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 203 毫秒
1.
Among the three closely related enzymes, lipoamide dehydrogenase, mercuric reductase, and glutathione reductase only the latter is inhibited by 2,4,6-trinitrobenzenesulfonate (TNBS). On the other hand, all three enzymes exhibit high rates of TNBS-dependent NADPH oxidation. In the case of glutathione reductase and mercuric reductase this TNBS-dependent activity displays substrate inhibition by excess of NADPH and is strongly stimulated by NADP+. The stimulation is especially pronounced with mercuric reductase, 25-fold under some conditions. Neither substrate inhibition nor stimulation by NAD+ is observed with lipoamide dehydrogenase.  相似文献   

2.
One- and two-electron reduction of quinones by glutathione reductase   总被引:1,自引:0,他引:1  
Yeast glutathione reductase (E.C. 1.6.4.2) catalyzes the oxidation of NADPH by p-quinones and ferricyanide with a maximal turnover number (TNmax) of 4-5 s-1.NADP+ stimulates the reaction and the TNmax/Km value of acceptors is reached at NADP+/NADPH greater than or equal to 100. TNmax is increased up to 30-33 s-1. The stimulatory effect of NADP+ may be associated with its complexation with the NADPH-binding site in the reduced enzyme (Kd = 40-60 microM). It is suggested that NADP+ shifts the electron density towards FAD in the two-electron-reduced enzyme and, evidently, changes its one-electron-reduction potentials, while quinones oxidize an equilibrium form of glutathione reductase containing reduced FAD. In the absence of NADP+ the reduction of quinones by glutathione reductase proceeds mainly in a two-electron manner. At NADP+/NADPH = 100 a one-electron reduction makes up 44% of the total process. At pH 6.0-7.0 the reduced forms of naphthoquinones undergo cyclic redox conversions. A hyperbolic dependence exists of the log TN/Km of quinones on their one-electron-reduction potentials.  相似文献   

3.
Takashi Iyanagi  Isao Yamazaki 《BBA》1970,216(2):282-294
The mitochondrial NADH dehydrogenase catalyzes a one-electron reduction of quinones. Semiquinones thus formed have the hyperfine structures of their free anion radicals and are suggested to be detached from the enzyme. In the presence of suitable electron acceptors electron transfer occurs from the semiquinone to the acceptor. The mechanism of quinone reduction by spinach ferredoxin-NADP reductase is the same as that by the NADH dehydrogenase.

On the other hand, the NAD(P)H dehydrogenase (DT-diaphorase) prepared from liver soluble fraction catalyzes a typical two-electron reduction of quinones such as p-benzoquinone and 2-methyl-1,4-naphthoquinone. The mechanisms of one-electron and two-electron reduction of quinones are readily distinguishable by the use of an electron spin resonance spectrometer equipped with a flow apparatus and also by the use of an appropriate set of electron acceptors.

It is concluded that the reduction of quinones and oxygen by flavoproteins falls into three mechanistic categories: one-electron, two-electron and mixed-type reactions.  相似文献   


4.
The nicotinamide nucleotide dimers (NAD)2 and (NADP)2, obtained by electrochemical reduction of NAD+ and NADP+, are able to reduce such single-electron acceptors as the proteins cytochrome c, azurin and methaemoglobin, though at different rates. Under the same conditions the reduced nicotinamide coenzymes NADH and NADPH are not able to reduce these proteins at measurable rates unless a catalyst (phenazine methosulphate or NADH-cytochrome c reductase in the case of cytochrome) is present. The redox mechanism seems to involve the formation of an NAD(P). radical that in the presence of O2 gives rise to superoxide (O2.-), since superoxide dismutase inhibited these reactions.  相似文献   

