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 46 毫秒
1.
alpha-Aminoadipate-semialdehyde dehydrogenase catalyzes the conversion of alpha-aminoadipate to alpha-aminoadipate-semialdehyde in the biosynthetic pathway of lysine in yeasts and molds. Mutants belonging to lys2 and lys5 loci of Saccharomyces cerevisiae lacked the alpha-aminoadipate-semialdehyde dehydrogenase activity. Complementation in vitro was demonstrated by combining the extracts from different lys2 and lys5 mutants. Some of the revertants of lys2 and lys5 mutants exhibited lower specific activity and higher thermolability of alpha-aminoadipate-semialdehyde dehydrogenase than the enzyme from wild-type cells. The enzyme was partially purified from wild-type cells and the molecular weight of the enzyme was estimated on a Sephacryl S-300 column at 180,000. Results from the revertant analysis and in vitro complementation indicated LYS2 and LYS5 as structural genes, each encoding a subunit of this large enzyme.  相似文献   

2.
Abstract We isolated from Candida maltosa mutants lacking saccharopine reductase ( lys9 ) and saccharopine dehydrogenase ( lys1 ). They accumulated α-aminoadipate-δ-semialdehyde (AASA) in the cell and excreted it into the culture medium. In the presence of 15 g glucose/l, 1.25 g NH4H2PO4/l and 50 mg l -lysine/l in a minimal salt medium C. maltosa G285 ( lys1 ) produced about 80–90 mg AASA/l within 48 h. It is the first report of lysine-requiring yeast mutants that accumulate and excrete AASA. In contrast, Pichia guilliermondii lys9 mutants lacked this AASA overproduction. The AASA accumulation by C. maltosa mutants may be explained by the low feedback regulation of their homocitrate synthase and the equilibrium of the enzyme reactions involved in the lysine biosynthesis.  相似文献   

3.
The alpha-aminoadipate (AA) pathway for the biosynthesis of lysine was investigated in the wild type and in lysine auxotrophs of the fission yeast Schizosaccharomyces pombe. Of the eight enzyme activities of the AA pathway that have been examined so far, six were present in the extract of wild-type S. pombe cells. Growth response to AA and accumulation studies indicated that three lysine auxotrophs, the lys2-97, lys4-95, and lys8-1 strains, were blocked before the AA step and that four lysine auxotrophs, the lys1-131, lys3-37, lys6-3, and lys7-2 strains, were blocked after the AA step. Among the mutants investigated, the lys2-97 mutant exhibited an enzyme lesion at the cis-homoaconitate hydratase step, the lys1-131 and lys7-2 mutants exhibited lesions at the AA reductase step, and lys3-37 exhibited a lesion at the saccharopine dehydrogenase step. These results demonstrated the basic similarity of the AA pathway in S. pombe and Saccharomyces cerevisiae.  相似文献   

4.
Three lysine auxotrophs, strains AU363, 7305d, and 8201-7A, were investigated genetically and biochemically to determine their gene loci, biochemical lesions, and roles in the lysine biosynthesis of Saccharomyces cerevisiae. These mutants were leaky and blocked after the alpha-aminoadipate step. Complementation studies placed these three mutations into a single, new complementation group, lys14. Tetrad analysis from appropriate crosses provided evidence that the lys14 locus represented a single nuclear gene and that lys14 mutants were genetically distinct from the other mutants (lys1, lys2, lys5, and lys9) blocked after the alpha-aminoadipate step. The lys14 strains, like lys9 mutants, accumulated alpha-aminoadipate-semialdehyde and lacked significant amounts of saccharopine reductase activity. On the bases of these results, it was concluded, therefore, that LYS9 and LYS14, two distinct genes, were required for the biosynthesis of saccharopine reductase in wild-type S. cerevisiae.  相似文献   

5.
A polyol dehydrogenase of broad specificity was purified 178-fold from extracts of the filamentous fungus Cephalosporium chrysogenum. The enzyme was found to act as an oxido-reductase in two substrate-coenzyme systems: D-sorbitol (or xylitol)-nicotinamide-adenine dinucleotide (NAD) and D-mannitol-nicotinamide adenine dinucleotide phosphate (NADP). The dehydrogenase was composed of five isozymes, which, as a mixture, exhibited these properties: Km to D-sorbitol and D-mannitol, 7.15 to 10(-2) M; PH optimum, 9 to 10; molecular weight, 300,000 subunit weight, 29,000; PI, 5.8 to 7.5. The NADP-linked activity was labile to treatment with heat or ethylenediaminetetraacetic acid. Mixed substrate assays support the hypothesis that both NAD-, and NADP-linked activities are associated with isozymes of a single dehydrogenase.  相似文献   

