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 203 毫秒
1.
Laser nephelometry is a suitable technique for the quantitative determination and differentiation of both lectins and glycoconjugates in the low-picomolar range. Simultaneously this method renders possible investigations on the specificity and mode of interaction between lectins and different ligands. The results demonstrate that the degree of co-operativity between concanavalin A and the respective glycoconjugate is dependent on the presence of hydrophobic binding sites and can be substantially altered by conformational changes of the ligand. The transition from apotransferrin to Fe3+-transferrin induces a transformation of the sigmoidal-shaped binding curve to a hyperbolic one. Hence, at low concentrations, Fe3+-transferrin is bound far better than apotransferrin, whereas maximal binding is nearly identical. After removal of N-acetylneuraminate, concanavalin A is less efficient in differentiating between the Fe3+-charged and Fe3+-free (apo) forms of transferrin.  相似文献   

2.
An obligatory role for barbonate (or other synergistic anions) in the specific binding of Fe3+ by transferrin has been a point of controversy for two decades. There are an equal number of confirmatory and negative reports of specific Fe3+-transferrin binary complexes. A criticism of previous studies is the use of only one synthetic route, and limited product testing. This study reports the development of several preparative routes aimed at the formation of a specific Fe3+-transferrin complex, and the characterization of the products by spectrophotometry and chemical reactivity. The preparative routes described include: (a) displacement of carbonate from Fe3+-transferrin-CO32- at low pH followed by removal of CO2 by several techniques; (b) addition of FeCl3 to apotransferrin under CO2-free conditions; (c) oxidation of Fe2+ in the presence of apotransferrin under CO2-free conditions; (d) reaction of apotransferrin with nonsubstituting Fe3+ complexes in the absence of CO2; and (e) attempts to displace anions from weak Fe3+-transferrin-anion complexes. The product were examined with regard to their visible spectra, and their examined with regard to their visible spectra, and their reactivity with: (a) NaHCO3, (b) Fe3+-nitrilotriacetic acid in NaHCO3, and (c) citrate. The results are compared with the characteristics of Fe3+-transferrin-anion complexes and nonspecific Fe3+, transferrin mixtures. The data indicate that in the absence of synergistic anions the affinity of the specific metal binding sites of transfe-rin for Fe3+ is so low as to not compete favorably with hydrolytic polymerization and nonspecific binding effects.  相似文献   

3.
Effects of protease inhibitors on liver regeneration   总被引:2,自引:0,他引:2  
The oxidation of Fe2+ to Fe3+ by oxygen at pH 7.45 is a first order reaction with a 25 minute half life. In the presence of apotransferrin the oxidation rate is greatly enhanced and Fe3+-transferrin is formed. The apotransferrin mediated reaction reaches 50% completion in one minute; it does not follow simple first order kinetics. Iron-saturated transferrin does not exhibit the rate enhancement effect suggesting that the specific metal binding sites are the loci of the iron oxidation. Addition of H2O2, an agent which rapidly oxidizes Fe2+ to Fe3+, during the reaction of Fe2+ with apotransferrin greatly decreases the yield of Fe3+-transferrin. It is postulated that the basis of the rate enhancement effect is the binding of Fe2+ to the metal binding site of the transferrin molecule, followed by a rapid oxidation of the iron to the trivalent form.  相似文献   

4.
Binding of vanadate to human serum transferrin   总被引:1,自引:0,他引:1  
Human serum transferrin specifically and reversibly binds 2 equiv of vanadate at the two metal-binding sites of the protein. The vanadium(V)-transferrin complex can be formed either by the addition of vanadate to apotransferrin or by the air oxidation of the vanadyl(IV)-transferrin complex. The formation of the vanadium complex can be blocked by loading the apotransferrin with iron(III), and bound vanadium can be displaced from the protein by the subsequent addition of either gallium(III) or iron(III). The binding constant for the second equiv of vanadate is 10(6.5) in 0.1 M hepes, pH 7.4 at 25 degrees C. The binding constant for the first equiv of vanadate is probably very similar, although no quantitative value could be determined. Although transferrin reacts with the vanadate anion, studies on the transferrin model compound ethylenebis(o-hydroxyphenylglycine) indicate that at pH 9.5, the vanadium is binding at the metal-binding site as a dioxovanadium(V) cation coordinated to two phenolic residues at each binding site. This bound cation appears to be protonated over the pH range 9.5-6.5, as shown by changes in the difference uv spectrum of the transferrin complex, to produce an oxohydroxo species. Further decreases in the pH lead to dissociation of the vanadium-transferrin complex.  相似文献   