5.
Geobacter sulfurreducens strain PCA oxidized acetate to CO2 via citric acid cycle reactions during growth with acetate plus fumarate in pure culture, and with acetate plus nitrate in coculture with Wolinella succinogenes. Acetate was activated by succinyl-CoA:acetate CoA-transferase and also via acetate kinase plus phosphotransacetylase. Citrate was formed by citrate synthase. Soluble isocitrate and malate dehydrogenases NADP+ and NAD+, respectively. Oxidation of 2-oxoglutarate was measured as benzyl viologen reduction and strictly CoA-dependent; a low activity was also observed with NADP+. Succinate dehydrogenase and fumarate ductase both were membrane-bound. Succinate oxidation was coupled to NADP+ reduction whereas fumarate reduction was coupled to NADPH and NADH Coupling of succinate oxidation to NADP+ or cytochrome(s) reduction required an ATP-dependent reversed electron transport. Net ATP synthesis proceeded exclusively through electron transport phosphorylation. During fumarate reduction, both NADPH and NADH delivered reducing equivalents into the electron transport chain, which contained a menaquinone. Overall, acetate oxidation with fumarate proceeded through an open loop of citric acid cycle reactions, excluding succinate dehydrogenase, with fumarate reductase as the key reaction for electron delivery, whereas acetate oxidation in the syntrophic coculture required the complete citric acid cycle.  相似文献   

6.
The lpdA (Rv3303c) gene from Mycobacterium tuberculosis encoding a new member of the flavoprotein disulfide reductases was expressed in Escherichia coli, and the recombinant LpdA protein was purified to homogeneity. LpdA is a homotetramer and co-purifies with one molecule of tightly but noncovalently bound FAD and NADP+ per monomer. Although annotated as a probable lipoamide dehydrogenase in M. tuberculosis, LpdA cannot catalyze reduction of lipoyl substrates, because it lacks one of two cysteine residues involved in dithiol-disulfide interchange with lipoyl substrates and a His-Glu pair involved in general acid catalysis. The crystal structure of LpdA was solved by multiple isomorphous replacement with anomalous scattering, which confirmed the absence of these catalytic residues from the active site. Although LpdA cannot catalyze reduction of disulfide-bonded substrates, it catalyzes the NAD(P)H-dependent reduction of alternative electron acceptors such as 2,6-dimethyl-1,4-benzoquinone and 5-hydroxy-1,4-naphthaquinone. Significant primary deuterium kinetic isotope effects were observed with [4S-2H]NADH establishing that the enzyme promotes transfer of the C4-proS hydride of NADH. The absence of an isotope effect with [4S-2H]NADPH, the low Km value of 0.5 microm for NADPH, and the potent inhibition of the NADH-dependent reduction of 2,6-dimethyl-1,4-benzoquinone by NADP+ (Ki approximately 6 nm) and 2'-phospho-ADP-ribose (Ki approximately 800 nm), demonstrate the high affinity of LpdA for 2'-phosphorylated nucleotides and that the physiological substrate/product pair is NADPH/NADP+ rather than NADH/NAD+. Modeling of NADP+ in the active site revealed that LpdA achieves the high specificity for NADP+ through interactions involving the 2'-phosphate of NADP+ and amino acid residues that are different from those in glutathione reductase.  相似文献   

7.
Nitrofurans inhibit the oxidation of NADPH by glutathione, catalyzed by yeast glutathione reductase (EC 1.6.4.2). acting as uncompetitive incomplete inhibitors for NADPH and glutathione. The quinoline-substituted nitrofurans were the most effective inhibitors. These compounds increased the turnover numbers of enzyme at fixed concentrations of reduced glutathione, in the reverse reaction of glutathione reductase, but in most cases diminished the affinity of the enzyme for NAD+. Nitrofurans are weak one-electron oxidants of glutathione reductase. Their reactivity is close to that of p-quinones possessing the analoguous one-electron reduction potential (Cénas, N.K., Rakauskiené, G.A. and Kulys, J.J. (1989) Biochim. Biophys. Acta 973, 399-404), and reaction is stimulated by NADP+. It is assumed, that nitrofurans bind to the 'regulative' site of glutathione reductase (Karplus, P.A., Pai, E.F. and Schulz, G.E. (1989) Eur. J. Biochem. 178, 693-703).  相似文献   