6.
Carbon monoxide-dependent production of H2, CO2, and CH4 was detected in crude cell extracts of acetate-grown Methanosarcina barkeri. This metabolic transformation was associated with an active methyl viologen-linked CO dehydrogenase activity (5 to 10 U/mg of protein). Carbon monoxide dehydrogenase activity was inhibited 85% by 10 microM KCN and was rapidly inactivated by O2. The enzyme was nearly homogeneous after 20-fold purification, indicating that a significant proportion of soluble cell protein was CO dehydrogenase (ca. 5%). The native purified enzyme displayed a molecular weight of 232,000 and a two-subunit composition of 92,000 and 18,000 daltons. The enzyme was shown to contain nickel by isolation of radioactive CO dehydrogenase from cells grown in 63Ni. Analysis of enzyme kinetic properties revealed an apparent Km of 5 mM for CO and a Vmax of 1,300 U/mg of protein. The spectral properties of the enzyme were similar to those published for CO dehydrogenase from acetogenic anaerobes. The physiological functions of the enzyme are discussed.  相似文献   

7.
The activity of glucose dehydrogenase present in resting spores of Bacillus subtilis varied strikingly with the conditions for disrupting the spores by sonic treatment, namely, the time and strength of sonication, and the type and pH of the solution used for suspending the spores. When the resting spores were sonicated for 30 min at a current of 1.45 A in 100 mM phosphate buffer in the range of pH 6.0 to 6.6 or in deionized water, the enzyme activity of the former suspension was approximately 10 times higher than that of the latter suspension. However, the enzyme activity of the latter was markedly stimulated in the presence of sodium chloride. The glucose dehydrogenase from resting spores disrupted in 100 mM phosphate buffer (pH 6.6) was a salt-independent, active enzyme with a molecular weight of about 120,000, whereas the enzyme from resting spores disrupted in deionized water was a salt-dependent, inactive one with a molecular weight of about 55,000. A high concentration of dipicolinic acid strongly inhibited activation by a salt of inactive glucose dehydrogenase from resting spores in deionized water, suggesting one of its several important roles in vivo.  相似文献   

8.
C1-Tetrahydrofolate synthase is a multifunctional enzyme which catalyzes three reactions in 1-carbon metabolism: 10-formyltetrahydrofolate synthetase; 5,10-methenyltetrahydrofolate cyclohydrolase; 5,10-methylenetetrahydrofolate dehydrogenase. A rapid 1-day purification procedure has been developed which gives 40 mg of pure enzyme from 10 rabbit livers. The 10-formyltetrahydrofolate synthetase activity of this trifunctional enzyme has a specific activity that is 4-fold higher than the enzyme previously purified from rabbit liver. Conditions have been developed for the rapid isolation of a tryptic fragment of the enzyme which contains the methylenetetrahydrofolate dehydrogenase and methenyltetrahydrofolate cyclohydrolase activities. This fragment is a monomer exhibiting a subunit and native molecular weight of 36,000 in most buffers. However, in phosphate buffers the native molecular weight suggests that the fragment is a dimer. Conditions are also given whereby chymotryptic digestion allows the simultaneous isolation from the native enzyme of a large fragment containing the 10-formyltetrahydrofolate synthetase activity and a smaller fragment containing the dehydrogenase and cyclohydrolase activities. The large fragment is a dimer with a subunit molecular weight of 66,000. The small fragment retains all of the dehydrogenase and cyclohydrolase activities of the native enzyme. The large fragment is unstable but retains most of the 10-formyltetrahydrofolate synthetase activity. Km values of substrates for the two fragments are the same as the values for the native enzyme. The 10-formyltetrahydrofolate synthetase activity of the native enzyme requires ammonium or potassium ions for expression of full catalytic activity. The effect of these two ions on the catalytic activity of the large chymotryptic fragment is the same as with the native enzyme. We have shown by differential scanning calorimetry that the native enzyme contains two protein domains which show thermal transitions at 47 and 60 degrees C. Evidence is presented that the two domains are related to the two protein fragments generated by proteolysis of the native enzyme. The larger of the two domains contains the active site for the 10-formyltetrahydrofolate synthetase activity while the smaller domain contains the active site which catalyzes the dehydrogenase and cyclohydrolase reactions. Replacement of sodium ion buffers with either ammonium or potassium ions results in an increase in stability of the large domain of the native enzyme. This change in stability is not accompanied by a change in the quaternary structure of the enzyme.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

