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.
Ye Q  Hyndman D  Li X  Flynn TG  Jia Z 《Proteins》2000,38(1):41-48
Chinese hamster ovary (CHO) reductase is an enzyme belonging to the aldo-keto reductase (AKR) superfamily that is induced by the aldehyde-containing protease inhibitor ALLN (Inoue, Sharma, Schimke, et al., J Biol Chem 1993;268: 5894). It shows 70% sequence identity to human aldose reductase (Hyndman, Takenoshita, Vera, et al., J Biol Chem 1997;272:13286), which is a target for drug design because of its implication in diabetic complications. We have determined the crystal structure of CHO reductase complexed with nicotinamide adenine dinucleotide phosphate (NADP)+ to 2.4 A resolution. Similar to aldose reductase and other AKRs, CHO reductase is an alpha/beta TIM barrel enzyme with cofactor bound in an extended conformation. All key residues involved in cofactor binding are conserved with respect to other AKR members. CHO reductase shows a high degree of sequence identity (91%) with another AKR member, FR-1 (mouse fibroblast growth factor-regulated protein), especially around the variable C-terminal end of the protein and has a similar substrate binding pocket that is larger than that of aldose reductase. However, there are distinct differences that can account for differences in substrate specificity. Trp111, which lies horizontal to the substrate pocket in all other AKR members is perpendicular in CHO reductase and is accompanied by movement of Leu300. This coupled with movement of loops A, B, and C away from the active site region accounts for the ability of CHO reductase to bind larger substrates. The position of Trp219 is significantly altered with respect to aldose reductase and appears to release Cys298 from steric constraints. These studies show that AKRs such as CHO reductase are excellent models for examining the effects of subtle changes in amino acid sequence and alignment on binding and catalysis.  相似文献   

2.
The yakC gene in Schizosaccharomyces pombe, which encodes yakC protein (YakC), a potential member of an aldo-keto reductase (AKR) family, was cloned and expressed in Escherichia coli cells. The recombinant YakC purified to homogeneity catalyzed the reduction of 2-nitrobenzaldehyde (k(cat), 44.1 s(-1), K(m), 0.185 +/- 0.018 mM), 2-phthalaldehyde (19.8, 0.333 +/- 0.032), and pyridine-2-aldehyde (7.64, 0.302 +/- 0.028). Neither pyridoxal nor other compounds examined acted as substrates. NADPH, but not NADH, was a hydrogen donor. The enzyme is a monomer with a molecular weight of 38,900 +/- 6,600 (SDS-PAGE). The amino acid sequence deduced from yakC showed the highest (34%) identity with that of pyridoxal reductase (AKR8A1) among the identified AKRs. Twenty-one function-unknown proteins showed 40% or higher identity to the deduced amino acid sequence: DR2261 protein of Deionococcus radiodurans showed the highest (50%) identity. The predicted secondary structure of YakC is similar to that of human aldose reductase, a representative AKR. The results establish YakC as the first member of a new AKR family, AKR13. The yeast cells contained enzyme(s) other than YakC and pyridoxal reductase with the ability to reduce 2-nitrobenzaldehyde: total (100%) activity in the crude extract consisted of about 23% YakC, about 44% pyridoxal reductase, and about 33% other enzyme(s).  相似文献   

3.
A novel aldo-keto reductase (AKR) from Escherichia coli has been cloned, expressed and purified. This protein, YghZ, is distantly related (<40%) to mammalian aflatoxin dialdehyde reductases of the aldo-keto reductase AKR7 family and to potassium channel beta-subunits in the AKR6 family. The enzyme has been placed in a new AKR family (AKR14), with the designation AKR14A1. Sequences encoding putative homologues of this enzyme exist in many other bacteria. The enzyme can reduce several aldehyde and diketone substrates, including the toxic metabolite methylglyoxal. The K(m) for the model substrate 4-nitrobenzaldehyde is 1.06 mM and for the endogenous dicarbonyl methylglyoxal it is 3.4 mM. Overexpression of the recombinant enzyme in E. coli leads to increased resistance to methylglyoxal. It is possible that this enzyme plays a role in the metabolism of methylglyoxal, and can influence its levels in vivo.  相似文献   

