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.
2.
3.
Of the major amino acid side chains that anchor pyridoxal 5'-phosphate at the coenzyme binding site of bacterial D-amino acid transaminase, two have been substituted using site-directed mutagenesis. Thus, Ser-180 was changed to an Ala (S180A) with little effect on enzyme activity, but replacement of Tyr-31 by Gln (Y31Q) led to 99% loss of activity. Titration of SH groups of the native Y31Q enzyme with DTNB proceeded much faster and to a greater extent than the corresponding titration for the native wild-type and S180A mutant enzymes. The stability of each mutant to denaturing agents such as urea or guanidine was similar, i.e., in their PLP forms, S180A and Y31Q lost 50% of their activities at a 5-15% lower concentration of urea or guanidine than did the wild-type enzyme. Upon removal of denaturing agent, significant activity was restored in the absence of added pyridoxal 5'-phosphate, but addition of thiols was required. In spite of its low activity, Y31Q was able to form the PMP form of the enzyme just as readily as the wild-type and the S180A enzymes in the presence of normal D-amino acid substrates. However, beta-chloro-D-alanine was a much better substrate and inactivator of the Y31Q enzyme than it was for the wild-type or S180A enzymes, most likely because the Y31Q mutant formed the pyridoxamine 5-phosphate form more rapidly than the other two enzymes. The stereochemical fidelity of the Y31Q recombinant mutant enzyme was much less than that of the S180A and wild-type enzymes because racemase activity, i.e., conversion of L-alanine to D-alanine, was higher than for the wild-type or S180A mutant enzymes, perhaps because the coenzyme has more flexibility in this mutant enzyme.  相似文献   

4.
The mutant IP7 of Escherichia coli B requires isoleucine or pyridoxine for growth as a consequence of a mutation in the gene coding for biosynthetic threonine deaminase. The mutation of IP7 was shown to be of the nonsense type by the following data: (1) reversion to isoleucine prototrophy involves the formation of external suppression at a high frequency, as shown by transduction experiments; and (ii) the isoleucine requirement is suppressed by lysogenization with a phage carrying the amber suppressor su-3. Cell extracts of the mutant strain contain a low activity of threonine deaminase. The possibility that this activity is biodegradative was ruled out by kinetic experiments. The mutant threonine deaminase was purified to homogeneity by conventional procedures. The enzyme is a dimer of identical subunits of an approximate molecular weight of 43,000 (Grimminger and Feldner, 1974), whereas the wild-type enzyme is a tetramer of 50,000-dalton subunits (Calhoun et al., 1973; Grimminger et al., 1973). The mutant enzyme is not inhibited by isoleucine and does not bind isoleucine, as shown by equilibrium dialysis experiments. Pyridoxal phosphate enhances the maximum catalytic activity of the mutant enzyme by a factor of five, whereas the wild-type enzyme is not affected. In wild-type and mutant threonine deaminase the ratio of protein subunits and bound pyridoxal phosphate is 2:1. The activation of threonine deaminase from strain IP7 is due to a second coenzyme binding site, as shown by (i) spectrophotometric titration of the enzyme with pyridoxal phosphate and by (ii) measurement the pyridoxal phosphate content of the enzyme after sodium borohydride reduction of the protein. The observation of one pyridoxal phosphate binding site per peptide dimer in the wild-type enzyme and of two binding sites per dimer in the mutant strongly suggests that one of the potential sites in the wild-type enzyme is masked by allosteric effects. The factors responsible for the half-of-the-sites reactivity of the coenzyme sites appear to be nonoperative in the mutant protein.  相似文献   

