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.
12-ketoursodeoxycholic acid (12-keto-UDCA) is a key intermediate for the synthesis of ursodeoxycholic acid (UDCA), an important therapeutic agent for non-surgical treatment of human cholesterol gallstones and various liver diseases. The goal of this study is to develop a new enzymatic route for the synthesis 12-keto-UDCA based on a combination of NADPH-dependent 7β-hydroxysteroid dehydrogenase (7β-HSDH, EC 1.1.1.201) and NADH-dependent 3α-hydroxysteroid dehydrogenase (3α-HSDH, EC 1.1.1.50). In the presence of NADPH and NADH, the combination of these enzymes has the capacity to reduce the 3-carbonyl- and 7-carbonyl-groups of dehydrocholic acid (DHCA), forming 12-keto-UDCA in a single step. For cofactor regeneration, an engineered formate dehydrogenase, which is able to regenerate NADPH and NADH simultaneously, was used. All three enzymes were overexpressed in an engineered expression host Escherichia coli BL21(DE3)Δ7α-HSDH devoid of 7α-hydroxysteroid dehydrogenase, an enzyme indigenous to E. coli, in order to avoid formation of the undesired by-product 12-chenodeoxycholic acid in the reaction mixture. The stability of enzymes and reaction conditions such as pH value and substrate concentration were evaluated. No significant loss of activity was observed after 5 days under reaction condition. Under the optimal condition (10 mM of DHCA and pH 6), 99 % formation of 12-keto-UDCA with 91 % yield was observed.  相似文献   

2.
Novel 4,5-allenic 3β-hydroxy-5,10-secosteroids have been synthesized by sodium borohydride reduction of the corresponding conjugated allenic 3-oxo-5, 10-secosteroids. The secosteroid allenic alcohols are substrates for bovine adrenal and human placental Δ5-3β-hydroxysteroid dehydrogenases, and the resulting electrophilic conjugated allenic ketones are shown to inactivate these dehydrogenases in a time-dependent manner. Inactivated enzyme did not recover activity after filtration through Sephadex G-25. In contrast, the secosteroid allenic alcohols were not oxidized at C-3 by the bacterial 3β(and 17β)-hydroxysteroid dehydrogenase from P. testosteroni, nor did the corresponding allenic ketones inactivate this enzyme when incubated directly.  相似文献   

3.
Glycyrrhetic acid, derived from a main component of liquorice, was converted to 3-ketoglycyrrhetic acid reversibly by rat liver homogenates in the presence of NADPH or NADP+. Glycyrrhetic acid-oxidizing and 3-ketoglycyrrhetic acid-reducing activities were localized in microsomes among the subcellular fractions of rat liver. Glycyrrhetic acid-oxidizing activity and 3-ketoglycyrrhetic acid-reducing activities showed pH optima at 6.3 and 8.5, respectively, and required NADP+ or NAD+ and NADPH or NADH, respectively, indicating that these activities were due to glycyrrhetinate dehydrogenase. The dehydrogenase was not solubilized from the membranes by the treatment with 1 M NaCl or sonication, indicating that the enzyme is a membrane component. The dehydrogenase was solubilized with detergents such as Emalgen 913, Triton X-100 and sodium cholate, and then separated from 3β-hydroxysteroid dehydrogenase (5β-androstan-3β-ol-17-one-oxidizing activity) by butyl-Toyopearl 650 M column chromatography. Partially purified enzyme catalyzed the reversible reaction between glycyrrhetic acid and 3-ketoglycyrrhetic acid, but was inactive toward 3-epiglycyrrhetic acid and other steroids having the 3β-hydroxyl group. The enzyme required NADP+ and NADPH for the highest activities of oxidation and reduction, respectively, and NAD+ and NADH for considerable activities, similar to the results with microsomes. From these results the enzyme is defined as glycyrrhetinate dehydrogenase, being quite different from 3β-hydroxysteroid dehydrogenase of Ruminococcus sp. from human intestine, which is active for both glycyrrhetic acid and steroids having the 3β-hydroxyl group.  相似文献   