4.
The AKRs (aldo-keto reductases) are a superfamily of enzymes which mainly rely on NADPH to reversibly reduce various carbonyl-containing compounds to the corresponding alcohols. A small number have been found with dual NADPH/NADH specificity, usually preferring NADPH, but none are exclusive for NADH. Crystal structures of the dual-specificity enzyme xylose reductase (AKR2B5) indicate that NAD+ is bound via a key interaction with a glutamate that is able to change conformations to accommodate the 2'-phosphate of NADP+. Sequence comparisons suggest that analogous glutamate or aspartate residues may function in other AKRs to allow NADH utilization. Based on this, nine putative enzymes with potential NADH specificity were identified and seven genes were successfully expressed and purified from Drosophila melanogaster, Escherichia coli, Schizosaccharomyces pombe, Sulfolobus solfataricus, Sinorhizobium meliloti and Thermotoga maritima. Each was assayed for co-substrate dependence with conventional AKR substrates. Three were exclusive for NADPH (AKR2E3, AKR3F2 and AKR3F3), two were dual-specific (AKR3C2 and AKR3F1) and one was specific for NADH (AKR11B2), the first such activity in an AKR. Fluorescence measurements of the seventh protein indicated that it bound both NADPH and NADH but had no activity. Mutation of the aspartate into an alanine residue or a more mobile glutamate in the NADH-specific E. coli protein converted it into an enzyme with dual specificity. These results show that the presence of this carboxylate is an indication of NADH dependence. This should allow improved prediction of co-substrate specificity and provide a basis for engineering enzymes with altered co-substrate utilization for this class of enzymes.  相似文献   

5.
In many organisms, glycogen gives rise to 1,5-anhydro-D-fructose (AF), which is reduced to 1,5-anhydro-D-glucitol (AG). AF reductase, which catalyzes the latter reaction, was purified from pig liver, but mouse ortholog has not yet been reported. In the database, aldo-keto reductase family 1, member E1 (AKR1E1) showed highest homology to pig enzyme. We confirmed that cloned AKR1E1 is mouse ortholog based on enzymatic properties of purified recombinant protein.  相似文献   

6.
Mitochondrial thioredoxin reductase was purified from bovine adrenal cortex. The enzyme is a first protein component in the mitochondrial thioredoxin-dependent peroxide reductase system. The purified reductase exhibited an apparent molecular mass of 56 kDa on SDS/PAGE, whereas the native protein was about 100 kDa, suggesting a homodimeric structure. It catalysed NADPH-dependent reduction of 5, 5'dithiobis(2-nitrobenzoic acid) and thioredoxins from various origins but not glutathione, oxidized dithiothreitol, DL-alpha-lipoic acid, or insulin. Amino acid and nucleotide sequence analyses revealed that it had a presequence composed of 21 amino acids which had features characteristic of a mitochondrial targeting signal. The amino acid sequence of the mature protein was similar to that of bovine cytosolic thioredoxin reductase (57%) and of human glutathione reductase (34%) and less similar to that of Escherichia coli (19%) or yeast (17%) enzymes. Human and bovine cytosolic thioredoxin reductase were recently identified to contain selenocysteine (Sec) as one of their amino acid constituents. We also identified Sec in the C-terminal region of mitochondrial (mt)-thioredoxin reductase by means of MS and amino acid sequence analyses of the C-terminal fragment. The four-amino acid motif, Gly-Cys-Sec-Gly, which is conserved among all Sec-containing thioredoxin reductases, probably functions as the third redox centre of the enzyme, as the mitochondrial reductase was inhibited by 1-chloro-2,4-dinitrobenzene, which was reported to modify Sec and Cys covalently. It is known that mammalian thioredoxin reductase is different from bacterial or yeast enzyme in, for example, their subunit molecular masses and domain structures. These two different types of enzymes with similar activity are suggested to have evolved convergently. Our data clearly show that mitochondria, which might have originated from symbiotic prokaryotes, contain thioredoxin reductase similar to the cytosolic enzyme and different from the bacterial one.  相似文献   

