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 589 毫秒
1.
Anionic phosphatidic acid (PA) has been shown to stabilize and bind stronger than phosphatidylglycerol via electrostatic and hydrogen bond interaction with the positively charged residues of potassium channel KcsA. However, the effects of these lipids on KcsA folding or secondary structure are not clear. In this study, the secondary structure analyses of KcsA potassium channel was carried out using circular dichroism spectroscopy. It was found that PA interaction leads to increases in α-helical and β-sheet content of KcsA protein. In PA, KcsA α-helical structure was further stabilized by classical membrane-active cosolvent trifluoroethanol followed by reduction in the β-sheet content indicating cooperative transformation from the β-sheet to an α-helical structure. The data further uncover the role of anionic PA in KcsA folding and provide mechanism by which strong hydrogen bonds/electrostatic interaction among PA headgroup and basic residues on lipid binding domains may induce high helical structure thereby altering the protein folding and increasing the stability of tetrameric assembly.  相似文献   

2.
Folding type-specific secondary structure propensities of 20 naturally occurring amino acids have been derived from α-helical, β-sheet, α/β, and α+β proteins of known structures. These data show that each residue type of amino acids has intrinsic propensities in different regions of secondary structures for different folding types of proteins. Each of the folding types shows markedly different rank ordering, indicating folding type-specific effects on the secondary structure propensities of amino acids. Rigorous statistical tests have been made to validate the folding type-specific effects. It should be noted that α and β proteins have relatively small α-helices and β-strands forming propensities respectively compared with those of α+β and α/β proteins. This may suggest that, with more complex architectures than α and β proteins, α+β and α/β proteins require larger propensities to distinguish from interacting α-helices and β-strands. Our finding of folding type-specific secondary structure propensities suggests that sequence space accessible to each folding type may have differing features. Differing sequence space features might be constrained by topological requirement for each of the folding types. Almost all strong β-sheet forming residues are hydrophobic in character regardless of folding types, thus suggesting the hydrophobicities of side chains as a key determinant of β-sheet structures. In contrast, conformational entropy of side chains is a major determinant of the helical propensities of amino acids, although other interactions such as hydrophobicities and charged interactions cannot be neglected. These results will be helpful to protein design, class-based secondary structure prediction, and protein folding. © 1998 John Wiley & Sons, Inc. Biopoly 45: 35–49, 1998  相似文献   

3.
Conformational transitions of calmodulin as studied by vacuum-uv CD   总被引:1,自引:0,他引:1  
CD measurements were made for calmodulin and its calcium (Ca2+) complexes at different ionic strengths and Ca2+ concentrations. Calmodulin at an ionic strength of 0.00M and in the absence of Ca2+ exists as an α-helical protein with a negligible amount of β-sheet. An increase in ionic strength, whether or not Ca2+ is present, increases α-helix at the expense of “other” (coil) structure. The changes in β-sheet and β-turns are insignificant. Binding of Ca2+ at low ionic strength occurs in stages with at least one folding intermediate before attaining the final stable state. Binding of Ca2+ at an ionic strength of 0.165M causes only a slight increase in α-helix, so that the secondary structure of the protein depends on ionic strength and is insensitive to the nature of the cation (i.e., Ca2+). Thus, the activation of calmodulin by Ca2+ must be due to a structural reorientation rather than to a major secondary structural alteration. The CD estimation of secondary structure with 4 mol Ca2+/calmodulin (61% α-helix, 2% antiparallel β-sheet, 2% parallel β-sheet, 21% β-turns, and 14% other) is in excellent agreement with the x-ray results.  相似文献   

4.
E A Kabat  T T Wu 《Biopolymers》1973,12(4):751-774
The influence of nearest-neighbor pairs of amino acids (n ? 1) and (n + 1) on the conformation of amino acid (n) in proteins has been studied. From experimental data on eleven proteins of known three dimensional structures, our definition of an α-helical domain in the Φ,Ψ plot has been reexamined and found to be satisfactory. On the same principle, a regular β-sheet domain has been delineated. We then revised our 20 × 20 table of frequencies of occurrences of various conformations tabulating three values: α-helical, β-sheet, and neither. These frequencies were then used to locate the helixbreaking positions in cytochrome b5, papain, thermolysin, and calcium-binding protein. In conjuction with the helical wheel method, they were useful for predicting the locations of most α-helical segments. Similarly the β-sheet breaking positions in papain were located and most of the β-sheets found by X-ray diffraction were close to or between these positions. Data on β-sheets are extremely sparse so that extensive tests were not possible. The application of this method to abnormal hemoglobins suggested possible distortions of helices and in several instances correlated with abnormal properties of the hemoglobins and association with disease. The variable region of human immunoglobin heavy chains was found to have a very low α-helical content though β-sheet structures might exist.  相似文献   

