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 15 毫秒
1.
2.
Flavokinase (ATP:riboflavin 5'-phosphotransferase) [EC 2.7.1.26] was purified to apparent homogeneity from rat intestinal mucosa by fractionation with ammonium sulfate, gel filtration, and flavin affinity chromatography. The addition of ATP to the enzyme solution was necessary for its binding to the affinity gel. The apparent molecular weight of the enzyme was estimated to be 13,500 by gel filtration on Sephadex G-100 and by SDS-PAGE. The properties of the enzyme, including its flavin specificity, were studied. Three types of riboflavin analogues were used for the flavin specificity study; namely, ones modified at the ribityl group, and at positions 3 and 8 of the isoalloxazine ring. Of the analogues modified at the ribityl group or position 3 of the isoalloxazine ring, only 2'-deoxyriboflavin was phosphorylated and then only weakly. On the other hand, most analogues modified at position 8 of the isoalloxazine ring were good substrates for the kinase, an appropriate increase in the substituent volume at position 8 of the isoalloxazine ring resulting in an increase in the Vmax value. In a previous paper on the mechanism of intestinal absorption of riboflavin, we proposed that one of the specific processes for the absorption of riboflavin is phosphorylation by flavokinase [Kasai, S. et al. (1988) J. Nutr. Sci. Vitaminol. 34, 265-280]. The present results support this conclusion because analogues that were absorbed at low concentrations through a process specific for riboflavin in our previous study were phosphorylated effectively by the enzyme, whereas those that were absorbed solely through simple diffusion at all concentrations were not phosphorylated or only phosphorylated weakly. The properties of the flavokinases from intestinal mucosa and liver were compared.  相似文献   

3.
The native flavin, FAD, was removed from chicken liver xanthine dehydrogenase and milk xanthine oxidase by incubation with CaCl2. The deflavoenzymes, still retaining their molybdopterin and iron-sulfur prosthetic groups, were reconstituted with a series of FAD derivatives containing chemically reactive or environmentally sensitive substituents in the isoalloxazine ring system. The reconstituted enzymes containing these artificial flavins were all catalytically active. With both the chicken liver dehydrogenase and the milk oxidase, the flavin 8-position was found to be freely accessible to solvent. The flavin 6-position was also freely accessible to solvent in milk xanthine oxidase, but was significantly less exposed to solvent in the chicken liver dehydrogenase. Pronounced differences in protein structure surrounding the bound flavin were indicated by the spectral properties of the two enzymes reconstituted with flavins containing ionizable -OH or -SH substituents at the flavin 6- or 8-positions. Milk xanthine oxidase either displayed no preference for binding of the neutral or anionic flavin (8-OH-FAD) or a slight preference for the anionic form of the flavin (6-hydroxy-FAD, 6-mercapto-FAD, and possibly 8-mercapto-FAD). On the other hand, the chicken liver dehydrogenase had a dramatic preference for binding the neutral (protonated) forms of all four flavins, perturbing the pK of the ionizable substituent greater than or equal to 4 pH units. These results imply the existence of a strong negative charge in the flavin binding site of the dehydrogenase, which is absent in the oxidase.  相似文献   

4.
The effect of riboflavin and iron on 6-hydroxy-2,4,5-triaminopyrimidine synthesis rate was investigated in the cultures of the yeast Pichia guilliermondii (rib2 mutants) with the blocked second reaction to flavinogenesis. It was shown that riboflavin inhibited the 6-hydroxy-2,4,5-triaminopyrimidine synthesis rate in iron-rich and iron-deficient cells of mutants with low riboflavin requirements. Cycloheximide did not prevent the stimulation of 6-hydroxy-2,4,5-triaminopyrimidine synthesis caused by riboflavin starvation. 7-methyl-8-trifluoromethyl-10-(1'-D-ribityl)isoalloxazine strongly inhibited the 6-hydroxy-2,4,5-triaminopyrimidine synthesis, while 7-methyl-8-trifluoro-methyl-10-(beta-hydroxyethyl)izoalloxazine and galactoflavin exerted only a slight effect on this process. The 6-hydroxy-2,4,5-triaminopyrimidine synthesis rate in iron-deficient cells was significantly higher than in iron-rich cells. The 2,2'-dipyridyl treatment of iron-rich cells caused the stimulation of 6-hydroxy-2,4,5-triaminopyrimidine synthesis and cycloheximide abolished this effect. The results suggest that the activity of the first enzyme of flavinogenesis (guanylic cyclohydrolase) is under the control of feedback inhibition by flavins and the biosynthesis of this enzyme is regulated by iron.  相似文献   

