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 crystal structure of the GCY1 gene product from Saccharomyces cerevisiae has been determined to 2.5 A and is being refined. The model includes two protein molecules, one apo and one holo, per asymmetric unit. Examination of the model reveals that the active site surface is somewhat flat when compared with the other aldo-keto reductase structures, possibly accommodating larger substrates. The K(m) for NADPH (28.5 microM) is higher than that seen for other family members. This can be explained structurally by the lack of the 'safety belt' of residues seen in other aldo-keto reductases with higher affinity for NADPH. Catalysis also differs from the other aldo-keto reductases. The tyrosine that acts as an acid in the reduction reaction is flipped out of the catalytic pocket. This implies that the protein must either undergo a conformational change before catalysis can take place or that there is an alternate acid moiety.  相似文献   

2.
AtCBR, a cDNA encoding NADH-cytochrome (Cyt) b5 reductase, and AtB5-A and AtB5-B, two cDNAs encoding Cyt b5, were isolated from Arabidopsis. The primary structure deduced from the AtCBR cDNA was 40% identical to those of the NADH-Cyt b5 reductases of yeast and mammals. A recombinant AtCBR protein prepared using a baculovirus system exhibited typical spectral properties of NADH-Cyt b5 reductase and was used to study its electron-transfer activity. The recombinant NADH-Cyt b5 reductase was functionally active and displayed strict specificity to NADH for the reduction of a recombinant Cyt b5 (AtB5-A), whereas no Cyt b5 reduction was observed when NADPH was used as the electron donor. Conversely, a recombinant NADPH-Cyt P450 reductase of Arabidopsis was able to reduce Cyt b5 with NADPH but not with NADH. To our knowledge, this is the first evidence in higher plants that both NADH-Cyt b5 reductase and NADPH-Cyt P450 reductase can reduce Cyt b5 and have clear specificities in terms of the electron donor, NADH or NADPH, respectively. This substrate specificity of the two reductases is discussed in relation to the NADH- and NADPH-dependent activities of microsomal fatty acid desaturases.  相似文献   

3.
A cytosolic aldo-keto reductase was purified from Saccharomyces cerevisiae ATCC 26602 to homogeneity by affinity chromatography, chromatofocusing, and hydroxylapatite chromatography. The relative molecular weights of the aldo-keto reductase as determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and size exclusion chromatography were 36,800 and 35,000, respectively, indicating that the enzyme is monomeric. Amino acid composition and N-terminal sequence analysis revealed that the enzyme is closely related to the aldose reductases of xylose-fermenting yeasts and mammalian tissues. The enzyme was apparently immunologically unrelated to the aldose reductases of other xylose-fermenting yeasts. The aldo-keto reductase is NADPH specific and catalyzes the reduction of a variety of aldehydes. The best substrate for the enzyme is the aromatic aldehyde p-nitrobenzaldehyde (Km = 46 microM; kcat/Km = 52,100 s-1 M-1), whereas among the aldoses, DL-glyceraldehyde was the preferred substrate (Km = 1.44 mM; kcat/Km = 1,790 s-1 M-1). The enzyme failed to catalyze the reduction of menadione and p-benzoquinone, substrates for carbonyl reductase. The enzyme was inhibited only slightly by 2 mM sodium valproate and was activated by pyridoxal 5'-phosphate. The optimum pH of the enzyme is 5. These data indicate that the S. cerevisiae aldo-keto reductase is a monomeric NADPH-specific reductase with strong similarities to the aldose reductases.  相似文献   

4.
The β-subunit of the voltage-sensitive K+ channels shares 15–30% amino acid identity with the sequences of aldo–keto reductases (AKR) genes. However, the AKR properties of the protein remain unknown. To begin to understand its oxidoreductase properties, we examine the pyridine coenzyme binding activity of the protein in vitro. The cDNA of Kvβ2.1 from rat brain was subcloned into a prokaryotic expression vector and overexpressed in Escherichia coli. The purified protein was tetrameric in solution as determined by size exclusion chromatography. The protein displayed high affinity binding to NADPH as determined by fluorometric titration. The KD values for NADPH of the full-length wild-type protein and the N-terminus deleted protein were 0.1±0.007 and 0.05±0.006 M, respectively — indicating that the cofactor binding domain is restricted to the C-terminus, and is not drastically affected by the absence of the N-terminus amino acids, which form the ball and chain regulating voltage-dependent inactivation of the α-subunit. The protein displayed poor affinity for other coenzymes and the corresponding values of the KD for NADH and NAD were between 1–3 μM whereas the KD for FAD was >10 μM. However, relatively high affinity binding was observed with 3-acetyl pyridine NADP, indicating selective recognition of the 2′ phosphate at the binding site. The selectivity of Kvβ2.1 for NADPH over NADP may be significant in regulating the K+ channels as a function of the cellular redox state.  相似文献   