7.
A soil bacterium capable of metabolizing organophosphorus compounds by reducing the P S group in the molecules was taxonomically identified as Klebsiella sp. strain F51-1-2. The gene involved in the reduction of organophosphorus compounds was cloned from this strain by the shotgun technique, and the deduced protein (named AKR5F1) showed homology to members of the aldo-keto reductase (AKR) superfamily. The intact coding region for AKR5F1 was subcloned into vector pET28a and overexpressed in Escherichia coli BL21(DE3). Recombinant His(6)-tagged AKR5F1 was purified in one step using Ni-nitrilotriacetic acid affinity chromatography. Assays for cofactor specificity indicated that reductive transformation of organophosphorus compounds by the recombinant AKR5F1 specifically required NADH. The kinetic constants of the purified recombinant AKR5F1 toward six thion organophosphorus compounds were determined. For example, the K(m) and k(cat) values of reductive transformation of malathion by the purified recombinant AKR5F1 are 269.5 +/- 47.0 microM and 25.7 +/- 1.7 min(-1), respectively. Furthermore, the reductive transformation of organophosphorus compounds can be largely explained by structural modeling.  相似文献   

8.
Succinic semialdehyde reductase, a NADP+-dependent enzyme, was purified from whole pig brain homogenates. The enzyme preparation migrates as a single protein and activity band on analytical gel electrophoresis. Succinic semialdehyde reductase (Mr 110,000) catalyzes the reduction of succinic semialdehyde to 4-hydroxybutyrate. The equilibrium constant of the reaction is Keq = 5.8 X 10(7) M-1 at pH 7 and 25 degrees C. The inhibition kinetic patterns obtained when 4-hydroxybutyrate or substrate analogs are used as inhibitors of the reaction catalyzed by the reductase are consistent with an ordered sequential mechanism, in which the coenzyme NADPH adds to the enzyme before the aldehyde substrate. A specific aldehyde reductase was also purified to homogeneity from brain mitochondria preparations. Its catalytic properties are identical to those of the enzyme isolated from whole brain homogenates. It is postulated that two enzymes, i.e. a NAD+-dependent dehydrogenase and a NADP+-dependent reductase, participate in the metabolism of succinic semialdehyde in the mitochondria matrix.  相似文献   

9.
10.
A new member of the aldo–keto reductase (AKR) superfamily with 3-dehydroecdysone reductase activity was found in the silkworm Bombyx mori upon induction by the insecticide diazinon. The amino acid sequence showed that this enzyme belongs to the AKR2 family, and the protein was assigned the systematic name AKR2E4. In this study, recombinant AKR2E4 was expressed, purified to near homogeneity, and kinetically characterized. Additionally, its ternary structure in complex with NADP+ and citrate was refined at 1.3 Å resolution to elucidate substrate binding and catalysis. The enzyme is a 33-kDa monomer and reduces dicarbonyl compounds such as isatin and 17α-hydroxy progesterone using NADPH as a cosubstrate. No NADH-dependent activity was detected. Robust activity toward the substrate inhibitor 3-dehydroecdysone was observed, which suggests that this enzyme plays a role in regulation of the important molting hormone ecdysone. This structure constitutes the first insect AKR structure determined. Bound NADPH is located at the center of the TIM- or (β/α)8-barrel, and residues involved in catalysis are conserved.  相似文献   

11.
Morphine 6-dehydrogenase, which catalyzes the dehydrogenation of morphine to morphinone, was purified 815-fold to a homogeneous protein from the soluble fraction of hamster liver with a yield of 15%. The enzyme was a monomeric protein with a molecular weight of 38 kDa and an isoelectric point of 5.6. Although both NAD and NADP served as cofactors, the enzyme activity with NADP was less than 5% that found with NAD at pH 7.4. With NAD, the enzyme gave the maximal activity at pH 9.3, and the K(m) and V(max) values toward morphine were 1.0 mM and 0.43 unit/mg protein, respectively. Among morphine congeners, normorphine exhibited higher activity than morphine, but codeine and ethylmorphine were poor substrates, and dihydromorphine and dihydrocodeine showed no detectable activity. The enzyme also exhibited significant activity for a variety of cyclic and alicyclic alcohols. In addition to xenobiotics, the enzyme catalyzed the dehydrogenation of 17beta-hydroxysteroids with much higher affinities than morphine. In the reverse reaction, the enzyme exhibited high activity for o-quinones, but morphinone, naloxone, and aromatic aldehydes and ketones were reduced at slow rates. Sulfhydryl reagents and ketamine strongly inhibited the enzyme, whereas pyrazole, barbital, and indomethacin had little effect on enzyme activity. 17beta-Hydroxysteroids inhibited the enzyme in a competitive manner against morphine. A total of 302 amino acid residues, which comprised approximately 94% of whole protein, were identified by sequencing of the peptides obtained by proteolytic digestion. This amino acid sequence of the enzyme showed significant homology to members of the aldo-keto reductase (AKR) superfamily and shared 63-64% identity with members of the AKR1C subfamily. These findings indicate that the enzyme is a new member of the AKR superfamily that is involved in steroid metabolism as 17beta-hydroxysteroid dehydrogenase as well as xenobiotic metabolism.  相似文献   