8.
Glutathione reductase from S. cerevisiae (EC 1.6.4.2) catalyzes the NADPH oxidation by glutathione in accordance with a "ping-pong" scheme. The catalytic constant kcat) is 240 s-1 (pH 7.0, 25 degrees C); kcat for the diaphorase reaction is 4-5 s-1. The enzyme activity does not change markedly at pH 5.5-8.0. At pH less than or equal to 7.0, NADP+ acts as a competitive inhibitor towards NADPH and as a noncompetitive inhibitor towards glutathione. NADP+ increases the diaphorase activity of the enzyme. The maximal activity is observed, when the NADP+/NADPH ratio exceeds 100. At pH 8.0, NADP+ acts as a mixed type inhibitor during the reduction of glutathione. High concentrations of NADP+ also inhibit the diaphorase activity due to the reoxidation of the reduced enzyme by NADP+ at pH 8.0. The redox potential of glutathione reductase calculated from the inhibition data is--306 mV (pH 8.0). Glutathione reductase reduces quinoidal compounds in an one-electron way. The hyperbolic dependence of the logarithm of the oxidation constant on the one electron reduction potential of quinone is observed. It is assumed that quinones oxidize the equilibtium fraction of the two-electron reduced enzyme containing reduced FAD.  相似文献   

9.
Low concentrations of NAD+ inhibit the NADH: acceptor reductase reactions catalyzed by soluble NADH dehydrogenase from bovine heart mitochondria. The degree of incomplete inhibition of the enzyme depends on the nature and concentration of artificial electron acceptors and is manifested only at low concentrations of the latter. Marked inhibition was demonstrated for the 2.6-dichlorophenolindophenol-, ferricyanide- and O2-reductase reactions, being weakly pronounced during the measurement of the NADH: cytochrome c reductase activity. The inhibition of the above reactions by oxidized NAD+ isn't competitive towards NADH. A kinetic scheme is proposed, which postulates NADH: acceptor reductase reactions occurrence via two mechanisms, namely, a ping-pong mechanism and oxidation of the product-enzyme complex by the acceptor. It was shown that low concentrations of NAD+ also inhibit the NADH oxidase reaction catalyzed by complex I.  相似文献   

10.
The reactions of NADPH oxidation by quinones and inorganic complexes catalyzed by NADPH: adrenodoxin reductase were studied. The catalytic constant for the enzyme at pH 7.0 is 20-25 s-1; the oxidative constants for the quinones vary from 5 X 10(5) to 1.1 X 10(3) M-1 s-1 and show an increase with a rise in the one-electron acceptor reduction potential. The mode of adrenodoxin reductase interaction with oxyquinones differs from that of the enzyme interaction with alkyl-substituted quinones and inorganic complexes. NADPH competitively inhibits electron acceptors, whereas NADP+ is a competitive inhibitor of NADPH and a uncompetitive inhibitor of electron acceptors. (Ki = 25 microM). The depth of FAD incorporation into the enzyme molecule as calculated according to the outer sphere electron transfer theory is 6.1 A.  相似文献   

11.
The S9 fraction of MCF-7 human breast carcinoma cells has NAD(P)H (quinone-acceptor) oxidoreductase activity as measured by the reduction of dichlorophenol-indophenol (DCPIP). This reduction is dependent on the activators Tween-20 and bovine serum albumin and it is inhibitable by dicumarol. The S9 fraction also has cytochrome c reductase activity which is approximately 29 times less than the two-electron reduction activity of NAD(P)H (quinone-acceptor) oxidoreductase. Diaziquone (AZQ) is a substrate for this NAD(P)H oxidoreductase active S9 fraction as judged by its enzymatic reduction detected spectrophotometrically and by electron spin resonance spectroscopy. Two-electron mediated enzymatic reduction of AZQ was evidenced by the formation of the colorless dihydroquinone (AZQH2) which could be followed at 340 nm. The production of the dihydroquinone was inhibitable by dicumarol implicating NAD(P)H oxidoreductase in its formation. Under aerobic conditions, electron spin resonance spectroscopy showed evidence for the production of AZQ semiquinone (AZQH) and oxygen radicals. Under anaerobic conditions no oxygen radicals were observed, but the semiquinone was stable for hours. These results are also inhibitable by dicumarol and suggest a two-step one-electron oxidation process of the dihydroquinone. The production of semiquinone and oxygen radicals as detected by electron spin resonance spectroscopy was more sensitive to dicumarol when NADPH was used as cofactor (68% inhibition of OH and 65% inhibition of AZQH) than when NADH was used (28% inhibition of OH and 5% inhibition of AZQH). This suggests that NADH flavin reductases play a more important role in the one-electron reduction pathway of AZQ in MCF-7 S9 fraction than NADPH reductases. The reduction of AZQ by NAD(P)H (quinone-acceptor) oxidoreductase may play an important role in the bioreductive alkylating properties of AZQ.  相似文献   