5.
1. Nitrofurantoin reductase which catalyzes the bioactivation of nitrofurantoin was purified to electrophoretic homogenity from sheep liver and lung microsomes, with a yield of 15% and 35%, respectively. The specific activity of both reductases was found to be similar (140 nmol/min/mg protein).2. The effects of nitrofurantoin and NADPH concentrations, pH, ionic strength, amount of enzyme and reaction period, on the enzyme activity were studied and the optimum conditions for maximum activity of purified liver and lung nitrofurantoin reductases were determined.3. The enzyme concentration was found proportional with the square root of the rate of nitrofurantoin reduction up to approximately 15 μg protein/ml and 25 μg protein/ml incubation mixture for liver and lung nitrofurantoin reductases, respectively.4. The plots of inverse of the nitrofurantoin concentration against the inverse of the square root of the velocity for the reduction of nitrofurantoin by liver and lung enzymes gave Km values as 27.78 μM and 32.25 μM, respectively.5. The purified liver and lung enzymes were also saturated by NADPH at similar concentrations and the Km values were calculated as 29.4 μM and 35.5 μM, respectively.6. The effects of magnesium, nickel, cadmium and copper ions on the nitrofurantoin reductase activity were examined. Magnesium ion was found to have almost no effect, whereas the other ions inhibited the activity of both liver and lung reductases.  相似文献   

6.
Total DNA of a population of uncultured organisms was extracted from soil samples, and by using PCR methods, the genes encoding two different 2,5-diketo-D-gluconic acid reductases (DKGRs) were recovered. Degenerate PCR primers based on published sequence information gave internal gene fragments homologous to known DKGRs. Nested primers specific for the internal fragments were combined with random primers to amplify flanking gene fragments from the environmental DNA, and two hypothetical full-length genes were predicted from the combined sequences. Based on these predictions, specific primers were used to amplify the two complete genes in single PCRs. These genes were cloned and expressed in Escherichia coli. The purified gene products catalyzed the reduction of 2,5-diketo-D-gluconic acid to 2-keto-L-gulonic acid. Compared to previously described DKGRs isolated from Corynebacterium spp., these environmental reductases possessed some valuable properties. Both exhibited greater than 20-fold-higher kcat/Km values than those previously determined, primarily as a result of better binding of substrate. The Km values for the two new reductases were 57 and 67 microM, versus 2 and 13 mM for the Corynebacterium enzymes. Both environmental DKGRs accepted NADH as well as NADPH as a cosubstrate; other DKGRs and most related aldo-keto reductases use only NADPH. In addition, one of the new reductases was more thermostable than known DKGRs.  相似文献   

7.
A yeast growing at 48°C was isolated from soil and the strain was identified as Cryptococcus lactativorus. The aldose reductase which the strain produced was purified 114-fold with an overall recovery of 36%. The stability of the enzyme was higher than that of other aldose reductases. The half life of the enzyme was 800 h and 14 h at 30°C and 50°C, respectively. The enzyme showed the best activity with d-xylose. l-Sorbose and d-fructose were also reduced by the enzyme. The enzyme was active with both NADPH and NADH as a conenzyme, and the activity with NADH was 1.25 times higher than that with NADPH. The Kmapp value for d-xylose was 8.6 mM and the Vmaxapp was 20.8 units/mg NADH was used as a coenzyme. The Kmapp values for NADPH and NADH were 6μM and 170 μM, respectively, when d-glucose was used as a substrate.  相似文献   