12.
Aldehyde reductase (AKR1A), a member of the aldo–keto reductase superfamily, suppresses diabetic complications via a reduction in metabolic intermediates; it also plays a role in ascorbic acid biosynthesis in mice. Because primates cannot synthesize ascorbic acid, a principle role of AKR1A appears to be the reductive detoxification of aldehydes. In this study, we isolated and immortalized mouse embryonic fibroblasts (MEFs) from wild-type (WT) and human Akr1a-transgenic (Tg) mice and used them to investigate the potential roles of AKR1A under culture conditions. Tg MEFs showed higher methylglyoxal- and acrolein-reducing activities than WT MEFs and also were more resistant to cytotoxicity. Enzymatic analyses of purified rat AKR1A showed that the efficiency of the acrolein reduction was about 20% that of glyceraldehyde. Ascorbic acid levels were quite low in the MEFs, and while the administration of ascorbic acid to the cells increased the intracellular levels of ascorbic acid, it had no affect on the resistance to acrolein. Endoplasmic reticulum stress and protein carbonylation induced by acrolein treatment were less evident in Tg MEFs than in WT MEFs. These data collectively indicate that one of the principle roles of AKR1A in primates is the reductive detoxification of aldehydes, notably acrolein, and protection from its detrimental effects.  相似文献   

13.
Proline reductase of Clostridium sticklandii is a membrane-bound protein and is released by treatment with detergents. The enzyme has been purified to homogeneity and is estimated by gel filtration and sedimentation equilibrium centrifugation to have a molecular weight of 298,000 to 327,000. A minimum molecular weight of 30,000 to 31,000 was calculated on the basis of sodium dodecyl sulfate-acrylamide gel electrophoresis and amino acid composition. Amino acid analysis showed a preponderance of acidic amino acids. No tryptophan was detected in the protein either spectrophotometrically or by amino acid analysis. A total of 20 sulfhydryl groups measured by titration of the reduced protein with 5,5'-dithiobis(2-nitrobenzoic acid) is in agreement with 20 cystic acid residues determined in hydrolysates of performic acid-oxidized protein. No molybdenum, iron, or selenium was found in the pure protein. Although NADH is the physiological electron donor for the proline reductase complex, the purified 300,000 molecular weight reductase component is inactive in the presence of NADH in vitro. Dithiothreitol, in contrast, can serve as electron donor both for unpurified (putative proline reductase complex) and purified proline reductase in vitro.  相似文献   

14.
15.
Cloning and expression of human aldose reductase   总被引:5,自引:0,他引:5  
The complete amino acid sequence of human retina and muscle aldose reductase was determined by nucleotide analysis of cDNA clones isolated using synthetic oligonucleotide probes based on partial amino acid sequences of purified human psoas muscle aldose reductase. The cDNA sequence differs substantially in the noncoding and coding regions of recently published sequences of this enzyme. The mRNA for aldose reductase was abundantly expressed in HeLa cells, but only scarcely in a neuroblastoma cell line. Recombinant baculovirus containing one of the muscle cDNA clones was constructed and used to infect Spodoptera frugiperda (SF9) cells. A prominent protein with an apparent molecular size of 36 kDa was identified by sodium dodecyl sulfate-polyacrylamide gel electrophoresis in the culture medium as well as in the homogenate of SF9 cells after 2 days of infection. Culture medium or the supernatant fraction of cell homogenates containing this protein had high aldose reductase activity which showed characteristics of the reported human enzyme. These findings indicate that the amino acid sequence reported in this paper represents human retina and muscle aldose reductase and that functional human aldose reductase can be expressed in large amounts in a baculovirus expression system. The result should facilitate refined structural analysis and the development of new specific aldose reductase inhibitors for the treatment of diabetic complications.  相似文献   

