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.
A (1→3),(1→4)-β-glucan synthase catalysing the synthesis of (1→3),(1→4)-β-glucan (mixed-linkage glucan) was investigated using microsomal membranes prepared from developing barley (Hordeum vulgare L. cv. Shikokuhadaka 97) endosperms harvested 21 days after flowering. The microsomal fraction produced (1→3),(1→4)-β-glucan by incorporation of [14C]Glc from UDP-[14C]Glc. The production of (1→3),(1→4)-β-glucan was ascertained by specific enzymatic digestion with endo-(1→3),(1→4)-β-glucanase (lichenase; EC 3.2.1.73) from Bacillus amyloliquefaciens, which released a radiolabelled trisaccharide (3-O-β-cellobiosyl-glucose) and a tetrasaccharide (3-O-β-cellotriosyl-glucose), the diagnostic oligosaccharides for the identification of (1→3),(1→4)-β-glucan. Digestion of the products with exo-(1→3)-β-glucanase (EC 3.2.1.58) from Basidiomycete QM806 released radiolabelled Glc, indicating that not only (1→3),(1→4)-β-glucans but also (1→3)-β-glucans (callose) had been formed due to the presence of (1→3)-β-glucan (callose) synthase (EC 2.4.1.34) in the microsomal fraction. The activity of (1→3),(1→4)-β-glucan synthase was maximal at pH 9.0 and at 25°C and in the presence of at least 2 mM Mg2+. The apparent Km and Vmax values for UDP-Glc were 0.33 mM and 480 pmol min−1 mg protein−1, respectively. Investigating the dependence of enzyme activity on developmental stage (7–35 days after flowering) of the endosperms, we found an increase of activity during the initial development reaching a maximum at 19 days, followed by a gradual decrease as the endosperms matured. The amount of (1→3),(1→4)-β-glucan in the cell walls of the endosperms, however, increased gradually towards maturation, even after 19 days. Analysing the relationship between enzyme activity and (1→3),(1→4)-β-glucan deposition in cell walls of endosperms prepared from 12 different barley varieties harvested 11–22 days after flowering showed that some varieties had both low activity and low glucan content, and in some both were high. But for several other varieties, the availability of donor substrate and other factors seem to influence the production of (1→3),(1→4)-β-glucan as well.  相似文献   

2.
A simple, rapid procedure has been developed to purify (1→3)-β-glucan synthase (UDP-Glc:(1→3)-β-glucan 3-β-d -glucosyl transferase (EC 2.4.1.34)) over 400-fold from membrane preparations of Italian ryegrass (Lolium multiflorum) and 60-fold from CHAPS-extracted membranes. When a CHAPS-extract of the membranes is treated with 8 mM CaCl2, proteinaceous material is precipitated. Although less than 10% of CHAPS-solubilized protein is removed in this step, the total activity recovered in the supernatant increases fourfold. Thus, CaCl2 precipitation appears to be important in removing inhibitors of the (1→3)-β-glucan synthase. In the presence of 1 mM UDP-glucose, the supernatant after CaCl2 treatment produces a high molecular weight, insoluble product that entraps a (1→3)-β-glucan synthase of high specific activity. The product-entrapped enzyme preparation contains six major polypeptides, and comparison of the SDS—PAGE pattern of this fraction with the polypeptide profile of an immunoprecipitated (1→3)-β-glucan synthase preparation suggests that polypeptides at 30–31 and 55–58 kDa are the most likely candidates for participation in (1→3)-β-glucan synthesis. When the reaction is performed on a larger scale, milligram quantities of product can be seen precipitating from the reaction mixture within 1 h of substrate addition. This product has been characterized by methylation analysis, 1H- and 13C-nmr spectroscopy, X-ray diffraction, electron microscopy, size exclusion chromatography, UV-induced fluorescence in the presence of the (1→3)-β-glucan-specific fluorochrome from aniline blue, and enzymic hydrolysis with a specific (1→3)-β-glucanase. These physical, chemical and enzymic analyses clearly demonstrate that the product is a microfibrillar (1→3)-β-glucan with a degree of polymerization of about 1500.  相似文献   

