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.
Phosphatidylserine decarboxylase of Escherichia coli is one of a small group of pyruvoyl-dependent enzymes (Satre, M., and Kennedy, E.P. (1978) J. Biol. Chem. 253, 479-483). The DNA sequence of the structural gene (psd) and partial protein sequence studies demonstrate that the enzyme contains two nonidentical subunits, alpha (Mr = 7,332) and beta (Mr = 28,579), which are derived from a single proenzyme. These two subunits are blocked at their respective amino termini. Reduction of the enzyme with NaCNBH3 in the presence of radiolabeled phosphatidylserine resulted in association of the label with the alpha subunit. Similar reduction in the presence of ammonium ions exposed a new amino terminus for the alpha subunit beginning with alanine. Therefore, the pyruvate prosthetic group is in amide linkage to the amino terminus of the alpha subunit. The amino terminus of the beta subunit was determined to be formylmethionine. The carboxyl terminus of the beta subunit was determined to be glycine as predicted by the DNA sequence. Comparison of the DNA sequence and protein sequence information revealed that the decarboxylase is made as a proenzyme (Mr = 35,893), and the predicted amino acid at the position of the pyruvate within the open reading frame of the proenzyme is serine. Therefore, as with other pyruvoyl-dependent decarboxylases, the prosthetic group is derived from serine through a post-translational cleavage of a proenzyme.  相似文献   

2.
The complete amino acid sequence of the alpha chain of histidine decarboxylase of Lactobacillus 30a has been established by isolation and analysis of the eight methionine-containing tryptic peptides of this chain. These peptides provide the overlaps required to order all nine peptides derived by complete cyanogen bromide cleavage of the alpha chain (Huynh, Q.K., Vaaler, G.L., Recsei, P.A., and Snell, E.E. (1984) J. Biol. Chem. 259, 2826-2832). Ordering of six of the latter peptides was confirmed by isolation and analysis of four peptides derived by incomplete cyanogen bromide cleavage. The alpha chain is composed of 226 residues and has a molecular weight of 24,892 calculated from the sequence. These results and the previously determined sequence of the beta chain (Vaaler, G.L., Recsei, P.A., Fox, J.L., and Snell, E.E. (1982) J. Biol. Chem. 257, 12770-12774) establish the complete amino acid sequence of the enzyme and of the pi chain of prohistidine decarboxylase. The latter is composed of 307 amino acids and has a calculated molecular weight of 33,731. Four segments of the pi chain sequence are repeated. The bond between Ser-81 and Ser-82 that is cleaved during proenzyme activation is in an uncharged portion of the sequence that is rich in serine and threonine residues and is predicted to be part of a beta sheet structure.  相似文献   

3.
Human S-adenosylmethionine decarboxylase is synthesized as a proenzyme that undergoes an autocatalytic cleavage reaction generating the alpha and beta subunits and forming the pyruvate prosthetic group, which is derived from an internal Ser residue (Ser-68). The mechanism of this processing reaction was studied using site-directed mutagenesis of conserved residues (His-243 and Ser-229) located close to the cleavage site. Mutant S229A failed to process, and mutant S229C cleaved very slowly, whereas mutant S229T processed normally, suggesting that the hydroxyl group of residue 229 is required for the processing reaction where Ser-229 may act as a proton acceptor. Mutant His-243A cleaved very slowly, forming a small amount of the correctly processed pyruvoyl enzyme but a much larger proportion of the alpha subunit with an amino-terminal Ser. The cleavage to form the latter was greatly enhanced by hydroxylamine. This result suggests that the N-O acyl shift needed for ester formation occurs normally in this mutant but that the next step, which is a beta-elimination reaction leading to the two subunits, does not occur. His-243 may therefore act as the basic residue that extracts the hydrogen of the alpha-carbon of Ser-68 in the ester in order to facilitate this reaction. The availability of the recombinant H243A S-adenosylmethionine decarboxylase proenzyme provides a useful model system to examine the processing reaction in vitro and test the design of specific inactivators aimed at blocking the production of the pyruvoyl prosthetic group.  相似文献   

