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.
3-Hexulose phosphate synthase and phospho-3-hexuloisomerase were purified 40- and 150-fold respectively from methane-grown Methylococcus capsulatus. The molecular weights of the enzymes were approximately 310000 and 67000 respectively, as determined by gel filtration. Dissociation of 3-hexulose phosphate synthase into subunits of molecular weight approx. 49000 under conditions of low pH or low ionic strength was observed. Within the range of compounds tested, 3-hexulose phosphate synthase is specific for formaldehyde and d-ribulose 5-phosphate (forward reaction) and d-arabino-3-hexulose 6-phosphate (reverse reaction), and phospho-3-hexuloisomerase is specific for d-arabino-3-hexulose 6-phosphate (forward reaction) and d-fructose 6-phosphate (reverse reaction). A bivalent cation is essential for activity and stability of 3-hexulose phosphate synthase; phospho-3-hexuloisomerase is inhibited by many bivalent cations. The pH optima of the two enzymes are 7.0 and 8.3 respectively and the equilibrium constants are 4.0x10(-5)m and 1.9x10(2)m respectively. The apparent Michaelis constants for 3-hexulose phosphate synthase are: d-ribulose 5-phosphate, 8.3x10(-5)m; formaldehyde, 4.9x10(-4)m; d-arabino-3-hexulose 6-phosphate, 7.5x10(-5)m. The apparent Michaelis constants for phospho-3-hexuloisomerase are: d-arabino-3-hexulose 6-phosphate, 1.0x10(-4)m; d-fructose 6-phosphate, 1.1x10(-3)m.  相似文献   

2.
Reduced nicotinamide adenine dinucleotide phosphate (NADPH)-aldehyde reductase was isolated in 24% yield and 66-fold purification from a dl-glyceraldehyde-grown Rhodotorula species. This enzyme was specific for NADPH, and d-, l-, or dl-glyceraldehyde were equally good substrates. Other substrates had activities as follows: methylglyoxal, 50%; fructose, 33%; d- and l-arabinose, 12%; d-xylose, 8%; d-glucose, 5%; d-erythrose and d-threose, 0 to 5%. The product from the reduction of dl-glyceraldehyde was glycerol, as shown by high voltage electrophoresis, paper chromatography, and direct enzymatic analysis. Kinetic studies gave K(m) values of 0.89 mm and 0.013 mm for dl-glyceraldehyde and NADPH, respectively. An optimal pH range of 6.3 to 6.7 was found for maximal activity. Reduction of NADP(+) by glycerol was not demonstrable. This Rhodotorula NADPH-aldehyde reductase activity was compared to similar enzymes from other sources.  相似文献   

3.
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   

4.
Aleem, M. I. H. (Research Institute for Advanced Studies, Baltimore, Md.). Thiosulfate oxidation and electron transport in Thiobacillus novellus. J. Bacteriol. 90:95-101. 1965.-A cell-free soluble enzyme system capable of oxidizing thiosulfate was obtained from Thiobacillus novellus adapted to grow autotrophically. The enzyme systems of autotrophically grown cells brought about the transfer of electrons from thiosulfate to molecular oxygen via cytochromes of the c and a types; the reactions were catalyzed jointly by thiosulfate oxidase and thiosulfate cytochrome c reductase. The levels of both of these enzymes were markedly reduced in the heterotrophically grown organism. Cell-free extracts from the autotrophically grown T. novellus catalyzed formate oxidation and enzymatically reduced cytochrome c with formate. Both formate oxidation and cytochrome c reduction activities were abolished under heterotrophic conditions. The thiosulfate-activating enzyme S(2)O(3) (-2)-cytochrome c reductase, as well as thiosulfate oxidase, was localized chiefly in the soluble cell-free fractions, and the former enzyme was purified more than 200-fold by ammonium sulfate fractionation and calcium phosphate gel adsorption procedures. Optimal activity of the purified enzyme occurred at pH 8.0 in the presence of 1.67 x 10(-1)m S(2)O(3) (-2) and 2.5 x 10(-4)m cytochrome c. The thiosulfate oxidase operated optimally at pH 7.5 and thiosulfate concentrations of 1.33 x 10(-3) to 3.33 x 10(-2)m in the presence of added cytochrome c at a concentration of 5 x 10(-4)m. Both enzymes were markedly sensitive to cyanide and to a lesser extent to some metal-binding agents. Although a 10(-3)m concentration of p-hydroxymercuribenzoate had no effect on S(2)O(3) (-2)-cytochrome c reductase, it caused a 50% inhibition of S(2)O(3) (-2) oxidase, which was completely reversed in the presence of 10(-3)m reduced glutathione. Carbon monoxide also inhibited S(2)O(3) (-2) oxidase; the inhibition was completely reversed by light.  相似文献   

