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.
Ribosome recycling factor (RRF), in concert with elongation factor EF-G, is required for disassembly of the posttermination complex of the ribosome after release of polypeptides. The crystal structure of Thermus thermophilus RRF was determined at 2.6 A resolution. It is a tRNA-like L-shaped molecule consisting of two domains: a long three-helix bundle (domain 1) and a three-layer beta/alpha/beta sandwich (domain 2). Although the individual domain structures are similar to those of Thermotoga maritima RRF (Selmer et al., Science, 1999, 286:2349-2352), the interdomain angle differs by 33 degrees in two molecules, suggesting that the hinge between two domains is potentially flexible and responsive to different conditions of crystal packing. The hinge connects hydrophobic junctions of domains 1 and 2. The structure-based genetic analysis revealed the strong correlation between the hinge flexibility and the in vivo function of RRF. First, altering the hinge flexibility by making alanine or serine substitutions for large-size residues conserved at the hinge loop and nearby in domain 1 frequently gave rise to gain of function except a Pro residue conserved at the hinge loop. Second, the hinge defect resulting from a too relaxed hinge structure can be compensated for by secondary alterations in domain 1 that seem to increase the hydrophobic contact between domain 1 and the hinge loop. These results show that the hinge flexibility is vital for the function of RRF and that the steric interaction between the hinge loop and domains 1 and 2 restricts the interdomain angle and/or the hinge flexibility. These results indicate that RRF possesses an architectural difference from tRNA regardless of a resemblance to tRNA shape: RRF has a "gooseneck" elbow, whereas the tRNA elbow is rigid, and the direction of flex of RRF and tRNA is at a nearly right angle to each other. Moreover, surface electrostatic potentials of the two RRF proteins are dissimilar and do not mimic the surface potential of tRNA or EF-G. These properties will add a new insight into RRF, suggesting that RRF is more than a simple tRNA mimic.  相似文献   

2.
Cooperative unfolding of Escherichia coli ribosome recycling factor (RRF) and its implication for function were investigated by comparing the in vitro unfolding and the in vivo activity of wild-type E. coli RRF and its temperature-sensitive mutant RRF(V117D). The experiments show that mutation V117D at domain I could perturb the domain II structure as evidenced in the near-UV CD and tyrosine fluorescence spectra though no significant globular conformation change occurred. Both equilibrium unfolding induced by heat or denaturant and kinetic unfolding induced by denaturant obey the two-state transition model, indicating V117D mutation does not perturb the efficient interdomain interaction, which results in cooperative unfolding of the RRF protein. However, the mutation significantly destabilizes the E. coli RRF protein, moving the thermal unfolding transition temperature range from 50-65 to 35-50 degrees C, which spans the non-permissive temperature for the growth of E. coli LJ14 strain (frr(ts)). The in vivo activity assays showed that although V117D mutation results in a temperature sensitive phenotype of E. coli LJ14 strain (frr(ts)), over-expression of mutant RRF(V117D) can eliminate the temperature sensitive phenotype at the non-permissive temperature (42 degrees C). Taking all the results into consideration, it can be suggested that the mechanism of the temperature sensitive phenotype of the E. coli LJ14 cells is due to inactivation of mutant RRF(V117D) caused by unfolding at the non-permissive temperatures.  相似文献   

3.
To investigate the ability of a protein to accommodate potentially destabilizing amino acid substitutions, and also to investigate the steric requirements for catalysis, proline was substituted at different sites within the long alpha-helix that connects the amino-terminal and carboxyl-terminal domains of T4 lysozyme. Of the four substitutions attempted, three yielded folded, functional proteins. The catalytic activities of these three mutant proteins (Q69P, D72P, and A74P) were 60-90% that of wild-type. Their melting temperatures were 7-12 degrees C less than that of wild-type at pH 6.5. Mutant D72P formed crystals isomorphous with wild-type allowing the structure to be determined at high resolution. In the crystal structure of wild-type lysozyme the interdomain alpha-helix has an overall bend angle of 8.5 degrees. In the mutant structure the introduction of the proline causes this bend angle to increase to 14 degrees and also causes a corresponding rotation of 5.5 degrees of carboxyl-terminal domain relative to the amino-terminal one. Except for the immediate location of the proline substitution there is very little change in the geometry of the interdomain alpha-helix. The results support the view that protein structures are adaptable and can compensate for potentially destabilizing amino acid substitutions. The results also suggest that the precise shape of the active site cleft of T4 lysozyme is not critical for catalysis.  相似文献   

