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 312 毫秒
1.
Preliminary evidence is presented that indicates that the dihydrofolate reductase activity of amethopterin-sensitive Streptococcus faecium var. durans ATCC 8043 is separable into two dihydrofolate reductases, one of which also reduces folate.  相似文献   

2.
Pediococcus cerevisiae/AMr, resistant to amethopterin, possesses a higher dihydrofolate reductase (5, 6, 7, 8-tetrahydrofolate: NADP+ oxidoreductase, EC 1.5.1.3) activity than the parent, a folate-permeable and thus amethopterin-susceptible strain and than the wild-type. The properties of dihydrofolate reductase from the three strains have been compared. Temperature, pH optima, heat stability, as well amethopterin binding did not reveal significant differences between the enzymes from the susceptible and resistant strains. The enzyme from the wild-type was 10 times more sensitive to inhibition by amethopterin and more susceptible to heat denaturation. The apparent Km values for dihydrofolate in enzymes from the three strains were in the range of 4.8--7.2 muM and for NADPH 6.5--8.0 muM. The amethopterin-resistant strain exhibited cross-resistance to trimethoprim and was about 40-fold more resistant to the latter than the sensitive parent and the wild-type. The resistance to trimethoprim appears to be a direct result of the increased dihydrofolate reductase activity. Inhibition of dihydrofolate reductase activity by this drug was similar in the three strains. 10--20 nmol caused 50% inhibition of 0.02 enzyme unit. Trimethoprim was about 10 000 times less effective inhibitor of dihydrofolate reductase than amethopterin. The cell extract of the AMr strain possessed a folate reductase activity three times higher than that of the sensitive strain. The activities of other folate-related enzymes like thymidylate synthetase and 10-formyltetrahydrofolate synthetase (formate: tetrahydrofolate ligase (ADP-forming), EC 6.3.4.3) were similar in the three strains studied.  相似文献   

3.
Pediococcus cerevisiae/AMr, resistant to amethopterin, possesses a higher dihydrofolate reductase (5, 6, 7, 8-tetrahydrofolate: NADP+ oxidoreductase, EC 1.5.1.3) activity than the parent, a folate-permeable and thus amethopterin-susceptible strain and than the wild-type. The properties of dihydrofolate reductase from the three strains have been compared. Temperature, pH optima, heat stability, as well amethopterin binding did not reveal significant differences between the enzymes from the susceptible and resistant strains. The enzyme from the wild-type was 10 times more sensitive to inhibition by amethopterin and more susceptible to heat denaturation. The apparent Km values for dihydrofolate in enzymes from the three strains were in the range of 4.8–7.2 μM and for NADPH 6.5–8.0 μM. The amethopterin-resistant strain exhibited cross-resistance to trimethoprim and was about 40-fold more resistant to the latter than the sensitive parent and the wild-type. The resistance to trimethoprim appears to be a direct result of the increased dihydrofolate reductase activity. Inhibition of dihydrofolate reductase activity by this drug was similar in the three strains. 10–20 nmol caused 50% inhibition of 0.02 enzyme unit. Trimethoprim was about 10 000 times less effective inhibitor of dihydrofolate reductase than amethopterin. The cell extract of the AMr strain possessed a folate reductase activity three times higher than that of the sensitive strain. The activities of other folate-related enzymes like thymidylate synthethase and 10-formyltetra-hydrofolate synthetase (formate: tetrahydrofolate ligase (ADP)-forming), EC 6.3.4.3) were similar in the three strains studied.  相似文献   

4.
Pediococcus cerevisiae/AMr, resistant to amethopterin, possesses a higher dihydrofolate reductase (5, 6, 7, 8-tetrahydrofolate: NADP+ oxidoreductase, EC 1.5.1.3) activity than the parent, a folate-permeable and thus amethopterin-susceptible strain and than the wild-type. The properties of dihydrofolate reductase from the three strains have been compared. Temperature, pH optima, heat stability, as well amethopterin binding did not reveal significant differences between the enzymes from the susceptible and resistant strains. The enzyme from the wild-type was 10 times more sensitive to inhibition by amethopterin and more susceptible to heat denaturation. The apparent Km values for dihydrofolate in enzymes from the three strains were in the range of 4.8–7.2 μM and for NADPH 6.5–8.0 μM. The amethopterin-resistant strain exhibited cross-resistance to trimethoprim and was about 40-fold more resistant to the latter than the sensitive parent and the wild-type. The resistance to trimethoprim appears to be a direct result of the increased dihydrofolate reductase activity. Inhibition of dihydrofolate reductase activity by this drug was similar in the three strains. 10–20 nmol caused 50% inhibition of 0.02 enzyme unit. Trimethoprim was about 10 000 times less effective inhibitor of dihydrofolate reductase than amethopterin. The cell extract of the AMr strain possessed a folate reductase activity three times higher than that of the sensitive strain. The activities of other folate-related enzymes like thymidylate synthethase and 10-formyltetra-hydrofolate synthetase (formate: tetrahydrofolate ligase (ADP)-forming), EC 6.3.4.3) were similar in the three strains studied.  相似文献   