5.
Epididymal 5alpha reductase activity was found distitributed in the crude nuclear fraction (44 percent) and microsomal fraction (41 percent). Spermatozoa contaminating the nuclear preparation accounted for only 3 percent of its activity. There were no regional differences in the distribution of total 5alpha reductase activity. However, the nuclear enzyme was more active in caput than in other regions. Maximal activity was found at pH 6.2 and at 32 degrees C. Both enzymes had an absolute requirement of reduced dinucleotides. The microsomal preparation could only us NADPH while the nuclear enzyme could use NADPH and NADH. The apparent Km for the microsomal preparation was 0.62 +/- 0.05 X 10(-6)M and Vmax was 555 +/- 38 pmoles/mg protein/hour. The nuclear enzyme presented similar values. The reaction was not inhibited by accumulation of product in the medium, but other steroids such as progesterone, epitestosterone (17alpha-hydroxy-4-androsten-3-one) and 3-oxo-4-androstene-17beta-carboxylic acid were potent competitive inhibitors. The reaction was strongly inhibited by Hg, Zn and Cu. The properties of the epididymal reductase are similar to those of the prostatic enzyme.  相似文献   

6.
Phosphoglycerate phosphomutase has been purified to homogeneity from vegetative cells and germinated spores of Bacillus megaterium, and the spore and cell enzymes appear identical. The enzyme is a monomer of molecular weight 61,000. The compound 2,3-diphosphoglyceric acid is not required for activity, but the enzyme has an absolute and specific requirement for Mn2+. The enzyme is inhibited by ethylenediaminetetraacetate and sulfhydryl reagents, has a pH optimum of about 8.0, and has Km values for 3-phosphoglyceric acid and Mn2+ of 5 x 10(-4) and 4 x 10(-5) M, respectively.  相似文献   

7.
The ADPglucose pyrophosphorylases of 7 plant-leaf tissues were partially purified and characterized. In all cases the enzymes showed stability to heat treatment at 65 degrees for 5 minutes in the presence of 0.02 m phosphate buffer, pH 7.0. The leaf ADPglucose pyrophosphorylases were activated 5 to 15-fold by 3-phosphoglycerate. Fructose-6-phosphate and fructose 1, 6-diphosphate stimulated ADPglucose pyrophosphorylase to lesser extents. The A(0.5) (conc of activator required to give 50% of the observed maximal activation) of 3-phosphoglycerate for the barley enzyme was 7 x 10(-6)m while for the sorghum enzyme it was 3.7 x 10(-4)m. Inorganic phosphate proved to be an effective inhibitor of ADPglucose synthesis. The I(0.5) (conc of inhibitor that gave 50% inhibition of activity for the various leaf enzymes varied from 2 x 10(-5)m (barley) to 1.9 x 10(-4)m (sorghum). This inhibition was reversed or antagonized by the activator 3-phosphoglycerate. These results form the basis for an hypothesis of the regulation of leaf starch biosynthesis.  相似文献   

8.
1. Biochemical properties of delta 1-pyrroline-5-carboxylate reductase from d. melanogaster have been investigated. 2. The enzyme is stable below 4 degrees C. 3. the pH optimum of the enzyme is 5.7. It is rapidly inactivated below pH 5.4. 4. The Km values for NADPH and delta 1-pyrroline-5-carboxylate are 1.6 x 10-5 and 2.5 x 10-6 M, respectively. 5. the estimated molecular weight of the enzyme is 225,000. 6. the enzyme is weakly inhibited by L-proline (Ki = 0.12 M).  相似文献   