5.
Riboflavin deficient mutant Pichia guilliermondii MS1 which requires approximately 1000-fold lower concentration of exogenous vitamin B2 for growth when compared with a non-adapted riboflavin deficient mutants of this species was isolated by means of of UV-irradiation. The growth of the mutant was strongly inhibited by actinomycin D and L-canavanine. The revertant MS8 and MS14 which synthesized riboflavin were selected from the strain MS1. These revertants posses a multiple sensitivity to actinomycin D, rifamycin, euflavine, mitomycin C, antimycin A, 8-azaadenine, 8-azaguanine, L-canavanine and 7-methyl-8-trifluoromethyl-10-(1'-D-ribityl)isoalloxazine. The ability to utilized glycerol and ethanol as a sole carbon source for growth was impaired in these mutants. The mutants which can utilize glycerol were isolated from the strain MS14. Such mutants were resistant to actonomycin D. Mutation (s) which determines a multiple sensitivity and inability to utilized glycerol was recessive.  相似文献   

6.
By surveying compounds having isoalloxazine derived from flavins on a high performance liquid chromatogram with fluorescence detection, two new flavin derivatives were found in human urine. These two compounds were purified by partition chromatography on a cellulose column and by paper chromatography with several solvent systems, and their structures were determined to be 7 alpha-hydroxyriboflavin and 8 alpha-hydroxyriboflavin. The relative distributions, measured by high performance liquid chromatography, of 7 alpha- and 8 alpha-hydroxyriboflavin, riboflavin, and hydroxyethylflavin and its derivative were calculated to be 31.1, 5.0, 25.6, 4.9, and 21.9%, respectively, to total flavins in normal human urine obtained in early morning. The excretion of 7 alpha- and 8 alpha-hydroxyriboflavin in human urine indicates the occurrence of a metabolic pathway of the isoalloxazine ring of flavin at its 7 alpha and 8 alpha positions.  相似文献   

7.
Murray TA  Foster MP  Swenson RP 《Biochemistry》2003,42(8):2317-2327
A mechanism has been proposed for the binding of flavin mononucleotide (FMN) and riboflavin to the apoflavodoxin from Desulfovibrio vulgaris [Murray, T. A., and Swenson, R. P. (2003) Biochemistry 42, 2307-2316]. In this model, the binding of the flavin isoalloxazine ring is dependent on the presence of a phosphate moiety in the phosphate-binding subsite, suggesting a cooperative interaction between that region and the ring-binding subsite. In the absence of inorganic phosphate, FMN can bind through the initial association of its 5'-phosphate group in the phosphate-binding subsite followed by insertion of the flavin ring. Because riboflavin lacks the 5'-phosphate group, it is unable to bind to this apoprotein in the absence of inorganic phosphate in solution. However, inorganic phosphate can enhance the rate of ring binding by occupying the phosphate-binding subsite. In this paper, NMR, near-UV circular dichroism (CD), and fluorescence spectroscopy provide evidence for a phosphate-induced conformational change within the isoalloxazine ring-binding subsite. Phosphate-dependent changes in the chemical shifts of 22 amide groups were observed in (1)H-(15)N HSQC NMR spectra. The majority of these groups are proximal to the phosphate-binding subsite or the loops that constitute the isoalloxazine ring-binding site. Also, a phosphate-dependent change in the environment or position of the Trp60 side chain was apparent in the NMR data and was confirmed by associated changes in the near-UV CD and tryptophan fluorescence spectra when compared to the spectra of the W60A mutant. These data suggest that phosphate, either the 5'-phosphate of the FMN or inorganic phosphate from solution, facilitates the movement of the side chain of Trp60 out of the isoalloxazine ring-binding site and other associated conformational changes, creating a population of apoflavodoxin that is capable of binding the isoalloxazine ring. This conformational switch may explain why some apoflavodoxins cannot bind riboflavin and also supports the "aromatic gate" model proposed from the crystal structure of the Anabaena apoflavodoxin [Genzor, C. G., Perales-Alcon, A., Sancho, J., and Romero, A. (1996) Nat. Struct. Biol. 3, 329-332].  相似文献   