5.
Citrate metabolism by Enterococcus faecium ET C9 and Enterococcus durans Ov 421 was studied as sole energy source and in presence of glucose or lactose. Both strains utilized citrate as the sole energy source. Enterococcus faecium ET C9 showed diauxic growth in the presence of a limiting concentration of glucose. Neither strain used citrate until glucose was fully metabolized. The strains showed co-metabolism of citrate and lactose. Lactate, acetate, formate, and flavour compounds (diacetyl, acetoin, and 2,3-butanediol) were detected in both strains. The highest production of flavour compounds was detected during growth of E. durans Ov 421 in media supplemented with citrate-glucose and citrate-lactose. Citrate lyase was inducible in both strains. Acetate kinase activities presented the highest values in LAPTc medium, with E. faecium ET C9 displaying a specific activity 2.4-fold higher than E. durans. The highest levels of alpha-acetolactate synthase specific activity were detected in E. durans grown in LAPTc+g, in accordance with the maximum production of flavour compounds detected in this medium. Diacetyl and acetoinreductases displayed lower specific activity values in the presence of citrate. Enterococcus faecium and E. durans displayed citrate lyase, acetate kinase, alpha-acetolactate synthase, and diacetyl and acetoin reductase activities. These enzymes are necessary for conversion of citrate to flavour compounds that are important in fermented dairy products.  相似文献   

6.
Interaction of several representative folate, quinazoline and pyridine nucleotide derivatives with dihydrofolate reductase from amethopterin-resistant Lactobacillus casei induces dramatic changes in its circular dichroic spectral properties. The binding of dihydrofolate induces a large extrinsic Cotton effect at 295 nm ([theta] = 113 800 deg . cm2 . dm-1). The generation of this band by dihydrofolate is strictly dependent on complex formation with a single substrate binding site and a KD = 7 . 10(-6) M. The other binary complexes examined include the enzyme . NADPH, enzyme . amethopterin, enzyme . folate, and enzyme . methasquin. All such complexes differ in spectral detail, the negative ellipticity at 330 nm being characteristic of the "folate site" complexes. The circular dichroic spectrum of the ternary complex of reductase . NADPH . methotrexate shows a positive symmetrical band centered at 360 nm ([theta] - 32 000 deg . cm2 . dm-1). Since both of the corresponding binary complexes exhibit negative bands in this region, this induced band represents a unique molecular property of the ternary complex. Chemical modification of a single tryptophan residue of the enzyme, as determined from magnetic circular dichroism spectra, results in a complete loss in the ability to bind either dihydrofolate or NADPH.  相似文献   