3.
The evolutionary relationships of two classes of plant β-glucan endohydrolases have been examined by comparison of their substrate specificities, their three-dimensional conformations and the structural features of their corresponding genes. These comparative studies provide compelling evidence that the (1→3)-β-glucanases and (1→3,1→4)-β-glucanases from higher plants share a common ancestry and, in all likelihood, that the (1→3,1→4)-β-glucanases diverged from the (1→3)-β-glucanases during the appearance of the graminaceous monocotyledons. The evolution of (1→3,1→4)-β-glucanases from (1→3)-β-glucanases does not appear to have invoked ‘modular’ mechanisms of change, such as those caused by exon shuffling or recombination. Instead, the shift in specificity has been acquired through a limited number of point mutations that have resulted in amino acid substitutions along the substrate-binding cleft. This is consistent with current theories that the evolution of new enzymic activity is often achieved through duplication of the gene encoding an existing enzyme which is capable of performing the required chemistry, in this case the hydrolysis of a glycosidic linkage, followed by the mutational alteration and fine-tuning of substrate specificity. The evolution of a new specificity has enabled a dramatic shift in the functional capabilities of the enzymes. (1→3)-β-Glucanases that play a major role, inter alia, in the protection of the plant against pathogenic microorganisms through their ability to hydrolyse the (1→3)-β-glucans of fungal cell walls, appear to have been recruited to generate (1→3,1→4)-β-glucanases, which quite specifically hydrolyse plant cell wall (1→3,1→4)-β-glucans in the graminaecous monocotyledons during normal wall metabolism. Thus, one class of β-glucan endohydrolase can degrade β-glucans in fungal walls, while the other hydrolyses structurally distinct β-glucans of plant cell walls. Detailed information on the three-dimensional structures of the enzymes and the identification of catalytic amino acids now present opportunities to explore the precise molecular and atomic details of substrate-binding, catalytic mechanisms and the sequence of molecular events that resulted in the evolution of the substrate specificities of the two classes of enzyme.  相似文献   

4.
An immunological assay has been used to investigate the synthesis of (1→3,1→4)-β-glucanase (EC 3.2.1.73) isoenzymes from isolated barley aleurone layers and scutella. Enzyme release from both tissues is enhanced by 1 micromolar gibberellic acid and 10 millimolar Ca2+, although increases induced by gibberellic acid are observed only in the presence of Ca2+. Isoenzyme I is synthesized predominantly in the scutellum, while isoenzyme II is synthesized exclusively in the aleurone. A third, putative isoenzyme III has been detected in significant proportions in scutellar secretions and may also be secreted from aleurone layers. Both gibberellic acid and Ca2+ appear to preferentially enhance isoenzyme II secretion from the aleurone and isoenzyme III secretion from scutella. The patterns of isoenzyme secretion are suggestive of tissue-specific differences in expression of the genes which code for (1→3,1→4)-β-glucanase isoenzymes. Qualitatively similar results were obtained with barley cultivars harvested in Australia and North America.  相似文献   

5.
Glycolipid antigen reacting to the monoclonal antibody directed to the developmentally regulated antigen SSEA-1 was isolated from human erythrocytes and colonic adenocarcinoma. The antigens have the Lex (Galβl→4[Fucα]→3]GlcNAcβl→R) or Ley (Fucαl→2Galβl→4[Fucαl→3]GlcNAcβl→R) structure at the termini of the branched polylactosaminolipid. In addition, a novel polyfucosyl structure locating exclusively at the internal GlcNAc was detected in the tumor antigen. The antibody reacts with a simple monovalent Lex glycolipid (Galβl→4[Fucαl→3]GlcNAcβl→3Galβl→4Glcβl→Cer) previously isolated from colonic carcinoma when presented at a high density on liposomes. The antibody therefore may react to the bivalent or multivalent Lex or Ley structure.  相似文献   