4.
The cholesterol metabolism pathway in Mycobacterium tuberculosis (M. tb) is a potential source of energy as well as secondary metabolite production that is important for survival of M. tb in the host macrophage. Oxidation and isomerization of 3β-hydroxysterols to 4-en-3-ones is requisite for sterol metabolism and the reaction is catalyzed by 3β-hydroxysteroid dehydrogenase (Rv1106c). Three series of 6-azasteroids and 4-azasteroids were employed to define the substrate preferences of M. tb 3β-hydroxysteroid dehydrogenase. 6-Azasteroids with large, hydrophobic side chains at the C17 position are the most effective inhibitors. Substitutions at C1, C2, C4 and N6 were poorly tolerated. Our structure-activity studies indicate that the 6-aza version of cholesterol is the best and tightest binding competitive inhibitor (Ki = 100 nM) of the steroid substrate and are consistent with cholesterol being the preferred substrate of M. tb 3β-hydroxysteroid dehydrogenase.  相似文献   

5.
Chenodeoxycholate (3α-, 7α-dihydroxy-5β-cholanoate) was linked to Sepharose 4B by an ethylenediamine bridge. When 3α-hydroxysteroid dehydrogenase and 7α-hydroxysteroid dehydrogenase preparations were applied to a column of covalently linked Chenodeoxycholate, both enzymes were retarded at pH 6.7; the 7α-OH oriented enzyme more than the 3α-OH enzyme. Approximately forty-fold purification of 7a-hydroxysteroid dehydrogenase was achieved in one step. Although no significant purification of 3α-hydroxysteroid dehydrogenase occurred, the background value in the fluorometric enzymatic estimation of bile acids by eluted 3α-hydroxysteroid dehydrogenase was markedly reduced. Molecular weight estimation by Sephadex G-200 gave the values of 47,000 for 3α-hydroxysteroid dehydrogenase and 105,000 for 7α-hydroxy-steroid dehydrogenase.  相似文献   

6.
Cytochromes P-450 and b5 were observed in the microsomal fraction of interstitial tissue of rat testes. Microsomal cytochrome b5 was reduced by the NADH coupled with the activities of Δ5-3β-hydroxysteroid dehydrogenase with Δ54 isomerase through conversion of pregnenolone to progesterone. Activities of NADPH-supported 17α-hydroxylase and C-17-C-20 lyase which converted progesterone to androstenedione were stimulated by either the presence of NADH or the oxidative reaction by the dehydrogenase upon Δ5-3β-hydroxysteroids. Androstenedione production enhanced by the reaction of the dehydrogenase was decreased by addition of the antibody against NADH-cytochrome b5 reductase which was purified from rat hepatic microsomes, suggesting the active participation of cytochrome b5 in the androgen synthesis.  相似文献   

7.
Using RACE techniques we have cloned and sequenced one of the hamster liver 3-hydroxy-hexobarbital dehydrogenases which catalyze not only cyclic alcohols but also 17β-hydroxy-steroids and 3α-hydroxysteroids. The gene specific primers to 3-hydroxyhexobarbital dehydrogenase 1 (G2) were synthesized on the basis of its partial peptide sequences. The sequence of full length cDNA generated by 3′- and 5′-RACE PCR consisted of 1225 nucleotides including an open reading frame of 972 nucleotides encoding a protein of 323 amino acids. The deduced amino acid sequence matched exactly with the partial peptide sequences of hamster liver 3-hydroxyhexobarbital dehydrogenase 1 (G2). The sequence showed 84.5% identity to mouse liver 17β-dehydrogenase(A-specific), and 74–76% identity to human liver bile acid binding protein/3α-hydroxysteroid dehydrogenase (DD2), human liver 3α-hydroxysteroid dehydrogenase type I (DD4) and type II (DD3), and rabbit ovary 20α-hydroxysteroid dehydrogenase. The protein contains catalytic residues of aldo-keto reductases, Asp50, Tyr55, Lys84, His117. These results suggest that the hamster liver 3-hydroxyhexobarbital/17β(3α)-hydroxysteroid dehydrogenase belongs to aldo-keto reductase superfamily. The insert containing the full-length cDNA of 3-hydroxyhexobarbital dehydrogenase and vector specific overhang produced by PCR was annealed with pET-32 Xa/LIC vector. The plasmid was transformed into BL21 (DE3) cells containing pLysS. The recombinant enzyme was induced 1 mM IPTG. The expressed enzyme was produced as fusion protein and purified by nickel chelating affinity chromatography followed by POROS CM column chromatography and superdex 75 gel filtration. Molecular weight of the recombinant enzyme fused thioredoxin and his•tag was about 55 000 and that was 35 000 after Factor Xa protease treatment. The recombinant enzyme dehydrogenated 3-hydroxy-hexobarbital, 1-acenaphthenol, 2-cyclohexen-1-ol, testosterone, glycolithocholic acid as well as the native enzyme purified from hamster liver.  相似文献   