8.
The effect of riboflavin and iron on 6-hydroxy-2,4,5-triaminopyrimidine synthesis rate was investigated in the cultures of the yeast Pichia guilliermondii (rib2 mutants) with the blocked second reaction of flavinogenesis.It was shown that riboflavin inhibited the 6-hydroxy-2,4,5-triaminopyrimidine synthesis rate in iron-rich and iron-deficient cells of mutants with low riboflavin requirements. Cycloheximide did not prevent the stimulation of 6-hydroxy-2,4,5-triaminopyrimidine synthesis caused by riboflavin starvation.7-methyl-8-trifluoromethyl-10-(1′- -ribityl)isoalloxazine strongly inhibited the 6-hydroxy-2,4,5-triaminopyrimidine synthesis, while 7-mithyl-8-trifluoromethyl-10-(β-hydroxyethyl) izoalloxazine and galactoflavin exerted only a slight effect on this process.The 6-hydroxy-2,4,5-triaminopyrimidine synthesis rate in iron-deficient cells was significantly higher than in iron-rich cells. The 2,2′-dipyridyl treatment of iron-rich cells caused the stimulation of 6-hydroxy-2,4,5-triaminopyrimidine synthesis and cycloheximide abolished this effect.The results suggest that the activity of the first enzyme of flavinogenesis (guanylic cyclohydrolase) is under the control of feedback inhibition by flavins and the biosynthesis of this enzyme is regulated by iron.  相似文献   

9.
10.
From a survey of flavin compounds using a high performance liquid chromatograph equipped with a fluorescence detector, 8 alpha-hydroxyriboflavin, its derivative, riboflavin, FMN, and FAD were found in rat organs. The derivative of 8 alpha-hydroxyriboflavin was purified from the kidney sequentially by DEAE column chromatography, paper chromatography, and high performance liquid chromatography, and was determined to be 8 alpha-hydroxyriboflavin 5'-phosphate. The amount of 8 alpha-hydroxyriboflavin in the brain, heart, liver, and kidney of the rat was calculated to be 0.010, 0.026, 0.172, and 0.103 nmol per g of wet tissue, respectively. The amount of 8 alpha-hydroxyriboflavin 5'-phosphate in the brain, heart, and liver was less than 0.002 nmol, while that in the kidney was 0.330 nmol, per g of wet tissue. On the other hand, we could not detect 7 alpha-hydroxyriboflavin in the organ extracts, even though it was excreted into human urine along with 8 alpha-hydroxyriboflavin. This is the first reported occurrence of 8 alpha-hydroxyriboflavin and 8 alpha-hydroxyriboflavin 5'-phosphate in higher animal organs.  相似文献   

11.
Metabolite profiling in succinate semialdehyde dehydrogenase (SSADH; Aldh5a1-/-) deficient mice previously revealed elevated gamma-hydroxybutyrate (GHB) and total GABA in urine and total brain and liver extracts. In this study, we extend our metabolic characterization of these mutant mice by documenting elevated GHB and total GABA in homogenates of mutant kidney, pancreas and heart. We quantified beta-alanine (a GABA homolog and putative neurotransmitter) to address its potential role in pathophysiology. We found normal levels of beta-alanine in urine and total homogenates of mutant brain, heart and pancreas, but elevated concentrations in mutant kidney and liver extracts. Amino acid analysis in mutant total brain homogenates revealed no abnormalities except for significantly decreased glutamine, which was normal in mutant liver and kidney extracts. Regional amino acid analysis (frontal cortex, parietal cortex, hippocampus and cerebellum) in mutant mice confirmed glutamine results. Glutamine synthetase protein and mRNA levels in homogenates of mutant mouse brain were normal. We profiled organic acid patterns in mutant brain homogenates to assess brain oxidative metabolism and found normal concentrations of Kreb's cycle intermediates but increased 4,5-dihydroxyhexanoic acid (a postulated derivative of succinic semialdehyde) levels. We conclude that SSADH-deficient mice represent a valid metabolic model of human SSADH deficiency, manifesting focal neurometabolic abnormalities which could provide key insights into pathophysiologic mechanisms.  相似文献   