4.
Conversion of the pi subunit of prohistidine decarboxylase to the alpha beta subunits of the active enzyme proceeds by a nonhydrolytic, monovalent cation-dependent, serinolysis reaction in which the hydroxyl oxygen of serine 82 of the pi chain is incorporated into the carboxyl group at the COOH terminus (serine 81) of the beta chain. Serine-82 becomes the pyruvate residue at the NH2 terminus of the alpha chain (Recsei, P.A., Huynh, Q. K., and Snell, E.E. (1983) Proc. Natl. Acad. Sci. U.S.A. 80, 973-977). The unusual reactivity of this particular -Ser-Ser- bond is demonstrated by its sensitivity to 1 M hydroxylamine, which cleaves the native proenzyme under mild conditions (pH 8.0, 37 degrees C) to yield a modified beta chain with serine hydroxamate at the COOH terminus (Ser-81) and a modified alpha chain containing serine (Ser-82 of the proenzyme) rather than pyruvate at the NH2 terminus. Neither an -Asn-Gly- bond nor other -Ser-Ser- bonds in the proenzyme were cleaved under these conditions. The reaction also did not occur with the denatured enzyme or with model peptides, indicating that the enhanced reactivity is a result of the particular conformation at this position in the native protein. The reaction with the native proenzyme proceeded optimally at pH 7.5-8.0 with a half-time (30 min) substantially less than that (3.5-4.5 h) required for the activation reaction and was not increased in rate by addition of K+. Correspondingly, preincubation of the proenzyme at pH 8.0 in the absence of both hydroxylamine and K+ modestly increased the rate of activation when K+ was subsequently added. Although these findings do not exclude other mechanisms, they are all consistent with and most easily explained by rearrangement of the pi chain to form an internal ester intermediate prior to the beta-elimination that occurs during activation to yield the alpha and beta chains of the mature enzyme.  相似文献   

5.
We have cloned and sequenced the Saccharomyces cerevisiae gene for S-adenosylmethionine decarboxylase. This enzyme contains covalently bound pyruvate which is essential for enzymatic activity. We have shown that this enzyme is synthesized as a Mr 46,000 proenzyme which is then cleaved post-translationally to form two polypeptide chains: a beta subunit (Mr 10,000) from the amino-terminal portion and an alpha subunit (Mr 36,000) from the carboxyl-terminal portion. The protein was overexpressed in Escherichia coli and purified to homogeneity. The purified enzyme contains both the alpha and beta subunits. About half of the alpha subunits have pyruvate blocking the amino-terminal end; the remaining alpha subunits have alanine in this position. From a comparison of the amino acid sequence deduced from the nucleotide sequence with the amino acid sequence of the amino-terminal portion of each subunit (determined by Edman degradation), we have identified the cleavage site of the proenzyme as the peptide bond between glutamic acid 87 and serine 88. The pyruvate moiety, which is essential for activity, is generated from serine 88 during the cleavage. The amino acid sequence of the yeast enzyme has essentially no homology with S-adenosylmethionine decarboxylase of E. coli (Tabor, C. W., and Tabor, H. (1987) J. Biol. Chem. 262, 16037-16040) and only a moderate degree of homology with the human and rat enzymes (Pajunen, A., Crozat, A., J?nne, O. A., Ihalainen, R., Laitinen, P. H., Stanley, B., Madhubala, R., and Pegg, A. E. (1988) J. Biol. Chem. 263, 17040-17049); all of these enzymes are pyruvoyl-containing proteins. Despite this limited overall homology the cleavage site of the yeast proenzyme is identical to the cleavage sites in the human and rat proenzymes, and seven of the eight amino acids adjacent to the cleavage site are identical in the three eukaryote enzymes.  相似文献   

6.
Mechanism-based inactivation of pyridoxal phosphate-dependent histidine decarboxylase by (S)-alpha-(fluoromethyl)histidine was studied. The molar ratio of inactivator to enzyme subunit required for complete inactivation increased from 1.63 at 10 degrees C to 3.00 at 37 degrees C. Two inactivation products were isolated by chromatographic fractionation of the reaction mixture and identified by NMR spectroscopy as 1-(4-imidazolyl)-3(5'-P-pyridoxylidene) acetone (I), the adduct formed between pyridoxal phosphate and inactivator, and 1-(4-imidazolyl) acetone (II), an intermediate compound formed during inactivation. Formation of these two products supports a previously proposed mechanism of inactivation (Hayashi, H., Tanase, S., and Snell, E. E. (1986) J. Biol. Chem. 261, 11003-11009), with minor modifications. A precursor of I was linked covalently to the enzyme by NaBH4 reduction if the reaction was carried out immediately after inactivation, before development of the 403 nm peak of I. A mutant histidine decarboxylase (S322A) in which Ser-322 was changed to Ala was also inactivated by alpha-fluoromethylhistidine demonstrating that Ser-322 is not essential for inactivation even though it is close to the active site and is derivatized by borohydride reduction of the inactivated wild-type enzyme. Following inactivation, both the wild-type and the S322A mutant enzyme could be partially reactivated by prolonged dialysis against buffer.  相似文献   