5.
The sedimentation behavior of 125I-labeled gastrin has been studied as a function of Fe3+ ion concentration and pH. Both sedimentation velocity and sedimentation equilibrium experiments indicated that high-molecular-weight Fe3+-gastrin complexes were formed at pH 5.0 and pH 7.4. Self-association of gastrin alone was observed at pH values below 5.0. 125I-labeled gastrin bound to human serum apotransferrin at pH 7.4. Scatchard analysis of the gastrin-apotransferrin complex gave a Kd of approximately 6.4 microM at 37 degrees C, with two binding sites per molecule of apotransferrin. No significant binding of gastrin to diferric transferrin was observed under the same conditions. The binding of gastrin to apotransferrin was inhibited by NaCl. The results are consistent with the hypothesis that gastrin and transferrin act synergistically in the uptake of dietary iron by the gastrointestinal tract.  相似文献   

6.
D C Harris 《Biochemistry》1977,16(3):560-564
Transferrin, the serum serum iron-transport protein which can bind two metal ions at physiologic pH, binds just one Fe3+, VO2+, or Cr3+ ion at pH 6.0. Fe3+ and VO2+ appear to be bound at the same site, designated A, based on electron paramagnetic resonance (EPR) spectra of VO2+-transferrin and (Fe3+)1(VO2+)1-transferrin. The EPR spectra of (Cr3+)1(VO2+)1-transferrin and of (Cr3+), (FE3+)1-transferrin indicate that that Cr3+ is bound to site B at pH 6.0. Transferrin was labeled at site A with 59Fe at pH 6.0 and at site B with 55Fe at pH 7.5. When the pH of the resulting preparation was lowered to 6.3 and the dissociated iron was separated by gel filtration, about ten times as much 55Fe as 59Fe was lost. The same EPR and isotopic-labeling experiments showed that Fe3+ added to transferrin at pH 7.5 binds to site A with about 90% selectivity.  相似文献   

7.
铽(Ⅲ)与人血清脱铁转铁蛋白结合的荧光光谱研究   总被引:5,自引:0,他引:5  
在pH7.40.1mol/LHepes及室温条件下,使用荧光光谱进行了Tb3+对人血清脱铁转铁蛋白的滴定.结果表明Tb3+与人血清脱铁转铁蛋白结合后,其549nm处的荧光强度增强约105倍.在549nm处Tb3+-脱铁转铁蛋白络合物的摩尔荧光强度是(9.65±0.05)×104mol-1L,Tb3+可占据脱铁转铁蛋白的两个金属离子结合部位,优先占据脱铁转铁蛋白的C端结合部位,条件平衡常数是lgKC=9.96±0.20,lgKN=6.37±0.16.Tb3+与R3+E(RE=Nd、Sm、Eu和Gd)间的线性自由能关系表明稀土离子占据脱铁转铁蛋白的C端结合部位时受离子大小的影响  相似文献   

8.
Structure of the human transferrin receptor-transferrin complex   总被引:9,自引:0,他引:9  
Cheng Y  Zak O  Aisen P  Harrison SC  Walz T 《Cell》2004,116(4):565-576
Iron, insoluble as free Fe(3+) and toxic as free Fe(2+), is distributed through the body as Fe(3+) bound to transferrin (Tf) for delivery to cells by endocytosis of its complex with transferrin receptor (TfR). Although much is understood of the transferrin endocytotic cycle, little has been uncovered of the molecular details underlying the formation of the receptor-transferrin complex. Using cryo-electron microscopy, we have produced a density map of the TfR-Tf complex at subnanometer resolution. An atomic model, obtained by fitting crystal structures of diferric Tf and the receptor ectodomain into the map, shows that the Tf N-lobe is sandwiched between the membrane and the TfR ectodomain and that the C-lobe abuts the receptor helical domain. When Tf binds receptor, its N-lobe moves by about 9 A with respect to its C-lobe. The structure of TfR-Tf complex helps account for known differences in the iron-release properties of free and receptor bound Tf.  相似文献   

