 8.0, and a nearly 22-fold lower S(0.5) (substrate concentration giving half-saturation under conditions where Michaelis-Menten kinetics does not apply) at pH 9.0. Coupling the A163G mutation with the K89L mutation markedly enhanced activity (100-1000-fold) over that of the single mutant K89L towards monocarboxylic amino acids, especially L-norleucine and L-methionine. The triple mutant K89L/S380A/A163G retained a level of activity towards monocarboxylic amino acids similar to that of the double mutant K89L/A163G, but could no longer use glutamate as substrate. In terms of natural amino-acid substrates, the triple mutant represents effective conversion of a glutamate dehydrogenase into a methionine dehydrogenase. Kinetic parameters for the reductive amination reaction are also reported. At pH 7 the triple mutant and K89L/A163G show 5 to 10-fold increased catalytic efficiency, compared with K89L, towards the novel substrates. In the oxidative deamination reaction, it is not possible to estimate k(cat) and K(m) separately, but for reductive amination the additional mutations have no significant effect on k(cat) at pH 7, and the increase in catalytic efficiency is entirely attributable to the measured decrease in K(m). At pH 8 the enhancement of catalytic efficiency with the novel substrates was much more striking (e.g. for norleucine approximately 2000-fold compared with wild-type or the K89L mutant), but it was not established whether this is also exclusively due to more favourable Michaelis constants.  相似文献   

10.
We have investigated reactions of the 5-phosphonoethyl and 5-phosphonoethenyl analogs of pyridoxal 5'-phosphate in the coenzyme site of cytosolic aspartate aminotransferase. Acid dissociation constants and equilibrium constants for hydration and for tautomerization have been evaluated for these compounds. In confirmation of previous results, both compounds are partially active. They bind to apoenzyme well and undergo conversion in the presence of glutamate to amine forms which show induced circular dichroism comparable to that of native enzyme. A normal "external" Schiff base is evidently formed with 2-methylaspartate, but the amounts of quinonoid intermediate formed with erythro-3-hydroxyaspartate are less than those formed with pyridoxal phosphate. The pKa of the imine group of the enzyme reconstituted with the phosphonoethyl analog is more than two units lower than that in the native enzyme. Binding of the dicarboxylates glutarate, 2-oxoglutarate, and succinate shifts the pKa upward. The absorption spectra of the resulting complexes indicate the existence of at least three low pH species. A shift of 2.3 to 2.9 ppm to a lower frequency was observed for the 31P NMR signal upon binding of these dicarboxylates or of 2-methylaspartate. Enzyme containing the analogs crystallizes. Polarized absorption spectra suggest that the coenzyme has an orientation similar to that of pyridoxal phosphate in the native enzyme.  相似文献   

11.
Nitrilase from Rhodococcus rhodochrous ATCC 33278 hydrolyses both aliphatic and aromatic nitriles. Replacing Tyr-142 in the wild-type enzyme with the aromatic amino acid phenylalanine did not alter specificity for either substrate. However, the mutants containing non-polar aliphatic amino acids (alanine, valine and leucine) at position 142 were specific only for aromatic substrates such as benzonitrile, m-tolunitrile and 2-cyanopyridine, and not for aliphatic substrates. These results suggest that the hydrolysis of substrates probably involves the conjugated pi-electron system of the aromatic ring of substrate or Tyr-142 as an electron acceptor. Moreover, the mutants containing charged amino acids such as aspartate, glutamate, arginine and asparagine at position 142 displayed no activity towards any nitrile, possibly owing to the disruption of hydrophobic interactions with substrates. Thus aromaticity of substrate or amino acid at position 142 in R. rhodochrous nitrilase is required for enzyme activity.  相似文献   