5.
The secondary structure implications of precipitation induced by a chaotropic salt, KSCN, and a structure stabilizing salt, Na2SO4, were studied for twelve different proteins. α-helix and β-sheet content of precipitate and native structures were estimated from the analysis of amide I band Raman spectra. A statistical analysis of the estimated perturbations in the secondary structure contents indicated that the most significant event is the formation of β-sheet structures with a concomitant loss of α-helix on precipitation with KSCN. The conformational changes for each protein were also analyzed with respect to elements of primary, secondary and tertiary structure existing in the native protein; primary structure was quantified by the fractions of hydrophobic and charged amino acids, secondary structure by x-ray estimates of α-helix and β-sheet contents of native proteins and tertiary structure by the dipole moment and solvent-accessible surface area. For the KSCN precipitates, factors affecting β-sheet content included the fraction of charged amino acids in the primary sequence and the surface area. Changes in α-helix content were influenced by the initial helical content and the dipole moment. The enhanced β-sheet contents of precipitates observed in this work parallel protein structural changes occurring in other aggregative phenomena.  相似文献   

6.
Various copolypeptides were prepared by benzylamine or tertiary amine-initiated copolymerizations of alanine–N-carboxyanhydride (Ala-NCA) and valine–N-carboxyanhydride (Val-NCA). The number-average molecular weights of these copolypeptides were detemined by 1H-nmr spectroscopic end-group analyses and viscosity measurements. The sequences were characterized by 15N-nmr spectra in solution, and the average lengths of the homogeneous blocks were determined from the signal intensities. The 50.3-and 75.4-MHz 13C-nmr CP/MAS spectra of the solid copolypeptides are not sensitive to sequence effects, but allow qualitative and quantitative analyses of the secondary structures. In contrast to other methods, the 13C-nmr spectra allow determination of the extent to which individual amino acids are incorporated into β-sheet or α-helix phases. Depending on primary structure and molecular weight, the secondary structure of (Ala/Val) copolypeptides may vary significantly. Both monomer units may be predominantly helical or predominantly β-sheet structure, or the Val units may prefer the β-sheet structure with most Ala-units forming β-helices. However, these secondary structures are more or less thermodynamically unstable and revert to the stable conformations on reprecipitation from trifluoroacetic acid/water.  相似文献   

7.
2,2,2-Trifuoroethanol (TFE)-induced conformational structure change of a β-sheet legume lectin, soybean agglutinin (SBA) has been investigated employing its exclusive structural forms in quaternary (tetramer) and tertiary (monomer) states, by far- and near-UV CD, FTIR, fluorescence, low temperature phosphorescence and chemical modification. Far-UV CD results show that, for SBA tetramer, native atypical β-conformation transforms to a highly α-helical structure, with the helical content reaching 57% in 95% TFE. For SBA monomer, atypical β-sheet first converts to typical β-sheet at low TFE concentration (10%), which then leads to a nonnative α-helix at higher TFE concentration. From temperature-dependent studies (5–60 °C) of TFE perturbation, typical β-sheet structure appears to be less stable than atypical β-sheet and the induced helix entails reduced thermal stability. The heat induced transitions are reversible except for atypical to typical β-sheet conversion. FTIR results reveal a partial α-helix conversion at high protein concentration but with quantitative yield. However, aggregation is detected with FTIR at lower TFE concentration, which disappears in more TFE. Near-UV CD, fluorescence and phosphorescence studies imply the existence of an intermediate with native-like secondary and tertiary structure, which could be related to the dissociation of tetramer to monomer. This has been further supported by concentration dependent far-UV CD studies. Chemical modification with N-bromosuccinimide (NBS) shows that all six tryptophans per monomer are solvent-exposed in the induced α-helical conformation. These results may provide novel and important insights into the perturbed folding problem of SBA in particular, and β-sheet oligomeric proteins in general.  相似文献   

