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.
1. Insoluble complexes, formed by electrostatic interaction between chondromucoprotein and chromoproteins (haemoglobin, methaemalbumin), were studied by measurement of precipitated pigment and by decrease in peroxidatic activity, maximum formation from aqueous solution occurring at pH 4·0–4·6. 2. Chondromucoprotein did not form complexes with plasma haptoglobins and haptohaemoglobins under these conditions, and high concentrations had no significant effect on colorimetric estimates of serum haptoglobin, although the peroxidatic activity of haemoglobinaemic serum was depressed owing to formation of chondromucoprotein–methaemalbumin complex. 3. The complexes formed by interaction between chondromucoprotein and plasma proteins contain two protein-bound biologically active components (plasminogen, haematin), as a result of co-precipitation after interaction between their carriers and chondromucoprotein. The possible presence of other biologically active trace components is discussed. 4. The results are related to complex-formation between other plasma proteins and chondromucoprotein, and possible implications arising from the complex-forming properties of tissue and urine chondromucoprotein are referred to. It is concluded that the inability of chondromucoprotein to form complexes with normal urine proteins is due to a deficiency of fibrinogen, β-lipoproteins and chromoproteins, which, in plasma, form a large proportion of the proteins involved in complex-formation.  相似文献   

2.
Hyaluronate appears to be covalently linked to the cell surface   总被引:1,自引:0,他引:1  
The purpose of this study was to examine the nature of the linkage between cell-surface hyaluronate and the plasma membrane. To accomplish this, rat fibrosarcoma cells were cultured in the presence of [3H]-acetate to isotopically label the hyaluronate, and then fixed with glutaraldehyde, which cross-links proteins but does not react directly with hyaluronate. The glutaraldehyde fixation stabilized the cells so that they could be manipulated in ways which would otherwise destroy cells. The fixed cells were then subjected to various treatments, and the amount of hyaluronate remaining on the cell surface was assayed via exhaustive digestion with Streptomyces hyaluronidase. Using this technique, we found that 1) cell-surface hyaluronate was quite stable for extended periods of time even in the presence of a large excess of non-labeled hyaluronate; 2) 4 M guanidine HCl and detergents did not extract a significant portion of cell-surface hyaluronate; 3) solutions of varying ionic strength (0-1 M NaCl) had no effect on the retention of hyaluronate; 4) the cell coat was stable in the range of pH 4-11, but outside this range a significant amount of hyaluronate was released; and 5) treatment with proteases released cell-surface hyaluronate. These results are consistent with the possibility that hyaluronate is covalently linked to a protein associated with the plasma membrane. Further support for this model came from experiments with the detergent Triton X-114, which can be used to separate soluble proteins from hydrophobic proteins. When nonfixed rat fibrosarcoma cells were extracted with this detergent and then partitioned by centrifugation, approximately 30 times as much hyaluronate was present in the detergent fraction which contained the hydrophobic proteins, as compared to the extracts pretreated with trypsin prior to phase separation. Again, these results suggest that cell-surface hyaluronate is directly linked to a hydrophobic core protein intercalated in the plasma membrane.  相似文献   

3.
Hyaluronate-mediated expansion of the extracellular matrix has been suggested as an important element of growth and morphogenesis in several developing systems. In vitro, various growth factors have been shown to stimulate hyaluronate synthesis as well as cell proliferation. A similar link between proliferation and hyaluronate production during in vivo growth is difficult to demonstrate, because in most systems the source of growth-promoting factors is either not known or not amenable to experimental manipulation. During amphibian limb regeneration, cell proliferation depends upon paracrine release of factors from axons in the limb stump, and the nerve supply can be eliminated or augmented experimentally for study of growth in this system. Denervated and amputated limbs of larval salamanders do not begin to regenerate until distal areas of the limb stumps are reinnervated. We have used such limbs to examine the effect exerted by the reappearance of nerves on the amount of hyaluronate in the tissue undergoing the growth response. Hyaluronate was demonstrated by the metachromatic dye Ethyl Stains-all, which stains hyaluronate blue while sulfated glycosaminoglycans (GAGs) and proteins in the extracellular matrix stain various shades of violet, and by microspectrophotometry of alcian-blue-stained GAGs in serial sections pretreated with buffer or with Streptomyces hyaluronidase (SH) to remove hyaluronate specifically. Both methods showed little hyaluronate in the distal region of limb stumps prior to reinnervation, while reinnervated stumps had amounts of hyaluronate similar to those of control blastemas. Autoradiography of 3H-glucosamine-labeled limbs indicated that hyaluronate in the blastemas of reinnervated limb stumps included material newly synthesized by cells throughout the growing tissue. The microspectrophotometric study revealed that the relative concentration of hyaluronate increased during the time distal limb areas were undergoing reinnervation, which was monitored by staining of nerve fibers. The increase in hyaluronate concentration was followed immediately by an increase in mitotic activity and a decrease in mesenchymal cell density, two changes leading to blastema formation that others have shown to be associated with reinnervation in this system. These observations indicate that the growth-promoting influence of nerves includes stimulation of hyaluronate production, an effect similar to that of serum or purified mitogens on many cultured cells. Hyaluronate synthesis appears to promote expansion of the limb stump, which occurs when denervated-amputated larval limbs are reinnervated.  相似文献   