8.
A reduced nicotinamide adenine dinucleotide phosphate (NADPH)-dependent reductase with the ability to reduce diacetyl has been isolated from Escherichia coli and has been purified 800-fold to near homogeneity. The product of the reduction of diacetyl was shown to be acetoin. The enzyme proved to catalyze the oxidation of NADPH in the presence of both uncharged α- and β-dicarbonyl compounds. Even monocarbonyl compounds showed slight activity with the enzyme. On the basis of its substrate specificity, it is suggested that the enzyme functions as a diacetyl reductase. In contrast to other diacetyl reductases, the one reported here is specific for NADPH and does not possess acetoin reductase activity. The pH optimum of this enzyme was found to be between 6 and 7. The maximal velocity for the NADPH-dependent reduction of diacetyl was determined to be 9.5 μmol per min per mg of protein and the Km values for diacetyl and NADPH were found to be 4.44 mM and 0.02 mM, respectively. The molecular weight was estimated by gel filtration on Sephadex G-100 to be approximately 10,000.  相似文献   

9.
An NADPH thioredoxin reductase C was co-purified with a 2-Cys peroxiredoxin by the combination of anion exchange chromatography and electroelution from gel slices after native PAGE from a thermophilic cyanobacterium Thermosynechococcus elongatus as an NAD(P)H oxidase complex induced by oxidative stress. The result provided a strong evidence that the NADPH thioredoxin reductase C interacts with the 2-Cys peroxiredoxin in vivo. An in vitro reconstitution assay with purified recombinant proteins revealed that both proteins were essential for an NADPH-dependent reduction of H2O2. These results suggest that the reductase transfers the reducing power from NADPH to the peroxiredoxin, which reduces peroxides in the cyanobacterium under oxidative stress. In contrast with other NADPH thioredoxin reductases, the NADPH thioredoxin reductase C contains a thioredoxin-like domain in addition to an NADPH thioredoxin reductase domain in the same polypeptide. Each domain contains a conserved CXYC motif. A point mutation at the CXYC motif in the NADPH thioredoxin reductase domain resulted in loss of the NADPH oxidation activity, while a mutation at the CXYC motif in the thioredoxin-like domain did not affect the electron transfer, indicating that this motif is not essential in the electron transport from NADPH to the 2-Cys peroxiredoxin.  相似文献   

10.
Nucleotide sequences were determined for cDNA clones for squash NADH:nitrate oxidoreductase (EC 1.6.6.1), which is one of the most completely characterized forms of this higher plant enzyme. An open reading frame of 2754 nucleotides began at the first ATG. The deduced amino acid sequence contains 918 residues, with a predicted Mr = 103,376. The amino acid sequence is very similar to sequences deduced for other higher plant nitrate reductases. The squash sequence has significant similarity to the amino acid sequences of sulfite oxidase, cytochrome b5, and NADH:cytochrome b5 reductase. Alignment of these sequences with that of squash defines domains of nitrate reductase that appear to bind its 3 prosthetic groups (molybdopterin, heme-iron, and FAD). The amino acid sequence of the FAD domain of squash nitrate reductase was aligned with FAD domain sequences of other NADH:nitrate reductases, NADH:cytochrome b5 reductases, NADPH:nitrate reductases, ferredoxin:NADP+ reductases, NADPH:cytochrome P-450 reductases, NADPH:sulfite reductase flavoproteins, and Bacillus megaterium cytochrome P-450BM-3. In this multiple alignment, 14 amino acid residues are invariant, which suggests these proteins are members of a family of flavoenzymes. Secondary structure elements of the structural model of spinach ferredoxin:NADP+ reductase were used to predict the secondary structure of squash nitrate reductase and the other related flavoenzymes in this family. We suggest that this family of flavoenzymes, nearly all of which reduce a hemoprotein, be called "flavoprotein pyridine nucleotide cytochrome reductases."  相似文献   

11.
Two 2,5-diketo-d-gluconate reductases, I and II, were purified respectively 918-fold and 28-fold from a mutant strain derived from Corynebacterium sp. SHS 0007. The enzymes appeared to be homogeneous on polyacrylamide gel electrophoresis. Both reductases converted 2,5-diketo-d-gluconate to 2-keto-l-gulonate in the presence of NADPH and seemed to be active only for reduction. The molecular weights of reductases I and II were estimated to be 29,000 and 34,000, respectively; and both were monomeric. Their isoelectric points were respectively pH 4.3 and pH 4.1. The optimum pH was 6.0 to 7.0 for reductase I, and 6.0 to 7.5 for reductase II. The Km values (pH 7.0, 30°C) of reductase I for 2,5-diketo-d-gluconate and for NADPH were 1.8 mM and 12 μM, respectively; and the corresponding values of reductase II were 13.5 mM and 13 μM. Both reductases converted 5-keto-d-fructose to l-sorbose in the presence of NADPH.  相似文献   