12.
Anaerobic reduction of the flavoprotein adrenodoxin reductase with NADPH yields a spectrum with long wavelength absorbance, 750 nm and higher. No EPR signal is observed. This spectrum is produced by titration of oxidized adrenodoxin reductase with NADPH, or of dithionite-reduced adrenodoxin reductase with NADP+. Both titrations yield a sharp endpoint at 1 NADP(H) added per flavin. Reduction with other reductants, including dithionite, excess NADH, and catalytic NADP+ with an NADPH generating system, yields a typical fully reduced flavin spectrum, without long wavelength absorbance. The species formed on NADPH reduction appears to be a two-electron-containing complex, with a low dissociation constant, between reduced adrenodoxin reductase and NADP+, designated ARH2-NADP+. Titration of dithionite-reduced adrenodoxin reductase with NADPH also produces a distinctive spectrum, with a sharp endpoint at 1 NADPH added per reduced flavin, indicating formation of a four-electron-containing complex between reduced adrenodoxin reductase and NADPH. Titration of adrenodoxin reductase with NADH, instead of NADPH, provides a curved titration plot rather than the sharp break seen with NADPH, and permits calculation of a potential for the AR/ARH2 couple of -0.291 V, close to that of NAD(P)H (-0.316 V). Oxidized adrenodoxin reductase binds NADP+ much more weakly (Kdiss=1.4 X 10(-5) M) than does reduced adrenodoxin reductase, with a single binding site. The preferential binding of NADP+ to reduced enzyme permits prediction of a more positive oxidation-reduction potential of the flavoprotein in the presence of NADP+; a change of about + 0.1 V has been demonstrated by titration with safranine T. From this alteration in potential, a Kdiss of 1.0 X 10(-8) M for binding of NADP+ to reduced adrenodoxin reductase is calculated. It is concluded that the strong binding of NADP+ to reduced adrenodoxin reductase provides the thermodynamic driving force for formation of a fully reduced flavoprotein form under conditions wherein incomplete reduction would otherwise be expected. Stopped flow studies demonstrate that reduction of adrenodoxin reductase by equimolar NADPH to form the ARH2-NADP+ complex is first order (k=28 s-1). When a large excess of NADPH is used, a second apparently first order process is observed (k=4.25 s-1), which is interpreted as replacement of NADPH for NADP+ in the ARH2-NADP+ complex. Comparison of these rate constants to catalytic flavin turnover numbers for reduction of various oxidants by NADPH, suggests an ordered sequential mechanism in which reduction of oxidant is accomplished by the ARH2-NADP+ complex, followed by dissociation of NADP+. The absolute dependence of NADPH-cytochrome c reduction on both adrenodoxin reductase and adrenodoxin is confirmed...  相似文献   

13.
Campylobacter species are rich in c-type cytochromes, including forms which bind carbon monoxide. The role of the various forms of cytochromes in Campylobacter fetus has been examined in cell-free preparations by using physiological electron donor and acceptor systems. Under anaerobic conditions, NADPH reduced essentially all of the cytochrome c in crude cell extracts, whereas the reduction level with succinate was 50 to 60%. The carbon monoxide spectrum with NADPH was predominated by the cytochrome c complex; evidence of a cytochrome o type was seen in the succinate-reduced extracts and in membrane fractions. Succinate-reduced cytochrome c was oxidized by oxygen via a cyanide-sensitive, membrane-associated system. NADPH-reduced cytochrome c was oxidized by a cyanide-insensitive system. Partially purified carbon monoxide-binding cytochrome c, isolated from the cytoplasm, could serve as electron acceptor for NADPH-cytochrome c oxidoreductase; the reduced cytochrome was oxidized by oxygen by a cyanide-insensitive system present in the cytoplasmic fraction. Horse heart cytochrome c was also reducible by NADPH and by succinate; the reduced cytochrome was oxidized by a cyanide-sensitive system in the membrane fraction. NADPH and NADH oxidase activities were observed aerobically and under anaerobic conditions with fumarate. NADPH was more active than NADH. NADP was also more effective than NAD as an electron acceptor for the coenzyme A-dependent pyruvate and alpha-ketoglutarate dehydrogenase activities found in crude extracts. These dehydrogenases used methyl viologen and metronidazole as electron acceptors; they could be loci for oxygen inhibition of growth. It is proposed that energy provision via the high-potential cytochrome c oxidase system in the cytoplasmic membrane is limited by oxygen-sensitive primary dehydrogenases and that the carbon monoxide-binding cytochrome c may have a role as an oxygen scavenger.  相似文献   