6.
Lichenan a β(1→ 3,1→ 4)β-glucan exhibited a broad antiviral activity against mechanically transmitted viruses of different taxonomic groups in different Nicoliana spp. As shown for tobacco mosaic virus in vitro . lichenan did not interfere with cell-free protein synthesis programmed with naked RNA or virus particles cotranslationally disassembled. Thus, inhibition of virus replication by lichenan is probably not attributed to direct interaction of lichenan and viral structural components, but rather to reduced sensitivity of the host tissue. Treatment of a suspension of cultured cells of Nicotiana tabacum with lichenan was accompanied by a slight increase in phenylalanine ammonia-lyase (PAL) activity. However, lichenan applied to leaf tissue did not significantly stimulate PAL-activity, therefore, antiviral action apparently did not depend on the induction of the phenylpropanoid pathway. According to experiments with lichenan and digitonin (used as a callose elicitor) restriction of viruses to initially-infected cells by means of callose deposition also appears to be an improbable mechanism of viral inhibition. In cell suspensions treated with lichenan, slight changes in the extracellular concentration of free K+ were observed that did not reflect a marked K+ leakage, as caused by digitonin. This finding requires further examination. According to the present results, it is likely that lichenan affects the early virus-cell interactions that occur after virus disassembly.  相似文献   

7.
Polyclonal antibodies raised against barley (1→3,1→4)-β-d-glucanase, α-amylase and carboxypeptidase were used to detect precursor polypeptides of these hydrolytic enzymes among the in vitro translation products of mRNA isolated from the scutellum and aleurone of germinating barley. In the scutellum, mRNA encoding carboxypeptidase appeared to be relatively more abundant than that encoding α-amylase or (1→3,1→4)-β-d-glucanase, while in the aleurone α-amylase and (1→3,1→4)-β-d-glucanase mRNAs predominated. The apparent molecular weights of the precursors for (1→3,1→4)-β-d-glucanase, α-amylase, and carboxypeptidase were 33,000, 44,000, and 35,000, respectively. In each case these are slightly higher (1,500-5,000) than molecular weights of the mature enzymes. Molecular weights of precursors immunoprecipitated from aleurone and scutellum mRNA translation products were identical for each enzyme.  相似文献   

8.
Four different H-type 1 (LedH) blood-group-active glycosphingolipids (LedH-I–IV) have been isolated from the plasma of blood-group O Le(a?b?) secretors. The agglutination of O Le(a?b?) erythrocytes from secretors by 50 μl of 4 hemagglutinating units of caprine anti-LedH (anti-H-type 1) serum was inhibited by 0.02 μg of each of all four glycolipids. No Lea or Leb activities or reaction against Ulex europaeus lectin could be found. LedH-I, -II, -III, and -IV at 0.05, 0.01, 0.01, and 0.02 μg each are sufficient for incubation in order to convert 9 × 107 O Le(a?b?) erythrocytes from nonsecretors into H-type 1 (LedH)-positive cells. Structural analysis of the H-type 1 glycolipids was performed in comparison to that of Lea- and Leb-blood-group-active glycolipids from human plasma isolated previously: Gas chromatography of peracetylated alditols revealed sugar composition. Combined gas chromatography-mass spectrometry established the glycosidic linkages. Together with the results obtained by direct inlet mass spectrometry of permethylated glycosphingolipids and by 360-MHz 1H nuclear magnetic resonance spectroscopy (Egge, H., and Hanfland, P., 1981, Arch. Biochem. Biophys., 210, 396–404; Dabrowski, J., Hanfland, P., Egge, H., and Dabrowski, U., 1981, Arch. Biochem. Biophys., 210, 405–411) the complete structures of the oligosaccharide chains of the Lea-, Leb-, and H-type 1-active glycolipids were established: Galβ1 → 3GlcNAc(4 ← 1αFuc)β1 → 3Galβ1 → 4Glcβ1 → 1 Cer for the Lea antigens; Fucα1 → 2Galβ1 → 3GlcNAc(4 ← 1αFuc)β1 → 3Galβ1 → 4Glcβ1 → 1 Cer for the Leb antigens; and Fucα1 → 2Galβ1 → 3GlcNAcβ1 → 3Galβ1 → 4Glcβ1 → 1 Cer for the H-type 1 (LedH) glycolipids. The diverse antigens of the same blood-group specificity obviously differ from one another in their lipid residue. In addition, plasmatic neolactotetraosylceramide could be identified, differing from that of human erythrocytes by a slower migration behavior in thin-layer chromatography.  相似文献   

9.
Blood group H antigen with globo-series structure, reacting with the monoclonal antibody MBrl, was isolated and characterized from human blood group O erythrocytes. The structure was identified by methylation analysis, direct probe mass spectrometry, and 1H-nuclear magnetic resonance spectroscopy as shown below: Fucαl → 2Galβl → 3GalNAcβl → 3Galαl → 4Galβl → 4Glcβl → 1Cer  相似文献   