12.
Since crystallographic studies on Escherichia coli aspartate transcarbamoylase (ATCase) indicate that Gln 231 is in the active site of the enzyme and participates in the binding of the substrate, aspartate, it seemed of interest to examine mutant enzymes in which Gln 231 was replaced by Asn or Ile. The two mutant forms containing amino acid substitutions were characterized by a combination of steady-state kinetics, hydrodynamic measurements, and equilibrium ligand binding techniques. Both mutant forms exhibited a dramatic reduction in the affinity of the protein for substrates and substrate analogues as well as a very large decrease in catalytic activity. Moreover, the amino acid substitutions introduced within the active site of the enzyme led to unusual allosteric properties in the mutant enzymes. Although the bisubstrate analogue N-(phosphonoacetyl)-L-aspartate promotes the characteristic global conformational change in the mutant forms that is observed with the wild-type enzyme, the combination of substrate and substrate analogue does not. Cooperativity with respect to substrate binding is largely reduced compared to wild-type ATCase. Also, the effector molecules ATP and CTP which bind to the regulatory chains have dramatic effects on the activity of the mutant enzymes containing replacements for Gln 231 in the catalytic chains. In stark contrast to the wild-type enzyme, in which effects of nucleotides are manifested primarily by changes in the K0.5 of the enzyme, ATP and CTP have large effects on the Vmax of the mutant enzymes.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
Glu-50 of aspartate transcarbamoylase from Escherichia coli forms a set of interdomain bridging interactions between the 2 domains of the catalytic chain; these interactions are critical for stabilization of the high-activity high-affinity form of the enzyme. The mutant enzyme with an alanine substituted for Glu-50 (Glu-50-->Ala) exhibits significantly reduced activity, little cooperativity, and altered regulatory behavior (Newton CJ, Kantrowitz ER, 1990, Biochemistry 29:1444-1451). A study of the structural consequences of replacing Glu-50 by alanine using solution X-ray scattering is reported here. Correspondingly, in the absence of substrates, the mutant enzyme is in the same, so-called T quaternary conformation as is the wild-type enzyme. In the presence of a saturating concentration of the bisubstrate analog N-phosphonacetyl-L-aspartate (PALA), the mutant enzyme is in the same, so-called R quaternary conformation as the wild-type enzyme. However, the Glu-50-->Ala enzyme differs from the wild-type enzyme, in that its scattering pattern is hardly altered by a combination of carbamoyl phosphate and succinate. Addition of ATP under these conditions does result in a slight shift toward the R structure. Steady-state kinetic studies indicate that, in contrast to the wild-type enzyme, the Glu-50-->Ala enzyme is activated by PALA at saturating concentrations of carbamoyl phosphate and aspartate, and that PALA increases the affinity of the mutant enzyme for aspartate. These data suggest that the enzyme does not undergo the normal T to R transition upon binding of the physiological substrates and verifies the previous suggestion that the interdomain bridging interactions involving Glu-50 are critical for the creation of the high-activity, high-affinity R state of the enzyme.  相似文献   

14.
Sialin, the protein coded by SLC17A5, is responsible for membrane potential (Δψ)-driven aspartate and glutamate transport into synaptic vesicles in addition to H+/sialic acid co-transport in lysosomes. Rodent sialin mutants harboring the mutations associated with Salla disease in humans did not transport aspartate and glutamate whereas H+/sialic acid co-transport activity was about one-third of the wild-type protein. In this study, we investigate the effects of various mutations on the transport activities of human sialin. Proteoliposomes containing purified heterologously expressed human sialin exhibited both Δψ-driven aspartate and glutamate transport activity and H+/sialic acid co-transport activity. Aspartate and glutamate transport was not detected in the R39C and K136E mutant forms of SLC17A5 protein associated with Salla disease, whereas H+/sialic acid co-transport activity corresponded to 30-50% of the recombinant wild-type protein. In contrast, SLC17A5 protein harboring the mutations associated with infantile sialic acid storage disease, H183R and Δ268SSLRN272 still showed normal levels of Δψ-driven aspartate and glutamate transport even though H+/sialic acid co-transport activity was absent. Human sialin carrying the G328E mutation that causes both phenotypes, and P334R and G378V mutations that cause infantile sialic acid storage disease showed no transport activity. These results support the idea that people suffering from Salla disease have been defective in aspartergic and glutamatergic neurotransmissions.  相似文献   

15.
Amino acids are key carbon and energy sources for the asaccharolytic food-borne human pathogen Campylobacter jejuni . During microaerobic growth in amino acid rich complex media, aspartate, glutamate, proline and serine are the only amino acids significantly utilized by strain NCTC 11168. The catabolism of aspartate and glutamate was investigated. An aspartase ( aspA ) mutant (unable to utilize any amino acid except serine) and a Cj0762 c ( aspB ) mutant lacking aspartate:glutamate aminotransferase (unable to utilize glutamate), were severely growth impaired in complex media, and an aspA sdaA mutant (also lacking serine dehydratase) failed to grow in complex media unless supplemented with pyruvate and fumarate. Aspartase was shown by activity and proteomic analyses to be upregulated by oxygen limitation, and aspartate enhanced oxygen-limited growth of C. jejuni in an aspA -dependent manner. Stoichiometric aspartate uptake and succinate excretion involving the redundant DcuA and DcuB transporters indicated that in addition to a catabolic role, AspA can provide fumarate for respiration. Significantly, an aspA mutant of C. jejuni 81-176 was impaired in its ability to persist in the intestines of outbred chickens relative to the parent strain. Together, our data highlight the dual function of aspartase in C. jejuni and suggest a role during growth in the avian gut.  相似文献   