9.
Abstract Two constitutive acetoacetyl-CoA (AcAc-CoA) reductases were purified from Alcaligenes eutrophus . Incorporation of [1-14C]-acetyl-CoA into poly-3-hydroxybutyrate (PHB) by systems reconstituted from purified preparations of either 3-ketothiolase, AcAc-CoA reductase and PHB synthase, occurred only when NADPH-AcAc-CoA reductase was present. The NADH reductase was active with all of the d (−)- and l (+)-3-hydroxyacyl-CoA substrates tested (C4-C10), whereas the NADPH reductase was only active with d (−)-3-hydroxyacyl-CoAs (C4-C6). The products of AcAc-CoA reduction by the NADH- and NADPH-linked enzymes were l (+)-3-hydroxybutyryl-CoA and d (−)-3-hydroxybutyryl-CoA, respectively. The NADH-linked enzyme had an M r of 150,000 (containing identical M r 30,000 sub-units) and the NADPH-linked enzyme appeared to be a tetramer ( M r 84,000) with identical sub-units ( M r 23,000). K mapp values of 22 μM and 5 μM for AcAc-CoA and 13 μM (NADH) and 19 μM (NADPH) for the coenzymes were determined for the NADH- and NADPH-linked enzymes, respectively.  相似文献   

10.
11.
Fungal metabolism of biphenyl.   总被引:9,自引:0,他引:9       下载免费PDF全文
gamma-Glutamyl phosphate reductase, the second enzyme of proline biosynthesis, catalyses the formation of l-glutamic acid 5-semialdehyde from gamma-glutamyl phosphate with NAD(P)H as cofactor. It was purified 150-fold from crude extracts of Pseudomonas aeruginosa PAO 1 by DEAE-cellulose chromatography and hydroxyapatite adsorption chromatography. The partially purified preparation, when assayed in the reverse of the biosynthetic direction, utilized l-1-pyrroline-5-carboxylic acid as substrate and reduced NAD(P)(+). The apparent K(m) values were: NAD(+), 0.36mm; NADP(+), 0.31mm; l-1-pyrroline-5-carboxylic acid, 4mm with NADP(+) and 8mm with NAD(+); P(i), 28mm. 3-(Phosphonoacetylamido)-l-alanine, a structural analogue of gamma-glutamyl phosphate, inhibited this enzyme competitively (K(i)=7mm). 1-Pyrroline-5-carboxylate reductase (EC 1.5.1.2), the third enzyme of proline biosynthesis, was purified 56-fold by (NH(4))(2)SO(4) fractionation, Sephadex G-150 gel filtration and DEAE-cellulose chromatography. It reduced l-1-pyrroline-5-carboxylate with NAD(P)H as a cofactor to l-proline. NADH (K(m)=0.05mm) was a better substrate than NADPH (K(m)=0.02mm). The apparent K(m) values for l-1-pyrroline-5-carboxylate were 0.12mm with NADPH and 0.09mm with NADH. The 3-acetylpyridine analogue of NAD(+) at 2mm caused 95% inhibition of the enzyme, which was also inhibited by thio-NAD(P)(+), heavy-metal ions and thiol-blocking reagents. In cells of strain PAO 1 grown on a proline-medium the activity of gamma-glutamyl kinase and gamma-glutamyl phosphate reductase was about 40% lower than in cells grown on a glutamate medium. No repressive effect of proline on 1-pyrroline-5-carboxylate reductase was observed.  相似文献   

