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.
Hamster sarcoma virus (HSV) transformation of Nil-8 fibroblasts is associated with an increase in the average size of N-acetyllactosamine (complex) type N-linked glycans due to an increase in both the average number of branches/chain and in the fraction of N-linked glycans containing poly(GlcNAc(beta 1,3) Gal-(beta 1,4)) (polylactosaminylglycan) chains. Analysis of glycopeptides from the envelope glycoproteins of Sindbis virus and vesicular stomatitis virus (VSV) grown in Nil-8 and Nil/HSV cells indicated that the transformation-associated shift to larger N-linked oligosaccharides selectively affects some glycosylation sites far more than others. Glycosylation of the Sindbis virus glycoproteins and of Asn-179 of VSV G was similar in Nil-8 and Nil/HSV cells; oligosaccharide processing generally did not proceed beyond the biantennary complex stage. In contrast, Asn-336 of VSV G carried primarily biantennary complex glycans in Nil-8-grown virus (ratio, triantennary, and larger to biantennary complex glycans (tri+/bi) = 0.5) but more highly branched structures in Nil/HSV-grown virus (tri+/bi = 8.1). All of the triantennary or larger oligosaccharides from Asn-336 of Nil/HSV-grown VSV G bound to leukoagglutinating phytohemagglutinin-agarose, indicating the presence of a branch attached to the Man3GlcNAc2 core via a beta 1,6-linked GlcNAc residue and suggesting that increased UDP-GlcNAc:alpha-D-mannoside beta 1,6-N-acetylglucosaminyl transferase V (GlcNAc transferase V) activity accompanied transformation. At least 20% of these leukoagglutinating phytohemagglutinin-binding oligosaccharides were sensitive to an enzyme specific for polylactosaminylglycan chains, Escherichia freundii endo-beta-galactosidase.  相似文献   

2.
Human lutropin or luteinizing hormone (hLH) is a heterodimeric glycoprotein, composed of two subunits. hLH alpha (N-glycosylated at Asn52 and Asn78) and hLH beta (N-glycosylated at Asn30). The sugar chains were liberated by hydrazinolysis from intact hLH beta and from glycopeptides obtained after tryptic digestion of hLH alpha, subsequently reduced and fractionated as alditols by anion-exchange and ion-suppression amine-adsorption HPLC and identified mainly by one-dimensional (1D) and two-dimensional (2D) 1H-NMR spectroscopy. The results indicate predominantly diantennary. N-acetyllactosamine-type structures at all three glycosylation sites. The oligosaccharides attached to Asn52 (hLH alpha) and Asn30 (hLH beta) show a remarkably similar pattern, with mainly chain-terminating 4-sulphated 2-deoxy-2-N-acetylamino-D-galactose (GalNAc) and a sulphated/sialylated structure as the major single component. However, virtually all N-glycans on the beta subunit bear a fucose residue alpha 1-6-linked to the proximal GlcNAc, whereas those at Asn52 (and Asn78) of the alpha subunit are predominantly non-fucosylated. The oligosaccharides at Asn78 (hLH alpha) are sialylated rather than sulphated and contain the unique sequence NeuAc alpha 2-6 GalNAc beta 1-4GlcNAc beta 1-2 Man alpha 1-3 as part of the majority of mono- and disialylated compounds. The major single constituent at Asn78 has the following structure: [formula, see text]  相似文献   

3.
Many eukaryotic proteins are modified by N-linked glycosylation, a process in which oligosaccharides are added to asparagine residues in the sequon Asn-X-Ser/Thr. However, not all such sequons are glycosylated. For example, rabies virus glycoprotein (RGP) contains three sequons, only two of which appear to be glycosylated in virions. To examine further the signals in proteins which regulate N-linked core glycosylation, the glycosylation efficiencies of each of the three sequons in the antigenic domain of RGP were compared. For these studies, mutants were generated in which one or more sequons were deleted by site-directed mutagenesis. Core glycosylation of these mutants was studied using two independent systems: 1) in vitro translation in rabbit reticulocyte lysate supplemented with dog pancreatic microsomes, and 2) transfection into glycosylation-deficient Chinese hamster ovary cells. Parallel results were obtained with both systems, demonstrating that the sequon at Asn37 is inefficiently glycosylated, the sequons at Asn247 and Asn319 are efficiently glycosylated, and the glycosylation efficiency of each sequon is not influenced by glycosylation at other sequons in this protein. High levels of cell surface expression of RGP in Chinese hamster ovary cells are seen with any mutant containing an intact sequon at Asn247 or Asn319, whereas low levels of cell surface expression are seen when the sequon at Asn37 is present alone; deletion of all three sequons completely blocks RGP cell surface expression. Thus, although core glycosylation at Asn37 is inefficient, it is still sufficient to support a biological function, cell surface expression. Future studies using mutagenesis of this model protein and its expression in these two well defined systems will aim to begin to unravel the rules governing core glycosylation of glycoproteins.  相似文献   