8.
A direct method for determination of Δ5 3β-hydroxysteroid dehydrogenase (3β-HSD) activity was employed in isolated Leydig cells (LC) derived from rats on fetal day 19 (F19) and postnatal (N) days 1,12,24, 34 and 45 and adults. The activity of 3β-HSD in the adult LC was 1.15 ± 0.02 (μmole/μg DNA/hr, mean ± SEM, n = 73). Activities in the other groups, expressed as a percentage of the respective adult control, were: F19-38%; N1-39%; N12-8%; N24-89%; N34-166%; and N45-118%. A good correlation was found between histochemical staining for 3β-HSD and the quantitive method employed. Using (3H)-DHA as a substrate, LC isolated from F19, n1 and N12 produced testosterone in appreciable amounts (41%, 55% and 20% of the toal products respectively) whereas at advanced stages of development (N24 to adulthood) the major product was androstenedione (93 ± 1%). These findings may be explained by the observed decrease in 17β-hydroxysteroid dehydrogenase (17β-HSD) activity, due to an insufficient supply of NADPH, in the older vs. earlier stages of development. This study indicates the presence of steroidogenic enzymatic activity in LC throughout development in the rat. It also provides a relatively simple in vitro model for studies of testicular regulation during development.  相似文献   

9.
Several oxidative enzymes in the testis of the teiid lizard Cnemidophorus tigris were studied histochemically. The cells of the circumtesticular sheath (Leydig cell tunic) are functionally equivalent to Leydig cells of the interstitium on the basis of similar histochemical reactions for the five enzyme systems studied. Both groups of cells were positive for 3β-hydroxysteroid dehydrogenase, 17β-hydroxysteroid dehydrogenase, NADH diaphorase, NADPH diaphorase, and glucose-6-phosphate dehydrogenase. These results support the hypothesis that the circumtesticular sheath has endocrine function as indicated by its vascularity and its ability to catalyze histochemical reactions involving steroid biosynthesis.  相似文献   

10.
The capability of granulosa and theca interna cells, from preovulatory follicles of the domestic hen, to metabolize steroid precursors was evaluated. Granulosa and theca interna cells were isolated from ovarian preovulatory follicles at three different developmental stages: F1, F3 and F5. Tritiated pregnenolone (P5), progesterone (P4), dehydroepiandrosterone (DHEA), androstenedione (A4) and testosterone (T) were employed as precursors and their metabolic products were evaluated. The major metabolite of P5 by granulosa cells was P4, but we also observed low amounts of 5β-pregnandione. DHEA metabolism by granulosa cells yielded mainly A4, and minute quantities of 5β-androstan-3,17-dione (5β-dione) were detected. The only significant metabolite obtained in granulosa cells from A4 was 5β-dione, whereas T was only transformed into A4. On the other hand, P5 metabolism by theca interna cells yielded A4 as the main product, also P4, 17α-OHP4, 17α-OHP5, 5β-pregnandione, and DHEA, were found. When DHEA was the precursor A4 was produced in higher amounts than 5β-dione. A4 was mainly transformed into 5β-dione. In similar conditions, T was transformed into A4. These results show that granulosa cells have enzymatic activities of 3β-hydroxysteroid dehydrogenase/5-4 isomerase (3β-HSD from P5 and DHEA), 17β-hydroxysteroid dehydrogenase (17β-HSD from T) and 5β-reductase (from P5, DHEA and A4). Whereas theca interna cells have enzymatic activities of cytochrome P450c17 (from P5 and P4), 3β-HSD (from P5 and DHEA), 17β-HSD (from T) and 5β-reductase (from P4, DHEA and A4). These data support the concept that theca interna cells have the ability to synthesize androgens from progestins produced in granulosa cells. In addition, since theca interna cells did not show the capacity to aromatize androgens suggests that interaction between theca interna and theca externa cells occurs in vivo, thus confirming the three cell model for estrogen production. Furthermore, the fact that other metabolites were produced both in granulosa and theca interna cells, but in a different extent, suggests that complex mechanisms are participating in the regulation of steroid synthesis in avian ovary follicles.  相似文献   