5.
6.
Lysyl-transfer ribonucleic acid (tRNA) synthetase activity was compared in three independently isolated Escherichia coli K-12 mutants of the enzyme S-adenosyl-L-methionine synthetase (metK mutants) and their isogenic parents. In all three cases the activity of the lysyl-tRNA synthetase was elevated two- to fourfold in the mutant strains. Glycyl-L-leucine (3 mM) usually enhanced lysyl-tRNA synthetase activity two- to threefold in wild-type cells but did not further stimulate the synthetase activity in metK mutants. By two other criteria, the lysyl-tRNA synthetase from wild-type cells grown with the peptide and from the metK mutant RG62, grown in minimal medium, were similar. These criteria are enhanced resistance to thermal inactivation and altered susceptibility to endogenous proteases when compared with the synthetase from wild-type cells grown in minimal medium. In a separate set of experiments, the activities of the lysyl-, arginyl-, seryl-, and valyl-tRNA synthetases were measured in an isogenic pair of relt and rel strains of E. coli grown in a relatively poor growth medium (acetate) and in enriched medium. In the rel+ strain the level of all four synthetases was higher (two- to fourfold) in the enriched medium as expected. In the rel strain the difference in the activities of the synthetases between the two media were diminished. In all four cases the activities of the synthetases were higher in acetate medium in the rel strain. Evidence is presented that these two modes of metabolic regulation act independently.  相似文献   

7.
Salmonella enterica serovar Typhimurium LT2 showed increased sensitivity to propionate when the 2-methylcitric acid cycle was blocked. A derivative of a prpC mutant (which lacked 2-methylcitrate synthase activity) resistant to propionate was isolated, and the mutation responsible for the newly acquired resistance to propionate was mapped to the citrate synthase (gltA) gene. These results suggested that citrate synthase activity was the source of the increased sensitivity to propionate observed in the absence of the 2-methylcitric acid cycle. DNA sequencing of the wild-type and mutant gltA alleles revealed that the ATG start codon of the wild-type gene was converted to the rare GTG start codon in the revertant strain. This result suggested that lower levels of this enzyme were present in the mutant. Consistent with this change, cell-free extracts of the propionate-resistant strain contained 12-fold less citrate synthase activity. This was interpreted to mean that, in the wild-type strain, high levels of citrate synthase activity were the source of a toxic metabolite. In vitro experiments performed with homogeneous citrate synthase enzyme indicated that this enzyme was capable of synthesizing 2-methylcitrate from propionyl-CoA and oxaloacetate. This result lent further support to the in vivo data, which suggested that citrate synthase was the source of a toxic metabolite.  相似文献   

8.
A new acetate-requiring mutant strain of Neurospora crassa, ace-9, has been isolated. The mutant gene was mapped between nuc-2 and arg-12 on the right arm of the second linkage group. The ace-9 mutant strain shows very weak activity of pyruvate dehydrogenase complex (PDHC). Three strains that show no activity of PDHC had already been found, i.e., ace-2, ace-3, and ace-4. Thus the ace-9 is the fourth gene that causes the deficiency in PDHC activity by a mutation. Deficiency of PDHC activity in ace-9 strain seems to be due to defective E1 component, because (1) the activity of E1 component enzyme is very weak in ace-9 mutant strain, and (2) normal PDHC activity was resumed when a preparation of ace-9 was mixed with E1-E2 fraction of wild type or with E1 component of wild type E. coli. Difference in thermostability of both E1 component enzyme and PDHC between ace-9 and the wild type strains supports this conclusion.  相似文献   

9.
The obligate intracellular bacterium Rickettsia prowazekii has recently been shown to transport the essential metabolite S-adenosylmethionine (SAM). The existence of such a transporter would suggest that the metK gene, coding for the enzyme that synthesizes SAM, is unnecessary for rickettsial growth. Genome sequencing has revealed that this is the case for the metK genes of the spotted fever group and the Madrid E strain of R. prowazekii, which contain recognizable inactivating mutations. However, several strains of the typhus group rickettsiae possess metK genes lacking obvious mutations. In order to determine if these genes code for a product that retains MAT function, an Escherichia coli metK deletion mutant was constructed in which individual rickettsial metK genes were tested for the ability to complement the methionine adenosyltransferase deficiency. Both the R. prowazekii Breinl and R. typhi Wilmington metK genes complemented at a level comparable to that of an E. coli metK control, demonstrating that the typhus group rickettsiae have the capability of synthesizing as well as transporting SAM. However, the appearance of mutations that affect the function of the metK gene products (a stop codon in the Madrid E strain and a 6-bp deletion in the Breinl strain) provides experimental support for the hypothesis that these typhus group genes, like the more degenerate spotted fever group orthologs, are in the process of gene degradation.  相似文献   