4.
Changes in CD of hyaluronates and chondroitins upon periodate oxidation   总被引:1,自引:0,他引:1  
N Ueno  B Chakrabarti 《Biopolymers》1987,26(8):1413-1420
Changes in conformation of hyaluronate and chondroitin sulfates following periodate oxidation were studied by CD. We monitored the progressive oxidation of these polymers by periodic acid at 4°C in pH 5.6 buffer. The negative CD band of hyaluronate at 208 nm decreased in intensity upon oxidation and changed its sign after 16 h of oxidation. In contrast, the 208-nm CD band of chondroitin sulfates decreased, but showed no change in sign even after 48 h of oxidation. A specific difference in solution conformation between hyaluronate and chondroitins may be responsible for the difference in oxidation-induced dichroic behavior. The results are discussed in terms of available x-ray diffraction analyses of these polymers.  相似文献   

5.
Proteoglycans were prepared from human femoral-head articular cartilage by using either guanidinium hydrochloride or MgCl2 as extractant, followed by density-gradient centrifugation. The proteoglycan subunit had a particle weight of 2.6 x 10(6), with a radius of gyration, RG, of 68.5 nm in 150 mM-NaCl/20 mM-sodium phosphate buffer, pH 7.0. The proteoglycan aggregate had a particle weight of 3.7 x 10(6) (RG 84 nm) for guanidinium hydrochloride extracts and 8.7 x 10(6) (RG 118 nm) for MgCl2 extracts in the same buffer. The addition of excess of high-molecular-weight hyaluronate did not significantly alter the particle size of the aggregate. The small increase in size probably reflects a rapid equilibrium between hyaluronate and proteoglycan monomers, and is not due to proteolytic cleavage producing non-aggregating units. Experiments that support the rapid-interaction hypothesis include analytical ultracentrifugation and column chromatography. This interaction does not appear to be pressure-sensitive at 20 degrees C, but is sensitive to temperature variation near the physiological range.  相似文献   

6.
Hyaluronate binding proteins also bind to fibronectin, laminin and collagen   总被引:4,自引:0,他引:4  
Small molecular weight proteins, isolated from the culture medium of embryonic chick heart fibroblasts and 3T3 cell lines by hyaluronate affinity chromatography, bind in order of apparent affinity, to hyaluronate, fibronectin, collagen and laminin. Such proteins isolated from the MSV-transformed 3T3 cell line bind in greater amounts to the nectins and hyaluronate than do similar proteins isolated from heart fibroblasts or 3T3 cells. These small hyaluronate binding proteins are immunologically distinct from other well characterized proteins such as laminin, fibronectin, bovine serum albumin and actin. Their relationship to other small, extracellular proteins and their possible role in structuring of extracellular matrix are discussed.  相似文献   

7.
8.
Ferricytochrome b5 was found to convert oxyhaemoglobin into methaemoglobin under conditions previously found to be optimal for complex-formation between ferricytochrome b5 and methaemoglobin [Mauk & Mauk (1982) Biochemistry 21, 4730-4734]. As this reaction is completely inhibited by CO, it is proposed that oxyhaemoglobin is oxidized after O2 dissociation, as has been suggested for the oxidation of oxyhaemoglobin by inorganic complexes. From the present analysis, ferricytochrome b5 seems unlikely to contribute significantly to methaemoglobin formation in vivo. Nevertheless, this observation provides a relatively convenient means of investigating the mechanism by which these two proteins interact.  相似文献   