7.
Activity analyses of pure dihydrofolate reductase from amethopterin-resistant Lactobacillus casei conducted with commercial sources of NADPH yielded a progression of nonlinear assay tracings whose shapes were both pH dependent and reminiscent of classical product inhibition. The extent of curving of the assay tracings was dependent on the source and age of the commercial NADPH and was enhanced as the pH was decreased from 7.5 to 5.0. Under these conditions a “pseudo”-pH-activity profile, exhibiting a maximal specific activity of 9 units/mg of protein between pH 7.0 and 7.5, was found. In contrast, freshly prepared NADPH provided strictly linear assay tracings over the pH range of 8.5 to 5.0, yielding uniformly higher specific activities than those observed with commercial NADPH. The new pH-activity profile was characterized by a broad optimum between pH 5.0 and 6.0, with a maximal specificity activity of 24.9 units/ mg in 0.1m potassium phosphate in the absence of added salt. The curving phenomenon and pseudo-pH optimum observed with commercial NADPH is attributed to the presence of minor but potent inhibitory impurities in these coenzyme preparations. Optimal concentrations of monovalent (~0.1 m) and divalent (~0.05 m) salts activated the enzyme between 1.5- and 1.7-fold, resulting in maximal specific activities in the range of 34 to 39 units/mg. A similar extent of activation was observed in 0.8 m Tris-acetate buffer, pH 5.5. At concentrations of monovalent salts above 0.5 m and of divalent salts above 0.2 m a reduction in salt-dependent activation and, in some cases, inhibition of activity were obtained. Substrate specificity studies indicated that the V for folate at saturating levels is 1% of that for dihydrofolate. Deamino-NADPH yielded V values 1.4-fold higher than that for NADPH, while acetylpyridine-NADPH and thio-NADPH provided values 6.5- and 235-fold lower, respectively, than the value with the natural coenzyme. Gel electrophoresis studies reflected a similar trend of selectivity in the interaction of NADPH and its analogs to form stable binary complexes. Stable ternary complexes of enzyme and amethopterin were formed with NADPH, deamino-NADPH, thio-NADPH, and acetylpyridine-NADPH. Although neither dihydrofolate nor NADP+ and its analog form stable complexes with L. casei dihydrofolate reductase, both NADP+ and deamino-NADP+ interact with enzyme and dihydrofolate to generate stable ternary complexes.  相似文献   

8.
We describe the development of resistance to trimetrexate and piritrexim (BW 301U) by a stepwise selection protocol in Chinese hamster ovary cells. Selection in trimetrexate resulted in initial resistance as a result of dihydrofolate reductase gene amplification. Several trimetrexate-resistant variants that display 250-340-fold and 25-50-fold resistance to lipophilic and hydrophilic antifolates, respectively, were established. Increased antifolate resistance was associated with a prominent overexpression of dihydrofolate reductase as determined from the elevated folate reductase activity, cellular labeling with fluorescein-methotrexate, and steady-state mRNA levels as a result of a consistent dihydrofolate reductase gene amplification. However, upon subsequent incremental increases in trimetrexate, further resistance was also associated with amplification of the multidrug resistance gene. This resulted in overexpression of P-glycoprotein and a subsequent 20-50-fold collateral resistance to pleiotropic drugs such as adriamycin, actinomycin D, vinca alkaloids, etoposide, and colchicine. In contrast, initial resistance following selection with low piritrexim concentrations resulted from an unknown mechanism(s) not involving overproduction of either dihydrofolate reductase or P-glycoprotein. This piritrexim resistance was shared with trimetrexate but not with methotrexate. Upon further selection with piritrexim, resistant variants emerge with amplified dihydrofolate reductase but not with multidrug resistance genes. These variants were subsequently resistant to both hydrophilic and lipophilic folate antagonists but retained sensitivity to pleiotropic drugs. The pattern of resistance with methotrexate, trimetrexate, and piritrexim shared a common mechanism, dihydrofolate reductase gene amplification, but differed regarding the additional amplification of the multidrug resistance gene in trimetrexate-resistant cells as well as the emergence of an additional unknown mechanism(s) of resistance to lipid-soluble antifolates upon initial selection in piritrexim.  相似文献   

9.
Differentiation of Some Enterococci by Gas Chromatography   总被引:7,自引:0,他引:7       下载免费PDF全文
Relative fatty acid compositions of 37 enterococci were examined by gas chromatography. Streptococcus faecalis, S. faecium, and S. faecium var. durans yielded similar fatty acid patterns. Strains of S. faecium var. casseliflavus, and a motile yellow-pigmented streptococcus, contained very low levels of C(19:0) cyclopropane fatty acid and four unidentified components, compared to the other strains of enterococci examined. There were no significant differences in the fatty acid patterns of enterococci grouped according to plant, animal, or human source.  相似文献   

10.
The metabolic regulation of the synthesis of the folate reductase and the specific dihydrofolate reductase of the amethopterin-resistant bacterium Streptococcus faecium var. durans Ak was investigated. Under growth conditions for maximum synthesis, the differential rate increased two fold during exponential growth. Synthesis of both reductases was coordinate and air-sensitive.  相似文献   