16.
Squalene synthetase (farnesyl diphosphate:farnesyl diphosphate farnesyltransferase; EC 2.5.1.21) is thought to represent a major control point of isoprene and sterol biosynthesis in eukaryotes. We demonstrate structural and functional conservation between the enzymes from humans, a budding yeast (Saccharomyces cerevisiae), and a fission yeast (Schizosaccharomyces pombe). The amino acid sequences of the human and S. pombe proteins deduced from cloned cDNAs were compared to those of the known S. cerevisiae protein. All are predicted to encode C-terminal membrane-spanning proteins of approximately 50 kDa with similar hydropathy profiles. Extensive sequence conservation exists in regions of the enzyme proposed to interact with its prenyl substrates (i.e., two farnesyl diphosphate molecules). Many of the highly conserved regions are also present in phytoene and prephytoene diphosphate synthetases, enzymes which catalyze prenyl substrate condensation reactions analogous to that of squalene synthetase. Expression of cDNA clones encoding S. pombe or hybrid human-S. cerevisiae squalene synthetases reversed the ergosterol requirement of S. cerevisiae cells bearing ERG9 gene disruptions, showing that these enzymes can functionally replace the S. cerevisiae enzyme. Inhibition of sterol synthesis in S. cerevisiae and S. pombe cells or in cultured human fibroblasts by treatment with the 3-hydroxy-3-methylglutaryl coenzyme A reductase inhibitor lovastatin resulted in elevated levels of squalene synthetase mRNA in all three cell types.  相似文献   

17.
The 2.2 Å X-ray crystal structure of Candida tenuis xylose reductase (AKR2B5) bound with NADP+ reveals that Phe-114 contributes to the substrate binding pocket of the enzyme. In the related human aldose reductase (AKR1B1), this phenylalanine is replaced by a tryptophan. The side chain of Trp was previously implicated in forming a hydrogen bond with bound substrate or inhibitor. The apparent Michaelis constant of AKR2B5 for xylose (Km≈90 mM) is 60 times that of AKR1B1, perhaps because critical enzyme–substrate interactions of Trp are not available to Phe-114. We, therefore, prepared a Phe-114→Trp mutant (F114W) of AKR2B5, to mimic the aldose reductase relationship in xylose reductase. Detailed analysis of the kinetic consequences in purified F114W revealed that the Km values for xylose and xylitol at pH 7.0 and 25°C were increased 5.1- and 4.4-fold, respectively, in the mutant compared with the wild-type. Turnover numbers (kcat) of F114W for xylose reduction and xylitol oxidation were half those of the wild-type. Apparent dissociation constants of NADH (KiNADH=44 µM) and NAD+ (KiNAD+=177 µM) were increased 1.6- and 1.4-fold in comparison with values of KiNADH and KiNAD+ for the wild-type, respectively. Catalytic efficiencies (kcat/Km) for NADH-dependent reduction of different aldehydes were between 3.1- and 31.5-fold lower than the corresponding kcat/Km values of the wild-type. Therefore, replacement of Phe-114 with Trp weakens rather than strengthens apparent substrate binding by AKR2B5, suggesting that xylose reductase exploits residue 114 in a different manner from aldose reductase.  相似文献   

18.
In order to understand more fully the structural features of aldo-keto reductases (AKRs) that determine their substrate specificities it would be desirable to obtain crystal structures of an AKR with a substrate at the active site. Unfortunately the reaction mechanism does not allow a binary complex between enzyme and substrate and to date ternary complexes of enzyme, NADP(H) and substrate or product have not been achieved. Previous crystal structures, in conjunction with numerous kinetic and theoretical analyses, have led to the general acceptance of the active site tyrosine as the general acid–base catalytic residue in the enzyme. This view is supported by the generation of an enzymatically inactive site-directed mutant (tyrosine-48 to phenylalanine) in human aldose reductase [AKR1B1]. However, crystallization of this mutant was unsuccessful. We have attempted to generate a trapped cofactor/substrate complex in pig aldehyde reductase [AKR1A2] using a tyrosine 50 to phenylalanine site-directed mutant. We have been successful in the generation of the first high resolution binary AKR–Y50F:NADP(H) crystal structure, but we were unable to generate any ternary complexes. The binary complex was refined to 2.2A and shows a clear lack of density due to the missing hydroxyl group. Other residues in the active site are not significantly perturbed when compared to other available reductase structures. The mutant binds cofactor (both oxidized and reduced) more tightly but shows a complete lack of binding of the aldehyde reductase inhibitor barbitone as determined by fluorescence titrations. Attempts at substrate addition to the active site, either by cocrystallization or by soaking, were all unsuccessful using pyridine-3-aldehyde, 4-carboxybenzaldehyde, succinic semialdehyde, methylglyoxal, and other substrates. The lack of ternary complex formation, combined with the significant differences in the binding of barbitone provides some experimental proof of the proposal that the hydroxyl group on the active site tyrosine is essential for substrate binding in addition to its major role in catalysis. We propose that the initial event in catalysis is the binding of the oxygen moiety of the carbonyl-group of the substrate through hydrogen bonding to the tyrosine hydroxyl group.  相似文献   