12.
In addition to 8alpha-(N3-histidyl)riboflavin, 8alpha-(N1-histidyl)riboflavin is also formed during the reaction of Nalpha-blocked histidine with 8alpha-bromotetraacetylriboflavin in a yield of 20-25% of the total histidylflavin fraction. The properties of 8alpha-(N1-histidyl)riboflavin are inditical with those of the histidylflavin isolated from thiamine dehydrogenase and beta-cyclopiazonate oxidocyclase but differ from those of 8alpha-(N3-histidyl)riboflavin. These properties include pKa of fluorescence quenching, electrophoretic mobility at pH 5.0, stability to storage, and reduction by NaBH4. Proof for 8alpha substitution is shown by the electron paramagnetic resonance and electron-nuclear double resonance spectra of the cationic semiquinone form, as well as by the proton magnetic resonance spectrum of the oxidized form. The site of histidine substitution by the 8alpha-methylene of the flavin moiety was shown by methylation of the imidazole ring with methyl iodide, cleavage of the methylhistidine-flavin bond by acid hydrolysis at 150 degrees C, and identification of the methylhistidine isomer by electrophoresis. 3-Methylhistidine is the product from the N1-histidylflavin isomer, while 1-methylhistidine is produced from the N3 isomer. The flavin product from reductive Zn cleavage of either isomer has been identified as riboflavin. The compound obtained on acid treatment of 8alpha-(N3-histidyl)riboflavin (previously thought to be the N1 isomer) differs from the parent compound only in the ribityl side chain, since chemical degradation studies show 1-methylhistidine as a product and a flavin product which differs from riboflavin only in mobility in thin-layer chromatography, but not in absorption, fluorescence, and electron paramagnetic resonance spectral properties. Proof that acid modification involves only the ribityl chain has come from the observations that alkaline irradiation of this flavin yields lumiflavin, that the proton magnetic resonance spectrum of the compound differs from that of riboflavin in the region of the ribityl proton resonance, and that its periodate titer is lower than that of authentic riboflavin. The identity of 8alpha-(N1-histidyl)riboflavin with the histidylflavin from thiamine dehydrogenase and beta-cyclopiazonate oxidocyclase shows that both isomeric forms of 8alpha-histidylflavin occur in nature.  相似文献   

13.
Ulrike Dohrmann 《Planta》1983,159(4):357-365
Several types of membrane-localized flavin binding sites were investigated in sporangiophores (spph) and mycelia of Phycomyces blakesleeanus. In-vitro binding of riboflavin, riboflavin-5′-phosphate, and flavin-adenine-dinucleotide was demonstrated with unfractionated membrane preparations by means of competition of [14C]riboflavin binding. Saturation of binding was only obtained with the highly water-soluble riboflavin-5′-phosphate, but by extrapolation it was shown that riboflavin showed the highest affinity towards the binding sites (KD about 4·10-6M). The number of binding sites was estimated to be 0.7 nmol g-1 fresh-weight equivalent. Analysis of endogenous soluble flavin revealed that only riboflavin, riboflavin-5′-phosphate, and flavin-adenine-dinucleotide occurred in Phycomyces, and at a concentration of at least 1 nmol g-1 fresh-weight equivalent in entire spph. Thus, the measured binding sites could reach saturation in-vivo. In the apical part of spph to which blue-light sensitivity is restricted, the amount of soluble flavin was three-fold higher. Exclusively in this zone, heat-labile riboflavin proteins were measured at a concentration of about 3 nmol g-1 fresh-weight equivalent. The amount of covalently bound flavin was higher in spph tips than in intact spph (8 nmol and 3 nmol g-1 fresh-weight equivalent, respectively). In either case, the concentrations of the flavin-membrane complexes were higher than the theoretical calculated concentration of (anisotropic) blue-light photoreceptor in Phycomyces (Bergman et al. 1969), and their involvement in blue-light photoreception is considered.  相似文献   