9.
The alpha-aminoadipate pathway for the biosynthesis of lysine is present only in fungi and euglena. Until now, this unique metabolic pathway has never been investigated in the opportunistic fungal pathogens Candida albicans, Cryptococcus neoformans, and Aspergillus fumigatus. Five of the eight enzymes (homocitrate synthase, homoisocitrate dehydrogenase, alpha-aminoadipate reductase, saccharopine reductase, and saccharopine dehydrogenase) of the alpha-aminoadipate pathway and glucose-6-phosphate dehydrogenase, a glycolytic enzyme used as a control, were demonstrated in wild-type cells of these organisms. All enzymes were present in Saccharomyces cerevisiae and the pathogenic organisms except C. neoformans 32608 serotype C, which exhibited no saccharopine reductase activity. The levels of enzyme activity varied considerably from strain to strain. Variation among organisms was also observed for the control enzyme. Among the pathogens, C. albicans exhibited much higher homocitrate synthase, homoisocitrate dehydrogenase, and alpha-aminoadipate reductase activities. Seven lysine auxotrophs of C. albicans and one of Candida tropicalis were characterized biochemically to determine the biochemical blocks and gene-enzyme relationships. Growth responses to alpha-aminoadipate- and lysine-supplemented media, accumulation of alpha-aminoadipate semialdehyde, and the lack of enzyme activity revealed that five of the mutants (WA104, WA153, WC7-1-3, WD1-31-2, and A5155) were blocked at the alpha-aminoadipate reductase step, two (STN57 and WD1-3-6) were blocked at the saccharopine dehydrogenase step, and the C. tropicalis mutant (X-16) was blocked at the saccharopine reductase step. The cloned LYS1 gene of C. albicans in the recombinant plasmid YpB1078 complemented saccharopine dehydrogenase (lys1) mutants of S. cerevisiae and C. albicans. The Lys1+ transformed strains exhibited significant saccharopine dehydrogenase activity in comparison with untransformed mutants. The cloned LYS1 gene has been localized on a 1.8-kb HindIII DNA insert of the recombinant plasmid YpB1041RG1. These results established the gene-enzyme relationship in the second half of the alpha-aminoadipate pathway. The presence of this unique pathway in the pathogenic fungi could be useful for their rapid detection and control.  相似文献   

10.
I purified a new dihydrolipoamide dehydrogenase from a lpd mutant of Escherichia coli deficient in the lipoamide dehydrogenase (EC 1.6.4.3) common to the pyruvate dehydrogenase (EC 1.2.4.1) and 2-oxoglutarate dehydrogenase complexes. The occurrence of the new lipoamide dehydrogenase in lpd mutants, including a lpd deletion mutant and the immunological properties of the enzyme, showed that it is different from the lpd gene product. The new dihydrolipoamide dehydrogenase had a molecular weight of 46,000, as determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. It was expressed in low amounts. It catalyzed the NAD+-dependent reduction of dihydrolipoamide with a maximal activity of 20 mumol/min per mg of protein and exhibited a hyperbolic dependence of catalytic activity on the concentration of both dihydrolipoamide and NAD+. The possible implication of the new dihydrolipoamide in the function of 2-oxo acid dehydrogenase complexes is discussed, as is its relation to binding protein-dependent transport.  相似文献   