7.
When Lactobacillus buchneri was grown in the presence of [hydroxyl-18O]serine and pyridoxamine, no 18O was found in its histidine decarboxylase (HisDCase). However, when pyridoxamine was omitted from the growth medium, the labeled serine was incorporated into the HisDCase without dilution. Internal serine residues of the enzyme contained 18O only in their hydroxyl group, while the COOH-terminal serine of the beta chain of HisDCase contained equal amounts of 18O in both its hydroxyl and carboxyl group. This enzyme, like the HisDCase from Lactobacillus 30a (Recsei, P. A., Huynh, Q. K., and Snell, E. E. (1983) Proc. Natl. Acad. Sci. U.S.A. 80, 973-977), therefore, arises by nonhydrolytic serinolysis of its proenzyme. This result, together with comparative sequence data (Huynh, Q. K., and Snell, E. E. (1985) J. Biol. Chem. 260, 2798-2803), makes it highly probable that all of the pyruvoyl-dependent HisDCases arise by a similar mechanism from inactive proenzymes.  相似文献   

8.
The nucleotide sequence of a 1.3-kilobase NaeI fragment from Morganella morganii AM-15 that contains the gene for histidine decarboxylase has been determined. The gene was initially identified among total chromosomal digests using a mixed sequence oligonucleotide probe corresponding to amino acids 11-16 of histidine decarboxylase and then cloned on a 5.5-kilobase PstI fragment. The structural gene contains 1131 nucleotides and encodes 377 amino acids with the sequence: (sequence: in text). The independently determined NH2-terminal sequence of this enzyme (Tanase, S., Guirard, B. M., and Snell, E. E. (1985) J. Biol. Chem. 260, 6738-6746) and the amino acid sequences of two tryptic peptides reported in the accompanying paper (Hayashi, H., Tanase, S., and Snell, E. E. (1986) J. Biol. Chem. 261, 11003-11009) are localized in the sequence presented here; the lysine that binds pyridoxal phosphate is situated at residue 232, whereas the serine that binds the adduct formed between pyridoxal phosphate and the inhibitor alpha-fluoromethylhistidine is positioned at residue 322.  相似文献   

9.
Mammalian S-adenosylmethionine decarboxylase was expressed at a high level in an Escherichia coli mutant deficient in this enzyme. The proenzyme form of this enzyme was cleaved and processed to the mature decarboxylase which contains two pairs of nonidentical subunits, the larger of which contains a pyruvate prosthetic group. In order to determine the site of formation of the pyruvate, two approaches were used. First, the mammalian S-adenosylmethionine decarboxylase produced in E. coli was purified to homogeneity and the pyruvate converted to alanine by a reductive amination. The large subunit was then isolated by reversed phase high pressure liquid chromatography and the amino-terminal sequence determined and compared with the sequence of the proenzyme derived from its cDNA. These results indicated that the bond between glutamic acid 67 and serine 68 was the site of cleavage. Second, each of the serine residues in portion of the proenzyme likely to contain the cleavage site were altered by site-directed mutagenesis and the RNA produced from plasmids containing these mutations was translated in a reticulocyte lysate. The translation products were tested for processing and for S-adenosylmethionine decarboxylase activity. Altering the serine residues at positions 50, 66, and 69 to alanines had little effect but changing serine at position 68 to alanine completely prevented both processing and activity. These results indicate that the serine residue at position 68 of the proenzyme which is in the underlined position in the sequence -Leu-Ser-Glu-Ser-Ser-Met- is the residue which is converted to the pyruvate prosthetic group in human S-adenosylmethionine decarboxylase.  相似文献   