9.
Effect of iron chelators on the transferrin receptor in K562 cells   总被引:16,自引:0,他引:16  
Delivery of iron to K562 cells by diferric transferrin involves a cycle of binding to surface receptors, internalization into an acidic compartment, transfer of iron to ferritin, and release of apotransferrin from the cell. To evaluate potential feedback effects of iron on this system, we exposed cells to iron chelators and monitored the activity of the transferrin receptor. In the present study, we found that chelation of extracellular iron by the hydrophilic chelators desferrioxamine B, diethylenetriaminepentaacetic acid, or apolactoferrin enhanced the release from the cells of previously internalized 125I-transferrin. Presaturation of these compounds with iron blocked this effect. These chelators did not affect the uptake of iron from transferrin. In contrast, the hydrophobic chelator 2,2-bipyridine, which partitions into cell membranes, completely blocked iron uptake by chelating the iron during its transfer across the membrane. The 2,2-bipyridine did not, however, enhance the release of 125I-transferrin from the cells, indicating that extracellular iron chelation is the key to this effect. Desferrioxamine, unlike the other hydrophilic chelators, can enter the cell and chelate an intracellular pool of iron. This produced a parallel increase in surface and intracellular transferrin receptors, reaching 2-fold at 24 h and 3-fold at 48 h. This increase in receptor number required ongoing protein synthesis and could be blocked by cycloheximide. Diethylenetriaminepentaacetic acid or desferrioxamine presaturated with iron did not induce new transferrin receptors. The new receptors were functionally active and produced an increase in 59Fe uptake from 59Fe-transferrin. We conclude that the transferrin receptor in the K562 cell is regulated in part by chelatable iron: chelation of extracellular iron enhances the release of apotransferrin from the cell, while chelation of an intracellular iron pool results in the biosynthesis of new receptors.  相似文献   

10.
Samples of human apotransferrin (apo . HTr) were saturated with Fe(III) by two different techniques, a method employing excess trisodium citrate to chelate Fe(III) and a nonchelating approach which involves the ferroxidase activity of ceruloplasmin to convert Fe(II)----Fe(III). The samples were radiolabelled with either 55Fe or 3H. Using an initial molar Fe/apo . HTr ratio of 2.0-2.1, preparations of human transferrin with bound Fe (Fe . HTr) using the citrate method invariably contained 2.2-2.4 atoms Fe/molecule, whereas Fe . HTr (ceruloplasmin method) contained 2.0 atoms/molecule as shown by spectrophotometric and radioactivity measurements. Uptake of Fe from these Fe . HTr preparations by K-562 cells grown in a serum-free medium was marginally, but consistently, more rapid from 55Fe . HTr (citrate) than from 55Fe . HTr (ceruloplasmin). Taking account of the different Fe contents of the Fe . HTr preparations, the rate measured over a 2-h period amounted to approximately 12,700 and 16,100 Fe atoms/(cell . min) for Fe . HTr (ceruloplasmin) and Fe . HTr (citrate), respectively. However, cell binding by the two Fe . [3H]HTr preparations did not differ significantly over the 8-h incubation period. Furthermore, from the 3H distribution, the quantities of Fe . HTr bound reversibly at the cell surface and contained within the cell were similar for the two Fe . HTr preparations. The results indicate that apo . HTr may bind Fe in different ways depending on the method of Fe presentation and that the Fe . HTr product can donate Fe to K-562 cells at a rate which may reflect the method used for Fe-complex formation.  相似文献   