14.
Abstract: We have identified succinic semialdehyde dehydrogenase protein in rat and human neural and nonneural tissues. Tissue localization was determined by enzymatic assay and by western immunoblotting using polyclonal antibodies raised in rabbit against the purified rat brain protein. Although brain shows the highest level of succinic semialdehyde dehydrogenase activity, substantial amounts of enzyme activity occur in mammalian liver, pituitary, heart, and ovary. We further demonstrate the absence of succinic semialdehyde dehydrogenase enzyme activity and protein in brain, liver, and kidney tissue samples from an individual affected with succinic semialdehyde dehydrogenase deficiency, thereby verifying the specificity of our antibodies.  相似文献   

15.
As in the case of the succinate and sarcosine dehydrogenases of liver mitochondria, the flavin prosthetic group of the bacterial sarcosine dehydrogenase can be released from the enzyme by proteolytic digestion with trypsin and chymotrypsin. The flavin, isolated in the dinucleotide form and covalently bound to a peptide fragment, is converted to the mononucleotide and purified by sequential chromatography on Sephadex G-25, DEAE-Sephadex A-25, followed by preparative paper chromatography and high voltage electrophoresis.The absorption maxima of the purified flavin at pH 7 are found at 268, 350, and 447 nm, with 268:447 nm and 350:447 nm ratios of 3.08 and 0.79, respectively. The fluorescence excitation and emission maxima, 450 and 530 nm, respectively, are similar to those of flavin mononucleotide. The fluorescence of the flavin-peptide is maximal at pH 3.0–3.1.Amino acid analysis of the flavin-peptide (riboflavin form) gave the following molar ratios of amino acids to flavin: Lys(1), Asp(2), Thr(1), Ser(1), Glu(1), Gly(1), and Ala(1). Aspartic acid was the N-terminal amino acid. Upon more extensive hydrolysis, histidine was obtained in 71–84% yields. Employing aminopeptidase M, the partial sequence of amino acids in the flavin-peptide was determined to be as follows: Flavin
-Asp-Lys-Ser-Glu-Gly-His-(Asp,Ala,Thr)-Evidence is presented that the isoalloxazine ring is linked covalently via its 8 α-methyl group to N-3 of histidine.  相似文献   

16.
Riboflavin uptake by washed cells of riboflavin deficient mutant MS1-3 of Pichia guilliermondii yeast was strongly depressed by D-glucose, L-sorbose, alpha-methyl-D-glucoside, sucrose, trehalose, maltose and salicin but not by D-mannose, D-galactose, D-fructose or ribitol. Glucose decreased also the initial uptake rate of riboflavin analogue, 8-piperidyl-10-(1'-D-galactityl) isoalloxazine; the inhibition having a competitive character (Ki==5,7 mM). Apparently riboflavin permease is able to accept not only riboflavin and its analogues but also glucose and some of glucose derivates. Cells preloaded with riboflavin and transferred into riboflavin-free medium excreted vitamin B2 into the medium. This excretion was strongly stimulated by D-glucose, D-fructose, D-mannose but not by citrate or succinate. In contrast to riboflavin, 8-piperidyl-10-(1'-D-galactityl) isoalloxazine was not excreted into the medium even in the presence of glucose. The rate of riboflavin excretion depended on temperature and pH of incubation medium (pH optimum approximately 7.0) and was decreased in the presence of different inhibitors of energy metabolism. It seems that the exit of riboflavin from the cells is accomplished by energy-dependent specific system of excretion (excretase) which in some properties is different from that of riboflavin permease.  相似文献   

17.
The flavoenzymes dimethylglycine dehydrogenase (EC 1.5.99.2) and sarcosine dehydrogenase (EC 1.5.99.1) contain covalently bound FAD linked via the 8 alpha-position of the isoalloxazine ring to the imidazole N(3) of a histidine residue (Cook, R. J., Misono, K. S., and Wagner, C. (1984) J. Biol. Chem. 259, 12475-12480). The flavin-peptides from tryptic digests of these two enzymes have been isolated and sequenced. Automated sequence analysis showed that the flavin-peptide from dimethylglycine dehydrogenase contained 25 amino acid residues in the following sequence: Ser-Glu-Leu-Thr-Ala-Gly-Ser- Thr-Trp-His(flavin)-Ala-Ala-Gly-Leu-Thr-Thr-Tyr-Phe-His-Pro-Gly-Ile-A sn-Leu-Lys. The sequence determined for the flavin-peptide from sarcosine dehydrogenase contained 14 amino acid residues Leu-Thr-Ser-Gly-Thr-Thr-Trp-His(flavin)-Thr-Ala-Gly-Leu-Gly-Arg.  相似文献   