4.
Yeast invertase contains 14 sequons, all of which are glycosylated to varying degrees except for sequon 5 which is marginally glycosylated, if at all. This sequon overlaps with sequon 4 in a sequence consisting of Asn92-Asn93-Thr94-Ser95(Reddy et al., 1988, J. Biol. Chem., 263, 6978-6985). To determine whether glycosylation at Asn93is sterically hindered by the oligosaccharide on Asn92, the latter amino acid was converted to a glutamine residue by site-directed mutagenesis of the SUC2 gene in a plasmid vector which was expressed in Saccharomyces cerevisiae. A glycopeptide encompassing sequons 3 through 6 was purified from a tryptic digest of the mutagenized invertase and sequenced by Edman degradation, which revealed that Asn93 of sequon 5 contained very little, if any, carbohydrate, despite the elimination of sequon 4. When Ser and Thr were inverted to yield Asn-Asn-Ser-Thr carbohydrate was associated primarily with the second sequon, in agreement with numerous studies indicating that Asn-X-Thr is preferred to Asn-X-Ser as an oligosaccharide acceptor. However, when the invertase overlapping sequons were converted to Asn-Asn-Ser-Ser, both sequons were clearly glycosylated, with the latter sequon predominating. These findings rule out steric hindrance as a factor involved in preventing the glycosylation of sequon 5 in invertase. Comparable results were obtained using an in vitro system with sequon-containing tri- and tetrapeptides acceptors, in addition to larger oligosaccharide acceptors.  相似文献   

5.
Factor I (FI) is a serine protease that inhibits all complement pathways by degrading activated complement components C3b and C4b. FI functions only in the presence of several cofactors, such as factor H, C4b-binding protein, complement receptor 1, and membrane cofactor protein. FI is composed of two chains linked by a disulfide bridge; the light chain comprises only the serine protease (SP) domain, whereas the heavy chain contains the FI membrane attack complex domain (FIMAC), CD5 domain, and low density lipoprotein receptor 1 (LDLr1) and LDLr2 domains. To better understand how FI inhibits complement, we used homology-based three-dimensional models of FI domains in an attempt to identify potential protein-protein interaction sites. Specific amino acids were then mutated to yield 20 recombinant mutants of FI carrying additional surface-exposed N-glycosylation sites that were expected to sterically hinder interactions. The Michaelis constant (K(m)) of all FI mutants toward a small substrate was not increased. We found that many mutations in the FIMAC and SP domains nearly abolished the ability of FI to degrade C4b and C3b in the fluid phase and on the surface, irrespective of the cofactor used. On the other hand, only a few alterations in the CD5 and LDLr1/2 domains impaired this activity. In conclusion, all analyzed cofactors form similar trimolecular complexes with FI and C3b/C4b, and the accessibility of FIMAC and SP domains is crucial for the function of FI.  相似文献   

