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 31 毫秒
1.
The turnover of dihydrofolate reductase has been studied in rapidly dividing cells of mouse lymphoma L1210 and Lactobacillus casei. Cells in culture were exposed to [14C]leucine for 24 hr and the subsequent decrease in radioactivity of the enzyme was followed as a function of time. The L1210 enzyme was isolated in pure form by subjecting the cell sonicate to affinity chromatography on amethopterin-Sepharose. The L. casei cells were processed by a multistep procedure which yielded the pure enzyme in both of its principal forms: (I), without TPNH; and (II), containing an equimolal amount of noncovalently bound TPNH. The half-lives (t12>) of dihydrofolate reductase in the amethopterin-sensitive L1210 cells (L1210/S) and in the cells of a partially resistant subline (L1210/R2), characterized by an 8-fold increase in enzyme level, were 18 and 19 hr. When these cells were grown in the presence of sublethal concentrations of amethopterin, t12 values were increased to 39 and 90 hr. These results suggest that the transient increase in dihydrofolate reductase activity, observed when cells are exposed to amethopterin, is due largely to a decreased susceptibility of the enzyme-inhibitor complex to degradation. Bound TPNH also increases the half-life of dihydrofolate reductase as shown by the fact that forms (I) and (II) of the L. casei enzyme had tx12 values of approximately 3 and 9 hr.  相似文献   

2.
When dihydrofolate reductase from a methotrexate-resistant strain of Escherichia coli B, MB 1428, is treated with approximately a 5 mol ratio of N-bromosuccinimide (NBS) to enzyme at pH 7.2 and assayed at the same pH, there is a 40% loss of activity due to the modification of 1 histidine residue and possibly 1 methionine residue before oxidation of tryptophan occurs. The initial modification is accompanied by a shift of the pH for maximal enzymatic activity from pH 7.2 to pH 5.5 Upon further treatment with N-bromosuccinimide, the activity is gradually reduced from 60 to 0% as tryptophan residues become oxidized. An NBS to enzyme mole ratio of approximately 20 results in 90% inactivation of the enzyme. When the enzyme is titrated with NBS in 6 M guanidine HCl, 5 mol of tryptophan react per mol of enzyme, a result in agreement with the total tryptophan content as determined by magnetic circular dichroism. The 40% NBS-inactivated sample posses full binding capacity for methotrexate and reduced triphosphopyridine nucleotide, and the Km values for dihydrofolate and TPNH are the same as for the native enzyme. After 90% inactivation, only half of the enzyme molecules bind methotrexate, and the dissociation constant for methotrexate is 40 nM as compared to 4 nM for native enzyme in solutions of 0.1 M ionic strength, pH 7.2 Also, TPNH is not bound as tightly to the modified enzyme-methotrexate complex as to the unmodified enzyme-methotrexate complex. Circular dichroism studies indicate the 90% NBS-inactivated enzyme has the same alpha helix content as the native enzyme but less beta structure, while the 40% inactivated enzyme is essentially the same as the native enzyme. Protection experiments were complicated by the fact that NBS reacts with the substrates and cofactors of the enzyme. Although protection of specific residues was not determined, it was clear that TPNH was partially protected from NBS reaction when bound to the enzyme, and the enzyme, and the enzyme was not inactivated by NBS until the TPNH had reacted.  相似文献   

3.
The TPNH- and O2-dependent drug hydroxylation system of liver microsomes has been studied using normal rats and rats in which the drug-hydroxylating activity has been enhanced by repeated injections of phenobarbital. The oxidative demethylation of aminopyrine is employed as an assay. Optimal conditions for the assay with regard to the concentrations of TPNH and aminopyrine are established. TPN inhibits the reaction in a competitive manner, similarly to its effect on the microsomal TPNH-cytochrome c reductase. Drug hydroxylation, but not the "TPNH oxidase," TPNH-cytochrome c, -2,6-dichlorophenolindophenol, or -neotetrazolium reductase reaction, or the TPNH-dependent lipid peroxidation, is blocked by carbon monoxide. Microsomes from phenobarbital-treated rats exhibit increased activities of the various TPNH-linked reductase reactions, parallel to the increased drug hydroxylation activity, whereas the "TPNH oxidase" activity does not change appreciably. Measurements with microsomes from drug-treated animals reveal a 1:1:1 stoichiometry of aminopyrine-dependent oxygen uptake, TPNH oxidation, and formaldehyde formation. Attempts to solubilize the drug-hydroxylating enzyme system are also presented. It is concluded that the drug-hydroxylating enzyme system involves the microsomal TPNH-cytochrome c reductase and CO-binding pigment, and a hypothetic reaction scheme accounting for the data presented is proposed.  相似文献   