10.
The two cysteinyl residues present in histidine decarboxylase from Lactobacillus 30a differ greatly in reactivity. One (class 1) reacts readily in the native state with dithiobis-(2-nitrobenzoate) with complete loss of enzyme activity; the other (class 2) reacts only after denaturation of the enzyme (Lane, R. S., and Snell, E. E. (1976) Biochemistry 15, 4175-4179). These differences in reactivity permitted use of covalent (disulfide) chromatography to isolate separate peptides that contain these two residues. Sequence analysis showed that the class 1 cysteinyl residue is at position 147 in a hydrophilic portion of the alpha chain (Huynh, Q. K., Recsei, P. A., Vaaler, G. L., and Snell, E. E. (1984) J. Biol. Chem. 259, 2833-2839), while the class 2 cysteinyl residue is present at position 71, adjacent to a hydrophobic portion of the same chain. Cysteinyl peptides identical with or homologous to the class 2 cysteinyl peptide of the Lactobacillus 30a enzyme were isolated from the alpha subunits of histidine decarboxylases from Lactobacillus buchneri and Clostridium perfringens, respectively. The L. buchneri enzyme also contained a peptide homologous to the class 1 cysteinyl peptide from Lactobacillus 30a. However, no corresponding peptide was present in the enzyme from C. perfringens, in which the second cysteinyl residue of the alpha chain occupies position 3, very near the essential pyruvoyl residue. This enzyme, unlike those from Lactobacillus 30a or L. buchneri, also contains one cysteinyl residue in its beta chain. Although Cys 147 is an active site residue in histidine decarboxylase from Lactobacillus 30a, the absence of a corresponding residue in the C. perfringens enzyme confirms previous indications (Recsei, P. A., and Snell, E. E. (1982) J. Biol. Chem. 257, 7196-7202) that this SH group is not essential for decarboxylase action.  相似文献   

11.
12.
A phosphatidylserine-auxotrophic mutant of cultured Chinese hamster ovary cells, PSA-3, manifests a defect in phosphatidylserine synthase I activity (Kuge, O., Nishijima, M., and Akamatsu, Y. (1986) J. Biol. Chem. 261, 5790-5794). We cloned a Chinese hamster gene, designated pssC, which was able to transform the PSA-3 cell line to a phosphatidylserine prototroph. The resultant transformant contained phosphatidylserine in normal amounts but remained defective in phosphatidylserine synthase I activity, indicating that pssC is a suppressor gene. Using the genomic fragment of pssC as a probe, a cDNA clone of pssC was isolated, and its nucleotide sequence was determined. A computer search through a protein data bank revealed that pssC had homology with the Escherichia coli psd gene encoding the proenzyme of phosphatidylserine decarboxylase at the amino acid level. Introduction of the cloned pssC gene into PSA-3 resulted in a 2-fold increase in phosphatidylserine decarboxylase activity. When the pssC cDNA was placed downstream of the yeast GAL1 promoter and introduced into yeast Saccharomyces cerevisiae cells, the phosphatidylserine decarboxylase activity increased in a galactose-dependent manner. These results indicate that pssC encodes phosphatidylserine decarboxylase. The mechanism by which pssC complements the defect of PSA-3 in phosphatidylserine biosynthesis is discussed.  相似文献   

13.
We have previously shown that the gene (speD) for S-adenosylmethionine decarboxylase is part of an operon that also contains the gene (speE) for spermidine synthase (Tabor, C. W., Tabor, H., and Xie, Q.-W. (1986) Proc. Natl. Acad. Sci. U. S. A. 83, 6040-6044). We have now determined the nucleotide sequence of this operon and have found that speD codes for a polypeptide of Mr = 30,400, which is considerably greater than the subunit size of the purified enzyme. Our studies show that S-adenosylmethionine decarboxylase is first formed as a Mr = 30,400 polypeptide and that this proenzyme is then cleaved at the Lys111-Ser112 peptide bond to form a Mr = 12,400 subunit and a Mr = 18,000 subunit. The latter subunit contains the pyruvoyl moiety that we previously showed is required for enzymatic activity. Both subunits are present in the purified enzyme. These conclusions are based on (i) pulse-chase experiments with a strain containing a speD+ plasmid which showed a precursor-product relationship between the proenzyme and the enzyme subunits, (ii) the amino acid sequence of the proenzyme form of S-adenosylmethionine decarboxylase (derived from the nucleotide sequence of the speD gene), and (iii) comparison of this sequence of the proenzyme with the N-terminal amino acid sequences of the two subunits of the purified enzyme reported by Anton and Kutny (Anton, D. L., and Kutny, R. (1987) J. Biol. Chem. 262, 2817-2822).  相似文献   