11.
Dihydrofolate reductase from bovine liver has been purified 5000-fold employing conventional techniques and methotrexate/aminohexyl/Sepharose affinity chromatography. Electrophoresis of the isolated enzyme on polyacrylamide gels resulted in the separation of two enzymatically active protein components which were not interconvertible by treatment with dihydrofolate and/or the coenzyme. The two forms, present in a ratio of 20:1, were found by isoelectric focusing to have isoelectric points of 7.15 and 5.94. They had identical specific activities toward dihydrofolate (26.1-27.0 U/mg) and folate (1.3-2.2 U/mg), and had identical molecular weights (23500) and amino acid compositions. Due to the small quantity of the acidic form and the similarity of the two forms, the amino-terminal sequence (19 residues) was determined on a mixture of carboxymethylated reductase. The single sulfhydryl group of the enzyme can be modified by several sulfhydryl reagents in the native enzyme without loss of activity. Modification of the same residue occurs in the denaturated state and partially inhibits renaturation to the fully acitve enzyme. One disulfide bridge was detected by reduction and alkylation. The cleavage of this bond did not effect the enzymatic activity.  相似文献   

12.
A strain of Streptococcus faecium (ATCC 8043) which is highly resistant to the antifolic acid compound, amethopterin, was gently ruptured by exposing protoplasts of the organism to a hypotonic solution. The crude lysate resulting there-from was treated by various chemical and physical techniques designed to separate folic acid reductase from dihydrofolic acid reductase. In the process, the enzyme was purified approximately 160-fold; however, throughout the process, the enzyme preparation maintained the ability to reduce folic acid to tetrahydrofolic acid. Attempts to isolate mutants showing a deficiency in either folic acid reductase or dihydrofolic acid reductase were unsuccessful. Based on these results, it is concluded that folic acid is reduced to tetrahydrofolic acid by one enzyme in S. faecium (ATCC 8043). The crude lysate was also subjected to ultracentrifugation. An analysis of the supernatant fluid and the sediment indicated that the reductive activity is located in the soluble fraction of the cell.  相似文献   

13.
Dihydrofolate reductase from chicken liver has a single sulfhydryl group which reacts stoichiometrically and specifically with a wide variety of organic mercury compounds to yield an enzyme derivative which exhibits up to 10-fold the activity of the unmodified form when measured at pH 6.5, the optimum for the modified enzyme. The sulfhydryl group is apparently not at the active site since a 25-fold excess of either major cosubstrate, dihydrofolate or TPNH, affects neither the rate nor extent of the modification reaction. The reaction is essentially instantaneous and yields an enzyme with altered kinetic properties for all the substrate pairs examined (TPNH/dihydrofolate, TPNH/ folate, and DPNH/dihydrofolate) when tested near their pH optima. V values increased 3- to 10-fold when TPNH was cofactor; Km values increased 10- to 15-fold for the TPNH/dihydrofolate pair. The mercurial-activated enzyme, unlike the native form, exhibits a markedly increased sensitivity to heat, proteolysis, and the ionic environment, losing approximately 50% of its activity under conditions where there is no loss of activity in the native form. However, substrates can afford protection, the order of effectiveness being identical with the relative affinities of the substrates for the native enzyme (Subramanian, S., and Kaufman, B. T. (1978) Proc. Nat. Acad. Sci. USA75, 3201). Thus, dihydrofolate, with the largest binding constant is the most efficient, protecting completely against trypsin digestion when present at a 1:1 ratio with enzyme. Heating the mercury enzyme in the absence of substrates gives rise to a stable but altered conformation characterized by a time course which shows marked hysteresis. The striking similarity of the properties of the mercurial-activated dihydrofolate reductase to the reductase activated by 4 m urea, a reagent known to affect the tertiary structure of proteins, suggests that covalent binding of organic mercurials to the sulfhydryl group results in a similar conformational change characterized by a marked facilitation of the dihydrofolate reductase reaction.  相似文献   