6.
The Sindbis virus glycoproteins, E1 and E2, comprise a useful model system for evaluating the effects of local protein structure on the processing of N-linked oligosaccharides by Golgi enzymes. The conversion of oligomannose to N-acetyllactosamine (complex) oligosaccharides is hindered to different extents at the four glycosylation sites, so that the complex/oligomannose ratio decreases in the order E1-Asn139 greater than E2-Asn196 greater than E1-Asn245 greater than E2-Asn318. The processing steps most susceptible to interference were deduced from the oligosaccharide compositions at hindered sites in virus from baby hamster kidney cells (BHK), chick embryo fibroblasts (CEF), and normal and hamster sarcoma virus (HSV)-transformed hamster fibroblasts (Nil-8). Persistence of Man6-9GlcNAc2 was taken to indicate interference with alpha 2-mannosidase(s) I (alpha-mannosidase I), Man5GlcNAc2, with UDP-GlcNAc:alpha-D-mannoside beta 1----2-N-acetylglucosaminyltransferase I (GlcNAc transferase I), and unbisected hybrid glycans, with GlcNAc transferase I-dependent alpha 3(alpha 6)-mannosidase (alpha-mannosidase II). Taken together, the results indicate that all four sites acquire a precursor oligosaccharide with equally high efficiency, but alpha-mannosidase I, GlcNAc transferase I, and alpha-mannosidase II are all impeded at E2-Asn318 and, to a lesser extent, at E1-Asn245. In contrast, sialic acid and galactose transfer to hybrid glycans (in BHK cells) is virtually quantitative even at E2-Asn318. E2-Asn318 carried no complex oligosaccharides, but the structures of those at E1-Asn245 indicate almost complete GlcNAc transfer by UDP-GlcNAc:alpha-D-mannoside beta 1----2-N-acetylglucosaminyltransferase II (GlcNAc transferase II), galactosylation, and sialylation. Because the E2-Asn318 and E1-Asn245 glycans have previously been shown to be less accessible to a steric probe than those at E2-Asn196 or E1-Asn139, a simple explanation for these results would be that alpha-mannosidase I, GlcNAc transferase I, and alpha-mannosidase II are more susceptible to steric hindrance than are the later processing steps examined. Finally, in addition to these site-specific effects, the overall extent of viral oligosaccharide processing varied with host and cellular growth status. For example, alpha-mannosidase I processing is more complete in BHK cells compared to CEF, and in confluent Nil-8 cells compared to subconfluent or HSV-transformed Nil-8 cells.  相似文献   

7.
Endo-beta-N-acetylglucosaminidase from Arthrobacter protophormiae was tested for its capacity to release N-linked sugar chains from native yeast invertase. The enzyme liberated about 80% of the sugar chains from the native invertase. Deglycosylated invertase was digested by chymotrypsin or pepsin, and twelve N-acetylglucosamine-containing glycopeptides were isolated. The amino acid sequences of these glycopeptides were analyzed by a protein sequencer, and the elution position of 4-L-aspartylglycosylamine was directly identified by conventional sequencing. The endo-beta-N-acetylglucosaminidase was found to remove mainly nine sugar chains from native invertase.  相似文献   

8.
Concanavalin A (ConA) kills the procyclic (insect) form of Trypanosoma brucei by binding to its major surface glycoprotein, procyclin. We previously isolated a mutant cell line, ConA 1-1, that is less agglutinated and more resistant to ConA killing than are wild-type (WT) cells. Subsequently we found that the ConA resistance phenotype in this mutant is due to the fact that the procyclin either has no N-glycan or has an N-glycan with an altered structure. Here we demonstrate that the alteration in procyclin N-glycosylation correlates with two defects in the N-linked oligosaccharide biosynthetic pathway. First, ConA 1-1 has a defect in activity of polyprenol reductase, an enzyme involved in synthesis of dolichol. Metabolic incorporation of [3H]mevalonate showed that ConA 1-1 synthesizes equal amounts of dolichol and polyprenol, whereas WT cells make predominantly dolichol. Second, we found that ConA 1-1 synthesizes and accumulates an oligosaccharide lipid (OSL) precursor that is smaller in size than that from WT cells. The glycan of OSL in WT cells is apparently Man9GlcNAc2, whereas that from ConA 1-1 is Man7GlcNAc2. The smaller OSL glycan in the ConA 1-1 explains how some procyclin polypeptides bear a Man4GlcNAc2 modified with a terminal N-acetyllactosamine group, which is poorly recognized by ConA.  相似文献   