10.
The temperature-sensitive Bacillus subtilis tms-26 mutant strain was characterized biochemically and shown to be defective in N-acetylglucosamine 1-phosphate uridyltransferase activity. At the permissive temperature (34 degrees C), the mutant strain contained about 15% of the wild-type activity of this enzyme, whereas at the nonpermissive temperature (48 degrees C), the mutant enzyme was barely detectable. Furthermore, the N-acetylglucosamine 1-phosphate uridyltransferase activity of the tms-26 mutant strain was much more heat labile in vitro than that of the wild-type strain. The level of N-acetylglucosamine 1-phosphate, the substrate of the uridyltransferase activity, was elevated more than 40-fold in the mutant strain at the permissive temperature compared with the level in the wild-type strain. During a temperature shift, the level of UDP-N-acetylglucosamine, the product of the uridyltransferase activity, decreased much more in the mutant strain than in the wild-type strain. An Escherichia coli strain harboring the wild-type version of the tms-26 allele on a plasmid contained increased N-acetylglucosamine 1-phosphate uridyltransferase activity compared with that in the haploid strain. It is suggested that the gene for N-acetylglucosamine 1-phosphate uridyltransferase in B. subtilis be designated gcaD.  相似文献   

11.
The role of collagenase as a mechanism of bacterial pathogenicity in some pathogenic bacteria has been reported. The information on the role of collagenase in Aeromonas spp. pathogenesis is scant. In the present study, a mutant Aeromonas veronii RY001 that is deficient in the putative collagenase gene acg was constructed and compared with the wild-type strain for virulence factors. Bacterial cells and cell-free extracellular products of the mutant had significantly less collagenolytic activity, but there were not significant differences in caseinolytic, gelatinolytic, and elastolytic activities. Adhesion and invasion abilities of the mutant strain on epithelioma papillosum of carp cells was only 56% of that of the wild-type strain, and the cytotoxicity of the mutant strain to epithelioma papillosum of carp cells was only 42% of that of the wild-type strain. The LD50 values of the wild-type strain were determined as 1.6 x 10(6) and 3.5 x 10(5) cfu in goldfish and mice, respectively, whereas the mutant RY001 strain showed slightly higher values (i.e., 2.8 x 10(6) and 1.4 x 10(6) cfu in goldfish and mice, respectively). These results indicated the involvement of the collagenase gene in the pathogenesis of A. veronii.  相似文献   

12.
J A Kreps  T Ponappa  W Dong    C D Town 《Plant physiology》1996,110(4):1159-1165
A mutant of Arabidopsis thaliana, amt-1, was previously selected for resistance to growth inhibition by the tryptophan analog alpha-methyltryptophan. This mutant had elevated tryptophan levels and exhibited higher anthranilate synthase (AS) activity that showed increased resistance to feedback inhibition by tryptophan. In this study, extracts of the mutant callus exhibited higher AS activity than wild-type callus when assayed with either glutamine or ammonium sulfate as amino donor, thus suggesting that elevated AS activity in the mutant was due to an alteration in the alpha subunit of the enzyme. The mutant also showed cross-resistance to 5-methylanthranilate and 6-methylanthranilate and mapped to chromosome V at or close to ASA1 (a gene encoding the AS alpha subunit). ASA1 mRNA and protein levels were similar in mutant and wild-type leaf extracts. Levels of ASA1 mRNA and protein were also similar in callus cultures of mutant and wild type, although the levels in callus were higher than in leaf tissue. Sequencing of the ASA1 gene from amt-1 revealed a G to A transition relative to the wild-type gene that would result in the substitution of an asparagine residue in place of aspartic acid at position 341 in the predicted amino acid sequence of the ASA1 protein. The mutant allele in strain amt-1 has been renamed trp5-1.  相似文献   