8.
The secondary structure of the major neurotoxin from the sea snake Lapemis hardwickii was investigated by several methods of conformational analysis: structure prediction, circular dichroism, and laser Raman spectroscopy. From the primary structure, secondary structure prediction yielded two regions of β-sheet structure at residues 1–7 and 41–45. β-Turns were predicted at residues 14–17, 20–23, 30–33, 37–40, and 46–49. From the predictions, the toxin appears to be composed of approximately 20% β-sheet and 33% β-turn. The CD spectrum of the native toxin appears to be a hybrid of model spectra for β-sheet and β-turn proteins. The pH perturbation studies on the toxin observed by CD demonstrated that the toxin is a very stable molecule except at extremely high or low pH values. The Raman data indicated that the toxin contains both antiparallel β-sheet and β-turn structure. Using two methods of secondary structure quantitation from Raman spectra the molecule was calculated to contain 35% β-sheet from one method and 27% from the other. Overall, the various methods demonstrate that the toxin is composed of β-sheet and β-turn structure with little or no α-helix present. From the comparison of these different techniques appreciation can be gained for the necessity of several methods when identifying and quantitating secondary structure.  相似文献   

9.
A multiscale simulation method of protein folding is proposed, using atomic representation of protein and solvent, combing genetic algorithms to determine the key protein structures from a global view, with molecular dynamic simulations to reveal the local folding pathways, thus providing an integrated landscape of protein folding. The method is found to be superior to previously investigated global search algorithms or dynamic simulations alone. For secondary structure formation of a selected peptide, RN24, the structures and dynamics produced by this method agree well with corresponding experimental results. Three most populated conformations are observed, including hairpin, β-sheet and α-helix. The energetic barriers separating these three structures are comparable to the kinetic energy of the atoms of the peptide, implying that the transition between these states can be easily triggered by kinetic perturbations, mainly through electrostatic interactions between charged atoms. Transitions between α-helix and β-sheet should jump over at least two energy barriers and may stay in the energetic trap of hairpin. It is proposed that the structure of proteins should be jointly governed by thermodynamic and dynamic factors; free energy is not the exclusive dominant for stability of proteins.  相似文献   

10.
The venom of the North African scorpion Androctonus mauretanicus mauretanicus possesses numerous highly active neurotoxins that specifically bind to various ion channels. One of these, P05, has been found to bind specifically to calcium-activated potassium channels and also to compete with apamin, a toxin extracted from bee venom. Besides the highly potent ones, several of these peptides (including that of P01) have been purified and been found to possess only a very weak, although significant, activity in competition with apamin. The amino acid sequence of P01 shows that it is shorter than P05 by two residues. This deletion occurs within an α-helix stretch (residues 5–12). This α-helix has been shown to be involved in the interaction of P05 with its receptor via two arginine residues. These two arginines are absent in the P01 sequence. Furthermore, a proline residue in position 7 of the P01 sequence may act as an α-helix breaker. We have determined the solution structure of P01 by conventional two-dimensional 1H nuclear magnetic resonance and show that 1) the proline residue does not disturb the α-helix running from residues 5 to 12; 2) the two arginines are topologically replaced by two acidic residues, which explains the drop in activity; 3) the residual binding activity may be due to the histidine residue in position 9; and 4) the overall secondary structure is conserved, i.e., an α-helix running from residues 5 to 12, two antiparallel stretches of β-sheet (residues 15–20 and 23–27) connected by a type I′ β-turn, and three disulfide bridges connecting the α-helix to the β-sheet.  相似文献   