9.
Effect of glycosylation on yeast invertase oligomer stability   总被引:5,自引:0,他引:5  
Yeast external invertase is a glycoprotein that exists as a dimer that can associate to form tetramers, hexamers, and octamers (Chu, F., Watorek, W., and Maley, F. (1983) Arch. Biochem. Biophys. 223, 543-555; Esmon, P. C., Esmon, B. E., Schauer, I. E., Taylor, A., and Schekman, R. (1987) J. Biol. Chem., 262, 4395-4401), a process that is facilitated by the attached oligosaccharide chains. We have studied this association by high performance liquid chromatography on a gel filtration matrix, by which procedure wild-type bakers' yeast invertase gives two peaks, and invertase from a core mutant (mnn1 mnn9) of Saccharomyces cerevisiae X2180 gives three peaks. Concentration of an invertase solution by freezing drives the dimers into higher aggregates that, at 30 degrees C, re-equilibrate to a mixture of smaller forms, the composition of which depends on pH, concentration, and time. The invertase from a mutant, mnn1 mnn9 dpg1, which underglycosylates its glycoproteins and produces invertase with 4-7 oligosaccharide chains, forms oligomers of much lower stability than the mnn1 mnn9 invertase, which has 8-11 carbohydrate chains. Both of these mutants release external invertase from the periplasm into the medium during growth, but we conclude that defects in the cell wall structure may be more important in this release than an altered tendency of the invertases to aggregate. Investigation of aggregate formation by electron microscopy revealed that all invertases, including the internal nonglycosylated enzyme, form octamers under appropriate conditions.  相似文献   

10.
Nonglycosylated murine and human granulocyte-macrophage colony-stimulating factor have a molecular mass of approximately 14.5 kDa predicted from the primary amino acid sequence. The expression of both proteins in COS cells leads to a heterogeneous population of molecules that differ in the degree of glycosylation. Both human and murine molecules contain two N-linked glycosylation sites that are situated in nonhomologous locations along the linear sequence. Despite this difference both proteins show a similar size distribution among the glycosylation variants. These studies analyze the effects of introducing in the murine protein novel N-linked glycosylation sites corresponding to those sites found in the human molecule. A panel of molecules composed of various combinations of human N-linked glycosylation sites in either the presence or the absence of murine N-linked glycosylation was compared. Substitution of a proper human N-linked glycosylation consensus sequence at Asn 24 did not result in N-linked glycosylation, nor was there any considerable effect on bioactivity. Replacement of the N-linked glycosylation consensus sequence at Asn 34 results in glycosylation similar to that found in the human molecule and causes a significant decrease in bioactivity. These data suggest that the position of N-linked glycosylation is critical for maximal bioactivity in a particular species and that the changes in position of these sites in different species probably occurred during evolution in response to changes in their receptors.  相似文献   

11.
Human alpha1,3 fucosyltransferases (FucTs) contain four highly conserved cysteine (Cys) residues, in addition to a free Cys residue that lies near the binding site for GDP-fucose (Holmes, E. H., Xu, Z. , Sherwood, A. L., and Macher, B. A. (1995) J. Biol. Chem. 270, 8145-8151). The participation of the highly conserved Cys residues in disulfide bonds and their functional significance were characterized by mass spectrometry (MS) analyses and site-directed mutagenesis, respectively. Among the human FucTs is a subset of enzymes (FucT III, V, and VI) having highly homologous sequences, especially in the catalytic domain, and Cys residues in FucT III and V were characterized. The amino acid sequence of FucT III was characterized. Peptides containing the four conserved Cys residues were detected after reduction and alkylation, and found to be involved in disulfide bonds. The disulfide bond pattern was characterized by multiple stage MS analysis and the use of Glu-C protease and MS/MS analysis. Disulfide bonds in FucT III occur between Cys residues (Cys(81) to Cys(338) and Cys(91) to Cys(341)) at the N and C termini of the catalytic domain, bringing these ends close together in space. Mutagenesis of highly conserved Cys residues to Ser in FucT V resulted in proteins lacking enzymatic activity. Three of the four mutants have molecular weights similar to wild type enzyme and maintained an ability to bind GDP, whereas the other (Cys(104)) produced a series of lower molecular weight bands when characterized by Western blot analysis, and did not bind GDP. FucTs have highly conserved, potential N-linked sites, and our mass spectrometry analyses demonstrated that both N-linked sites are modified with oligosaccharides.  相似文献   