10.
Antiserum against galactosyl(α1 → 4)galactosyl(β1 → 4)glucosylceramide (globotriaosylceramide, Gb3) was raised in rabbits by the administration of four weekly intramuscular injections of 1.5 mg of the purified glycolipid along with bovine serum albumin and Freund's complete adjuvant. AntiGb3 activity was quantitated initially by immunoprecipitation employing Gb3 mixed with 100-fold excess of lecithin and cholesterol (1 : 1 or 1 : 2, by wt.) as antigen. Subsequently, complement fixation tests done with antigen preparations containing Gb3/lecithin/cholesterol (1 : 6 : 20, by wt.) showed antiGb3 titres of up to 1 : 8192. Fractionation of the antiserum by BioGel A5m chromatography indicated the antibody was an IgM immunoglobulin. The partially purified antibody stimulated complement-dependent release of glucose from glucose-containing liposomes prepared with sphingomyelin/cholesterol/dicetylphosphate/Gb3 (molar ratio, 100 : 75 : 11 : 1). The antibody crossreacted with the trisaccharide, Gal(α1 → 4)Gal(β1 → 4)Glc, but not with galactosylceramide, lactosylceramide, GM1 ganglioside, globotetraosylceramide, digalactosyldiglyceride or a number of low molecular weight saccharides.  相似文献   

11.
12.
13.
The action on tamarind seed xyloglucan of the pure, xyloglucan-specific endo-(1→4)-β-D-glucanase from nasturtium (Tropaeolum majus L.) cotyledons has been compared with that of a pure endo-(1→4)-β-D-glucanase (‘cellulase’) of fungal origin. The fungal enzyme hydrolysed the polysaccharide almost completely to a mixture of the four xyloglucan oligosaccharides: Exhaustive digestion with the nasturtium enzyme gave the same four oligosaccharides plus large amounts of higher oligosaccharides and higher-polymeric material. Five of the product oligosaccharides (D,E,F,G,H) were purified and shown to be dimers of oligosaccharides A to C. D (glc8xyl6) had the structure A→A, H (glc8xyl6gal4) was C→C, whereas E (glc8xyl6gal), F (glc8xyl6gal2) and G (glc8xyl6gal3) were mixtures of structural isomers with the appropriate composition. For example, F contained B2→B2 (30%), A→C (30%), C→A (20%), B2B1 (15%) and others (about 5%). At moderate concentration (about 3 mM) oligosaccharides D to H were not further hydrolysed by the nasturtium enzyme, but underwent transglycosylation to give oligosaccharides from the group A, B, C, plus higher oligomeric structures. At lower substrate concentrations, hydrolysis was observed. Similarly, tamarind seed xyloglucan was hydrolysed to a greater extent at lower concentrations. It is concluded that the xyloglucan-specific nasturtium-seed endo-(1→4)-β-D-glucanase has a powerful xyloglucan-xyloglucan endo-transglycosylase activity in addition to its known xyloglucan-specific hydrolytic action. It would be more appropriately classified as a xyloglucan endo-transglycosylase. The action and specificity of the nasturtium enzyme are discussed in the context of xyloglucan metabolism in the cell walls of seeds and in other plant tissues.  相似文献   

14.
Using anion-exchange chromatography on Source 15Q followed by hydrophobic interaction chromatography on Source 15 Isopropyl, a lichenase-like endo-(1→4)-β-glucanase (BG, 28 kDa, pI 4.1) was isolated from a culture filtrate of Aspergillus japonicus. The enzyme was highly active against barley β-glucan and lichenan (263 and 267 U/mg protein) and had much lower activity toward carboxymethylcellulose (3.9 U/mg). The mode of action of the BG on barley β-glucan and lichenan was studied in comparison with that of Bacillus subtilis lichenase and endo-(1→4)-β-glucanases (EG I, II, and III) of Trichoderma reesei. The BG behaved very similar to the bacterial lichenase, except the tri- and tetrasaccharides formed as the end products of β-glucan hydrolysis with the BG contained the β-(1→3)-glucoside linkage at the non-reducing end, while the lichenase-derived oligosaccharides had the β-(1→3)-linkage at the reducing end. The BG was characterized by a high amino acid sequence identity to the EG of Aspergillus kawachii (UniProt entry Q12679) from a family 12 of glycoside hydrolases (96% in 162 identified aa residues out of total 223 residues) and also showed lower sequence similarity to the EglA of Aspergillus niger (O74705).  相似文献   