14.
We studied changes induced by cold on electron transfer pathways (linked to NADH or NADPH oxidation) in endoplasmic reticulum of rapeseed hypocotyls (Brassica napus L.) from a freezing-sensitive variety (ISL) and freezing-tolerant variety (Tradition). Plantlets were grown at 22 degrees C then submitted to a cold shock of 13 or 35 days at 4 degrees C. We measured the content in NADH, NADPH, NAD and NADP of the hypocotyls and the redox power was estimated by the reduced versus oxidized nucleotide ratio. The contents in cytochromes b (5) and P-450, electron acceptors of NADH and NADPH respectively, were determined by differential spectrophotometry. Finally, activity of both NADH-cytochrome b (5) reductase (E.C.1.6.2.2) and NADPH cytochrome P-450 reductase (E.C.1.6.2.4) was determined by reduction of exogenous cytochrome c. Results show that during cold shock, along with an increase of linolenic acid content, there was a general activation of the NADPH pathway which was observed more quickly in Tradition plantlets than in ISL ones. Due to transfer of electrons that can occur between NADPH reductase and cytochrome b (5), this could favor fatty acid desaturation in Tradition, explaining why linolenic acid accumulation was more pronounced in this variety. Besides, more cytochrome P-450 accumulated in ISL that could compete for electrons needed by the FAD3 desaturase, resulting in a relative slower enrichment in 18:3 fatty acid in these plantlets.  相似文献   

15.
Yeast glutathione reductase exists in a single molecular form which exhibits preferred NADPH and weak NADH linked multifunctional activities. Kinetic parameters for the NADPH and NADH linked reductase, transhydrogenase, electron transferase and diaphorase reactions have been determined. The functional preference for the NADPH linked reductase reaction is kinetically related to the high catalytic efficiency and low dissociation constants for substrates. NADP+ and NAD+ may interact with two different sites or different kinetic forms of the enzyme. The active site disulfide and histidine are required for the reductase activity but are not essential to the transhydrogenase, electron transferase and diaphorase activities. Amidation of carboxyl groups and Co(II) chelation of glutathione reductase facilitate the electron transferase reaction presumably by encouraging the formation of an anionic flavosemiquinone.  相似文献   

16.
17.
Monodehydroascorbate reductase (EC 1.6.5.4) was purified from cucumber fruit to a homogeneous state as judged by polyacrylamide gel electrophoresis. The cucumber monodehydroascorbate reductase was a monomer with a molecular weight of 47,000. It contained 1 mol of FAD/mol of enzyme which was reduced by NAD(P)H and reoxidized by monodehydroascorbate. The enzyme had an exposed thiol group whose blockage with thiol reagents inhibited the electron transfer from NAD(P)H to the enzyme FAD. Both NADH and NADPH served as electron donors with Km values of 4.6 and 23 microM, respectively, and Vmax of 200 mol of NADH and 150 mol of NADPH oxidized mol of enzyme-1 s-1. The Km for monodehydroascorbate was 1.4 microM. The amino acid composition of the enzyme is presented. In addition to monodehydroascorbate, the enzyme catalyzed the reduction of ferricyanide and 2,6-dichloroindophenol but showed little reactivity with calf liver cytochrome b5 and horse heart cytochrome c. The kinetic data suggested a ping-pong mechanism for the monodehydroascorbate reductase-catalyzed reaction. Cucumber monodehydroascorbate reductase occurs in soluble form and can be distinguished from NADPH dehydrogenase, NADH dehydrogenase, DT diaphorase, microsome-bound NADH-cytochrome b5 reductase, and NADPH-cytochrome c reductase by its molecular weight, amino acid composition, and specificity of electron acceptors and donors.  相似文献   