12.
A novel short-chain (S)-1-phenyl-1,2-ethanediol dehydrogenase (SCR) from Candida parapsilosis exhibits coenzyme specificity for NADPH over NADH. It catalyzes an anti-Prelog type reaction to reduce 2-hydroxyacetophenone into (S)-1-phenyl-1,2-ethanediol. The coding gene was overexpressed in Escherichia coli and the purified protein was crystallized. The crystal structure of the apo-form was solved to 2.7 Å resolution. This protein forms a homo-tetramer with a broken 2-2-2 symmetry. The overall fold of each SCR subunit is similar to that of the known structures of other homologous alcohol dehydrogenases, although the latter usually form tetramers with perfect 2-2-2 symmetries. Additionally, in the apo-SCR structure, the entrance of the NADPH pocket is blocked by a surface loop. In order to understand the structure–function relationship of SCR, we carried out a number of mutagenesis–enzymatic analyses based on the new structural information. First, mutations of the putative catalytic Ser-Tyr-Lys triad confirmed their functional role. Second, truncation of an N-terminal 31-residue peptide indicated its role in oligomerization, but not in catalytic activity. Similarly, a V270D point mutation rendered the SCR as a dimer, rather than a tetramer, without affecting the enzymatic activity. Moreover, the S67D/H68D double-point mutation inside the coenzyme-binding pocket resulted in a nearly 10-fold increase and a 20-fold decrease in the kcat/KM value when NADH and NADPH were used as cofactors, respectively, with kcat remaining essentially the same. This latter result provides a new example of a protein engineering approach to modify the coenzyme specificity in SCR and short-chain dehydrogenases/reductases in general.  相似文献   

13.
Bile acid deficiency is a serious syndrome in newborns that can result in death if untreated. 5β-Reductase deficiency is one form of bile acid deficiency and is characterized by dramatically decreased levels of physiologically active 5β-reduced bile acids. AKR1D1 (aldo-keto reductase 1D1) is the only known human enzyme that stereo-specifically reduces the Δ4 double bond in 3-keto steroids and sterols to yield the 5β-hydrogenated product. Analysis of the AKR1D1 gene in five patients with 5β-reductase deficiency revealed five different mutations resulting in an amino acid substitution in the protein. To investigate a causal role for these observed point mutations in AKR1D1 in 5β-reductase deficiency, we characterized their effect on enzymatic properties. Attempts to purify mutant enzymes by overexpression in Escherichia coli only yielded sufficient amounts of the P133R mutant for further characterization. This enzyme displayed a highly reduced Km and Vmax reminiscent of uncompetitive kinetics with 4-cholesten-7α-ol-3-one as substrate. In addition, this mutant displayed no change in cofactor affinity but was more thermolabile in the absence of NADPH as judged by CD spectroscopy. All mutants were compared following expression in HEK 293 cells. Although these enzymes were equally expressed based on mRNA levels, protein expression and functional activity were dramatically reduced. Cycloheximide treatment also revealed that several of the expressed mutants were less stable. Our findings show that the reported mutations in AKR1D1 in patients with 5β-reductase lead to significantly decreased levels of active enzyme and could be causal in the development of bile acid deficiency syndrome.  相似文献   

14.
The nitrate reductase activity of 5-day-old whole corn roots was isolated using phosphate buffer. The relatively stable nitrate reductase extract can be separated into three fractions using affinity chromatography on blue-Sepharose. The first fraction, eluted with NADPH, reduces nearly equal amounts of nitrate with either NADPH or NADH. A subsequent elution with NADH yields a nitrate reductase which is more active with NADH as electron donor. Further elution with salt gives a nitrate reductase fraction which is active with both NADH and NADPH, but is more active with NADH. All three nitrate reductase fractions have pH optima of 7.5 and Stokes radii of about 6.0 nanometers. The NADPH-eluted enzyme has a nitrate Km of 0.3 millimolar in the presence of NADPH, whereas the NADH-eluted enzyme has a nitrate Km of 0.07 millimolar in the presence of NADH. The NADPH-eluted fraction appears to be similar to the NAD(P)H:nitrate reductase isolated from corn scutellum and the NADH-eluted fraction is similar to the NADH:nitrate reductases isolated from corn leaf and scutellum. The salt-eluted fraction appears to be a mixture of NAD(P)H: and NADH:nitrate reductases.  相似文献   