14.
粪肠、屎肠球菌及相近种部分持家基因的系统发育分析   总被引:1,自引:0,他引:1  
【目的】利用16S rRNA、clpX和recA基因分子标记研究Enterococcus faecalis、Enterococcus faecium及相近种间的种系发育关系,并比较这些基因序列对E.faecalis、E.faecium及相近种的区分能力。【方法】以分离自传统乳制品中的9株E.faecium和1株E.durans分离株为研究对象,以clpX和recA基因片段为标记,通过PCR扩增、测序,结合已公布的近缘种相应序列构建系统发育树并与16S rRNA基因进行比较。【结果】在基于clpX和recA基因的进化树中,10株试验菌株与E.faecalis始终处于同一分支。与该物种这两个基因的平均相似性为99.6%和98.6%,与另一分支的Faecium-group(E.durans和E.faecium)的平均相似性仅为61.5%和33.5%。相近种E.durans和E.hirae间这两个基因的差异性为20.3%和39.0%;在基于16S rRNA基因的进化树中,试验菌株与Faecium-group(E.lactis、E.faecium、E.durans、E.hirae)处于同一分支。与这些成员间该基因的相似性大于99.6%,与E.faecalis基因的平均相似性可达98.4%。相近种间该基因相似性无明显差异。【结论】按照10株试验菌株clpX和recA基因的分析结果可将由传统生理生化和16S rRNA基因序列鉴定的9株E.faecium和1株E.durans归类为E.faecalis,clpX和recA基因可用于部分相近种的分类鉴定。  相似文献   

15.
The methylenetetrahydrofolate dehydrogenase of the amethopterin-resistant strain Streptococcus faecium var. durans A(k) was purified 100-fold. Because it is extremely labile, this enzyme required protection by 1 mm nicotinamide adenine dinucleotide phosphate (NADP(+)) during purification; 0.01 mm EADP(+) with 0.1% bovine plasma albumin stabilized the purified enzyme during storage at -20 C. Although the enzyme has properties of sulfhydryl enzymes, thiol compounds were not stabilizers. Oxidation of methylenetetrahydrofolate, catalyzed by the purified enzyme preparation, is NADP(+)-specific and yields methenyltetrahydrofolate and the reduced pyridine nucleotide. K(m) values for NADP(+) and for 5,10-methylenetetrahydrofolate (prepared as the formaldehyde adduct of biologically synthesized l,l-tetrahydrofolate) were calculated to be 0.021 and 0.026 mm, respectively. Neither purine bases and their derivatives nor serine inhibited the reaction. In growing cultures, the differential rate of synthesis of the methylenetetrahydrofolate dehydrogenase was dependent upon the composition of the medium. A medium which contained acid-hydrolyzed casein, and thus an exogenous source of serine, was repressive for this enzyme. In a serine-free, completely defined medium, the amount of folate added (for serine synthesis de novo) affected the duration of the initial exponential growth phase. At the termination of this phase, which primarily reflected the onset of a decreased rate of serine biosynthesis, synthesis of the methylenetetrahydrofolate dehydrogenase was derepressed. Exogenous serine in the completely defined medium prevented the derepression. Furthermore, physiological concentrations of l-serine were repressive not only for the dehydrogenase but also for the methenyltetrahydrofolate cyclohydrolase and the serine hydroxymethyl-transferase. Concomitantly, the differential rate of synthesis of the formyltetrahydrofolate synthetase of S. faecium var. durans A(k) was increased. Apparently, serine regulates the differential rates of syntheses of these enzymes.  相似文献   

16.
A single subcutaneous injection of folate, homofolate or MTX resulted in the inhibition of the activity of dihydrofolate reductase in homogenates prepared from the kidneys of normal mice. Stimulation of 3H-thymidine uptake occurred in the kidneys of treated animals approximately 30 hr after administration of either folate or homofolate, and reached a peak 72 hr after administration. The effects of folate and MTX on dihydrofolate reductase activity invivo were also determined. One hr after administration of 15 mg/kg methotrexate (MTX) or 300 mg/kg folate, enzyme activity invivo was inhibited by 90%.3H-deoxyuridine uptake was neither stimulated nor depressed after treatment with MTX. After administration of folate, uptake of 3H-deoxyuridine was stimulated at approximately 30 hr after drug-treatment and reached a peak at 72 hr after folate administration. Treatment with xanthopterin had no effect on the activity of dihydrofolate reductase invitro. Xanthopterin stimulated uptake of both deoxyuridine and thymidine in an identical manner.The increased DNA synthesis that occurs in animals after treatment with agents that cause renal damage is distinct from the effect these agents have upon dihydrofolate reductase. Nucleoside incorporation after treatment with folate, homofolate, MTX or xanthopterin cannot be predicted on the basis of enzyme inhibition. Treatment with MTX, folate or homofolate results in enzyme inhibition which is not correlated with the uptake of deoxyuridine into DNA.  相似文献   