4.
Ribosome recycling factor, referred to as RRF, is essential for bacterial growth because of its activity of decomposition of the post-termination complex of the ribosome after release of polypeptides. In this study, we isolated a conditionally lethal amber mutation, named frr-3, in the Escherichia coli RRF gene at amino acid position 161, showing that the truncation of the C-terminal 25 amino acids of RRF is lethal to E. coli. An RRF gene cloned from Thermus thermophilus, whose protein is 44% identical and 68% similar to E. coli RRF, failed to complement the frr-3(Am) allele. However, truncation of the C-terminal five amino acids conferred intergeneric complementation activity on T. thermophilus RRF, demonstrating the modulator activity of the C-terminal tail. Rapid purification of T. thermophilus RRF was achieved by T7-RNA polymerase-driven overexpression for crystallography.  相似文献   

5.
12/15-Lipoxygenases (12/15-LOXs) have been implicated in inflammatory and hyperproliferative diseases but the structural biology of these enzymes is not well developed. Most LOXs constitute single polypeptide chain proteins that fold into a two-domain structure. In the crystal structure the two domains are tightly associated, but small angle X-ray scattering data and dynamic fluorescence studies suggested a high degree of structural flexibility involving movement of the N-terminal domain relative to catalytic subunit. When we inspected the interdomain interface we have found a limited number of side-chain contacts which are involved in interactions of these two structural subunits. One of such contact points involves tyrosine 98 of N-terminal domain. This aromatic amino acid is invariant in vertebrate LOXs regardless of overall sequence identity. To explore in more detail the role of aromatic interactions in interdomain association we have mutated Y98 to various residues and quantified the structural and functional consequences of these alterations. We have found that loss of an aromatic moiety at position 98 impaired the catalytic activity and membrane binding capacity of the mutant enzymes. Although CD and fluorescence emission spectra of wild-type and mutant enzyme species were indistinguishable, the mutation led to enlargement of the molecular shape of the enzyme as detected by analytic gel filtration and this structural alteration was shown to be associated with a loss of protein thermal stability. The possible role of tight interdomain association for the enzyme's structural performance is discussed.  相似文献   

6.
Elongation factor G (EF-G) is a G protein with motor function that drives two target molecules, a tRNA in the translating ribosome and the ribosome recycling factor (RRF) in the post-termination complex. How G protein motor action is transmitted to RRF is unknown. Thermus thermophilus RRF is nonfunctional in Escherichia coli. It became functional upon introducing a plasmid expressing E. coli EF-G with surface changes in its tRNA-mimic domain or by replacing the E. coli EF-G tRNA-mimic domain by the Thermus domain. Thermus RRF could also be activated by introducing surface substitutions in its anticodon arm-mimic region. These gain-of-function phenotypes depend on the combination of heterologous EF-G and RRF alleles. These mutational studies suggest that EF-G motor action is transmitted to RRF by specific surface contacts between the domains that mimic the anticodon arm.  相似文献   

7.
Shin R  Tzou YM  Krishna NR 《Biochemistry》2011,50(44):9457-9467
The capsid protein (CA) of HIV-1 plays a significant role in the assembly of the immature virion and is the critical building block of its mature capsid. Thus, there has been significant interest in the CA protein as a target in the design of inhibitors of early and late stage events in the HIV-1 replication cycle. However, because of its inherent flexibility from the interdomain linker and the monomer-dimer equilibrium in solution, the HIV-1 wild-type CA monomer has defied structural determinations by X-ray crystallography and nuclear magnetic resonance spectroscopy. Here we report the detailed solution structure of full-length HIV-1 CA using a monomeric mutant that, though noninfective, preserves many of the critical properties of the wild-type protein. The structure shows independently folded N-terminal (NTD) and C-terminal domains (CTD) joined by a flexible linker. The CTD shows some differences from that of the dimeric wild-type CTD structures. This study provides insights into the molecular mechanism of the wild-type CA dimerization critical for capsid assembly. The monomeric mutant allows investigation of interactions of CA with human cellular proteins exploited by HIV-1, directly in solution without the complications associated with the monomer-dimer equilibrium of the wild-type protein. This structure also permits the design of inhibitors directed at a novel target, viz., interdomain flexibility, as well as inhibitors that target multiple interdomain interactions critical for assembly and interactions of CA with host cellular proteins that play significant roles within the replication cycle of HIV-1.  相似文献   