13.
We have purified DNA polymerase alpha from a temperature-sensitive mutant cell line of mouse FM3A cells, tsFT20, that shows temperature-sensitive activity of DNA polymerase alpha (Murakami, Y., Yasuda, H., Miyazawa, H., Hanaoka, F., and Yamada, M. (1985) Proc. Natl. Acad. Sci. U.S.A. 82, 1761-1765). The purified enzyme was composed of two polypeptides with the same apparent molecular weights as those of purified DNA polymerase alpha from the parental strain, FM3A (Mr 180,000 and 68,000). Heat inactivation experiments revealed that this purified DNA polymerase alpha from tsFT20 cells was more heat-labile than the wild-type enzyme. We have also purified primase from both ts-FT20 cells and wild-type cells. Both primase fractions consist of two polypeptides with molecular weights of 54,000 and 46,000. No difference was observed between the heat labilities of the primases from tsFT20 cells and wild-type cells. Comparisons of wild-type and mutant polymerase indicated that the temperature-sensitive mutation in DNA polymerase alpha from tsFT20 cells affect the dCTP-binding site of the enzyme. The mutation also changed the optimum pH and the optimum KCl concentration of the enzyme.  相似文献   

14.
The enzyme serine transhydroxymethylase (EC 2.1.2.1; L-serine:tetrahydrofolate-5,10-hydroxymethyltransferase) is responsible both for the synthesis of glycine from serine and production of the 5,10-methylenetetrahydrofolate necessary as a methyl donor for methionine synthesis. Two mutants selected for alteration in serine transhydroxymethylase regulation also have phenotypes characteristic of metK (methionine regulatory) mutants, including ethionine, norleucine, and alpha-methylmethionine resistance and reduced levels of S-adenosylmethionine synthetase (EC 2.5.1.6; adenosine 5'-triphosphate:L-methionine S-adenosyltransferase) activity. Because this suggested the existence of a common regulatory component, the regulation of serine transhydroxymethylase was examined in other methionine regulatory mutants (metK and metJ mutants). Normally, serine transhydroxymethylase levels are repressed three- to sixfold in cells grown in the presence of serine, glycine, methionine, adenine, guanine, and thymine. This does not occur in metK and metJ mutants; thus, these mutations do affect the regulation of both serine transhydroxymethylase and the methionine biosynthetic enzymes. Lesions in the metK gene have been reported to reduce S-adenosylmethionine synthetase levels. To determine whether the metK gene actually encodes for S-adenosylmethionine synthetase, a mutant was characterized in which this enzyme has a 26-fold increased apparent Km for methionine. This mutation causes a phenotype associated with metK mutants and is cotransducible with the serA locus at the same frequency as metK lesions. Thus, the affect of metK mutations on the regulation of glycine and methionine synthesis in Salmonella typhimurium appears to be due to either an altered S-adenosylmethionine synthetase or altered S-adenosylmethionine pools.  相似文献   