11.
小麦多聚半乳糖醛酸酶抑制蛋白的部分结构   总被引:4,自引:0,他引:4  
为了弄清小麦多聚半乳糖醛酸酶抑制蛋白 (polygalacturonase inhibitingprotein ,PGIP)的作用机制 ,并为其在基因工程中的应用提供依据 ,对其结构进行了研究 .用Edman降解法测得小麦PGIP的N端序列为Lys Pro Leu Leu Thr Lys Ile Thr Lys Gly Ala Ala Ser Thr .用CD谱研究其二级结构 ,发现小麦PGIP天然态含有 4 3 7%的 β折叠和 13 1%的α螺旋 .酸碱和温度变性引起了二级结构改变 .不完全变性阶段 ,二级结构的变化表现为α螺旋无明显变化 ,β折叠遭到破坏 ;活性完全丧失阶段 ,β折叠变化很小 ,α螺旋含量明显减少 .用NR R(非还原 还原 )双向对角线SDS PAGE鉴定出小麦PGIP含有链内二硫键 .用去糖基化法确证了小麦PGIP的糖含量为 2 2 %.小麦PGIP与双子叶植物PGIP相比 ,一级结构差异较大 ,同源性由 36 %变为 9%;二级结构相似 ,都是高 β 折叠的蛋白 ;均具有链内二硫键 ;在糖含量上也相似 .研究结果为进一步弄清小麦PGIP作用机理打下了基础 ,同时对于植物抗赤霉病基因工程具有重要意义 .  相似文献   

12.
Dynorphins, endogeneous opioid neuropeptides, function as ligands to the opioid kappa receptors and also induce non-opioid effects in neurons, probably related to direct membrane interactions. We have characterized the structure transitions of dynorphins (big dynorphin, dynorphin A and dynorphin B) induced by the detergent sodium dodecyl sulfate (SDS). In SDS titrations monitored by circular dichroism, we observed secondary structure conversions of the peptides from random coil to α-helix with a highly aggregated intermediate. As determined by Fourier transform infrared spectroscopy, this intermediate exhibited β-sheet structure for dynorphin B and big dynorphin. In contrast, aggregated dynorphin A was α-helical without considerable β-sheet content. Hydrophobicity analysis indicates that the YGGFLRR motif present in all dynorphins is prone to be inserted in the membrane. Comparing big dynorphin with dynorphin A and dynorphin B, we suggest that the potent neurotoxicity of big dynorphin could be related to the combination of amino acid sequences and secondary structure propensities of dynorphin A and dynorphin B, which may generate a synergistic effect for big dynorphin membrane perturbing properties. The induced aggregated α-helix of dynorphin A is also correlated with membrane perturbations, whereas the β-sheet of dynorphin B does not correlate with membrane perturbations.  相似文献   

13.
Poly-L -lysine exists as an α-helix at high pH and a random coil at neutral pH. When the α-helix is heated above 27°C, the macromolecule undergoes a conformational transition to a β-sheet. In this study, the stability of the secondary structure of poly-L -lysine in solutions subjected to shear flow, at temperatures below the α-helix to β-sheet transition temperature, were examined using Raman spectroscopy and CD. Solutions initially in the α-helical state showed time-dependent increases in viscosity with shearing, rising as much as an order of magnitude. Visual observation and turbidity measurements showed the formation of a gel-like phase under flow. Laser Raman measurements demonstrated the presence of small amounts of β-sheet structure evidenced by the amide I band at 1666 cm−1. CD measurements indicated that solutions of predominantly α-helical conformation at 20°C transformed into 85% α-helix and 15% β-sheet after being sheared for 20 min. However, on continued shearing the content of β-sheet conformation decreased. The observed phenomena were explained in terms of a “zipping-up” molecular model based on flow enhanced hydrophobic interactions similar to that observed in gel-forming flexible polymers. © 1998 John Wiley & Sons, Inc. Biopoly 45: 239–246, 1998  相似文献   

14.
A new approach for evaluating the secondary structure of proteins by CD spectroscopy of overlapping peptide segments is applied to porcine adenylate kinase (AK1) and yeast guanylate kinase (GK3). One hundred seventy-six peptide segments of a length of 15 residues, overlapping by 13 residues and covering the complete sequences of AK1 and GK3, were synthesized in order to evaluate their secondary structure composition by CD spectroscopy. The peptides were prepared by solid phase multiple peptide synthesis method using the 9-fluorenylmethoxycarbonyl/tert-butyl strategy. The individual peptide secondary structures were studied with CD spectroscopy in a mixture of 30% trifluoroethanol in phosphate buffer (pH 7) and subsequently compared with x-ray data of AK1 and GK3. Peptide segments that cover α-helical regions of the AK1 or GK3 sequence mainly showed CD spectra with increasing and decreasing Cotton effects that were typical for appearing and disappearing α-helical structures. For segments with dominating β-sheet conformation, however, the application of this method is limited due to the stability and clustering of β-sheet segments in solution and due to the difficult interpretation of random-coiled superimposed β-sheet CD signals. Nevertheless, the results of this method especially for α-helical segments are very impressive. All α-helical and 71% of the β-sheet containing regions of the AK1 and GK3 could be identified. Moreover, it was shown that CD spectra of consecutive peptide content reveal the appearance and disappearance of α-helical secondary structure elements and help localizing them on the sequence string. © 1997 John Wiley & Sons, Inc. Biopoly 41: 213–231, 1997  相似文献   