11.
The cnx- group of mutants of Aspergillus nidulans lacks xanthine dehydrogenase (xanthine: NAD+ oxidoreductase, EC 1.2.1.37) and nitrate reductase (EC 1.6.6.3) activities and are thought to be defective in the synthesis of a molybdenum-containing cofactor, 'cnx', common to xanthine dehydrogenase and nitrate reductase [Pateman, J.A., Rever, B.M., Cove, D.J. and Roberts, D.B. (1964) Nature (Lond.) 201, 58-60]. The cnx cofactor has a role in maintaining the aggregated multimeric structure of nitrate reductase [MacDonald, D.W., Cove, D.J. and Coddington, A. (1974) Mol. Gen. Genet. 128, 187-199]. We report here that, in cnx- mutants grown under conditions inducing xanthine dehydrogenase I, a species cross-reacting with antisera to the native enzyme and of half its molecular weight is present, together with cross-reacting molecules of similar molecular weight to the native enzyme. This suggests that the cnx cofactor has a role in maintaining the aggregated structure of xanthine dehydrogenase I. Both cross-reacting species are capable of passing reducing equivalents from NADH to a tetrazolium salt, showing that the cnx cofactor is not necessary for enzymic activity towards NADH.  相似文献   

12.
Xanthine dehydrogenase (EC 1.2.1.37) is the first enzyme in the degradative pathway by which fungi convert purines to ammonia. In vivo, the activity is induced 6-fold by growth in uric acid. Hypoxanthine, xanthine, adenine, or guanine also induce enzyme activity but to a lesser degree. Immunoelectrophoresis using monospecific antibodies prepared against Neurospora crassa xanthine dehydrogenase shows that the induced increase in enzyme activity results from increased numbers of xanthine dehydrogenase molecules, presumably arising from de novo enzyme synthesis. Xanthine dehydrogenase has been purified to homogeneity by conventional methods followed by immunoabsorption to monospecific antibodies coupled to Sepharose 6B. Electrophoresis of purified xanthine dehydrogenase reveals a single protein band which also exhibits enzyme activity. The average specific activity of purified enzyme is 140 nmol of isoxanthopterine produced/min/mg. Xanthine dehydrogenase activity is substrate-inhibited by xanthine (0.14 mM), hypoxanthine (0.3 mM), and pterine (10 micron), is only slightly affected by metal binding agents such as KCN (6 mM), but is strongly inhibited by sulfhydryl reagents such as p-hydroxymercuribenzoate (2 micron). The molecular weight of xanthine dehydrogenase is 357,000 as calculated from a sedimentation coefficient of 11.8 S and a Stokes radius of 6.37 nm. Sodium dodecyl sulfate-gel electrophoresis of the enzyme reveals a single protein band having a molecular weight of 155,000. So the xanthine dehydrogenase protein appears to be a dimer. In contrast to xanthine dehydrogenases from animal sources which typically possess as prosthetic groups 2 FAD molecules, 2 molybdenum atoms, 8 atoms of iron, and 8 acid-labile sulfides, the Neurospora enzyme contains 2 FAD molecules, 1 molybdenum atom, 12 atoms of iron, and 14 eq of labile sulfide/molecule. The absorption spectrum of the enzyme shows maxima between 400 and 500 nm typical of a non-heme iron-containing flavoprotein.  相似文献   

13.
NAD-dependent 1,2-propanediol dehydrogenase (EC 1.1.1.4) activity was detected in cell-free crude extracts of various propane-grown bacteria. The enzyme activity was much lower in 1-propanol-grown cells than in propane-grown cells of Pseudomonas fluorescens NRRL B-1244, indicating that the enzyme may be inducible by metabolites of propane subterminal oxidation. 1,2-Propanediol dehydrogenase was purified from propane-grown Ps. fluorescens NRRL B-1244. The purified enzyme fraction shows a single-protein band upon acrylamide gel electrophoresis and has a molecular weight of 760,000. It consists of 10 subunits of identical molecular weight (77,600). It oxidizes diols that possess either two adjacent hydroxy groups, or a hydroxy group with an adjacent carbonyl group. Primary and secondary alcohols are not oxidized. The pH and temperature optima for 1,2-propanediol dehydrogenase are 8.5 and 20-25 degrees C, respectively. The activation energy calculated is 5.76 kcal/mol. 1,2-Propanediol dehydrogenase does not catalyze the reduction of acetol or acetoin in the presence of NADH (reverse reaction). The Km values at 25 degrees C, pH 7.0, buffer solution for 1,2-propan1,2-propanediol dehydrogenase are 8.5 and 20-25 degrees C, respectively. The activation energy calculated is 5.76 kcal/mol. 1,2-Propanediol dehydrogenase does not catalyze the reduction of acetol or acetoin in the presence of NADH (reverse reaction). The Km values at 25 degrees C, pH 7.0, buffer solution for 1,2-propan1,2-propanediol dehydrogenase are 8.5 and 20-25 degrees C, respectively. The activation energy calculated is 5.76 kcal/mol. 1,2-Propanediol dehydrogenase does not catalyze the reduction of acetol or acetoin in the presence of NADH (reverse reaction). The Km values at 25 degrees C, pH 7.0, buffer solution for 1,2-propanediol and NAD are 2 X 10(-2) and 9 X 10(-5) M, respectively. The 1,2-propanediol dehydrogenase activity was inhibited by strong thiol reagents, but not by metal-chelating agents. The amino acid composition of the purified enzyme was determined. Antisera prepared against purified 1,2-propanediol dehydrogenase from propane-grown Ps. fluorescens NRRL B-1244 formed homologous precipitin bands with isofunctional enzymes derived from propane-grown Arthrobacter sp. NRRL B-11315, Nocardia paraffinica ATCC 21198, and Mycobacterium sp. P2y, but not from propane-grown Pseudomonas multivorans ATCC 17616 and Brevibacterium sp. ATCC 14649, or 1-propanol-grown Ps. fluorescens NRRL B-1244. Isofunctional enzymes derived from methane-grown methylotrophs also showed different immunological and catalytic properties.  相似文献   