15.
From an Escherichia coli purine auxotroph a mutant defective in phosphoribosylpyrophosphate (PRib-PP) synthetase has been isolated and partially characterized. In contrast to the parental strain, the mutant was able to grow on nucleosides as purine source, whereas growth on purine bases was reduced. Kinetic analysis of the mutant PRib-PP synthetase revealed an apparent Km for ATP and ribose 5-phosphate of 1.0 mM and 240 muM respectively, compared to 60 muM and 45 muM respectively for the wild-type enzyme. ADP, which inhibits the wild-type enzyme at a concentration of 0.5 mM ribose 5-phosphate, stimulated the mutant enzyme. The activity of PRib-PP synthetase in crude extract was higher in the mutant than in the parent. When starved for purines an accumulation of PRib-PP was observed in the parent strain, while the pool decreased in the mutant. During pyrimidine starvation derepression of PRib-PP synthetase activity was observed in both strains, although to a lesser extent in the mutant. Our data suggest that the mutant harbors a mutation in the structural gene for PRib-PP synthetase. The mutation responsible for the altered PRib-PP synthetase was located in the purB-hemA region at 26 min on the recalibrated linkage map.  相似文献   

16.
Most prototrophic strains of Escherichia coli become restricted for methionine at 44 degrees C. A mutant strain (RG62 metK) in which the level of S-adenosylmethionine synthetase activity is only 10 to 20% of normal shows constitutive expression of one of the heat shock proteins, the lysU gene product, lysyl-tRNA synthetase form II, at 37 degrees C. These findings suggested a possible linkage between methionine metabolism and heat shock. We examined the induction of heat shock polypeptides in strain RG62 (metK) and in its parent, RG (metK+), from which it was derived by spontaneous mutation. Exponential-phase cultures of the two strains were pulse-labeled with [3H]leucine shortly after a shift from 37 to 44 degrees C, and the total cellular polypeptides were examined by two-dimensional electrophoresis. The results confirmed the constitutive production of the lysU gene product previously reported for strain RG62, but also revealed that the induction of 2 of the 17 heat shock polypeptides, C14.7 and G13.5, was markedly depressed. Otherwise the heat shock induction pattern was similar in timing and magnitude in the two strains. Transformation of the mutant strain with a plasmid, pK8, containing the metK coding sequence and promoter region as a 1.8-kilobase insert into pBR322 restored normal induction of C14.7 and G13.5, but did not prevent constitutive expression of the lysU gene product in the medium required for growth of this strain. The three heat shock polypeptides abnormally controlled in strain RG62 are the three polypeptides which are not induced when rapid synthesis of the htpR gene product is induced by isopropyl-beta-D-thiogalactopyranoside at 28 degree C (R. A. VanBogelen, M. A. Acton, and F. C. Neidhardt, Genes Dev. 1:525-531, 1987). We postulate that induction of these three polypeptides involves metabolic signals in addition to the synthesis of the htpR gene product and that strain RG62 (metK) fails to produce the signals involved in induction of C14.7 and G13.5 on a shift-up in temperature and produces the signal related to lysU induction even at 37 degree C.  相似文献   

17.
A mutated version of the hygromycin B phosphotransferase (hph(mut)) gene from Escherichia coli, isolated by directed evolution at 75 degrees C in transformants of a thermophilic strain of Sulfolobus solfataricus, was characterized with respect to its genetic stability in both the original mesophilic and the new thermophilic hosts. This gene was demonstrated to be able to express the hygromycin B resistance phenotype and to be steadily maintained and propagated also in other, more thermophilic strains of S. solfataricus, i.e., up to 82 degrees C. Furthermore, it may be transferred to S. solfataricus cells by cotransformation with pKMSD48, another extrachromosomal element derived from the virus SSV1 of Sulfolobus shibatae, without any loss of stability and without affecting the replication and infectivity of this viral DNA. The hph(mut) and the wild-type gene products were expressed at higher levels in E. coli and purified by specific affinity chromatography on immobilized hygromycin B. Comparative characterization revealed that the mutant enzyme had acquired significant thermoresistance and displayed higher thermal activity with augmented catalytic efficiency.  相似文献   