15.
A major bottleneck in the field of biochemistry is our limited understanding of the processes by which a protein folds into its native conformation. Much of the work on this issue has focused on the conserved core of the folded protein. However, one might imagine that a ubiquitous motif for unaided folding or for the recognition of chaperones may involve regions on the surface of the native structure. We explore this possibility by an analysis of the spatial distribution of regions with amphiphilic α-helical potential on the surface of β-sheet proteins. All proteins, Including β-sheet proteins, contain regions with amphiphilic α-helical potential. That is, any α-helix formed by that region would be amphiphilic, having both hydrophobic and hydrophilic surfaces. In the three-dimensional structure of all β-sheet proteins analyzed, we have found a distinct pattern in the spatial distribution of sequences with amphiphilic α-helical potential. The amphiphilic regions occur in ring shaped clusters approximately 20 to 30 Å in diameter on the surface of the protein. In addition, these regions have a strong preference for positively charged amino acids and a lower preference for residues not favorable to α-helix formation. Although the purpose of these amphiphilic regions which are not associated with naturally occurring α-helix is unknown, they may play a critical role in highly conserved processes such as protein folding. © 1996 Wiley-Liss, Inc.  相似文献   

16.
Dynamic Monte Carlo studies have been performed on various diamond lattice models of β-proteins. Unlike previous work, no bias toward the native state is introduced; instead, the protein is allowed to freely hunt through all of phase space to find the equilibrium conformation. Thus, these systems may aid in the elucidation of the rules governing protein folding from a given primary sequence; in particular, the interplay of short- vs long-range interaction can be explored. Three distinct models (A? C) were examined. In model A, in addition to the preference for trans (t) over gauche states (g+ and g?) (thereby perhaps favoring β-sheet formation), attractive interactions are allowed between all nonbonded, nearest neighbor pairs of segments. If the molecules possess a relatively large fraction of t states in the denatured form, on cooling spontaneous collapse to a well-defined β-barrel is observed. Unfortunately, in model A the denatured state exhibits too much secondary structure to correctly model the globular protein collapse transition. Thus in models B and C, the local stiffness is reduced. In model B, in the absence of long-range interactions, t and g states are equally weighted, and cooperativity is introduced by favoring formation of adjacent pairs of nonbonded (but not necessarily parallel) t states. While the denatured state of these systems behaves like a random coil, their native globular structure is poorly defined. Model C retains the cooperativity of model B but allows for a slight preference of t over g states in the short-range interactions. Here, the denatured state is indistinguishable from a random coil, and the globular state is a well-defined β-barrel. Over a range of chain lengths, the collapse is well represented by an all-or-none model. Hence, model C possesses the essential qualitative features observed in real globular proteins. These studies strongly suggest that the uniqueness of the globular conformation requires some residual secondary structure to be present in the denatured state.  相似文献   

17.
The solid state secondary structure of myoglobin, RNase A, concanavalin A (Con A), poly(L -lysine), and two linear heterooligomeric peptides were examined by both far-uv CD spectroscopy1 and by ir spectroscopy. The proteins associated from water solution on glass and mica surfaces into noncrystalline, amorphous films, as judged by transmission electron microscopy of carbon-platinum replicas of surface and cross-fractured layer. The association into the solid state induced insignificant changes in the amide CD spectra of all α-helical myoglobin, decreased the molar ellipticity of the α/β RNase A, and increased the molar ellipticity of all-β Con A with no change in the positions of the bands' maxima. High-temperature exposure of the films induced permanent changes in the conformation of all proteins, resulting in less α-helix and more β-sheet structure. The results suggest that the protein α-helices are less stable in films and that the secondary structure may rearrange into β-sheets at high temperature. Two heterooligomeric peptides and poly (L -lysine), all in solution at neutral pH with “random coil” conformation, formed films with variable degrees of their secondary structure in β-sheets or β-turns. The result corresponded to the protein-derived Chou-Fasman amino acid propensities, and depended on both temperature and solvent used. The ir and CD spectra correlations of the peptides in the solid state indicate that the CD spectrum of a “random” structure in films differs from random coil in solution. Formic acid treatment transformed the secondary structure of the protein and peptide films into a stable α-helix or β-sheet conformations. The results indicate that the proteins aggregate into a noncrystalline, glass-like state with preserved secondary structure. The solid state secondary structure may undergo further irreversible transformations induced by heat or solvent. © 1993 John Wiley & Sons, Inc.  相似文献   