12.
The addition of asparagine (N)-linked polysaccharide chains (i.e., glycans) to the gp120 and gp41 glycoproteins of human immunodeficiency virus type 1 (HIV-1) envelope is not only required for correct protein folding, but also may provide protection against neutralizing antibodies as a “glycan shield.” As a result, strong host-specific selection is frequently associated with codon positions where nonsynonymous substitutions can create or disrupt potential N-linked glycosylation sites (PNGSs). Moreover, empirical data suggest that the individual contribution of PNGSs to the neutralization sensitivity or infectivity of HIV-1 may be critically dependent on the presence or absence of other PNGSs in the envelope sequence. Here we evaluate how glycan–glycan interactions have shaped the evolution of HIV-1 envelope sequences by analyzing the distribution of PNGSs in a large-sequence alignment. Using a “covarion”-type phylogenetic model, we find that the rates at which individual PNGSs are gained or lost vary significantly over time, suggesting that the selective advantage of having a PNGS may depend on the presence or absence of other PNGSs in the sequence. Consequently, we identify specific interactions between PNGSs in the alignment using a new paired-character phylogenetic model of evolution, and a Bayesian graphical model. Despite the fundamental differences between these two methods, several interactions are jointly identified by both. Mapping these interactions onto a structural model of HIV-1 gp120 reveals that negative (exclusive) interactions occur significantly more often between colocalized glycans, while positive (inclusive) interactions are restricted to more distant glycans. Our results imply that the adaptive repertoire of alternative configurations in the HIV-1 glycan shield is limited by functional interactions between the N-linked glycans. This represents a potential vulnerability of rapidly evolving HIV-1 populations that may provide useful glycan-based targets for neutralizing antibodies.  相似文献   

13.
The asparagine-linked glycosylation sites in rabbit high and low Mr secretory components (SC) have been determined for the three known allotypes, t61, t62, and t63. Purified SC polypeptides were subjected to mild periodate oxidation of terminal nonreducing sugars followed by selective reduction with [3H]sodium borohydride, SC polypeptides were further proteolytically cleaved, and the 3H-labeled peptides were isolated and characterized. Both high and low Mr SCs of the three allotypes possess a common glycosylation site at the asparagine residue position 400, whereas the second site, in the amino-terminal domain of SC, was found to be variable: the t61 and t63 allotypes contained a glycosylation site at positions 70 and 90, respectively. Moreover, although the t62 allotype was found to contain a triplet acceptor site (N-X-S) at positions 90-92, analyses showed that less than 30% of the t62 allotype peptides encompassing this region were glycosylated at residue 90. Furthermore, the amino acid sequence of the t61 SC in the region of residues 69-90 varies by 8 and 10 amino acid substitutions when compared with the t62 and t63 allotype sequences, respectively. However, neither the variation in amino acid sequence nor the variation in degree or site of glycosylation measurably affected the non-covalent binding of domain 1 to dimeric IgA.  相似文献   

14.
Highly purified, trypsin/detergent-solubilized thyroid peroxidase (TPO), prepared from pig thyroid tissue, was subjected to reduction and alkylation followed by trypsin digestion. The resulting peptides were fractionated using HPLC. Corresponding carbohydrate positive regions from three separate HPLC experiments were pooled and further chromatography was carried out to yield purified peptide suitable for sequence analysis and complete carbohydrate composition analysis. Four of the five putative sites for N-linked glycosylation were found to carry oligosaccharide units in which mannose and glucosamine were the sole or predominant sugars. Three of the four glycosylations occur at asparagine residues which are likely to be at beta turns or bends. The fifth putative glycosylation site could not be confirmed and may either be poorly glycosylated or escape glycosylation. All of the confirmed glycosylated sites occur in the N-terminal third of the TPO polypeptide chain, in the portion of the molecule believed to be extracellular. The isolation of at least two chromatographic forms of glycopeptide derived from each of the confirmed sites suggests microheterogeneity in the structure of the oligosaccharide units of thyroid peroxidase similar to that observed in many other glycoproteins.  相似文献   