8.
Translation termination at UAG is influenced by the nature of the 5′ flanking codon inEscherichia coli. Readthrough of the stop codon is always higher in a strain with mutant (prfA1) as compared to wild-type (prfA+) release factor one (RF1). Isocodons, which differ in the last base and are decoded by the same tRNA species, affect termination at UAG differently in strains with mutant or wild-type RF1. No general preference of the last codon base to favour readthrough or termination can be found. The data suggest that RF1 is sensitive to the nature of the wobble base anticodon-codon interaction at the ribosomal peptidyl-tRNA binding site (P-site). For some isoaccepting P-site tRNAs (tRNA3ProversustRNA2Pro, tRNA4ThrversustRNA1,3Thr) the effect is different on mutant and wild-type RF1, suggesting an interaction between RF1 at the aminoacyl-tRNA acceptor site (A-site) and the P-site tRNA itself. The glycine codons GGA (tRNA2Gly) and GGG (tRNA2,3Gly) at the ribosomal P-site are associated with an almost threefold higher readthrough of UAG than any of the other 42 codons tested, including the glycine codons GGU/C, in a strain with wild-type RF1. This differential response to the glycine codons is lost in the strain with the mutant form of RF1 since readthrough is increased to a similar high level for all four glycine codons. High α-helix propensity of the last amino acid residue at the C-terminal end of the nascent peptide is correlated with an increased termination at UAG. The effect is stronger on mutant compared to wild-type RF1. The data suggest that RF1-mediated termination at UAG is sensitive to the nature of the codon-anticodon interaction of the wobble base, the last amino acid residue of the nascent peptide chain, and the tRNA at the ribosomal P-site.  相似文献   

9.
The mutM (fpg) gene, which encodes a DNA glycosylase that excises an oxidatively damaged form of guanine, was cloned from an extremely thermophilic bacterium, Thermus thermophilus HB8. Its nucleotide sequence encoded a 266 amino acid protein with a molecular mass of approximately 30 kDa. Its predicted amino acid sequence showed 42% identity with the Escherichia coli protein. The amino acid residues Cys, Asn, Gln and Met, known to be chemically unstable at high temperatures, were decreased in number in T.thermophilus MutM protein compared to those of the E.coli one, whereas the number of Pro residues, considered to increase protein stability, was increased. The T.thermophilus mutM gene complemented the mutability of the E.coli mutM mutY double mutant, suggesting that T. thermophilus MutM protein was active in E.coli. The T.thermophilus MutM protein was overproduced in E.coli and then purified to homogeneity. Size-exclusion chromatography indicated that T. thermophilus MutM protein exists as a more compact monomer than the E.coli MutM protein in solution. Circular dichroism measurements indicated that the alpha-helical content of the protein was approximately 30%. Thermus thermophilus MutM protein was stable up to 75 degrees C at neutral pH, and between pH 5 and 11 and in the presence of up to 4 M urea at 25 degrees C. Denaturation analysis of T.thermophilus MutM protein in the presence of urea suggested that the protein had at least two domains, with estimated stabilities of 8.6 and 16.2 kcal/mol-1, respectively. Thermus thermophilus MutM protein showed 8-oxoguanine DNA glycosylase activity in vitro at both low and high temperatures.  相似文献   

10.
Prokaryotic class I release factors (RFs) respond to mRNA stop codons and terminate protein synthesis. They interact with the ribosomal decoding site and the peptidyl-transferase centre bridging these 75 A distant ribosomal centres. For this an elongated RF conformation, with partially unfolded core domains II.III.IV is required, which contrasts the known compact RF crystal structures. The crystal structure of Thermus thermophilus RF2 was determined and compared with solution structure of T. thermophilus and Escherichia coli RF2 by microcalorimetry, circular dichroism spectroscopy and small angle X-ray scattering. The structure of T. thermophilus RF2 in solution at 20 degrees C is predominantly compact like the crystal structure. Thermodynamic analysis point to an initial melting of domain I, which is independent from the melting of the core. The core domains II.III.IV melt cooperatively at the respective physiological temperatures for T. thermophilus and E. coli. Thermodynamic analyses and the X-ray scattering results for T. thermophilus RF2 in solution suggest that the compact conformation of RF2 resembles a physiological state in absence of the ribosome.  相似文献   