16.
Comparison of the farnesyl diphosphate (FPP) synthase amino acid sequences from four species with amino acid sequences from the related enzymes hexaprenyl diphosphate synthase and geranylgeranyl diphosphate synthase show the presence of two aspartate rich highly conserved domains. The aspartate motif ((I, L, or V)XDDXXD) of the second of those domains has homology with at least 9 prenyl transfer enzymes that utilize an allylic prenyl diphosphate as one substrate. In order to investigate the role of this second aspartate-rich domain in rat FPP synthase, we mutated the first or third aspartate to glutamate, expressed the wild-type and mutant enzymes in Escherichia coli, and purified them to apparent homogeneity using a single chromatographic step. Approximately 12 mg of homogeneous protein was isolated from 120 mg of crude bacterial extract. The kinetic parameters of the purified wild-type recombinant FPP synthase containing the DDYLD motif were as follows: Vmax = 0.84 mumol/min/mg; GPP Km = 1.0 microM; isopentenyl diphosphate (IPP) Km = 2.7 microM. Substitution of glutamate for the first aspartate (EDYLD) decreased the Vmax by over 90-fold. The Km for IPP increased, whereas the Km for GPP remained the same in this D243E mutant. Substitution of glutamate for the third aspartate (DDYLE) did not result in altered enzyme kinetics in the D247E mutant. These results suggest that the first aspartate in the second domain is involved in the catalysis by FPP synthase.  相似文献   

17.
The reaction of Escherichia coli aspartate aminotransferase (AspAT) with glutamate and other C5-dicarboxylates was analyzed in order to compare its mechanism of action toward C5 substrates with that toward C4 substrates, which had been extensively characterized. The association of the amino-group protonated and unprotonated forms of glutamate (SH(+) and S, respectively) with the Schiff-base protonated and unprotonated forms of the enzyme (E(L)H(+) and E(L), respectively) yields at least three forms of the Michaelis complex, whereas in the case of aspartate, only two species of this complex exist, E(L).SH(+) and E(L)H(+).S. The reaction of AspAT with 2-methylglutamate can be explained only when we consider all the protonation states of the Michaelis complex. Based on the previous crystallographic studies [Miyahara et al. (1994) J. Biochem. 116, 1001-1012], we consider that glutamate binds to the open form of AspAT and takes an extended conformation in the Michaelis complex, with the alpha-amino group of glutamate oriented in the opposite direction to the Schiff base. This is in contrast to the Michaelis complex of aspartate, in which a strong interaction of the alpha-amino group of aspartate and the Schiff base excludes the presence of the species E(L)H(+).SH(+). It is concluded that AspAT recognizes the two types of dicarboxylates with different chain lengths by changing the gross conformation of the enzyme protein.  相似文献   

18.
The type I regulatory subunit (R-I) of rat brain cAMP-dependent protein kinase was expressed in E. coli and site-directed mutagenesis was used to substitute amino acids in the putative cAMP-binding sites. The wild-type recombinant R-I bound 2 mol of cAMP/mol subunit, while two mutant R-Is with a single amino acid substitution in one of the two intrachain cAMP-binding sites (clone N153:a glutamate for Gly-200, and clone C254:an aspartate for Gly-324) bound 1 mol of cAMP/mol subunit. When these two substitutions were made in one mutant, cAMP did not bind to this mutant, indicating that binding of cAMP to N153 or C254 was to their nonmutated sites. Competition experiments with site-selective analogs and dissociation of bound cAMP from mutant R-Is provided evidence for strong intrachain interactions between the two classes of cAMP-binding sites in R-I.  相似文献   

19.
A mutant (A204) of Chinese hamster ovary cells (CHO-K1), which is deficient in dihydroorotate (DHO) dehydrogenase (E.C. 1.3,3.1) activity, has been isolated by a replica plating procedure. The mutant does not show a requirement for exogenously added pyrimidines. Examination of intact cells shows that the mutant accumulates a large amount of carbamyl aspartate and is markedly but not totally deficient in biosynthesis of orotate from earlier precursors of pyrimidine biosynthesis, including aspartate and dihydroorotic acid, when compared to wild-type cells. Analysis of cell-free extracts of mutant and wild-type cells shows that the mutant is deficient in DHO dehydrogenase activity, possessing ca. 5% of the wild-type activity. this evidence leads to the conclusion that this mutant, A204, is in fact partially deficient in DHO dehydrogenase, and that in these cells it is this enzyme which carries out the fourth step of de novo pyrimidine biosynthesis.  相似文献   

20.
The rate of transport of L-amino acids by Saccharomyces cerevisiae epsilon 1278b increased with time in response to nitrogen starvation. This increase could be prevented by the addition of ammonium sulfate or cycloheximide. A slow time-dependent loss of transport activity was observed when ammonium sulfate (or ammonium sulfate plus cycloheximide) was added to cells after 3 h of nitrogen starvation. This loss of activity was not observed in the presence of cycloheximide alone. In a mutant yeast strain which lacks the nicotinamide adenine dinucleotide phosphate-dependent (anabolic) glutamate dehydrogenase, no significant decrease in amino acid transport was observed when ammonium sulfate was added to nitrogen-starved cells. A double mutant, which lacks the nicotinamide adenine dinucleotide phosphate-dependent enzyme and in addition has a depressed level of the nicotinamide adenine dinucleotide-dependent (catabolic) glutamate dehydrogenase, shows the same sensitivity to ammonium ion as the wild-type strain. These data suggest that the inhibition of amino acid transport by ammonium ion results from the uptake of this metabolite into the cell and its subsequent incorporation into the alpha-amino groups of glutamate and other amino acids.  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