18.
Flavin reductases use flavins as substrates and are distinct from flavoenzymes which have tightly bound flavins. The reduced flavin can serve to reduce ferric complexes and iron proteins. In Escherichia coli, reactivation of ribonucleotide reductase is achieved by reduced flavins produced by flavin reductase. The crystal structure of E. coli flavin reductase reveals that the enzyme structure is similar to the structures of the ferredoxin reductase family of flavoproteins despite very low sequence similarities. The main difference between flavin reductase and structurally related flavoproteins is that there is no binding site for the AMP moiety of FAD. The direction of the helix in the flavin binding domain, corresponding to the phosphate binding helix in the flavoproteins, is also slightly different and less suitable for phosphate binding. Interactions for flavin substrates are instead provided by a hydrophobic isoalloxazine binding site that also contains a serine and a threonine, which form hydrogen bonds to the isoalloxazine of bound riboflavin in a substrate complex.  相似文献   

19.
Chemical synthesis and some properties of 6-substituted flavins   总被引:2,自引:0,他引:2  
A number of derivatives of riboflavin and of 3-methyllumiflavin substituted in the 6 position have been synthesized starting with 6-nitro flavins, reduction to the 6-amino flavin, and diazotization, followed by reaction with the appropriate nucleophile. The absorption spectra, oxidation-reduction potentials, and the electron spin resonance spectra of the radical cationic forms of several of these synthetic compounds have been determined, including 6-S-cysteinyl-3-methyllumiflavin and 6-S-cysteinylriboflavin. The latter has been shown to be identical with the dephosphorylated form of the aminoacyl flavin isolated from trimethylamine dehydrogenase [Steenkamp, D. J., Kenney, W. C. & Singer, T. P. (1978) J. Biol. Chem. 253, 2812-2817; Steenkamp, D. J., McIntire, W., & Kenney, W. C. (1978) J. Biol. Chem. 253, 2818-2824] in regard to absorption specturm, photochemical properties, and mobility in high-voltage electrophoresis and in thin-layer chromatography. An unusually pronounced interaction between the amino group and the isoalloxazine ring system was deduced from the absorption spectra of 6-amino-3-methyllumiflavin and 6-aminoriboflavin.  相似文献   

20.
Murray TA  Swenson RP 《Biochemistry》2003,42(8):2307-2316
The pathway(s) by which the flavin cofactor binds to the apoflavoprotein is the subject of some debate. The crystal and NMR structures of several different flavodoxins have provided some insight, although there is disagreement about the location of the initial interaction between the flavin mononucleotide (FMN) and the apoflavodoxin and the degree of protein conformational change associated with cofactor binding [Genzor, C. G., Perales-Alcon, A., Sancho, J., and Romero, A. (1996) Nat. Struct. Biol. 3, 329-332; Steensma, E., and van Mierlo, C. P. M. (1998) J. Mol. Biol. 282, 653-666]. Binding kinetics using stopped-flow spectrofluorimetry and phosphate competition studies were used to develop a model for flavin binding to the flavodoxin from Desulfovibrio vulgaris. In the presence of phosphate, the time course of fluorescence quenching associated with FMN binding to apoflavodoxin was biphasic, whereas riboflavin, which lacks the 5'-phosphate group of FMN, displayed monophasic binding kinetics. When the concentration of phosphate in solution was increased, the FMN binding rates of the two phases behaved differently; the rate of one phase decreased, while the rate of the other increased. A similar increase in the single phase associated with riboflavin binding was also observed. This has led to the following model. The binding of the flavin isoalloxazine ring to its subsite is dependent on the presence of a phosphate group in the phosphate-binding subsite. When phosphate is in the buffer solution, FMN can bind in either of two ways: by the initial insertion of the 5'-phosphate group followed by ring binding or, when inorganic phosphate from solution is bound, the insertion of the isoalloxazine ring first. Riboflavin, which lacks the phosphate moiety of FMN, binds only in the presence of inorganic phosphate, presumably due to the binding of this group in the phosphate-binding subsite. These results suggest that cooperative interactions exist between the phosphate subsite and the ring-binding region in the D. vulgaris flavodoxin that are necessary for isoalloxazine ring binding.  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