4.
An unusual allantoinase from Dolichos biflorus has been purified 62-fold. The purified enzyme has an unusual pH activity profile with a shoulder at pH 4 and a peak at pH 7.5. This is due to a single enzyme which does not need metal ions for activation. In the fully reduced state the enzyme exhibits a single sharp peak at 7.5; when it is not in the sulfhydryl form (in the fully oxidized SS form?) the enzyme shows a single pH optimum at pH 4. Km values for (±)-allantoin were 5.5 mM at pH 4 and 1.43 mM at pH 7.5. Allantoinase activity has been demonstrated in the resting seed, and increased linearly with time during the first 5 days of seedling growth.  相似文献   

5.
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   

6.
Thiosulfate reductase has been purified approximately 70-fold from an extract of bakers' yeast. An enzyme with a molecular weight of 17,000, a Stokes radius of 19 Å, and a pI of 5.1 was obtained. Initial velocity and inhibition studies indicate that the substrates add in a random fashion. Further evidence suggests that the rapid-equilibrium assumption is not totally applicable. The enzyme has two distinct but closely situated substrate binding sites—one for compounds with an RSO3? structure and one for the sulfhydryl substrate.  相似文献   

7.
Diamine oxidase from Euphorbia characias latex contains two sulfhydryl groups per mole of dimeric enzyme. The sulfhydryl groups are unreactive in the native enzyme but can be readily titrated by 4,4′-dithiodipyridine after protein denaturation, or anaerobically in the presence of the amine substrate. In the presence of both substrates (diamine and oxygen) they react sluggishly. The sulfhydryl groups show different reactivity toward various reagents, but in every case their modification inhibits catalytic activity. The insensitivity of the native enzyme to specific reagents suggests that the sulfhydryl groups are positioned in the interior of the protein and shielded from the solvent. Their reactivity in the presence of the amine substrate could be attributed to a conformational change occurring upon substrate binding or after substrate oxidation.  相似文献   

8.
Cryptosporidium is the causative agent of a gastrointestinal disease, cryptosporidiosis, which is often fatal in immunocompromised individuals and children. Thymidylate synthase (TS) and dihydrofolate reductase (DHFR) are essential enzymes in the folate biosynthesis pathway and are well established as drug targets in cancer, bacterial infections, and malaria. Cryptosporidium hominis has a bifunctional thymidylate synthase and dihydrofolate reductase enzyme, compared to separate enzymes in the host. We evaluated lead compound 1 from a novel series of antifolates, 2-amino-4-oxo-5-substituted pyrrolo[2,3-d]pyrimidines as an inhibitor of Cryptosporidium hominis thymidylate synthase with selectivity over the human enzyme. Complementing the enzyme inhibition compound 1 also has anti-cryptosporidial activity in cell culture. A crystal structure with compound 1 bound to the TS active site is discussed in terms of several van der Waals, hydrophobic and hydrogen bond interactions with the protein residues and the substrate analog 5-fluorodeoxyuridine monophosphate (TS), cofactor NADPH and inhibitor methotrexate (DHFR). Another crystal structure in complex with compound 1 bound in both the TS and DHFR active sites is also reported here. The crystal structures provide clues for analog design and for the design of ChTS–DHFR specific inhibitors.  相似文献   