15.
16.
Factors affecting the oligomeric structure of yeast external invertase   总被引:4,自引:0,他引:4  
It has been assumed that yeast external invertase is a dimer, with each subunit composed of a 60-kDa polypeptide chain. We now present evidence that at its optimal pH of 5.0, the predominant form of external invertase is an octamer with an average size of 8 X 10(5) Da. During ultracentrifugation the octamer dissociated to lower molecular weight forms, including a hexamer, tetramer, and dimer. All forms of the enzyme were shown to possess identical specific activities and to contain a similar carbohydrate to protein ratio. Although the monomer subunits (1 X 10(5) Da) were heterogenous in carbohydrate content, each subunit possessed nine oligosaccharide chains. When stained for protein and enzyme activity following sodium dodecyl sulfate-polyacrylamide gel electrophoresis, only the oligomeric form of the enzyme appeared to be active. Thus, on partially inactivating invertase with 4 M guanidine hydrochloride both octamer and monomer were evident on the gels but only the former was active. Similarly, incubating at pH 2.5 in the presence of sodium dodecyl sulfate yielded only inactive monomer. The monomer, unlike the active oligomeric aggregate, was unable to hydrolyze sucrose after sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Consistent with the in vitro studies, freshly prepared yeast lysate was shown to contain the octameric species of external invertase as the major active form of this enzyme. From these studies and others which employed deglycosylated invertase, it is concluded that the carbohydrate component of external invertase contributes not only to stabilizing enzyme activity, but also to maintaining its oligomeric structure.  相似文献   

17.
P S Appukuttan  H C Wu 《FEBS letters》1989,255(1):139-142
The relationship between protein glycosylation and fatty acylation of glycoproteins was studied in the wild-type and asparagine-linked glycosylation-deficient mutants (alg1 and alg2) of Saccharomyces cerevisiae. At the non-permissive temperature (37 degrees C), both mutant cells exhibited increased incorporation of [3H]palmitate into five polypeptides based on SDS-PAGE. In contrast, the wild-type yeast cells contained [3H]palmitate-labeled polypeptides of higher molecular weights, which were converted to the bands seen in the mutant cells upon treatment of the cell extract with endoglycosidase H prior to SDS-PAGE. In addition, labeling of the wild-type yeast cells with [3H]palmitate in the presence of tunicamycin revealed the incorporation of [3H]palmitate into the same five bands as found in the alg1 and alg2 mutants at the non-permissive temperature without tunicamycin. These results indicate that fatty acylation of glycoproteins proceeds independently of protein N-glycosylation in yeast cells.  相似文献   

18.
Endoplasmic reticulum (ER) stress responses comprising the unfolded protein response (UPR) are activated by conditions that disrupt folding and assembly of proteins inside the ER lumenal compartment. Conditions known to be proximal triggers of the UPR include saturation of chaperones with misfolded protein, redox imbalance, disruption of Ca2+ levels, interference with N-linked glycosylation, and failure to dispose of terminally misfolded proteins. Potentially, ER stress responses can reprogram cells to correct all of these problems and thereby restore ER function to normal. This article will review literature on stimulation of N-linked glycosylation by ER stress responses, focusing on metazoan systems. The mechanisms involved will be contrasted with those mediating stimulation of N-linked glycosylation by cytoplasmic stress responses. This information will interest readers who study the biological roles of stress responses, the functions of N-linked glycans, and potential strategies for treatment of genetic disorders of N-linked glycosylation.  相似文献   

19.
20.
N-Glycosylation occurs cotranslationally as soon as the growing polypeptide chain enters the endoplasmic reticulum, before the final native-like folded state is reached. We examined the role of the carbohydrate chains in the mechanism of protein folding. The in vitro folding and association of yeast invertase are used as an experimental system. External invertase contains approximately 50% carbohydrate, whereas cytoplasmic invertase is not glycosylated. The functional native state of both proteins is a homodimer. At pH greater than or equal to 6.5 and at protein concentrations below 3 micrograms/ml, the kinetics of reactivation and the final yields are similar for the two invertases. For both proteins, reactivation is a sequential reaction with a lag phase at the beginning. The nonglycosylated protein tends to aggregate during reactivation at low pH and at protein concentrations above 3 micrograms/ml. After separation of inactive material, the renatured protein is indistinguishable from the original native state by a number of physicochemical and functional criteria. The results suggest that the carbohydrate moiety does not affect the mechanism of folding and association of invertase. However, glycosylation improves the solubility of unfolded or partially folded invertase molecules and hence leads to a suppression of irreversible aggregation. Such a protective effect may also be important for the in vivo maturation of nascent glycosylated protein chains.  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