18.
A mutant strain of Escherichia coli K-12, designated 618, accumulates glycogen at a faster rate than wild-type strain 356. The mutation affects the ADPglucose pyrophosphorylase regulatory properties (N. Creuzat-Sigal, M. Latil-Damotte, J. Cattaneo, and J. Puig, p. 647-680, in R. Piras and H. G. Pontis, ed., Biochemistry of the Glycocide Linkage, 1972). The enzyme is less dependent on the activator, fructose 1,6 bis-phosphate for activity and is less sensitive to inhibition by the inhibitor, 5'-AMP. The structural gene, glgC, for this allosteric mutant enzyme was cloned into the bacterial plasmid pBR322 by inserting the chromosomal DNA at the PstI site. The glycogen biosynthetic genes were selected by cotransformation of the neighboring asd gene into an E. coli mutant also defective in branching enzyme (glgB) activity. Two recombinant plasmids, pEBL1 and pEBL3, that had PstI chromosomal DNA inserts containing glgC and glgB were isolated. Branching enzyme and ADPglucose pyrophosphorylase activities were increased 240- and 40-fold, respectively, in the asd glgB mutant, E. coli K-12 6281. The E. coli K-12 618 mutant glgC gene product was characterized after transformation of an E. coli B ADPglucose pyrophosphorylase mutant with the recombinant plasmid pEBL3. The kinetic properties of the cloned ADPglucose pyrophosphorylase were similar to those of the E. coli K-12 618 enzyme. The inserted DNA in pEBL1 was arranged in opposite orientation to that in pEBL3.  相似文献   

19.
1. A Km mutant of Saccharomyces cerevisiae with a lesion in CDPdiacylglycerol-inositol 3-phosphatidyltransferase was isolated. The mutant required a high concentration of myo-inositol for growth. 2. The CDPdiacylglycerol-inositol 3-phosphatidyltransferase in the mutant cells showed an apparent Km for myo-inositol over 200-times higher than that of the enzyme in wild-type cells. The maximum velocity of the mutant enzyme was comparable to that of the wild-type enzyme. 3. In mutant cells, labelled myo-inositol, phosphate and acetate were incorporated into phosphatidylinositol at much slower rates than in wild-type cells. The phosphatidylinositol content of mutant cells was markedly lower than that observed in wild-type cells. 4. Genetic analysis showed that the growth phenotype of the mutant arose from a single nuclear gene mutation in a gene coding for CDPdiacylglycerol-inositol 3-phosphatidyltransferase. 5. The mutant showed a normal level of phosphatidylserine synthase activity. The phosphatidylserine synthase gene was located between ura3 and hom3 on chromosome V, whereas the CDPdiacylglycerol-inositol 3-phosphatidyltransferase gene showed no linkage with ura3. 6. Labelled acetate was incorporated into various lipids including triacylglycerols, diacylglycerols, sterol esters and phospholipids other than phosphatidylinositol at faster rates in mutant cells than in wild-type cells. Incorporation into both the fatty acid and the sterol moieties was facilitated in the mutant. 7. A striking change in the cell-division process was observed when phosphatidylinositol synthesis was limited. The results showed that phosphatidylinositol synthesis is involved in the cell-division cycle of yeast.  相似文献   

20.
Gene-Enzyme Relationships in Neurospora Invertase   总被引:8,自引:4,他引:4       下载免费PDF全文
A spontaneous, single-gene mutation responsible for a total lack of invertase activity in Neurospora crassa is described. The mutation is believed to lie in the structural gene for invertase, since an immunologically cross-reacting protein is made by the mutant strain. In addition, there was no evidence for a defect in regulation of invertase activity or synthesis by the following criteria. (i) The invertaseless condition was recessive in heterokaryons; (ii) no invertase inhibitor was found in mutant extracts by mixing experiments; and (iii) none of the several sugars able to induce activity in wild-type strains was able to induce activity in the mutant strain. It was also discovered that most of the wild-type enzyme (55 to 75%) cannot be washed free from the rapidly sedimenting cell debris. This finding provided additional support for the hypothesis that Neurospora invertase is located within or about the cell wall.  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