12.
Wang H  Lei B  Tu SC 《Biochemistry》2000,39(26):7813-7819
Luminous bacteria contain three types of NAD(P)H-FMN oxidoreductases (flavin reductases) with different pyridine nucleotide specificities. Among them, the NADPH-specific flavin reductase from Vibrio harveyi exhibits a uniquely high preference for NADPH. In comparing the substrate specificity, crystal structure, and primary sequence of this flavin reductase with other structurally related proteins, we hypothesize that the conserved Arg203 residue of this reductase is critical to the specific recognition of NADPH. The mutation of this residue to an alanine resulted in only small changes in the binding and reduction potential of the FMN cofactor, the K(m) for the FMN substrate, and the k(cat). In contrast, the K(m) for NADPH was increased 36-fold by such a mutation. The characteristic perturbation of the FMN cofactor absorption spectrum upon NADP(+) binding by the wild-type reductase was abolished by the same mutation. While the k(cat)/K(m,NADPH) was reduced from 1990 x 10(5) to 46 x 10(5) M(-1) min(-1) by the mutation, the mutated variant showed a k(cat)/K(m,NADH) of 4 x 10(5) M(-1) min(-1), closely resembling that of the wild-type reductase. The deuterium isotope effects (D)V and (D)(V/K) for (4R)-[4-(2)H]-NADPH were 1.7 and 1.4, respectively, for the wild-type reductase but were increased to 3.8 and 4.0, respectively, for the mutated variant. Such a finding indicates that the rates of NADPH and NADP(+) dissociation in relation to the isotope-sensitive redox steps were both increased as a result of the mutation. These results all provide support to the critical role of the Arg203 in the specific recognition and binding of NADPH.  相似文献   

13.
Abstract— The enzymes for the biosynthesis of phosphatidic acid from acyl dihydroxyacetone phosphate were shown to be present in rat brain. These enzymes were mainly localized in the microsomal fraction of 12–14 day old rat brains. The brain microsomal acyl CoA: dihydroxyacetone phosphate acyl transferase (EC 2.3.1.42), exhibited a broad pH optimum between pH 5 and 9 with maximum activity at pH 5.4. K m for DHAP at pH 5.4 was 0.1 m m and V max was 0.86nmol/min/mg of microsomal protein. The corresponding microsomal enzyme for the glycerophosphate pathway (acyl CoA: sn -glycerol-3-phosphate acyl transferase EC 2.3.1.15) was shown to have a different pH optimum (pH 7.6). On the basis of the differences in pH optima, differential effects of sodium cholate in the enzymes and a common substrate competition study, these acyl transferases were postulated to be two different microsomal enzymes.
Acyl DHAP:NADPH oxidoreductase (EC 1.1.1.101) in brain microsomes was found to be quite specific for NADPH as cofactor, being able to utilize NADH only at very high concentrations. This enzyme exhibited a K m of 8.6 μ m with NADPH and V mx of 0.81 nmol/min/mg protein. The presence of these two enzymes and the known presence of l-acyl- sn -glycerol-3-phosphate: acyl CoA acyl transferase in brain (F leming & H ajra , 1977) demonstrated the biosynthesis of phosphatidic acid in brain via acyl dihydroxyacetone phosphate. Phosphatidic acid was shown to form when dihydroxyacetone phosphate, acyl CoA, NADPH and other cofactors were incubated together with brain microsomes. Further properties of the enzymes and the probable importance of the presence of this pathway in brain were discussed.  相似文献   

14.
Streptomyces antibioticus NF-18 is a hyperproducing strain of a Streptomyces hormone, virginiae butanolide A (VB-A), that induces virginiamycin production of S. virginiae at nanomolar concentrations. To characterize the biosynthetic pathway of VB-A, we identified and characterized for the first time the 6-dehydro VB-A reductase that is responsible for the final reduction step in the biosynthesis. Assay protocols and stabilization conditions were established. The 6-dehydro VB-A reductase was found to require NADPH, not NADH, as a coenzyme. The K(m) values of the enzyme for NADPH and (+/-)-6-dehydro VB-A were determined to be 50 +/- 2 microM and 100 +/- 5 microM, respectively. Ultracentrifugation experiments revealed that 6-dehydro VB-A reductase was present almost exclusively in the 100,000 x g supernatant fraction, indicating that the enzyme is a cytoplasmic-soluble protein. The M(r) of the native 6-dehydro VB-A reductase was estimated to be 82,000 +/- 3000 by molecular sieve HPLC. The optimal pH was found to be 6.7 +/- 0.2.  相似文献   