15.
Langerin mediates the carbohydrate-dependent uptake of pathogens by Langerhans cells in the first step of antigen presentation to the adaptive immune system. Langerin binds to an unusually diverse number of endogenous and pathogenic cell surface carbohydrates, including mannose-containing O-specific polysaccharides derived from bacterial lipopolysaccharides identified here by probing a microarray of bacterial polysaccharides. Crystal structures of the carbohydrate-recognition domain from human langerin bound to a series of oligomannose compounds, the blood group B antigen, and a fragment of β-glucan reveal binding to mannose, fucose, and glucose residues by Ca2+ coordination of vicinal hydroxyl groups with similar stereochemistry. Oligomannose compounds bind through a single mannose residue, with no other mannose residues contacting the protein directly. There is no evidence for a second Ca2+-independent binding site. Likewise, a β-glucan fragment, Glcβ1-3Glcβ1-3Glc, binds to langerin through the interaction of a single glucose residue with the Ca2+ site. The fucose moiety of the blood group B trisaccharide Galα1-3(Fucα1-2)Gal also binds to the Ca2+ site, and selective binding to this glycan compared to other fucose-containing oligosaccharides results from additional favorable interactions of the nonreducing terminal galactose, as well as of the fucose residue. Surprisingly, the equatorial 3-OH group and the axial 4-OH group of the galactose residue in 6SO4-Galβ1-4GlcNAc also coordinate Ca2+, a heretofore unobserved mode of galactose binding in a C-type carbohydrate-recognition domain bearing the Glu-Pro-Asn signature motif characteristic of mannose binding sites. Salt bridges between the sulfate group and two lysine residues appear to compensate for the nonoptimal binding of galactose at this site.  相似文献   

16.
A murine monoclonal antibody recognizing (1→6)-β-d -glucopyranosyl laminaritriose (G4) was prepared by immunizing BALB/c mice with G4-bovine serum albumin conjugate and fusing the splenocytes with mouse myeloma cells. The monoclonal antibody (IgM) provoked by the cloned cells showed low reactivity with schizophyllan, an antitumor polysaccharide, but notable reactivity with some low-molecular-weight schizophyllans. This antibody was useful for determination of the epitope of several polysaccharides. The extent of reactivity of this monoclonal antibody was related only to the molecular weight of schizophyllan.  相似文献   

17.
Expression patterns of barley β - d -glucan glucohydrolase genes were monitored using cDNAs encoding isoenzymes ExoI and ExoII. The cDNAs were isolated from 5-day-old seedling libraries. The enzymes are encoded by a small gene family, in which marked differences in codon usage are evident. The cDNAs can be used as specific probes for two subfamilies of β - d -glucan glucohydrolase genes. Genes of both subfamilies are transcribed in the scutellum of germinated grain, in elongating coleoptiles, and in young roots and leaves. Low levels of mRNA for the isoenzyme ExoI gene subfamily could be detected in aleurone layers of germinated grain. Most of the β - d -glucan glucohydrolase activity can be extracted from tissues with dilute aqueous buffers. Enzyme activity is highest in young leaves and elongating coleoptiles, but is not well-correlated with mRNA levels. The expression patterns are consistent with proposed roles for β -glucan glucohydrolases in the turnover or modification of cell-wall (1→3,1→4)- β - d -glucans in elongating coleoptiles and in young vegetative tissues.  相似文献   