14.
15.
An NAD-linked formate dehydrogenase (EC 1.2.1.2.) from methanol-grown Pichia pastoris NRRL Y-7556 has been purified. The purification procedure involved ammonium sulfate fractionation, hollow-fiber H1P10 filtration, ion-exchange chromatography, and gel filtration. Both dithiothreitol (10 mm) and glycerol (10%) were required for stability of the enzyme during purification. The final enzyme preparation was homogeneous as judged by polyacrylamide gel electrophoresis and by sedimentation pattern in an ultracentrifuge. The enzyme has a molecular weight of 94,000 and consists of two subunits of identical molecular weight. Formate dehydrogenase catalyzes specifically the oxidation of formate. No other compounds tested can replace NAD as the electron acceptor. The Michaelis constants were 0.14 mm for NAD and 16 mm for formate (pH 7.0, 25 °C). Optimum pH and temperature for formate dehydrogenase activity were around 6.5–7.5 and 20–25 °C, respectively. Amino acid composition of the enzyme was also studied. Antisera prepared against the purified enzyme from P. pastoris NRRL Y-7556 form precipitin bands with isofunctional enzymes from different strains of methanol-grown yeasts, but not bacteria, on immunodiffusion plates. Immunoglobulin fraction prepared against the enzyme from yeast strain Y-7556 inhibits the catalytic activity of the isofunctional enzymes from different strains of methanol-grown yeasts.  相似文献   

16.
A NADP(+)-dependent 3 beta-hydroxysteroid dehydrogenase activity was localized in the microsomal fraction of rat liver. This enzyme was solubilized and separated completely from 3 alpha-hydroxysteroid dehydrogenase by Matrex red A column chromatography. Partially purified 3 beta-hydroxysteroid dehydrogenase catalyzed the oxidation and reduction between the 3 beta-hydroxyl and 3-ketonic group of steroids or bile acids having no double bond in the A/B ring, but was inactive toward 3 alpha-hydroxyl group. The enzyme required NADP+ for oxidation and NADPH for reduction. The activity was inhibited by p-chloromercuribenzoic acid or p-chloromercuribenzenesulfonic acid at the concentration of 10(-4) M. The molecular weight of the enzyme was estimated to be about 43,000 by Sephadex G-200 column chromatography. From these results, it is concluded that the enzyme is a new type of microsomal NADP+:3 beta-hydroxysteroid dehydrogenase.  相似文献   

17.
The concentration-dependent association-dissociation equilibrium of the bifunctional enzyme aspartokinase I-homoserine dehydrogenase I of Escherichia coli K12 has been investigated at pH 7.6 in the presence of 10 mM L-threonine and 0.1 M KCl by equilibrium gel permeation monitored by a single-photon counting spectrophotometer. The results obtained are consistent with the existence of a dimer-tetramer equilibrium with the association constant of 2.6 X 10(7) M-1 (deltaG0 = -9.9 kcal/mol of dimer). The limiting partition cross-sections estimated by a three-parameter least squares minimization procedure indicate that the molecular radii of the dimer and tetramer are 53.8 A and 70 A, respectively. Both the dimeric and tetrameric forms of the enzyme possess dehydrogenase activity. Treatment of the enzyme with the chaotropic salts, potassium thiocyanate or potassium trichloroacetate, generates a monomeric form that is devoid of dehydrogenase activity. The catalytically inactive monomeric form of the enzyme has a molecular radius between 43 and 45.5 A and a molecular weight of approximately 80,000 as determined by small zone gel chromatography and sedimentation equilibrium studies.  相似文献   