9.
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   

10.
Reduction of purothionin by the wheat seed thioredoxin system   总被引:1,自引:1,他引:0       下载免费PDF全文
Thioredoxin h, the thioredoxin characteristic of heterotrophic plant tissues, was purified to homogeneity from wheat endosperm (flour) and found to resemble its counterpart from carrot cell cultures. In the presence of NADPH, homogeneous thioredoxin h and partially purified wheat endosperm thioredoxin reductase (NADPH), (EC 1.6.4.5), purothionin promoted the activation of chloroplast fructose-1,6-bisphosphatase (EC 3.1.3.11). Under these conditions, NADPH provided the reducing equivalents for a series of thiol reactions in which (a) thioredoxin reductase reduced thioredoxin h thereby converting it from disulfide (S-S) to sulfhydryl (SH) form; (b) the sulfhydryl form of thioredoxin h reduced the disulfide form of purothionin—a 5 kilodalton seed storage protein with 4 S-S bridges; and (c) the sulfhydryl form of purothionin reductively activated fructose-1,6-bisphosphatase. The results show that, since thioredoxin h does not react effectively with fructose-1,6-bisphosphatase, the thioredoxin system can activate an enzyme through purothionin by secondary thiol redox control. In a related type reaction, purothionin, inhibited the activity of either Escherichia coli or calf thymus ribonucleotide reductase with reduced thioredoxin as hydrogen donor. The results suggest that purothionin competes with ribonucleotide reductase for reducing equivalents from thioredoxin. Thus, inhibition of deoxyribonucleotide synthesis should be considered a possible mechanism when examining the toxic effects of purothionin on mammalian cells in S-phase.  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.
Targeting dihydrofolate reductase, here, we report the tumor growth inhibitory activity of substituted acridones. The screening of the molecules over 60 cell line panel of human cancer cells identified (S)-oxiran-2-ylmethyl 9-oxo-9,10-dihydroacridine-4-carboxylate (19) with average GI50 0.3 μM. The specificity of the compound to CCRF-CEM, MOLT-4 and SR cell lines of leukemia and SW-620, SF268, LOXIMVI, ACHN and MCF7 cancerous cells exhibiting GI50 in the nM range was observed. C6 Glioma cells treated with compound 19 showed differentiated cell morphology and cell cycle arrest in G2/M phase. The interactions of the compound with dihydrofolate reductase were ascertained with the help of enzyme immunoassays, molecular docking and molecular dynamic studies.  相似文献   

13.
The extramitochondrial isozyme of aspartate aminotransferase (l-aspartate:2-oxoglutarate aminotransferase EC 2.6.1.1) contains a cysteinyl residue (cysteine-390) which, in the presence of substrate, displays enhanced reactivity toward sulfhydryl reagents. To gain insight into the structural similarity of the enzyme in solution compared to its crystalline state and into the type of structural change induced by substrates, the reactivity of Cys-390 in the crystalline enzyme has been studied. The flat yellow plates, crystallized from polyethylene glycol, form spectroscopically detectable enzyme-substrate complexes (C. M. Metzler, D. E. Metzler, D. S. Martin, R. Newman, A. Arnone, and P. Rodgers, 1978, J. Biol. Chem. 253, 5251–5254). The crystals, both in the presence and absence of the substrate pair, glutamate and α-ketoglutarate, were treated with N-ethylmaleimide or N-ethyl[1-14C]maleimide and the extent of the reaction was monitored by the colorimetric sulfhydryl reaction with 5,5′-dithiobis-2-nitrobenzoic acid, by amino acid analysis, by radioactivity incorporated, and by the measurement of enzyme activity. A cysteine residue was modified only in the presence of substrate; the crystals remained undamaged. Since, any large conformational change in the enzyme would either be prevented by the crystalline lattice or would disrupt its integrity, it is concluded that the enhanced reactivity of cysteine-390 in the presence of substrates must be due to only a small local conformational change in the substrate binding region.  相似文献   

14.
The presence of a stable allantoinase in Lathyrus sativas and its de novo synthesis at a maximal rate in the first 48 hr of germination have been demonstrated. The plumule and radicle together exhibited highest enzyme activity. L. sativas allantoinase has been purified nearly 35-fold. The purified enzyme was optically active around pH 7.5, did not require any metal ion for activity and exhibited a Km of 2·56 mM for (±)-allantoin, and an activation energy of 5·6 kcal/mol. Unlike other plant allantoinases, the L. sativus enzyme is highly specific for (±)-allantoin and is shown to be a sulfhydryl enzyme which apparently exists in a stable form in vivo obviating the need for added sulfhydryl compounds for maximal activity.  相似文献   