17.
Escherichia coli DNA photolyase was overproduced and purified from each of two mutant E. coli strains lacking dihydrofolate reductase. The extent of over-production in the mutants was comparable to that seen in the wild type strain. Examination of the isolated photolyase from these strains revealed that the folate cofactor, 5,10-methenyltetrahydrofolate, was present in these proteins at a level of 60-80% compared to that purified from the wild type strain. Further examination of the dihydrofolate reductase-deficient strains revealed the presence of other tetrahydrofolate derivatives. These findings demonstrate that dihydrofolate reductase is not essential for the production of tetrahydrofolates in E. coli.  相似文献   

18.
In the x-ray structure of the human dihydrofolate reductase, phenylalanine 31 and phenylalanine 34 have been shown to be involved in hydrophobic interactions with bound substrates and inhibitors. Using oligonucleotide-directed mutagenesis and a bacterial expression system producing the wild-type and mutant human dihydrofolate reductases at levels of 10% of the bacterial protein, we have constructed, expressed, and purified a serine 31 (S31) mutant and a serine 34 (S34) mutant. Fluorescence titration experiments indicated that S31 bound the substrate H2folate 10-fold tighter and the coenzyme NADPH 2-fold tighter than the wild-type human dihydrofolate reductase. The serine 31 mutation had little effect on the steady-state kinetic properties of the enzyme but produced a 100-fold increase in the dissociation constant (Kd) for the inhibitor methotrexate. The serine 34 mutant had much greater alterations in its properties than S31; specifically, S34 had a 3-fold reduction in the Km for NADPH, a 24-fold increase in the Km for H2folate, a 3-fold reduction in the overall reaction rate kcat, and an 80,000-fold increase in the Kd for methotrexate. In addition, the pH dependence of the steady-state kinetic parameters of S34 were different from that of the wild-type enzyme. These results suggest that phenylalanine 31 and phenylalanine 34 make very different contributions to ligand binding and catalysis in the human dihydrofolate reductase.  相似文献   

19.
R plasmid dihydrofolate reductase with a dimeric subunit structure   总被引:5,自引:0,他引:5  
Dihydrofolate reductase specified by plasmid R483 from a trimethoprim-resistant strain of Escherichia coli has been purified 2,000-fold to homogeneity using dye-ligand chromatography, gel filtration, and polyacrylamide gel electrophoresis. The protein migrated as a single band on nondenaturing polyacrylamide gel electrophoresis and had a specific activity of 250 mumol/mg min(-1). The molecular weight was estimated to be 32,000 by gel filtration and 39,000 by Ferguson analysis of polyacrylamide gel electrophoresis. When subjected to electrophoresis in the presence of sodium dodecyl sulfate, the protein migrated as a single 19,000-molecular weight species, a fact that suggests that the native enzyme is a dimer of similar or identical subunits. Antibody specific for R483-encoded dihydrofolate reductase did not cross-react with dihydrofolate reductase encoded by plasmid R67, T4 phage, E. coli RT500, or mouse L1210 leukemia cells. The amino acid sequence of the first 34 NH2-terminal residues suggests that the R483 plasmid dihydrofolate reductase is more closely related to the chromosomal dihydrofolate reductase than is the enzyme coded by plasmid R67.  相似文献   

20.
A purification procedure is reported for obtaining bovine liver dihydrofolate reductase in high yield and amounts of 100-200 mg. A key step in the procedure is the use of an affinity gel prepared by coupling pteroyl-L-lysine to Sepharose. The purified reductase has a specific activity of about 100 units/mg and is homogeneous as judged by analytical ultracentrifugation, polyacrylamide gel electrophoresis, and titration with methotrexate. The products of the first step of Edman degradation indicated a minimum purity of 79%. The reductase has a molecular weight of about 21500 on the basis of amino acid composition and 22100 +/- 300 from equilibrium sedimentation. It is not inhibited by antiserum to the Streptococcus faecium reductase (isoenzyme 2). Unlike the reductase of many other vertebrate tissues, the bovine enzyme is inhibited by mercurials rather than activated and it has a single pH optimum at both low and high ionic strength. However, the position of the pH optimum is shifted and the activity increased by increasing ionic strength. Automatic Edman degradation has been used to determine 34 of the amino-terminal 37 amino acid residues. Considerable homology exists between this region and the corresponding regions of the reductase from S. faecium and from Escherichia coli. This strengthens the idea that this region contributes to the structure of the binding site for dihydrofolate.  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