18.
Characterization of xylitol-utilizing mutants of Erwinia uredovora.   总被引:3,自引:3,他引:0       下载免费PDF全文
Of the four pentitols ribitol, xylitol, D-arabitol, and L-arabitol, Erwinia uredovora was able to utilize only D-arabitol as a carbon and energy source. Although attempts to isolate ribitol- or L-arabitol-utilizing mutants were unsuccessful, mutants able to grow on xylitol were isolated at a frequency of 9 X 10(-8). Xylitol-positive mutants constitutively synthesized both a novel NAD-dependent xylitol-4-dehydrogenase, which oxidized xylitol to L-xylulose, and an L-xylulokinase. The xylitol dehydrogenase had a Km for xylitol of 48 mM and showed best activity with xylitol and D-threitol as substrates. However, D-threitol was not a growth substrate for E. uredovora, and its presence did not induce either dehydrogenase or kinase activity. Attempts to determine the origin of the xylitol catabolic enzymes were unsuccessful; neither enzyme was induced on any growth substrate or in the presence of any polyol tested. Analysis of xylitol-negative mutants isolated after Tn5 mutagenesis suggested that the xylitol dehydrogenase and the L-xylulokinase structural genes were components of two separate operons but were under common regulatory control.  相似文献   

19.
The regulatory isoform of glyceraldehyde-3-phosphate dehydrogenase (GAPDH) is a light-activated enzyme constituted by subunits GapA and GapB. The NADPH-dependent activity of regulatory GAPDH from spinach chloroplasts was affected by the redox potential (E(m,7.9), -353 +/- 11 mV) through the action of thioredoxin f. The redox dependence of recombinant GapB (E(m,7.9), -347 +/- 9 mV) was similar to native GAPDH, whereas GapA was essentially redox-insensitive. GapB mutants having one or two C-terminal cysteines mutated into serines (C358S, C349S, C349S/C358S) were less redox-sensitive than GapB. Different mutants with other cysteines substituted by serines (C18S, C274S, C285S) still showed strong redox regulation. Fully active GapB was a tetramer of B-subunits, and, when incubated with NAD, it associated to a high molecular weight oligomer showing low NADPH-dependent activity. The C-terminal GapB mutants (C358S, C349S, C349S/C358S) were active tetramers unable to aggregate to higher oligomers in the presence of NAD, whereas other mutants (C18S, C274S, C285S) again behaved like GapB. We conclude that a regulatory disulfide, between Cys-349 and Cys-358 of the C-terminal extension of GapB, does form in the presence of oxidized thioredoxin. This covalent modification is required for the NAD-dependent association into higher oligomers and inhibition of the NADPH-activity. By leading to GAPDH autoinhibition, thioredoxin and NAD may thus concur to the dark inactivation of the enzyme in vivo.  相似文献   

20.
Regulation of Proline Degradation in Salmonella typhimurium   总被引:30,自引:22,他引:8       下载免费PDF全文
The pathway for proline degradation in Salmonella typhimurium appears to be identical to that found in Escherichia coli and Bacillus subtilis. Delta(1)-Pyrroline-5-carboxylic acid (P5C) is an intermediate in the pathway; its formation consumes molecular oxygen. Assays were devised for proline oxidase and the nicotinamide adenine dinucleotide phosphate-specific P5C dehydrogenase activities. Both proline-degrading enzymes, proline oxidase and P5C dehydrogenase, are induced by proline and are subject to catabolite repression. Three types of mutants were isolated in which both enzymes are affected: constitutive mutants, mutants with reduced levels of enzyme activity, and mutants unable to produce either enzyme. Most of the mutants isolated for their lack of P5C dehydrogenase activity have a reduced level of proline oxidase activity. All the mutations are cotransducible. A genetic map of some of the mutations is presented. The actual effector of the pathway appears to be proline.  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