11.
Reticulocyte binding of Fe(III)_-transferrin and transferrin complexes with other metal ions have been compared by different investigators. The functional relevance of this comparison is not clear, therefore transferrin complexes with Fe(III), Cu(II), Mn(II) and Zn(II) have been studied further by DEAE-cellulose chromatography and by measurement of transferrin and metal uptakes by rabbit reticulocytes.Human Fe-transferrin behaved as a weaker anion than apotransferrin during DEAE-cellulose chromatography; since Fe-transferrin has a higher negative charge than apotransferrin and behaves a as stronger anion in electrophoretic systems, the chromatographic result was the opposite of that anticipated. The lower affinity of human Fe-transferrin for DEAE-cellulose is probably caused by a redistribution of charged groups on the surface of transferrin molecules when Fe(III) ions are bound and is therefore considered to be dependent on molecular conformation. Apotransferrin and divalent metal-transferrin complexes were found to have nearly equal affinities for DEAE-cellulose, thus the effect on surface charge of human transferrin molecules induced by binding Fe(III) appeared to be limited to that metal ion.Iron uptake by reticulocytes was associated with increased binding of transferrin to the cell surface: uptake of divalent metals occured without a concomitant increase in transferrin uptake or evidence of a specific metal-transfer process. Cu-transferrin was rapidly dissociated during incubation with cells.The effect of Fe(III)_binding on human transferrin molecules was to alter the molecular affinity for charged surfaces, namely DEAE-cellulose and reticulocyte membranes. This was less apparent with rabbit transferrin. Transferrin complexes with divalent metals behaved as apotransferrin in the process of association with reticulocytes.  相似文献   

12.
Transferrin and iron uptake by rat reticulocytes   总被引:1,自引:0,他引:1  
The uptake of transferrin labeled with 3H and 59Fe by rat reticulocytes was studied to clarify the characteristics of the uptake process and intracellular transport. Rat reticulocytes took up transferrin in a saturable, time- and temperature-dependent manner. Scatchard analysis of the binding parameters indicated that transferrin molecules were bound to cell-surface receptors with high affinity. Monodansyl- cadaverine, a potent inhibitor of transglutaminase, reduced the amount of internalized transferrin but has no effect on the total amount of cell-associated transferrin, suggesting that transferrin is taken up by rat reticulocytes via receptor-mediated endocytosis. About 50% of the internalized 3H label was released from the cells after reincubation for 1 h in fresh medium. In contrast, no release of 59Fe label was observed. By immunoprecipitation and subsequent SDS-PAGE the released 3H-labeled product was identified as apotransferrin. Lysosomotropic reagents and a proton ionophore reduced the uptake of 59Fe. These results indicated that iron was removed from transferrin at an intracellular site in an acidic environment. The released iron was found not to associate with any intermediate ligands before it was utilized for heme synthesis in mitochondria.  相似文献   

13.
A method is described in which the glycoprotein transferrin was double labeled. Its sialic acid residues were labeled with 3H through a consecutive oxidation-reduction technique utilizing tritiated NaBH4. Its protein moiety was labeled with either 125I or 59Fe. Incubation of this double-labeled molecule at 4 degrees C with K562 cells gave overlapping curves, indicating identical patterns of binding for all labels. At 37 degrees C, 3H and 125I demonstrated identical patterns while 59Fe was cummulatively retained. This method can be used to follow the fate of other glycoproteins and their possible desialation in vivo.  相似文献   

14.
The chemotherapeutic agent, bleomycin, forms a 1:1complex with both Fe(III) and Fe(II). The rate offerric ion transfer from bleomycin toapotransferrin is rather slow. However, when ascorbate was added toFe(III)-bleomycin priorto exposure to apotransferrin, the transfer rate was markedly increased. Ascorbatereadilyreduces Fe(III)-bleomycin to Fe(II)-bleomycin. A second order rate constant of 2.4 mM min wasestimated for this reaction. Fe(II)-bleomycinimmediately combines with O 2 , generating the so-called'acti-vatedbleomycin' complex. The data suggest that a reduced form of iron-bleomycin more readilydonatesits iron ion to apotransferrin. Reoxidation of ferrous ions, andFe(III)-transferrin formation occur rapidly.  相似文献   