18.
The similar three-dimensional structures of barley (1-->3)-beta-glucan endohydrolases and (1-->3,1-->4)-beta-glucan endohydrolases indicate that the enzymes are closely related in evolutionary terms. However, the (1-->3)-beta-glucanases hydrolyze polysaccharides of the type found in fungal cell walls and are members of the pathogenesis-related PR2 group of proteins, while the (1-->3,1-->4)-beta-glucanases function in plant cell wall metabolism. The (1-->3)-beta-glucanases have evolved to be significantly more stable than the (1-->3,1-->4)-beta-glucanases, probably as a consequence of the hostile environments imposed upon the plant by invading microorganisms. In attempts to define the molecular basis for the differences in stability, eight amino acid substitutions were introduced into a barley (1-->3,1-->4)-beta-glucanase using site-directed mutagenesis of a cDNA that encodes the enzyme. The amino acid substitutions chosen were based on structural comparisons of the barley (1-->3)- and (1-->3,1-->4)-beta-glucanases and of other higher plant (1-->3)-beta-glucanases. Three of the resulting mutant enzymes showed increased thermostability compared with the wild-type (1-->3,1-->4)-beta-glucanase. The largest increase in stability was observed when the histidine at position 300 was changed to a proline (mutant H300P), a mutation that was likely to decrease the entropy of the unfolded state of the enzyme. Furthermore, the three amino acid substitutions which increased the thermostability of barley (1-->3,1-->4)-beta-glucanase isoenzyme EII were all located in the COOH-terminal loop of the enzyme. Thus, this loop represents a particularly unstable region of the enzyme and could be involved in the initiation of unfolding of the (1-->3,1-->4)-beta-glucanase at elevated temperatures.  相似文献   

19.
Pseudomonas aeruginosa employs pili to mediate adherence to epithelial cell surfaces. The pilus adhesin of P. aeruginosa strains PAK and PAO has been shown to bind to the glycolipid asialo-GM1 (Lee et al., 1994 —accompanying article). PAK and PAO pili were examined for their abilities to bind to the synthetic βGalNAc(1–4)βGal (a minimal structural carbohydrate receptor sequence of asialo-GM1 and asialo-GM2 proposed by Krivan et al., 1988a) using solid-phase binding assays. Both pill specifically bound to βGalNAc(1–4)βGal. The binding of βGal-NAc(1–4)βGal-Biotin to the Immobilized PAK and PAO pili was inhibited by corresponding free pili. The receptor binding domain of the PAK pilus resides in the C-terminal disulphide-looped region (residues 128–144) of the pilin structural subunit (Irvin et al., 1989). Biotinylated synthetic peptides corresponding the C-terminal residues 128–144 of P. aeruginosa PAK and PAO pilin molecules were shown to bind to the βGalNAc(1–4)βGal-(bovine serum albumin (BSA)). The binding of biotinylated peptides to βGalNAc-(1–4)βGal-BSA was inhibited by PAK pili, Ac-KCTSDQDEOFIPKGCSK-OH (AcPAK(128–144)ox-OH) and Ac-ACKSTQDPMFTPKGCDN-OH (AcPAO(128–144)ox-OH) peptides. (In these peptides Ac denotes Nα -acetylation of the N-terminus, -OH means a peptide with a free a-carboxyl group at the C-terminus and the‘ox’denotes the oxidation of the sulphhydryl groups of Cys–129 and Cys–142.) Both acetylated peptides were also able to inhibit the binding of βGalNAc(1–4)βGal-biotin to the corresponding BSA-Peptide(128–144)ox-OH conjugates. The βGlcNAc(1–3)βGal(1–4)βGlc-biotin conjugate was unable to specifically bind to either Immobilized PAK and PAO pili or the respective C-termlnal peptides. The data above demonstrated that the P. aeruginosa pili recognize asialo-GM1 receptor analogue and that βGalNAc(1–4)βGal disaccharlde is sufficient for binding. Furthermore, the binding to βGalNAc(1–4)βGal was mediated by residues 128–144 of the pilin subunit.  相似文献   

20.
13C-nmr has been employed to probe the molecular conformation and crystal structure of (1 → 6)-β-D -glucan (pustulan) in the solution, gel, and solid states. CP/MAS 13C-nmr spectra recorded for partially crystalline solid pustulan display a resonance near 82 ppm that is absent in solution spectra. The intensity and peak width of this resonance were found to depend on relative crystallinity as determined by x-ray diffraction. CP/MAS spectra of aqueous pustulan gels also exhibit the 82-ppm resonance, suggesting that the gelation mechanism may involve microcrystalline junction zones. Since the 82-ppm resonance is absent in the CP/MAS spectrum of the (1 → 6)-β-linked dimer gentiobiose, we tentatively conclude the crystal structure of this dimer does not adequately model the yet undetermined structure of pustulan.  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