9.
A new method was developed to purify the streptococcal hyaluronate synthase in active form to electrophoretic homogeneity. The method is based on the extraction of protoplast membranes with digitonin and a phase separation into an aqueous and a detergent phase induced by addition of polyethylene glycol 6000 at 0°C. Proteins bound to hyaluronate were enriched in the aqueous phase, whereas other membrane proteins resided in the detergent phase. Final purification of the hyaluronate synthase was achieved by ion exchange chromatography.  相似文献   

10.
Hyaluronate is produced in high amount during the initial stages of regeneration of the tail and limbs of lizards, newts, and frog tadpoles. The fine distribution of hyaluronate in the regenerating tail blastemas has been assessed by ultrastructural immunolocalization of the Hyaluronate Binding Protein (HABP), a protein that indirectly reveals the presence of hyaluronate in tissues. The present electron microscopic study shows that HABP is detected in the cytoplasm but this proteins is mainly localized on the surfaces of cells in the wound epidermis and mesenchymal cells of the blastema. HABP appears, therefore, accumulated along the cell surface, indicating that hyaluronate coats these embryonic‐like cells and their antigens. The high level of hyaluronate in the blastema, aside favoring tissue hydration, cell movements, and remodeling for blastema formation and growth, likely elicits a protection from the possible immune‐reaction of lymphocytes and macrophages to embryonic‐fetal‐like antigens present on the surface of blastema and epidermal cells. Their survival, therefore, allows the continuous multiplication of these cells in regions rich in hyaluronate, promoting the regeneration of a new tail or limbs. The study suggests that organ regeneration in vertebrates is only possible in the presence of high hyaluronate content and hydration. These two conditions facilitate cell movement, immune‐protection, and activate the Wnt signaling pathway, like during development.  相似文献   

11.
E A Turley  D Moore  L J Hayden 《Biochemistry》1987,26(11):2997-3005
A hyaluronic acid binding fraction was purified from the supernatant media of both 3T3 and murine sarcoma virus (MSV) transformed 3T3 cultures by hyaluronate and immunoaffinity chromatography. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis resolved the hyaluronate affinity-purified fraction into three major protein bands of estimated molecular weight (Mr,e) 70K, 66K, and 56K which contained hyaluronate binding activity and which were termed hyaluronate binding proteins (HABP). Hyaluronate affinity chromatography combined with immunoaffinity chromatography, using antibody directed against the larger HABP, allowed a 20-fold purification of HABP. Fractions isolated from 3T3 supernatant medium also contained additional binding molecules in the molecular weight range of 20K. This material was present in vanishingly small amounts and was not detected with a silver stain or with [35S]methionine label. The three protein species isolated by hyaluronate affinity chromatography (Mr,e 70K, 66K, and 56K) were related to one another since they shared antigenic determinants and exhibited similar pI values. In isocratic conditions, HABP occurred as aggregates of up to 580 kilodaltons. Their glycoprotein nature was indicated by their incorporation of 3H-sugars. Enzyme-linked immunoadsorbent assay showed they were antigenically distinct from other hyaluronate binding proteins such as fibronectin, cartilage link protein, and the hyaluronate binding region of chondroitin sulfate proteoglycan. The apparent dissociation constant of HABP for hyaluronate was approximately 10(-8) M, and kinetic analyses showed these binding interactions were complex and of a positive cooperative nature.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
Aggregates formed by the interaction of cartilage proteoglycan monomers and fragments thereof with hyaluronate were studied by electron microscopy by use of rotary shadowing [Wiedemann, Paulsson, Timpl, Engel & Heinegård (1984) Biochem. J. 224, 331-333]. The differences in shape and packing of the proteins bound along the hyaluronate strand in aggregates formed in the presence and in the absence of link protein were examined in detail. The high resolution of the method allowed examination of the involvement in hyaluronate binding of the globular core-protein domains G1, G2 and G3 [Wiedemann, Paulsson, Timpl, Engel & Heinegård (1984) Biochem. J. 224, 331-333; Paulsson, Mörgelin, Wiedemann, Beardmore-Gray, Dunham, Hardingham, Heinegård, Timpl & Engel (1987) Biochem. J. 245, 763-772]. Fragments comprising the globular hyaluronate-binding region G1 form complexes with hyaluronate with an appearance of necklace-like structures, statistically interspaced by free hyaluronate strands. The closest centre-to-centre distance found between adjacent G1 domains was 12 nm. Another fragment comprising the binding region G1 and the adjacent second globular domain G2 attaches to hyaluronate only by one globule. Also, the core protein obtained by chondroitinase digestion of proteoglycan monomer binds only by domain G1, with domain G3 furthest removed from the hyaluronate. Globule G1 shows a statistical distribution along the hyaluronate strands. In contrast, when link protein is added, binding is no longer random, but instead uninterrupted densely packed aggregates are formed.  相似文献   