11.
The crystal structure of Thermus thermophilus elongation factor G (EF-G) carrying the point mutation His573Ala was determined at a resolution of 2.8 A. The mutant has a more closed structure than that previously reported for wild-type EF-G. This is obtained by a 10 degrees rigid rotation of domains III, IV and V with regard to domains I and II. This rotation results in a displacement of the tip of domain IV by approximately 9 A. The structure of domain III is now fully visible and reveals the double split beta-alpha-beta motif also observed for EF-G domain V and for several ribosomal proteins. A large number of fusidic acid resistant mutations found in domain III have now been possible to locate. Possible locations for the effector loop and a possible binding site for fusidic acid are discussed in relation to some of the fusidic acid resistant mutations.  相似文献   

12.
Bauer F  Sticht H 《FEBS letters》2007,581(8):1555-1560
Loop flexibility is discussed as a factor that affects ligand binding affinity of SH3 domains. To test this hypothesis, we designed a mutant in which a proline in the RT-loop of the human Lck SH3-domain is replaced by glycine. The dynamics and ligand binding properties of wild-type and mutant LckSH3 were studied by fluorescence and NMR spectroscopy as well as molecular dynamics simulations. Although the mutated residue does not form direct contacts with the ligand, the mutation increases ligand affinity by a factor of eight. The mutant exhibits increased loop flexibility and enhanced sampling of binding-competent conformations. This effect is expected to facilitate ligand binding itself and might also allow formation of tighter contacts in the complex thus resulting in an increased binding affinity.  相似文献   

13.
Limited thermostability of antibiotic resistance markers has restricted genetic research in the field of extremely thermophilic Archaea and bacteria. In this study, we used directed evolution and selection in the thermophilic bacterium Thermus thermophilus HB27 to find thermostable variants of a bleomycin-binding protein from the mesophilic bacterium Streptoalloteichus hindustanus. In a single selection round, we identified eight clones bearing five types of double mutated genes that provided T. thermophilus transformants with bleomycin resistance at 77 degrees C, while the wild-type gene could only do so up to 65 degrees C. Only six different amino acid positions were altered, three of which were glycine residues. All variant proteins were produced in Escherichia coli and analyzed biochemically for thermal stability and functionality at high temperature. A synthetic mutant resistance gene with low GC content was designed that combined four substitutions. The encoded protein showed up to 17 degrees C increased thermostability and unfolded at 85 degrees C in the absence of bleomycin, whereas in its presence the protein unfolded at 100 degrees C. Despite these highly thermophilic properties, this mutant was still able to function normally at mesophilic temperatures in vivo. The mutant protein was co-crystallized with bleomycin, and the structure of the binary complex was determined to a resolution of 1.5 A. Detailed structural analysis revealed possible molecular mechanisms of thermostabilization and enhanced antibiotic binding, which included the introduction of an intersubunit hydrogen bond network, improved hydrophobic packing of surface indentations, reduction of loop flexibility, and alpha-helix stabilization. The potential applicability of the thermostable selection marker is discussed.  相似文献   

14.
Ribosome recycling factor (RRF) of Thermotoga maritima was expressed in Escherichia coli from the cloned T. maritima RRF gene and purified. Expression of T. maritima RRF inhibited growth of the E. coli host in a dose-dependent manner, an effect counteracted by the overexpression of E. coli RRF. T. maritima RRF also inhibited the E. coli RRF reaction in vitro. Genes encoding RRFs from Streptococcus pneumoniae and Helicobacter pylori have been cloned, and they also impair growth of E. coli, although the inhibitory effect of these RRFs was less pronounced than that of T. maritima RRF. The amino acid sequence at positions 57 to 62, 74 to 78, 118 to 122, 154 to 160, and 172 to 176 in T. maritima RRF differed totally from that of E. coli RRF. This suggests that these regions are important for the inhibitory effect of heterologous RRF. We further suggest that bending and stretching of the RRF molecule at the hinge between two domains may be critical for RRF activity and therefore responsible for T. maritima RRF inhibition of the E. coli RRF reaction.  相似文献   