15.
NADPH-cytochrome P-450 reductases from pig liver and kidney and rabbit liver microsomes were purified to a specific activity of 50–62 μmol cytochrome c reduced/min/mg. All reductase preparations were separated into one major and one minor fraction on Sephadex G-200 columns. The molecular weights of the major fractions of the reductases were estimated to be 74,000, 75,000, and 75,500 for rabbit liver, pig kidney, and liver reductases, respectively, whereas the molecular weight of the minor fractions of these reductases, 67,000, was the same as that of the steapsin-solubilized pig liver reductase on SDS-polyacrylamide gel electrophoresis. Km values for NADPH and cytochrome c were: 20 and 29 μm or 14 and 28 μm for the pig kidney or liver reductase, respectively. Immunochemical studies, including Ouchterlony double diffusion experiments and inhibition of benzphetamine N-demethylation activity in microsomes by antibody against pig liver NADPH-cytochrome P-450 reductase, indicated the similarity of the purified liver and kidney reductases. There were no differences in the ability to reconstitute NADPH-mediated benzphetamine N-demethylation and laurate hydroxylation in reconstituted systems between the pig liver and kidney reductases, indicating that the reductase did not determine substrate specificity in these systems.  相似文献   

16.
We utilized the budding yeast Saccharomyces cerevisiae as a model to systematically explore physiological roles for yeast and mammalian aldo-keto reductases. Six open reading frames encoding putative aldo-keto reductases were identified when the yeast genome was queried against the sequence for human aldose reductase, the prototypical mammalian aldo-keto reductase. Recombinant proteins produced from five of these yeast open reading frames demonstrated NADPH-dependent reductase activity with a variety of aldehyde and ketone substrates. A triple aldo-keto reductase null mutant strain demonstrated a glucose-dependent heat shock phenotype which could be rescued by ectopic expression of human aldose reductase. Catalytically-inactive mutants of human or yeast aldo-keto reductases failed to effect a rescue of the heat shock phenotype, suggesting that the phenotype results from either an accumulation of one or more unmetabolized aldo-keto reductase substrates or a synthetic deficiency of aldo-keto reductase products generated in response to heat shock stress. These results suggest that multiple aldo-keto reductases fulfill functionally redundant roles in the stress response in yeast.  相似文献   

17.
Sphingomonads DC-6 and DC-2 degrade the chloroacetanilide herbicides alachlor, acetochlor, and butachlor via N-dealkylation. In this study, we report a three-component Rieske non-heme iron oxygenase (RHO) system catalyzing the N-dealkylation of these herbicides. The oxygenase component gene cndA is located in a transposable element that is highly conserved in the two strains. CndA shares 24 to 42% amino acid sequence identities with the oxygenase components of some RHOs that catalyze N- or O-demethylation. Two putative [2Fe-2S] ferredoxin genes and one glutathione reductase (GR)-type reductase gene were retrieved from the genome of each strain. These genes were not located in the immediate vicinity of cndA. The four ferredoxins share 64 to 72% amino acid sequence identities to the ferredoxin component of dicamba O-demethylase (DMO), and the two reductases share 62 to 65% amino acid sequence identities to the reductase component of DMO. cndA, the four ferredoxin genes, and the two reductases genes were expressed in Escherichia coli, and the recombinant proteins were purified using Ni-affinity chromatography. The individual components or the components in pairs displayed no activity; the enzyme mixture showed N-dealkylase activities toward alachlor, acetochlor, and butachlor only when CndA-His6 was combined with one of the four ferredoxins and one of the two reductases, suggesting that the enzyme consists of three components, a homo-oligomer oxygenase, a [2Fe-2S] ferredoxin, and a GR-type reductase, and CndA has a low specificity for the electron transport component (ETC). The N-dealkylase utilizes NADH, but not NADPH, as the electron donor.  相似文献   