13.
TSG-6 cDNA was isolated by differential screening of a lambda cDNA library prepared from tumor necrosis factor (TNF)-treated human diploid FS-4 fibroblasts. We show that TSG-6 mRNA was not detectable in untreated cells, but became readily induced by TNF in normal human fibroblast lines and in peripheral blood mononuclear cells. In contrast, TSG-6 mRNA was undetectable in either control or TNF-treated human vascular endothelial cells and a variety of tumor-derived or virus-transformed cell lines. The sequence of full-length TSG-6 cDNA revealed one major open reading frame predicting a polypeptide of 277 amino acids, including a typical cleavable signal peptide. The NH2-terminal half of the predicted TSG-6 protein sequence shows a significant homology with a region implicated in hyaluronate binding, present in cartilage link protein, proteoglycan core proteins, and the adhesion receptor CD44. The most extensive sequence homology exists between the predicted TSG-6 protein and CD44. Western blot analysis with an antiserum raised against a TSG-6 fusion protein detected a 39-kD glycoprotein in the supernatants of TNF-treated FS-4 cells and of cells transfected with TSG-6 cDNA. Binding of the TSG-6 protein to hyaluronate was demonstrated by coprecipitation. Our data indicate that the inflammatory cytokine (TNF or IL-1)-inducible, secretory TSG-6 protein is a novel member of the family of hyaluronate binding proteins, possibly involved in cell-cell and cell-matrix interactions during inflammation and tumorigenesis.  相似文献   

14.
Chondrocytes produce large pericellular coats in vitro that can be visualized by the exclusion of particles, e.g., fixed erythrocytes, and that are removed by treatment with Streptomyces hyaluronidase, which is specific for hyaluronate. In this study, we examined the kinetics of formation of these coats and the relationship of hyaluronate and proteoglycan to coat structure. Chondrocytes were isolated from chick tibia cartilage by collagenase-trypsin digestion and were characterized by their morphology and by their synthesis of both type II collagen and high molecular weight proteoglycans. The degree of spreading of the chondrocytes and the size of the coats were quantitated at various times subsequent to seeding by tracing phase-contrast photomicrographs of the cultures. After seeding, the chondrocytes attached themselves to the tissue culture dish and exhibited coats within 4 h. The coats reached a maximum size after 3-4 d and subsequently decreased over the next 2-3 d. Subcultured chondrocytes produced a large coat only if passaged before 4 d. Both primary and first passage cells, with or without coats, produced type II collagen but not type I collagen as determined by enzyme-linked immunosorbent assay. Treatment with Streptomyces hyaluronidase (1.0 mU/ml, 15 min), which completely removed the coat, released 58% of the chondroitin sulfate but only 9% of the proteins associated with the cell surface. The proteins released by hyaluronidase were not digestible by bacterial collagenase. Monensin and cycloheximide (0.01-10 microM, 48 h) caused a dose-dependent decrease in coat size that was linearly correlated to synthesis of cell surface hyaluronate (r = 0.98) but not chondroitin sulfate (r = 0.2). We conclude that the coat surrounding chondrocytes is dependent on hyaluronate for its structure and that hyaluronate retains a large proportion of the proteoglycan in the coat.  相似文献   