18.
Chemical modification of rat hepatic NADPH-cytochrome P-450 reductase by sodium 2,4,6-trinitrobenzenesulfonate (TNBS) resulted in a time-dependent loss of the reducing activity for cytochrome c. The inactivation exhibited pseudo-first-order kinetics with a reaction order approximately one, and a second-order constant of 4.8 min-1 X M-1. The reducing activities for 2,6-dichloroindophenol and K3Fe(CN)6 were also decreased by TNBS. Almost complete protection of the NADPH-cytochrome P-450 reductase from inactivation by TNBS was achieved by NADP(H), while partial protection was obtained with a high concentration of NADH. NAD, FAD and FMN showed no effect against the inactivation. 3-Acetylpyridine-adenine dinucleotide phosphate, adenosine 2',5'-bisphosphate and 2'AMP protected the enzyme against the chemical modification. Stoichiometric studies showed that the complete inactivation was caused by modification of three lysine residues per molecule of the enzyme. But, under the conditions where the inactivation was almost protected by NADPH, two lysine residues were modified. From those results, we propose that one residue of lysine is located at the binding site of the 2'-phosphate group on the adenosine ribose of NADP(H), and plays an essential role in the catalytic function of the NADPH-cytochrome P-450 reductase.  相似文献   

19.
Evidence suggesting that Bacillus polymyxa has an active ferredoxin-NADP(+) reductase (EC 1.6.99.4) was obtained when NADPH was found to provide reducing power for the nitrogenase of this organism; direct evidence was provided when it was shown that B. polymyxa extracts could substitute for the native ferredoxin-NADP(+) reductase in the photochemical reduction of NADP(+) by blue-green algal particles. The ferredoxin-NADP(+) reductase was purified about 80-fold by a combination of high-speed centrifugation, ammonium sulfate fractionation, and chromatography on Sephadex G-100 and diethylaminoethyl-cellulose. The molecular weight was estimated by gel filtration to be 60,000. A small amount of the enzyme was further purified by polyacrylamide gel electrophoresis and shown to be a flavoprotein. The reductase was specific for NADPH in the ferredoxin-dependent reduction of cytochrome c and methyl viologen diaphorase reactions; furthermore, NADP(+) was the acceptor of preference when the electron donor was photoreduced ferredoxin. The reductase also has an irreversible NADPH-NAD(+) transhydrogenase (reduced-NADP:NAD oxidoreductase, EC 1.6.1.1) activity, the rate of which was proportional to the concentration of NAD (K(m) = 5.0 x 10(-3)M). The reductase catalyzed electron transfer from NADPH not only to B. polymyxa ferredoxin but also to the ferredoxins of Clostridium pasteurianum, Azotobacter vinelandii, and spinach chloroplasts, although less effectively. Rubredoxin from Clostridium acidi-urici and azotoflavin from A. vinelandii also accept electrons from the B. polymyxa reductase. The pH optima for the various reactions catalyzed by the B. polymyxa ferredoxin-NADP reductase are similar to those of the chloroplast reductase. NAD and acetyl-coenzyme A, which obligatorily activate NADPH- and NADH-ferredoxin reductases, respectively, in Clostridium kluyveri, have no effect on B. polymyxa reductase.  相似文献   

20.
Rat liver 3-hydroxy-3-methylglutaryl-CoA (HMG-CoA) reductase catalyzes, in addition to its normal biosynthetic or forward reaction (HMG-CoA + 2 NADPH + 2H+----mevalonate + 2 NAD+ + CoASH), the reverse reaction (mevalonate + CoASH + 2 NADP+----HMG-CoA + 2 NADPH + 2H+) and two "half-reactions" that involve the presumed intermediate mevaldate (mevaldate + CoASH + NADP+----HMG-CoA + NADPH + H+ and mevaldate + NADPH + H+----mevalonate + NADP+). These reactions were studied using both enzyme solubilized by the traditional freeze-thaw method and enzyme solubilized with a nonionic detergent in the presence of inhibitors of proteolysis. All four reactions were inhibited by mevinolin, a known inhibitor of the forward (biosynthetic) reaction catalyzed by HMG-CoA reductase. When the enzyme was inactivated by ATP and a cytosolic, ADP-dependent HMG-CoA reductase kinase, the rates of both the forward reaction and the half-reactions decreased to comparable extents. Although coenzyme A is not a stoichiometric participant in the second half-reaction (mevaldate + NADPH + H+----mevalonate + NADP+), it was required as an activator of this reaction. This observation implies that coenzyme A may remain bound to the enzyme throughout the normal catalytic cycle of HMG-CoA reductase.  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