15.
NAD+-dependent and NADP+-dependent glyceraldehyde-3-phosphate (G-3-P) dehydrogenases were isolated from Euglena gracilis and characterized as to their physical and chemical parameters. NAD+-G-3-P dehydrogenase was found to have a strong resemblance to similar enzymes from muscle tissue. It has a molecular weight of about 140,000, four subunits of identical size and charge, and a single species of NH2-terminal amino acid. Two sulfhydryl groups per subunit are present, one of which is directly involved in the catalytic activity and is rapidly titratable. The enzyme also exhibits the “half the sites reactivity” of sulfhydryl groups as defined by O. P. Malhotra and S. A. Bernhard ((1968) J. Biol. Chem. 243, 1243). The pH and temperature optima are also similar to those of the enzymes from muscle tissue, as are the reaction kinetics and the strict specificity for NAD+.NADP+-dependent G-3-P dehydrogenase is different in many respects. Its molecular weight is slightly lower (~136,000) than that of the NAD+ enzyme, though it also consists of four subunits. It has a higher affinity for the reverse reaction substrates, in line with its probable function in vivo in CO2 fixation. There is only one sulfhydryl group per subunit, and that is not involved in activity, suggesting a difference in reaction mechanisms between the two enzymes. The NADP+-dependent enzyme exhibits activation by ATP, whereas the NAD+-dependent enzyme is competitively inhibited by this nucleotide.The greatest difference observed is in the physical characteristics of the enzymes. NADP+-G-3-P dehydrogenase was highly hydrophobic. Its solubility in a 10% aqueous solution of p-dioxane was approximately four to five times that of the NAD+-enzyme. Isolation of the enzyme was accomplished by fractionation in 1,2-dimethoxyethane, which also stabilized the enzymatic activity, as did aqueous p-dioxane. The high axial ratio of the NADP+-enzyme (~9) coupled with its very low degree of hydration as well as the high degree of amidation of the dicarboxylic amino acids (>90%) indicates that the exterior of the enzyme molecule is probably hydrophobic in nature. This is in agreement with its in vivo hydrophobic environment in the chloroplast membrane and explains the lability of the enzyme once extracted into an aqueous environment as well as its stabilization in solvents.  相似文献   

16.
An aldehyde reductase catalyzing the NADPH-dependent reduction of long-chain aldehydes has been purified 690-fold from bovine cardiac muscle. Based on the results obtained during gel filtration, this enzyme has an apparent molecular weight of 34,000. The pI of the aldehyde reductase was 6.1 and the enzymatic activity had a sharp pH optimum at 6.4. The enzyme catalyzed the reduction of aromatic aldehydes and aliphatic aldehydes having eight or more carbon atoms. Short-chain aldehydes, aldoses, or ketoses or long-chain methyl ketones were not utilized as substrates by this enzyme. However, the methyl ketone, pentadecan-2-one, was a competitive inhibitor of this enzyme with an apparent Ki = 10 μm when tetradecanal was the variable substrate. The reaction was not reversible when ethanol or hexadecanol was employed as substrate, utilizing either NAD+, or NADP+ as a cofactor. The addition of 10 mm pyrazole to the incubation medium had no effect on the enzymatic activity.  相似文献   

17.
The chemical modification of cysteine and tyrosine residues in formyltetrahydrofolate synthetase from Clostridium thermoaceticum has been examined relative to enzymatic activity and reactivity of these groups in the native protein. 4,4′-Dipyridyl disulfide, dansylaziridine, and fluorescein mercuric acetate all reacted with just one of six sulfhydryls per enzyme subunit, resulting in activities of 100, 95 and 70%, respectively. The Km values for MgATP, formate, and tetrahydrofolate were unaltered in the modified enzymes. ATP did produce a 2.5-fold reduction in the rate of reaction between the enzyme and 4,4′-dipyridyl disulfide. Tetranitromethane reacted most rapidly with a single sulfhydryl group per subunit to produce a 20–30% loss in activity. Subsequent additions of tetranitromethane modified 2.2 tyrosines per subunit which was proportional to the loss of the remaining enzymatic activity. Folic acid, a competitive inhibitor, protected against modification of the tyrosines and the associated activity losses; however, the oxidation of the single sulfhydryl group and the initial 20–30% activity loss were unaffected. In the presence of folic acid, higher concentrations of tetranitromethane produced a loss of the remaining activity proportional to the modification of 1.2 tyrosines per subunit. It is proposed that at least 1 tyrosine critical for enzymatic activity is located at or near the folic acid/tetrahydrofolate binding site.  相似文献   