15.
In the present study, we have examined the distribution of the hyaluronate receptor as well as hyaluronate itself in a variety of adult tissues. The hyaluronate receptor was localized with a monoclonal antibody, termed K-3, while hyaluronate was localized using proteolytic fragments of cartilage proteoglycan. Staining with the K-3 monoclonal antibody revealed that the hyaluronate receptor was present in a variety of epithelia including the skin, cheek, tongue, esophagus, vagina, intestines, oviduct, and bladder. However, it was notably absent from epithelial cells of the cornea and stomach as well as from endothelial cells of blood vessels. When present, the hyaluronate receptor was preferentially located in regions of active cell growth, such as in the basal layers of stratified epithelium and at the base of the crypts of Lieberkuhn in intestinal epithelium. A similar phenomenon was observed in cultured 3T3 cells. Cultures of 3T3 cells that were actively proliferating were found to have greater amounts of the receptor than their nonproliferating counterparts. When the various tissues were examined for hyaluronate, it was found to have a widespread distribution, being present in most of the basement membranes and between the cells in stratified epithelium. Indeed, in many cases, the distribution of hyaluronate closely paralleled that of the hyaluronate receptor. These results suggest that the interaction between hyaluronate and its receptor is involved in cell-to-substratum adhesion.  相似文献   

16.
When link protein binds to hyaluronate in the absence of proteoglycan monomer a high molecular weight complex is formed. Two assay procedures have been developed to examine the formation of the complex and the rate and stoichiometry of binding of link protein to hyaluronate in the complex. In the first, the complex is isolated by differential centrifugation, and the stoichiometry of binding of link protein to hyaluronate in the sedimented complex is determined. In the second assay, which involves turbidimetry, the rate of complex formation (delta A420/min) is determined, and the amount of complex formed is determined in terms of the maximum turbidity (A420,max) attained. The effects of temperature, pH, initial total solute concentration, and the ratio by weight of link protein to hyaluronate on the amount of complex formed and on the rate of complex formation were examined. There is a linear correlation between the amount of complex formed as determined by turbidity and by differential centrifugation. Using these assays, we examined the specificity of the binding of link protein to hyaluronate and the capacity of hyaluronate oligosaccharides to competitively inhibit the binding of link protein to hyaluronate. Hyaluronate decasaccharide is the oligosaccharide of minimum size that strongly inhibits the binding of link protein to hyaluronate. Proteoglycan monomers dissociate from hyaluronate as the pH is decreased from pH 7 to pH 5. Turbidimetric studies show that the rate of binding of link protein to hyaluronate increases with decreasing pH. The binding affinity of proteoglycan monomers for hyaluronate is decreased at pH 5, whereas the binding affinity of link protein for hyaluronate is not. This difference in the effect of pH on the stability of binding of link protein to hyaluronate, compared with proteoglycan monomer, explains in part the capacity of link protein to stabilize the binding of proteoglycan monomer to hyaluronate at pH 5.  相似文献   

17.
A relatively quick and simple assay for hyaluronate was developed using the specific binding protein, hyaluronectin. The hyaluronection was obtained by homogenizing the brains of Sprague-Dawley rats, and then centrifuging the homogenate. The resulting supernatant was used as a source of crude hyaluronectin. In the binding assay, the hyaluronectin was mixed with [3H]hyaluronate, followed by an equal volume of saturated (NH4)2SO4, which precipitated the hyaluronectin and any [3H]hyaluronate associated with it, but left free [3H]hyaluronate in solution. The mixture was then centrifuged, and the amount of bound [3H]hyaluronate in the precipitate was determined. Using this assay, we found that hyaluronectin specifically bound hyaluronate, since other glycosaminoglycans failed to compete for the binding protein. In addition, the interaction between hyaluronectin and hyaluronate was of relatively high affinity (Kd = 5.7 X 10(-10) M), and the size of the hyaluronate did not appear to substantially alter the amount of binding. To determine the amount of hyaluronate in an unknown sample, we used a competition assay in which the binding of a set amount of [3H]hyaluronate was blocked by the addition of unlabeled hyaluronate. By comparing the degree of competition of the unknown samples with that of known amounts of hyaluronate, it was possible to determine the amount of hyaluronate in the unknowns. We have found that this method is sensitive to 1 microgram or less of hyaluronate, and is unaffected by the presence of proteins.  相似文献   