15.
We improved the thermal stability of 3-isopropylmalate dehydrogenase from Bacillus subtilis by an in vivo evolutionary technique using an extreme thermophile, Thermus thermophilus, as a host cell. The leuB gene encoding B. subtilis 3-isopropylmalate dehydrogenase was integrated into the chromosome of a leuB-deficient strain of T. thermophilus. The resulting transformant showed a leucine-autotrophy at 56 degrees C but not at 61 degrees C and above. Phenotypically thermostabilized strains that can grow at 61 degrees C without leucine were isolated from spontaneous mutants. Screening temperature was stepwise increased from 61 to 66 and then to 70 degrees C and mutants that showed a leucine-autotrophic growth at 70 degrees C were obtained. DNA sequence analyses of the leuB genes from the mutant strains revealed three stepwise amino acid replacements, threonine-308 to isoleucine, isoleucine-95 to leucine, and methionine-292 to isoleucine. The mutant enzymes with these amino acid replacements were more stable against heat treatment than the wild-type enzyme. Furthermore, the triple-mutant enzyme showed significantly higher specific activity than that of the wild-type enzyme.  相似文献   

16.
We have performed random mutagenesis coupled with selection to isolate mutant enzymes with high catalytic activities at low temperature from thermophilic 3-isopropylmalate dehydrogenase (IPMDH) originally isolated from Thermus thermophilus. Five cold-adapted mutant IPMDHs with single-amino-acid substitutions were obtained and analyzed. Kinetic analysis revealed that there are two types of cold-adapted mutant IPMDH: k(cat)-improved (improved in k(cat)) and K(m)-improved (improved in k(cat)/K(m)) types. To determine the mechanisms of cold adaptation of these mutants, thermodynamic parameters were estimated and compared with those of the Escherichia coli wild-type IPMDH. The Delta G(m) values for Michaelis intermediate formation of the k(cat)-improved-type enzymes were larger than that of the T. thermophilus wild-type IPMDH and similar to that of the E. coli wild-type IPMDH. The Delta G(m) values of K(m)-improved-type enzymes were smaller than that of the T. thermophilus wild-type IPMDH. Fitting of NAD(+) binding was improved in the K(m)-improved-type enzymes. The two types of cold-adapted mutants employed one of the two strategies of E. coli wild-type IPMDH: relative destabilization of the Michaelis complex in k(cat)-improved-type, and destabilization of the rate-limiting step in K(m)-improved type mutants. Some cold-adapted mutant IPMDHs retained thermostability similar to that of the T. thermophilus wild-type IPMDH.  相似文献   

17.
Escherichia coli endonuclease VIII (Nei) excises oxidized pyrimidines from DNA. It shares significant sequence homology and similar mechanism with Fpg, a bacterial 8-oxoguanine glycosylase. The structure of a covalent Nei-DNA complex has been recently determined, revealing critical amino acid residues which are important for DNA binding and catalysis. Several Fpg structures have also been reported; however, analysis of structural dynamics of Fpg/Nei family proteins has been hindered by the lack of structures of uncomplexed and DNA-bound enzymes from the same source. We report a 2.8 A resolution structure of free wild-type Nei and two structures of its inactive mutants, Nei-E2A (2.3 A) and Nei-R252A (2.05 A). All three structures are virtually identical, demonstrating that the mutations did not affect the overall conformation of the protein in its free state. The structures show a significant conformational change compared with the Nei structure in its complex with DNA, reflecting a approximately 50 degrees rotation of the two main domains of the enzyme. Such interdomain flexibility has not been reported previously for any DNA glycosylase and may present the first evidence for a global DNA-induced conformational change in this class of enzymes. Several local but functionally relevant structural changes are also evident in other parts of the enzyme.  相似文献   