11.
The 3-α-hydroxysteroid dehydrogenase and the 3-β-hydroxysteroid dehydrogenase of Pseudomonas testosteroni were purified to homogeneity by polyacrylamide gel electrophoresis using the following stages: DEAE cellulose chromatography, affinity chromatography on oestrone-aminocaproate sepharose and Sephadex gel filtration.The pure 3-α-hydroxysteroid dehydrogenase was completely devoid of 3-β-hydroxysteroid dehydrogenase activity but could oxidize estradiol 17-β at an appreciable rate. This activity accounts for about 40 per cent of the total 17-β-estradiol dehydrogenase of the crude bacterial extract.Affinity labelling of pure 3-α-hydroxysteroid dehydrogenase was carried out using 5-β-pregnane 3,20-dione-12-α-iodoacetate and 5-α-androstane 3-one-17-β-bromoacetate. With both reagents, inactivation was obtained only in the presence of coenzyme, the substrate protected against inactivation and the enzyme was fully inhibited with covalent binding of 1 mole of reagent per mole of subunit suggesting an active site directed inhibition. Histidine and methionine were identified as the labelled aminoacid residues.  相似文献   

12.
11β-hydroxysteroid dehydrogenases regulate glucocorticoid concentrations and 17β-hydroxysteroid dehydrogenases regulate estrogen and androgen concentrations in mammals. Phylogenetic analysis of the sequences from two 11β-hydroxysteroid dehydrogenases and four mammalian 17β-hydroxysteroid dehydrogenases indicates unusual evolution in these enzymes. Type 1 11β- and 17β-hydroxysteroid dehydrogenases are on the same branch; Type 2 enzymes cluster on another branch with β-hydroxybutyrate dehydrogenase, 11-cis retinol dehydrogenase and retinol dehydrogenase; Type 3 17β-hydroxysteroid dehydrogenase is on a third branch; while the pig dehydrogenase clusters with a yeast multifunctional enzyme on a fourth branch. Pig 17β-hydroxysteroid dehydrogenase appears to have evolved independently from the other three 17β-hydroxysteroid dehydrogenase dehydrogenases; in which case, the evolution of 17β-hydroxysteroid dehydrogenase activity is an example of functional convergence. The phylogeny also suggests that independent evolution of specificity toward C11 substituents on glucocorticoids and C17 substituents on androgens and estrogens has occurred in Types 1 and 2 11β- and 17β-hydroxysteroid dehydrogenases.  相似文献   

13.
The proton magnetic resonance spectra of the dihydronicotinamide ring of αNADH3 and the nicotinamide ring of αNAD+ are reported and the proton absorptions assigned. The absolute assignment of the C4 methylene protons of αNADH is based on the generation of specifically deuterium-labeled (pro-S) B-deuterio-αNADH from enzymatically prepared B-deuterio-βNADH. The C4 proton absorption of αNAD+ is assigned by oxidation of B-deuterio-αNADH by the A specific, yeast alcohol dehydrogenase to yield 4-deuterio-αNAD+.The epimerization of either αNADH or βNADH yields an equilibrium ratio of approximately 9:1 βNADH to αNADH. The rate of epimerization of αNADH to βNADH at 38 °C in 0.05, pH 7.5, phosphate buffer is 3.1 × 10?3 min?1, corresponding to a half-life of 4 hr. Four related dehydrogenases, yeast and horse liver alcohol dehydrogenase and chicken M4 and H4 lactate dehydrogenase, are shown to oxidize αNADH to αNAD+ at rates three to four orders of magnitude slower than for βNADH. By using specifically labeled B-deuterio-αNADH the enzymatic oxidation by yeast alcohol dehydrogenase has been shown to occur with the identical stereospecificity as the oxidation of βNADH. The nonenzymatic epimerization of αNADH to βNADH and the enzymatic oxidation αNADH are discussed as a possible source of αNAD+in vivo.  相似文献   