18.
Asparagine synthetase was purified 240-fold from soybean (Glycine max (L.) Merr.) root nodules with a final recovery of 5% using Reactive Blue 2-crossed linked Agarose affinity gel chromatography. High levels of sulfhydryl protectants were required and the inclusion to glycerol and substrates in the extraction buffer helped to stabilize the enzyme. The final preparation had a specific activity of 3.77 mkat/kg protein when assayed at 30°C and was free of contaminating asparaginase activity. The enzyme had a broad pH maximum around pH 8.0 and apparent Km values for the substrates aspartate, Mg · ATP, and glutamine were 1.24 mM, 0.076 mM and 0.16 mM, respectively. Ammonium ion could partially replace glutamine as the nitrogen donor. Initial velocity patterns yielded parallel inverse plots with all substrate pairs suggesting an overall ping-pong reaction mechanism. Product inhibition patterns provided evidence that glutamine was the first substrate to bind to the enzyme and asparagine was the last product released.  相似文献   

19.
Sulfhydryl groups of bovine liver rhodanese (thiosulfate: cyanide sulfurtransferase, EC 2.8.1.1) were modified by treatment with tetrathionate. There was a linear relationship between loss of enzyme activity and the amount of tetrathionate used. At a ratio of one tetrathionate per mole of rhodanese, 100% of enzyme activity was lost in the sulfur-free E-form as compared with a 70% loss for the sulfur-containing ES-form of the enzyme. Addition of up to a 100-fold molar excess of tetrathionate to ES gave no further inactivation. Addition of cyanide to the maximally inactivated ES-tetrathionate complex gave complete loss of activity. Kinetic studies of maximally inactivated ES and partially inactivated E gave Km (K5) values that were essentially the same as native enzyme, indicating that the active enzyme, in all cases, bound thiosulfate-similarly. Reactivation was faster with the ES-form than with the E-form. The substrate, thiosulfate, could reactivate the enzyme up to 70% in 1 h with ES as compared to 24 h with E. Tetrathionate modification of rhodanese could be correlated with the changes in intrinsic fluorescence and with the binding of the active site reporter 2-anilinonaphthalene-8-sulfonic acid (2,8-ANS). Circular dichroism spectra of the protein suggested increased ordered secondary structure in the protein after reaction with tetrathionate. Cadmium chloride and phenylarsine oxide totally inactivated the enzyme at levels usually associated with their effect on enzymes containing vicinal sulfhydryl groups. Further, cadmium inhibition could be reserved by EDTA. Tetrathionate modification of rhodanese may proceed through the formation of sulfenylthiosulfate intermediates at sulfhydryl groups, close to but not identical with the active-site sulfhydryl group, which then can react further with the active-site sulfhydryl group to form disulfide bridges.  相似文献   

20.
Synthesis of long-chain fatty alcohols in preputial glands of mice is catalyzed by an NADPH-dependent acyl coenzyme A (CoA) reductase located in microsomal membranes; sensitivity to trypsin digestion indicates that the reductase is on the cytoplasmic side of the membrane. Results with pyrazole and phenobarbital demonstrate the reaction is not catalyzed by a nonspecific alcohol dehydrogenase or an aldehyde reductase. Acyl-CoA reductase activity is sensitive to sulfhydryl and serine reagent modification, is stimulated by bovine serum albumin, and produces an aldehyde intermediate. The activity is extremely detergent sensitive and cannot be restored even after removal of the detergents. Phospholipase C or asolectin treatment does not release the acyl-CoA reductase from microsomal membranes, but causes a significant decrease in the activity recovered in the membrane pellet. Glycerol does not solubilize the reductase activity, nor does 3.0 m NaCl; however, the combination of glycerol and 3.0 m NaCl did release about 50% of the acyl-CoA reductase from the microsomal pellet. Substrate concentration curves obtained in the presence or absence of bovine serum albumin show significant differences in enzyme activities. The reductase is sensitive to the concentration of palmitoyl-CoA and is progressively inhibited at levels beyond the critical micellar concentration of the substrate. The apparent Km for acyl-CoA reductase is 14 μm; however, the maximum velocity varies with the concentration of albumin used. Expression of enzyme activity in delipidated microsomes requires specific phospholipids, which suggests that in vivo regulation of acyl-CoA reductase activity could be achieved through modifications in membrane lipid composition.  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