18.
1. The stable complex formed between mammalian cytochrome c and phosvitin at low ionic strength was studied by partition in an aqueous two-phase system. Oxidized cytochrome c binds to phosvitin with a higher affinity than reduced cytochrome c. The difference was equivalent to a decrease of the redox potential by 22 mV on binding. 2. Complex-formation with phosvitin strongly inhibited the reaction of cytochrome c with reagents that react as negatively charged species, such as ascorbate, dithionite, ferricyanide and tetrachlorobenzoquinol. Reaction with uncharged reagents such as NNN'N'-tetramethylphenylenediamine and the reduced form of the N-methylphenazonium ion (present as the methylsulphate) was little affected by complex-formation, whereas oxidation of the reduced cytochrome by the positively charged tris-(phenanthroline)cobalt(III) ion was greatly stimulated. 3. A similar pattern of inhibition and stimulation of reaction rates was observed when phosvitin was replaced by other macromolecular polyanions such as dextran sulphate and heparin, indicating that the results were a general property of complex-formation with polyanions. A weaker but qualitatively similar effect was observed on addition of inositol hexaphosphate and ATP. 4. It is suggested that the effects of complex-formation with polyanions on the reactivity of cytochrome c with redox reagents are mainly the result of replacing the positive charge on the free cytochrome by a net negative charge. Any steric effects on polyanion binding are small in comparison with such electrostatic effects.  相似文献   

19.
Proteoglycans were extracted under nondissociative conditions from superficial and deeper layers of dog normal articular cartilage. The purified a-A1 preparations were characterized by velocity gradient centrifugation. Superficial specimens exhibited an abundant population of slow sedimenting aggregates whereas the aggregates of deeper preparations sedimented as two well-defined families of molecules. These dissimilarities in the size distribution of the aggregates observed between superficial and deeper a-A1 preparations derived most of all from differences in their content of hyaluronate and link proteins: (a) superficial preparations contained twice as much hyaluronate as deeper specimens; (b) superficial aggregates were link-free and unstable at pH 5.0 whereas deeper preparations contained link-proteins and their faster sedimenting aggregates were stabilized against dissociation at pH 5.0. In these proteoglycan preparations from different cartilage layers, the monomers exhibited an identical capacity for aggregation and the hyaluronate molecules displayed quite similar molecular weight (Mr = 5 x 10(5] and aggregating capacity. These observations as well as aggregating studies conducted with highly purified link protein and purified hyaluronate specimens of different molecular weights support the following conclusions: (a) link protein not only stabilizes proteoglycan aggregates but also enhances the aggregating capacity of hyaluronate; (b) for all practical purposes, the slow sedimenting aggregates represent a secondary complex of hyaluronate and proteoglycan monomers whereas the fast sedimenting aggregates may be considered as a ternary complex wherein link protein stabilizes the hyaluronate-proteoglycans interaction; (c) the distinctive heterogeneity of articular cartilage can be related to structurally different proteoglycan aggregates. The structural dissimilarities observed between superficial and deeper aggregates could reflect the different macromolecular organization of the proteoglycan molecules in the territorial and interterritorial matrices, respectively.  相似文献   

20.
SorLA/LR11 (250 kDa) is the largest and most composite member of the Vps10p-domain receptors, a family of type 1 proteins preferentially expressed in neuronal tissue. SorLA binds several ligands, including neurotensin, platelet-derived growth factor-bb, and lipoprotein lipase, and via complex-formation with the amyloid precursor protein it downregulates generation of Alzheimer's disease-associated Abeta-peptide. The receptor is mainly located in vesicles, suggesting a function in protein sorting and transport. Here we examined SorLA's trafficking using full-length and chimeric receptors and find that its cytoplasmic tail mediates efficient Golgi body-endosome transport, as well as AP-2 complex-dependent endocytosis. Functional sorting sites were mapped to an acidic cluster-dileucine-like motif and to a GGA binding site in the C terminus. Experiments in permanently or transiently AP-1 mu1-chain-deficient cells established that the AP-1 adaptor complex is essential to SorLA's transport between Golgi membranes and endosomes. Our results further implicate the GGA proteins in SorLA trafficking and provide evidence that SNX1 and Vps35, as parts of the retromer complex or possibly in a separate context, are engaged in retraction of the receptor from endosomes.  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