14.
The synthesis and activity of a new series of non-steroidal inhibitors of 17β-hydroxysteroid dehydrogenase that are based on a 1,5-benzodiazepine scaffold are presented. Their inhibitory potential was screened against 17β-hydroxysteroid dehydrogenase from the fungus Cochliobolus lunatus (17β-HSDcl), a model enzyme of the short-chain dehydrogenase/reductase superfamily. Some of these compounds are potent inhibitors of 17β-HSDcl activity, with IC50 values in the low micromolar range and represent promising lead compounds that should be further developed and investigated as inhibitors of human 17β-HSD isoforms, which are the enzymes associated with the development of many hormone-dependent and neuronal diseases.  相似文献   

15.
Incubation of 3α, 20β-hydroxysteroid dehydrogenase (3α, 20β-HSD; E.C.1.1.1.53) with the nucleoside 5'-p-fluorosulfonylbenzoyladenosine (FSA) caused a time-dependent and irreversible loss in enzyme activity. Both 3α- and 20β-hydroxysteroid oxidoreductase activities decreased at equal rates by a first order kinetic process (in 0.05m phosphate buffer at pH 6.0 and 25°C, t12 = 170 min). Incubation of 3α, 20β-HSD was quenched by addition of 2-mercaptoethanol which instantaneously reacts with the fluorosulfonyl group of FSA. The cofactor NADH protected 3α, 20β-HSD against inactivation by FSA, in a concentration-dependent manner. However, progesterone did not protect 3α, 20β-HSD against inactivation by FSA. Evidently, FSA causes inactivation of the enzyme by irreversibly binding to the NADH-binding region at the active site of 3α, 20β-HSD. Both 3α- and 20β-hydroxysteroid oxidoreductase activities disappeared at equal rates under a variety of enzyme-inactivating conditions. These results suggest that both 3α- and 20β-activites occur at the same active site of 3α, 20β-HSD.  相似文献   

16.
Several NAD(P)+-dependent hydroxysteroid dehydrogenases, namely 3α-hydroxysteroid dehydrogenase, β-hydroxysteroid dehydrogenase, 7α-hydroxysteroid dehydrogenase, and 12α-hydroxysteroid dehydrogenase were separately immobilized on nylon tubes for the continuous-flow automated assay of hydroxysteroids. 3α-Hydroxysteroid dehydrogenase was also immobilized on pore glass. Spectrophotometric monitoring in the visible region, where blank values were markedly reduced, was achieved through the Meldola blue catalyzed transfer of hydrogen from NAD(P)H to a tetrazolium salt. Nylon-tube-immobilized enzymes maintained 45–55% of the original activity after 1 month of intermittent use. The operational range, using the “end point” approach, was 1–25 nmol of steroid and the assay speed 10–15 samples/h. Reliable results were obtained in the determination of 3α-hydroxysteroids and 3β,17β-hydroxysteroids in urine and total bile acids in serum.  相似文献   