15.
Affinity labeling studies of NADP(+)-glutamate dehydrogenase from Salmonella typhimurium have shown that the peptide Leu-282-Lys-286 is located near the coenzyme site [Haeffner-Gormley et al. (1991) J. Biol. Chem. 266, 5388-5394]. The present study was undertaken to evaluate the role of lysine-286. The mutant enzymes K286R, K286Q, and K286E were prepared by site-directed mutagenesis, expressed in Escherichia coli, and purified. The Vmax values (micromoles of NADPH per minute per milligram of protein) were similar for WT (270), K286R (529), K296Q (409), and K286E (382) enzymes. As measured at pH 7.9, the Km value for NADPH was much greater for K286E (280 microM) than for WT (9.8 microM), K286R (30 microM), or K286Q (66 microM) enzymes. The efficiencies (kcat/Km) of the WT and K286R mutant were similar (1.2 x 10(3) min-1 microM-1 and 1.0 x 10(3) min-1 microM-1, respectively) while those of K286Q (0.30 x 10(3) min-1 microM-1) and K286E (0.07 x 10(3) min-1 microM-1) were greatly reduced. The decreased efficiency of the K286E mutant results from the increase in Km-NADPH, consistent with a role for a basic residue at position 286 which enhances the binding of NADPH. Plots of Vmax vs pH showed the pH optima to be 8.1-8.3 for all enzymes at saturating NADPH concentrations. A 40-fold increase in Km-NADPH for K286E was observed as the pH increased from 5.98 to 8.08, from which a unique pKe of 6.5 was calculated.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
1. Aldose reductase and aldehyde reductase were purified to homogeneity from human testis. 2. The molecular weight of aldose reductase and aldehyde reductase were estimated to be 36,000 and 38,000 by SDS-PAGE, and the pI values of these enzymes were found to be 5.9 and 5.1 by chromatofocusing, respectively. 3. Aldose reductase had activity for aldo-sugars, whereas aldehyde reductase was virtually inactive for aldo-sugars. The Km values of aldose reductase for D-glucose, D-galactose and D-xylose were 57, 49 and 6.2 mM, respectively. Aldose reductase utilized both NADPH and NADH as coenzymes, whereas aldehyde reductase only NADPH. 4. Sulfate ion caused 3-fold activation of aldose reductase, but little for that of aldehyde reductase. 5. Sodium valproate inhibited significantly aldehyde reductase, but not aldose reductase. Aldose reductase was inhibited strongly by aldose reductase inhibitors being in clinical trials at concentrations of the order of 10(-7)-10(-9) M. Aldehyde reductase was also inhibited by these inhibitors, but its susceptibility was less than aldose reductase. 6. Reaction of aldose reductase with pyridoxal 5'-phosphate (PLP) resulted ca 2.5-fold activation, but aldehyde reductase did not cause the activation. PLP-treated aldose reductase has lost the susceptibility to aldose reductase inhibitor.  相似文献   

17.
Random sequencing of a peppermint essential oil gland secretory cell cDNA library revealed a large number of clones that specified redox-type enzymes. Full-length acquisitions of each type were screened by functional expression in Escherichia coli using a newly developed in situ assay. cDNA clones encoding the monoterpene double-bond reductases (-)-isopiperitenone reductase and (+)-pulegone reductase were isolated, representing two central steps in the biosynthesis of (-)-menthol, the principal component of peppermint essential oil, and the first reductase genes of terpenoid metabolism to be described. The (-)-isopiperitenone reductase cDNA has an open reading frame of 942 nucleotides that encodes a 314 residue protein with a calculated molecular weight of 34,409. The recombinant reductase has an optimum pH of 5.5, and K(m) values of 1.0 and 2.2 microM for (-)-isopiperitenone and NADPH, respectively, with k(cat) of 1.3s(-1) for the formation of the product (+)-cis-isopulegone. The (+)-pulegone reductase cDNA has an open reading frame of 1026 nucleotides and encodes a 342 residue protein with a calculated molecular weight of 37,914. This recombinant reductase catalyzes the reduction of the 4(8)-double bond of (+)-pulegone to produce both (-)-menthone and (+)-isomenthone in a 55:45 ratio, has an optimum pH of 5.0, and K(m) values of 2.3 and 6.9 microM for (+)-pulegone and NADPH, respectively, with k(cat) of 1.8s(-1). Deduced sequence comparison revealed that these two highly substrate specific double-bond reductases show less than 12% identity. (-)-Isopiperitenone reductase is a member of the short-chain dehydrogenase/reductase superfamily and (+)-pulegone reductase is a member of the medium-chain dehydrogenase/reductase superfamily, implying very different evolutionary origins in spite of the similarity in substrates utilized and reactions catalyzed.  相似文献   