15.
The ability of transferrin to potentiate oxygen free radical-mediated endothelial cell injury was assessed. 51Cr-labeled endothelial cells derived from rat pulmonary arteries (RPAECs) were incubated with hydrogen peroxide (H2O2) in the presence and absence of holosaturated human transferrin, and the effect of transferrin on H2O2-mediated endothelial cell toxicity was determined. Addition of holosaturated transferrin potentiated H2O2-mediated RPAEC cytotoxicity at concentrations of H2O2 greater than 10 microM, suggesting that transferrin may provide a source of iron for free radical-mediated endothelial cell injury. Free radical-mediated injury is dependent on non-protein-bound iron. The ability of RPAECs to facilitate the release of iron from transferrin was assessed. We determined that RPAECs facilitate the release of transferrin-derived iron by reduction of transferrin-bound ferric iron (Fe3+) to ferrous iron (Fe2+). The reduction and release of transferrin-derived Fe2+ were inhibited by apotransferrin and chloroquine, indicating a dependence on receptor-specific binding of transferrin to the RPAEC cell surface, with subsequent endocytosis, acidification, and reduction of transferrin-bound Fe3+ to Fe2+. The release of transferrin-derived Fe2+ was potentiated by diethyldithiocarbamate, an inhibitor of intracellular superoxide dismutase (SOD). In contrast, exogenous SOD did not alter iron release, suggesting that intracellular superoxide anion (O2-) may play an important role in mediating the reduction and release of transferrin-derived iron. Results of this study suggest that transferrin may provide a source of iron for oxygen free radical-mediated endothelial cell injury and identify a novel mechanism by which endothelial cells may mediate the reduction and release of transferrin-derived iron.  相似文献   

16.
Liu R  Guan JQ  Zak O  Aisen P  Chance MR 《Biochemistry》2003,42(43):12447-12454
Human transferrin, a bilobal protein, with each lobe bearing a single iron-binding site, functions to transport iron into cells. While the N-terminal lobe alone does not measurably bind cellular transferrin receptors or serve as an iron donor for cells, the C-lobe is capable of both functions. We used hydroxyl radical-mediated protein footprinting and mass spectrometry to reveal the conformational changes that occur upon complex formation for the human transferrin C-lobe (residues 334-679) bound to the ectodomain of human transferrin receptor 1 (residues 121-760). Oxidation rates for proteolytic peptides in the C-lobe, the receptor, and their complex have been measured by mass spectrometry; upon formation of the complex, a dramatic decrease in modification rates, indicating protection of specific side chain groups, can be seen in C-lobe sequences corresponding to residues 381-401, 415-433, and 457-470. Peptide sequences experiencing modification rate decreases in the transferrin receptor upon C-lobe binding include residues 232-240, 365-371, 496-508, 580 and 581, 614-623, 634-646, 647-681, and 733-760. In addition, several peptides in the receptor exhibit enhancements in the rate of modification consistent with allosteric effects of complex formation. Using tandem mass spectrometry, the sites of modification with altered reactivity in the complex include Met382, Met389, Trp460, Met464, and Phe427 in the C-lobe and Tyr503, Pro581, Tyr611, Leu619, Met635, Phe650, Trp740, Trp754, and Phe760 within the transferrin receptor. Using available genetic, biochemical, and structural data, we confirm that the conserved RGD sequence (residues 646-648) in the helical domain of the transferrin receptor, including residues from Leu619 to Phe650, is a primary binding site for the transferrin C-lobe.  相似文献   

17.
The interaction of hydroxypyridinones with human serum transferrin and ovotransferrin has been studied by analyzing the distribution of iron between the chelator and the proteins as a function of both ligand concentration and transferrin saturation. The kinetics of iron removal by 3-hydroxypyridin-4-ones from both transferrins is slow; in ovotransferrin it appears to be monophasic, in contrast to that observed for serum transferrin. After 24 hours incubation at a 40:1 chelator:protein molar ratio, the percentage of iron removed from Fe(III)-ovotransferrin is 50%-60%, and is somewhat higher in the case of serum transferrin, in line with the respective affinity constants for the metal. The 3-hydroxypyridin-2-ones and the 3-hydroxypyran-4-ones, both of which have lower affinities for Fe(III), remove smaller proportions of the metal. The percentage of desaturation obtained with bidentate and hexadentate pyridinones appears to be similar for both transferrin classes at chelator:protein molar ratios from 40:1. The degree of transferrin saturation influences the extent of chelator mediated iron mobilization in the case of serum transferrin, but not of ovotransferrin. 59Fe competition studies demonstrate that bidentate pyridin-4-ones are capable of donating iron to serum apotransferrin; the relative concentrations of ligand and protein influence the distribution of iron because their effective binding constants (at pH 7.4) for Fe(III) are similar.  相似文献   