18.
A computer program is used to analyse automatically and objectively the atomic co-ordinates of a large number of globular proteins in order to identify the regions of α-helix, β-sheet and reverse-turn secondary structure. Several different criteria for the assignment of secondary structure are tested for accuracy, reproducibility and efficiency. The most successful criterion, which is based on patterns of peptide hydrogen bonds, inter-Cα distances and inter-Cα torsion angles, is used to find the secondary structure of all the proteins studied. The accuracy of the derived assignments is assessed by comparing them with the secondary structure reported in the literature for each protein. The reliability of the methods is assessed by comparing the secondary structures derived from the independently determined sets of co-ordinates available for some proteins.We provide the first objective and consistent compilation of α-helix, β-sheet and reverse-turn secondary structure in almost all globular proteins of known tertiary structure. These data will be invaluable for analysing the relative tendencies of different amino acids to occur in different types of secondary structure, for analysing the regularity of the secondary structure itself, and for analysing how the pieces of secondary structure fit together to form the globular tertiary structure of each protein.  相似文献   

19.
Folding of the ribosomal protein S6 is a malleable process controlled by two competing, and partly overlapping, folding nuclei. Together, these nuclei extend over most of the S6 structure, except the edge strand β2, which is consistently missing in the folding transition states; despite being part of the S6 four-stranded sheet, β2 seems not to be part of the cooperative unit of the protein. The question is then whether β2 can be removed from the S6 structure without compromising folding cooperativity or native state integrity. To investigate this, we constructed a truncated variant of S6 lacking β2, reducing the size of the protein from 96 to 76 residues (S6(Δβ2)). The new S6 variant expresses well in Escherichia coli and has a well dispersed heteronuclear single quantum correlation spectrum and a perfectly wild-type-like crystal structure, but with a smaller three-stranded β-sheet. Moreover, S6(Δβ2) displays an archetypical v-shaped chevron plot with decreased slope of the unfolding limb, as expected from a protein with maintained folding cooperativity and reduced size. The results support the notion that foldons, as defined by the structural distribution of the folding nuclei, represent a property-based level of hierarchy in the build-up of larger protein structures and suggest that the role of β2 in S6 is mainly in intermolecular binding, consistent with the position of this strand in the ribosomal assembly.  相似文献   

20.
Alcohol-induced conformational transitions of erv C, a highly stable cysteine protease, were followed by CD, fluorescence, and activity. At acidic pH, the addition of different alcohols caused two types of conformational transitions. Increasing the concentration of nonfluorinated alkyl alcohols induced a conformational switch from α-helix to β-sheet. Under these conditions, the protein lost its proteolytic activity and tertiary structure. The switch was a sudden one, observed in 50% methanol, 45% ethanol, and 40% propanol. Under similar conditions of pH and concentration, however, glycerol and TFE enhanced the α-helicity of the protein. Methanol-induced denaturation was observed to occur in two stages; the first is the β-sheet state stabilized at low alcohol concentrations, and the other is the β-sheet state with enhanced ellipticity stabilized at high alcohol concentrations. This β-sheet conformation can be attained from the native as well as 6 M GuHCl-denatured state by addition of methanol and exhibits properties different from the native or unfolded state. This state shows loss of tertiary structure and activity, enhanced nonnative secondary structure, noncooperative temperature unfolding, and higher stability toward denaturants as compared to the native state, which are characteristic of the molten globule-like state or O-state, and thus this state may be functioning as an intermediate in the folding pathway of erv C.  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