18.
The 5alpha reductase activity ofthe monkey epididymis was studied. The enzyme was found in particulate subcellular fractions, its distribution closely resembling that of the microsomal marker enzyme NADPH: cytochrome c reductase, suggesting an association of 5alpha reductase with membranes of the endoplasmic reticulum. Maximal enzyme activity was found at pH 5.4 and at 32--37 C. The crude nuclear preparation had a Km: 0.315 x 10(-6)M and Vmax: 168 pmoles/mg protein/h. The microsomal enzyme had a Km: 0.243 x 10(-6)M and Vmax: 828 pmoles/mg protein/h. Neither enzyme preparation was affected by addition to the incubation media of dihydrotestosterone (DHT) or 5alpha-androstane-3alpha,17beta-diol. The endogenous androgen concentration in the epididymides of 2 different monkeys, in ng/g wet weight was: DHT 20.81 +/- 1.98; T: 9.0L +/- 2.83; diol: 3.03 +/- 0.41.  相似文献   

19.
1. Kinetic experiments suggested the possible existence of at least two different NAD(+)-dependent aldehyde dehydrogenases in rat liver. Distribution studies showed that one enzyme, designated enzyme I, was exclusively localized in the mitochondria and that another enzyme, designated enzyme II, was localized in both the mitochondria and the microsomal fraction. 2. A NADP(+)-dependent enzyme was also found in the mitochondria and the microsomal fraction and it is suggested that this enzyme is identical with enzyme II. 3. The K(m) for acetaldehyde was apparently less than 10mum for enzyme I and 0.9-1.7mm for enzyme II. The K(m) for NAD(+) was similar for both enzymes (20-30mum). The K(m) for NADP(+) was 2-3mm and for acetaldehyde 0.5-0.7mm for the NADP(+)-dependent activity. 4. The NAD(+)-dependent enzymes show pH optima between 9 and 10. The highest activity was found in pyrophosphate buffer for both enzymes. In phosphate buffer there was a striking difference in activity between the two enzymes. Compared with the activity in pyrophosphate buffer, the activity of enzyme II was uninfluenced, whereas the activity of enzyme I was very low. 5. The results are compared with those of earlier investigations on the distribution of aldehyde dehydrogenase and with the results from purified enzymes from different sources.  相似文献   

20.
Kim MJ  Kim Y 《Molecules and cells》1999,9(5):470-475
Reduced form of nicotineamide adenine dinucleotide phosphate (NADPH)-cytochrome P450 reductase was solubilized from a microsomal fraction of Gentiana triflora flowers by 3-[(3 Cholamidopropyl)-dimethylammonio]-1-propane sulfonate detergent and purified to electrophoretic homogeneity. The purification was achieved by adenosine 2', 5'-bisphosphate-Sepharose chromatography, followed by high-performance anion-exchange chromatography. A Mr value of 82,000 was obtained by SDS/polyacrylamide-gel electrophoresis. Western blot analysis showed that the purified protein cross-reacted with polyclonal antibody raised against rabbit anti-Gentiana triflora NADPH-cytochrome P450 reductase antibodies. The temperature and pH optimum for reduction of cytochrome c was 25 degrees C and 7.4 respectively. The Km values for the binding of NADPH and cytochrome c were 9.4 and 3.2 microM, respectively. In this paper, we present some results of the purification and partial characterization of microsomal NADPH-cytochrome P450 reductase from Gentiana triflora flowers.  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