18.
We have studied the mobility of the multidomain folding catalyst, protein disulfide isomerase (PDI), by a coarse‐graining approach based on flexibility. We analyze our simulations of yeast PDI (yPDI) using measures of backbone movement, relative positions and orientations of domains, and distances between functional sites. We find that there is interdomain flexibility at every interdomain junction but these show very different characteristics. The extent of interdomain flexibility is such that yPDI's two active sites can approach much more closely than is found in crystal structures—and indeed hinge motion to bring these sites into proximity is the lowest energy normal mode of motion of the protein. The flexibility predicted for yPDI (based on one structure) includes the other known conformation of yPDI and is consistent with (i) the mobility observed experimentally for mammalian PDI and (ii) molecular dynamics. We also observe intradomain flexibility and clear differences between the domains in their propensity for internal motion. Our results suggest that PDI flexibility enables it to interact with many different partner molecules of widely different sizes and shapes, and highlights considerable similarities of yPDI and mammalian PDI. Proteins 2016; 84:1776–1785. © 2016 Wiley Periodicals, Inc.  相似文献   

19.
Aimes RT  Hemmer W  Taylor SS 《Biochemistry》2000,39(28):8325-8332
The glycine-rich loop, one of the most important motifs in the conserved protein kinase catalytic core, embraces the entire nucleotide, is very mobile, and is exquisitely sensitive to what occupies the active site cleft. Of the three conserved glycines [G(50)TG(52)SFG(55) in cAMP-dependent protein kinase (cAPK)], Gly(52) is the most important for catalysis because it allows the backbone amide of Ser(53) at the tip of the loop to hydrogen bond to the gamma-phosphate of ATP [Grant, B. D. et al. (1998) Biochemistry 37, 7708]. The structural model of the catalytic subunit:ATP:PKI((5)(-)(24)) (heat-stable protein kinase inhibitor) ternary complex in the closed conformation suggests that Ser(53) also might be essential for stabilization of the peptide substrate-enzyme complex via a hydrogen bond between the P-site carbonyl in PKI and the Ser(53) side-chain hydroxyl [Bossemeyer, D. et al. (1993) EMBO J. 12, 849]. To address the importance of the Ser(53) side chain in catalysis, inhibition, and P-site specificity, Ser(53) was replaced with threonine, glycine, and proline. Removal of the side chain (i.e., mutation to glycine) had no effect on the steady-state phosphorylation of a peptide substrate (LRRASLG) or on the interaction with physiological inhibitors, including the type-I and -II regulatory subunits and PKI. However, this mutation did affect the P-site specificity; the glycine mutant can more readily phosphorylate a P-site threonine in a peptide substrate (5-6-fold better than wild-type). The proline mutant is compromised catalytically with altered k(cat) and K(m) for both peptide and ATP and with altered sensitivity to both regulatory subunits and PKI. Steric constraints as well as restricted flexibility could account for these effects. These combined results demonstrate that while the backbone amide of Ser(53) may be required for efficient catalysis, the side chain is not.  相似文献   

20.
Based on its genome sequence, the pathway of beta-oxidative fatty acid degradation in Salmonella enterica serovar Typhimurium LT2 has been thought to be identical to the well-characterized Escherichia coli K-12 system. We report that wild-type strains of S. enterica grow on decanoic acid, whereas wild-type E. coli strains cannot. Mutant strains (carrying fadR) of both organisms in which the genes of fatty acid degradation (fad) are expressed constitutively are readily isolated. The S. enterica fadR strains grow more rapidly than the wild-type strains on decanoic acid and also grow well on octanoic and hexanoic acids (which do not support growth of wild-type strains). By contrast, E. coli fadR strains grow well on decanoic acid but grow only exceedingly slowly on octanoic acid and fail to grow at all on hexanoic acid. The two wild-type organisms also differed in the ability to grow on oleic acid when FadR was overexpressed. Under these superrepression conditions, E. coli failed to grow, whereas S. enterica grew well. Exchange of the wild-type fadR genes between the two organisms showed this to be a property of S. enterica rather than of the FadR proteins per se. This difference in growth was attributed to S. enterica having higher cytosolic levels of the inducing ligands, long-chain acyl coenzyme As (acyl-CoAs). The most striking results were the differences in the compositions of CoA metabolites of strains grown with octanoic acid or oleic acid. S. enterica cleanly converted all of the acid to acetyl-CoA, whereas E. coli accumulated high levels of intermediate-chain-length products. Exchange of homologous genes between the two organisms showed that the S. enterica FadE and FadBA enzymes were responsible for the greater efficiency of beta-oxidation relative to that of E. coli.  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