14.
E Diaz  D L Anton 《Biochemistry》1991,30(16):4078-4081
S-Adenosylmethionine decarboxylase from Escherichia coli is a member of a small class of enzymes that uses a pyruvoyl prosthetic group. The pyruvoyl group is proposed to form a Schiff base with the substrate and then act as an electron sink facilitating decarboxylation. We have previously shown that once every 6000-7000 turnovers the enzyme undergoes an inactivation that results in a transaminated pyruvoyl group and the formation of an acrolein-like species from the methionine moiety. The acrolein then covalently alkylates the enzyme [Anton, D. L., & Kutny, R. (1987) Biochemistry 26, 6444]. After reduction of the alkylated enzyme with NaBH4, a tryptic peptide with the sequence Ala-Asp-Ile-Glu-Val-Ser-Thr-[S-(3-hydroxypropyl)Cys]-Gly-Val-Ile-Ser-Pro - Leu-Lys was isolated. This corresponds to acrolein alkylation of a cysteine residue in the second tryptic peptide from the NH2 terminal of the alpha-subunit [Anton, D. L., & Kutny, R. (1987) J. Biol. Chem. 262, 2817-2822]. The modified residue derived is from Cys-140 of the proenzyme [Tabor, C. W., & Tabor, H. (1987) J. Biol. Chem. 262, 16037-16040] and lies in the only sequence conserved between rat liver and E. coli S-adenosylmethionine decarboxylase [Pajunen et al. (1988) J. Biol. Chem. 263, 17040-17049]. We suggest that the alkylated Cys residue could have a role in the catalytic mechanism.  相似文献   

15.
16.
A cyclic nucleotide-independent protein kinase, protease-activated kinase II, which incorporates up to four phosphates into 40 S ribosomal protein S6, has been purified from the postribosomal supernatant of rabbit reticulocytes. Protease-activated kinase II was purified as an inactive proenzyme by chromatography on DEAE-cellulose, phosphocellulose, Sephadex G-150, and hydroxylapatite. The enzyme was activated in vitro by limited digestion with trypsin or chymotrypsin. No other mode of activation for protease-activated kinase II in vitro was identified. The proenzyme had a molecular weight of 80,000 as measured by gel filtration; following tryptic digestion, the molecular weight of the activated protein kinase was 45,000-55,000. Protease-activated kinase II required Mg2+ for activity but was inhibited by other divalent cations, monovalent cations, and fluoride ion. ATP was the phosphoryl donor in the phosphorylation reaction; GTP had no effect. In vitro, multiple phosphorylation of S6 was observed with some phosphate incorporated into S10. Phosphorylation of S6 by protease-activated kinase II has been shown to be stimulated in serum-starved 3T3-L1 cells by insulin (Perisic, O., and Traugh, J. A. (1983) J. Biol. Chem. 258, 9589-9592) and in reticulocytes by altering the pH of the incubation medium (Perisic, O., and Traugh, J. A. (1983) J. Biol. Chem. 258, 13998-14002.  相似文献   

17.
The alpha chain of histidine decarboxylase contains eight internal methionine residues. After reductive amination to convert the NH2-terminal pyruvoyl residue to an alanyl residue and cyanogen bromide treatment, 13 pure peptides were isolated. Four of these are incomplete cleavage products. The sequence of each of the remaining nine peptides was established by automated and manual degradation of the intact peptides and of smaller peptides obtained from tryptic, chymotryptic, and staphylococcal protease digests of the cyanogen bromide peptides. These results, together with the data on overlapping peptides reported in the accompanying paper (Huynh, Q. K., Recsei, P. A., Vaaler, G. L., and Snell, E. E. (1984) J. Biol. Chem. 259, 2833-2839), establish the complete amino acid sequence of the alpha chain.  相似文献   