17.
The synthesis of 4-ethenylidene-5α-androstane-3β, 17β-diol (5) and of 4-ethenylidene-5α-androstane-3,17-dione (4) is described. Compound 5 is a competitive inhibitor of solubilized bovine microsomal adrenal Δ5-3β-hydroxysteroid dehydrogenase, with Ki =2.7μM, and is converted by the enzyme to the corresponding 3-ketone. Compound 4 shown to irreversibly inactivate the enzyme in a time-dependent manner (t12 =31 min; 55μM; pH =7.0). The substrate, dehydroepiandrosterone, protects against inactivation by compound 4. In contrast, compound 5 is not oxidized at the 3-position by the 3β-(and 17β)-hydroxysteroid dehydrogenase from P. testosteroni, but is oxidized at the 17-position. Nevertheless, the 4-ethenylidene-3,17-diketone (4) causes irreversible time-dependent inactivation (t12 =28min; 64μM; pH =7.0) when incubated directly with this bacterial enzyme, acting as an affinity label.  相似文献   

18.
We (3,4) previously observed the reduction of 21-dehydrocorticosteroids in the presence of 20β-hydroxysteroid dehydrogenase proceeded at a faster rate than the reduction of the corresponding corticosteroids. The presence of adjacent carbonyl groups suggested the possibility that the increased rate of reduction of the 20-one,21-a1 steroid analogs resulted from a lack of specificity of the enzyme 20β-hydroxysteroid dehydrogenase for either the aldehyde or ketone group. Nuclear magnetic resonance spectroscopy indicated that the angular methyl groups of the steroid were sensitive probes for the constituents on the basic steroid skeleton. The C18 methyl resonance of 17,21-dihydroxy-4-pregnene-3,20-dione and 17-hydroxy-3,20-dioxo-4-pregnene-21-a1 were 0.722 ppm and 0.728 ppm respectively. The magnitude and sign of the change in chemical shift of the C18 methyl resonance for the enzymatic products of 17,21-dihydroxy-4-pregnene-3,20-dione and 17-hydroxy-3,20-dioxo-4-pregnene-21-a1 (+0.135 ppm and +0.144 ppm respectively) were consistent with a stereochemical assignment of 20β-hydroxyl.  相似文献   

19.
Ovarian slices of the European eel at the silver stage were incubated with 4 tritiated precursors (pregnenolone, progesterone, androstenedione, testosterone) in the presence or not of an inhibitor of 11β-hydroxylase activity, metopirone. Ether extracts were submitted to a gradient elution chromatography on celite columns. Isolated peaks were identified by isopolarity on TLC, microchemical reactions and recrystallization to constant specific activity. Interpretation of the results shows that the ovary of the European eel contains the following enzymes: a 3β-hydroxysteroid dehydrogenase, 5→4-ene-isomerase complex, a 17α-hydroxylase, a C21-C19 desmolase, a 17β-hydroxysteroid oxidoreductase, a 5α-reductase, a 3β-hydroxysteroid oxidoreductase and an aromatase complex. Metopirone effect indicates the presence of an 11β-hydroxylase activity. At this stage, 5β-reductase, 20β-reductase and 21-hydroxylase activities are not detected in the ovary. Water-soluble steroids were formed from all the precursors used. It appears that the ovarian biosynthesis is orientated towards the production of 5α-reduced compounds and that this might limit the production of 17β-estradiol by lowering the amount of disposable endogenous precursors.  相似文献   

20.
Perfluorooctane acid (PFOA) is classified as a persistent organic pollutant and as an endocrine disruptor. The mechanism by which PFOA causes reduced testosterone production in males is not known. We tested our hypothesis that PFOA interferes with Leydig cell steroidogenic enzymes by measuring its effect on 3β-hydroxysteroid dehydrogenase (3β-HSD) and 17β-hydroxysteroid dehydrogenase 3 (17β-HSD3) activities in rat testis microsomes and Leydig cells. The IC50s of PFOA and mode of inhibition were assayed. PFOA inhibited microsomal 3β-HSD with an IC50 of 53.2 ± 25.9 μM and 17β-HSD3 with an IC50 17.7 ± 6.8 μM. PFOA inhibited intact Leydig cell 3β-HSD with an IC50 of 146.1 ± 0.9 μM and 17β-HSD3 with an IC50 of 194.8 ± 1.0 μM. The inhibitions of 3β-HSD and 17β-HSD3 by PFOA were competitive for the substrates. In conclusion, PFOA inhibits 3β-HSD and 17β-HSD3 in rat Leydig cells.  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