18.
This study systematically examined the characteristics of specific binding of adult diferric transferrin to its receptor using a Triton X-100 solubilized preparation from human placentas as the receptor source. The following information was obtained. The ionic strength for maximal binding is in the range of 0.1-0.3 M NaCl. The pH optimum for specific binding extends over the range, from pH 6.0-10.0. Specific binding of diferric transferrin is not affected by 2.5 approximately 50 mM CaCl2 or by 10 mM EDTA. Triton X-100 in the concentration range of 0.02-3.0% does not affect specific binding. Specific binding is saturated within 10 min at 25 or 37 degrees C in the presence of excess amounts of diferric transferrin. The binding is reversible and the dissociation of diferric transferrin from the transferrin receptor is complete within 40 min at 25 degrees C. Apotransferrin, both adult and fetal, showed less binding than the holotransferrin species by competitive binding assay in the presence of 10 mM EDTA independent of up to 20 mM CaCl2. A 1500-fold molar excess of adult and fetal apotransferrin is required to give 40% inhibition for 125I-labeled diferric transferrin binding. Since calcium ion is not a factor, and since apotransferrin has such high binding affinity for iron (Ka = 1 X 10(24], this experiment suggests that the EDTA was necessary to prevent conversion of apotransferrin to holotransferrin from available iron in the reaction system. The specificity of the transferrin receptor for transferrin was examined by competitive binding studies in which 125I-diferric transferrin binding was measured in the presence of a series of other proteins. The proteins tested in the competitive binding studies were classified into three groups; in the first group were human serum albumin and ovalbumin; in the second group were proteins containing iron ions, such as hemoglobin, hemoglobin-haptoglobin complex, heme-hemopexin complex, ferritin, and diferric lactoferrin; in the third group were the metal-binding serum proteins, ceruloplasmin and metallothionein. None of these proteins except ferritin showed inhibition of diferric transferrin binding to the receptor. The effect of ferritin was small since a 700- to 1500-fold molar excess of ferritin is required for 50% inhibition of binding of diferric transferrin to the receptor.  相似文献   

19.
Zak O  Aisen P 《Biochemistry》2002,41(5):1647-1653
Eukaryotic transferrins comprise a class of bilobal iron-binding proteins in which each lobe carries a single binding site. Although expression of full-length transferrins and their N-terminal lobes, in wild-type and mutated forms, has been successfully accomplished by several laboratories, expression of C-lobes has been much less satisfactory. A possible explanation of the difficulty is that proper folding of the C-lobe, with its 11 disulfide bonds, depends on prior synthesis and proper folding of the N-lobe. We have therefore developed a new strategy, introducing a specific factor Xa cleavage site in the interlobe-connecting strand to permit separation of the lobes after expression of the full-length protein. The resulting protein was expressed in satisfactory yield, >20 mg/L, and could be easily and completely cleaved to yield two distinguishable fragments representing N- and C-lobes, respectively. Retaining the glycosylation sites, found only in the C-lobe, made it possible to separate the fragments from each other by ConA affinity chromatography. The isolated C-lobe so obtained displayed spectroscopic and kinetic features of the C-lobe in native transferrin and was competent as an iron donor for K562 cells to which it bound in saturable fashion inhibitable by native diferric transferrin. Since the N-lobe by itself will neither bind nor donate iron to cells, the primary receptor-recognition site of transferrin resides in its C-lobe.  相似文献   

20.
The molecular basis of the transferrin (TF)-transferrin receptor (TFR) interaction is not known. The C-lobe of TF is required to facilitate binding to the TFR and both the N- and C-lobes are necessary for maximal binding. Several mAb have been raised against human transferrin (hTF). One of these, designated F11, is specific to the C-lobe of hTF and does not recognize mouse or pig TF. Furthermore, mAb F11 inhibits the binding of TF to TFR on HeLa cells. To map the epitope for mAb F11, constructs spanning various regions of hTF were expressed as glutathione S-transferase (GST) fusion proteins in Escherichia coli. The recombinant fusion proteins were analysed in an iterative fashion by immunoblotting using mAb F11 as the probe. This process resulted in the localization of the F11 epitope to the C1 domain (residues 365-401) of hTF. Subsequent computer modelling suggested that the epitope is probably restricted to a surface patch of hTF consisting of residues 365-385. Mutagenesis of the F11 epitope of hTF to the sequence of either mouse or pig TF confirmed the identity of the epitope as immunoreactivity was diminished or lost. In agreement with other studies, these epitope mapping studies support a role for residues in the C1 domain of hTF in receptor 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