18.
Xylose reductase is a homodimeric oxidoreductase dependent on NADPH or NADH and belongs to the largely monomeric aldo-keto reductase superfamily of proteins. It catalyzes the first step in the assimilation of xylose, an aldose found to be a major constituent monosaccharide of renewable plant hemicellulosic material, into yeast metabolic pathways. It does this by reducing open chain xylose to xylitol, which is reoxidized to xylulose by xylitol dehydrogenase and metabolically integrated via the pentose phosphate pathway. No structure has yet been determined for a xylose reductase, a dimeric aldo-keto reductase or a family 2 aldo-keto reductase. The structures of the Candida tenuis xylose reductase apo- and holoenzyme, which crystallize in spacegroup C2 with different unit cells, have been determined to 2.2 A resolution and an R-factor of 17.9 and 20.8%, respectively. Residues responsible for mediating the novel dimeric interface include Asp-178, Arg-181, Lys-202, Phe-206, Trp-313, and Pro-319. Alignments with other superfamily members indicate that these interactions are conserved in other dimeric xylose reductases but not throughout the remainder of the oligomeric aldo-keto reductases, predicting alternate modes of oligomerization for other families. An arrangement of side chains in a catalytic triad shows that Tyr-52 has a conserved function as a general acid. The loop that folds over the NAD(P)H cosubstrate is disordered in the apo form but becomes ordered upon cosubstrate binding. A slow conformational isomerization of this loop probably accounts for the observed rate-limiting step involving release of cosubstrate. Xylose binding (K(m) = 87 mM) is mediated by interactions with a binding pocket that is more polar than a typical aldo-keto reductase. Modeling of xylose into the active site of the holoenzyme using ordered waters as a guide for sugar hydroxyls suggests a convincing mode of substrate binding.  相似文献   

19.
NADH:nitrate reductase (EC 1.6.6.1) and NAD(P)H:nitrate reductase (EC 1.6.6.2) were purified from wild-type soybean (Glycine max [L.] Merr., cv Williams) and nr1-mutant soybean plants. Purification included Blue Sepharose- and hydroxylapatite-column chromatography using acetone powders from fully expanded unifoliolate leaves as the enzyme source.

Two forms of constitutive nitrate reductase were sequentially eluted with NADPH and NADH from Blue Sepharose loaded with extract from wild-type plants grown on urea as sole nitrogen source. The form eluted with NADPH was designated c1NR, and the form eluted with NADH was designated c2NR. Nitrate-grown nr1 mutant soybean plants yielded a NADH:nitrate reductase (designated iNR) when Blue Sepharose columns were eluted with NADH; NADPH failed to elute any NR form from Blue Sepharose loaded with this extract. Both c1NR and c2NR had similar pH optima of 6.5, sedimentation behavior (s20,w of 5.5-6.0), and electrophoretic mobility. However, c1NR was more active with NADPH than with NADH, while c2NR preferred NADH as electron donor. Apparent Michaelis constants for nitrate were 5 millimolar (c1NR) and 0.19 millimolar (c2NR). The iNR from the mutant had a pH optimum of 7.5, s20,w of 7.6, and was less mobile on polyacrylamide gels than c1NR and c2NR. The iNR preferred NADH over NADPH and had an apparent Michaelis constant of 0.13 millimolar for nitrate.

Thus, wild-type soybean contains two forms of constitutive nitrate reductase, both differing in their physical properties from nitrate reductases common in higher plants. The inducible nitrate reductase form present in soybeans, however, appears to be similar to most substrateinduced nitrate reductases found in higher plants.

  相似文献   

20.
NADPH-dependent d-xylose reductases (XRs) from Candida tropicalis IFO 0618 were purified and characterized. Mono Q HPLC revealed three XR isomers. The Km values of XR1, XR2 and XR3 for d-xylose were 37, 30 and 34 mM, and for NADPH 14, 18 and 9 μM, respectively. NADH did not act as a cofactor. The specificities of the three XRs for several aldoses were essentially the same. Gel filtration and cross-linking analysis showed that both XR1 and XR2 were dimers composed of identical subunits. The pI values of XR1 and XR2 were estimated to be 4.15 and 4.10, respectively. Comparison of the peptide maps of XR1 and XR2 showed that the molecular weights of 8 fragments of lysylendopeptidase-digested XR1 and XR2 were essentially the same as each other. The amino acid composition of each XR was also very similar. The molecular weights of XR1 and XR2 by mass spectra analysis were 36,497.91 and 36,539.68, respectively. The amino acid sequences of two XR1 peptide fragments (Nos. 4 and 5) were highly homologous with those of Pichia stipitis XR and mammalian aldose reductases.  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