19.
Recent studies have demonstrated that aldo-keto reductase family 1 B10 (AKR1B10), a novel protein overexpressed in human hepatocellular carcinoma and non-small cell lung carcinoma, may facilitate cancer cell growth by detoxifying intracellular reactive carbonyls. This study presents a novel function of AKR1B10 in tumorigenic mammary epithelial cells (RAO-3), regulating fatty acid synthesis. In RAO-3 cells, Sephacryl-S 300 gel filtration and DEAE-Sepharose ion exchange chromatography demonstrated that AKR1B10 exists in two distinct forms, monomers (approximately 40 kDa) bound to DEAE-Sepharose column and protein complexes (approximately 300 kDa) remaining in flow-through. Co-immunoprecipitation with AKR1B10 antibody and protein mass spectrometry analysis identified that AKR1B10 associates with acetyl-CoA carboxylase-alpha (ACCA), a rate-limiting enzyme of de novo fatty acid synthesis. This association between AKR1B10 and ACCA proteins was further confirmed by co-immunoprecipitation with ACCA antibody and pulldown assays with recombinant AKR1B10 protein. Intracellular fluorescent studies showed that AKR1B10 and ACCA proteins co-localize in the cytoplasm of RAO-3 cells. More interestingly, small interfering RNA-mediated AKR1B10 knock down increased ACCA degradation through ubiquitination-proteasome pathway and resulted in >50% decrease of fatty acid synthesis in RAO-3 cells. These data suggest that AKR1B10 is a novel regulator of the biosynthesis of fatty acid, an essential component of the cell membrane, in breast cancer cells.  相似文献   

20.
Glutathione reductase from the liver of DBA/2J mice was purified to homogeneity by means of ammonium sulfate fractionation and two subsequent affinity chromatography steps using 8-(6-aminohexyl)-amino-2'-phospho-adenosine diphosphoribose and N6-(6-aminohexyl)-adenosine 2',5'-biphosphate-Sephadex columns. A facile procedure for the synthesis of 8-(6-aminohexyl)-amino-2'-phospho-adenosine diphosphoribose is also presented. The purified enzyme exhibits a specific activity of 158 U/mg and an A280/A460 of 6.8. It was shown to be a dimer of Mr 105000 with a Stokes radius of 4.18 nm and an isoelectric point of 6.46. Amino acid composition revealed some similarity between the mouse and the human enzyme. Antibodies against mouse glutathione reductase were raised in rabbits and exhibited high specificity. The catalytic properties of mouse liver glutathione reductase have been studied under a variety of experimental conditions. As with the same enzyme from other sources, the kinetic data are consistent with a 'branched' mechanism. The enzyme was stabilized against thermal inactivation at 80 degrees C by GSSG and less markedly by NADP+ and GSH, but not by NADPH or FAD. Incubation of mouse glutathione reductase in the presence of NADPH or NADH, but not NADP+ or NAD+, produced an almost complete inactivation. The inactivation by NADPH was time, pH and concentration dependent. Oxidized glutathione protected the enzyme against inactivation, which could also be reversed by GSSG or other electron acceptors. The enzyme remained in the inactive state even after eliminating the excess NADPH. The inactive enzyme showed the same molecular weight as the active glutathione reductase. The spectral properties of the inactive enzyme have also been studied. It is proposed that auto-inactivation of glutathione reductase by NADPH and the protection as well as reactivation by GSSG play in vivo an important regulatory role.  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