18.
Previously reported inhibitions of heart lactate dehydrogenase (Guppy, M., and Hochachka, P.W. (1978) J. Biol. Chem. 253, 8465-8469) and muscle pyruvate kinase (Kemp, R.G. (1973) J. Biol. Chem. 248, 3963-3967) by creatine phosphate are due to oxalate which is a contaminant found in some commercial preparations of creatine phosphate.  相似文献   

19.
The role of juxtamembrane stalk glycosylation in modulating stalk cleavage and shedding of membrane proteins remains unresolved, despite reports that proteins expressed in glycosylation-deficient cells undergo accelerated proteolysis. We have constructed stalk glycosylation mutants of angiotensin-converting enzyme (ACE), a type I ectoprotein that is vigorously shed when expressed in Chinese hamster ovary cells. Surprisingly, stalk glycosylation did not significantly inhibit release. Introduction of an N-linked glycan directly adjacent to the native stalk cleavage site resulted in a 13-residue, proximal displacement of the cleavage site, from the Arg-626/Ser-627 to the Phe-640/Leu-641 bond. Substitution of the wild-type stalk with a Ser-/Thr-rich sequence known to be heavily O-glycosylated produced a mutant (ACE-JGL) in which this chimeric stalk was partially O-glycosylated; incomplete glycosylation may have been due to membrane proximity. Relative to levels of cell-associated ACE-JGL, rates of basal, unstimulated release of ACE-JGL were enhanced compared with wild-type ACE. ACE-JGL was cleaved at an Ala/Thr bond, 14 residues from the membrane. Notably, phorbol ester stimulation and TAPI (a peptide hydroxamate) inhibition of release-universal characteristics of regulated ectodomain shedding-were significantly blunted for ACE-JGL, as was a formerly undescribed transient stimulation of ACE release by 3, 4-dichloroisocoumarin. These data indicate that (1) stalk glycosylation modulates but does not inhibit ectodomain shedding; and (2) a Ser-/Thr-rich, O-glycosylated stalk directs cleavage, at least in part, by an alternative shedding protease, which may resemble an activity recently described in TNF-alpha convertase null cells [Buxbaum, J. D., et al. (1998) J. Biol. Chem. 273, 27765-27767].  相似文献   

20.
Clyne T  Kinch LN  Phillips MA 《Biochemistry》2002,41(44):13207-13216
S-Adenosylmethionine decarboxylase (AdoMetDC) is a pyruvoyl-dependent enzyme that is processed from a single polypeptide into two subunits creating the cofactor. In the human enzyme, both the proenzyme processing reaction and enzyme activity are stimulated by the polyamine putrescine. The processing reaction of Trypanosoma cruzi AdoMetDC was studied in an in vitro translation system. The enzyme was fully processed in the absence of putrescine, and the rate of this reaction was not stimulated by addition of the polyamine. Residues in the putrescine binding site of the human enzyme were evaluated for their role in processing of the T. cruzi enzyme. The E15A, I80K/S178E, D174A, and E256A mutant T. cruzi enzymes were fully processed. In contrast, mutation of R13 to Leu (the equivalent residue in the human enzyme) abolished processing of the T. cruzi enzyme, demonstrating that Arg at position 13 is a major determinant for proenzyme processing in the parasite enzyme. This amino acid change is a key structural difference that is likely to be a factor in the finding that putrescine has no role in processing of the T. cruzi enzyme. In contrast, the activity of T. cruzi AdoMetDC is stimulated by putrescine. Equilibrium sedimentation experiments demonstrated that putrescine does not alter the oligomeric state of the enzyme. The putrescine binding constant for binding to the T. cruzi enzyme (K(d) = 150 microM) was measured by a fluorescence assay and by ultrafiltration with a radiolabeled ligand. The mutant T. cruzi enzyme D174V no longer binds putrescine, and is not activated by the diamine. In contrast, mutation of E15, S178, E256, and I80 had no effect on putrescine binding. The k(cat)/K(m) values for E15A and E256A mutants were stimulated by putrescine to a smaller extent than the wild-type enzyme (2- and 4-fold vs 11-fold, respectively). These data suggest that the putrescine binding site on the T. cruzi enzyme contains only limited elements (D174) in common with the human enzyme and that the diamine plays different roles in the function of the mammalian and parasite enzymes.  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