首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The pH dependence of amide proton exchange rates have been measured for trp-repressor. One class of protons exchanges too fast to be measured in these experiments. Among the protons that have measurable hydrogen-deuterium exchange rates, two additional classes may be distinguished. The second class of protons are in elements of secondary structure that are mostly on the surface of the protein, and exchange linearly with increasing base concentration (log kex versus pH). The third class of amide protons is characterized by much higher protection against exchange at higher pH. These protons are located in the core of the protein, in helices B and C. The exchange rate in the core region does not increase linearly with pH, but rather goes through a minimum around pH 6. The mechanism of exchange for the slowly exchanging core protons is interpreted in terms of the two-process model of Hilton and Woodward (1979, Biochemistry 18:5834-5841), i.e., exchange through both a local mechanism that does not require unfolding of the protein, and a mechanism involving global unfolding of the protein. The increase in exchange rates at low pH is attributed to a partial unfolding of the repressor. It is concluded that the formation of secondary structure alone is insufficient to account for the high protection factors seen in the core of native proteins at higher pH, and that tertiary interactions are essential to stabilize the structure.  相似文献   

2.
Acid-induced unfolding of proteins often results in an intermediate structure, called the molten globule structure or "A" state, which retains at least partial secondary structure but lacks a rigid tertiary structure. Acid-induced unfolding has been studied extensively for alpha-helical proteins, while few studies have been done on proteins containing only beta-strands. Tumor necrosis factor-alpha (TNF-alpha) is a trimer in which the individual subunits consist of antiparallel beta-sheet, organized into a jellyroll beta-sandwich. We have found previously [Narhi et al. (1996) Biochemistry 35, 11447-11453] that thermal denaturation of TNF-alpha results in an aggregate which contains a substantial amount of alpha-helix and that the addition of trifluoroethanol induces alpha-helix in both murine and human TNF-alpha. Here we show that acid also can induce alpha-helix in these proteins. At acidic pH (below 4), both human and murine TNF-alpha convert to a monomeric form, as determined by sedimentation and diffusion constants obtained from sedimentation velocity experiments. The sedimentation coefficient indicated that this monomer was only slightly expanded relative to the native state. Near-UV circular dichroic (CD) analysis showed a loss of tertiary structure. These structural features coincide with the notion that the acid-induced structure of TNF-alpha is a molten globule. What is unique in this protein is that TNF-alpha acquires alpha-helical structure, which is not present in the native structure as determined by both CD and Fourier transform infrared spectroscopy. Even more surprising is that TNF-alpha at pH 3.3 undergoes a very gradual noncooperative change in secondary structure upon heating, which results in an increase in alpha-helical content. At pH 2.2 in the absence of salt, TNF-alpha shows considerable alpha-helix, although heating does not change the spectrum. At pH 2.2, physiological salt decreases the amount of alpha-helix at ambient temperature, and upon heating, we see the noncooperative increase in alpha-helix as observed at pH 3.3 with low salt. The addition of salt at low pH induces reassociation but to a range of oligomers rather than a unique trimer structure. This acid-induced formation of an alpha-helical monomer of TNF-alpha may be related to its known interaction with lipid bilayers.  相似文献   

3.
Molecular dynamics simulations of alpha-lactalbumin were performed under conditions of neutral pH and low pH in order to study the acid-induced molten globule state. Through the use of experimental techniques such as NMR and CD spectroscopy, molten globules have been characterized as being compact intermediates with secondary structure similar to that of the native protein but with tertiary structure that is disordered. The detailed structure of the molten globule state is unknown, however. Through the use of computer simulations we can study the structural changes which occur upon lowering pH. The simulations presented here differ from previous unfolding simulations in two important ways: the electrostatic interactions are treated more accurately than ever before, and artificially high temperatures are not used to force the protein to unfold. Simulations of 880 psec each were run at pH 7 (control simulation) and pH 2. We concentrate on the interesting changes in the tertiary interactions within the protein with lowering of pH. In particular, there is a loss of native tertiary contacts in the beta domain and interdomain region, and a large decrease in interdomain hydrogen bonds.  相似文献   

4.
The thermostability of manganese stabilizing protein of photosystem II was examined by biochemical and spectroscopic techniques. Samples of both native and recombinant spinach manganese stabilizing protein incubated at 90 degreesC and then cooled to 25 degreesC were capable of rebinding to, and of reactivating, the O2-evolution activity of photosystem II membranes from which the native protein had been removed. Far-UV circular dichroism and FT-IR spectroscopies were used to analyze the structural consequences of heating manganese stabilizing protein. The data obtained from these techniques show that heating causes a complete loss of the protein's secondary structure, and that this is a reversible, noncooperative phenomenon. Upon cooling, the secondary structures of the heat-treated proteins return to a state similar to, but not identical with, that of the native, unheated controls. Restoration of a near-native tertiary structure is confirmed both by size-exclusion chromatography and by near-UV circular dichroism. The functional and structural thermostability of manganese stabilizing protein reported here, in conjunction with additional known properties of this protein (acidic pI, high random coil and turn content, anomalous hydrodynamic behavior), identifies manganese stabilizing protein as a natively unfolded protein [Weinreb et al. (1996) Biochemistry 35, 13709-13715]. Although these proteins lack amino acid sequence identity, their functional solution conformations under physiological conditions are said to be "natively unfolded". We suggest that, as with other members of this family of proteins, the natively unfolded structure of manganese stabilizing protein facilitates the highly effective protein-protein interactions that are necessary for its assembly into photosystem II.  相似文献   

5.
Pulmonary surfactant protein A (SP-A) is synthesized by type II cells and stored intracellularly in secretory granules (lamellar bodies) together with surfactant lipids and hydrophobic surfactant proteins B and C (SP-B and SP-C). We asked whether the progressive decrease in pH along the exocytic pathway could influence the secondary structure and lipid binding and aggregation properties of porcine SP-A. Conformational analysis from CD spectra of SP-A at various pH values indicated that the percentage of alpha-helix progressively decreased and that of beta-sheet increased as the pH was reduced. The protein underwent a marked self-aggregation at mildly acidic pH in the presence of Ca2+, conditions thought to resemble those existing in the trans-Golgi network. Protein aggregation was greater as the pH was reduced. We also found that both neutral and acidic vesicles either with or without SP-B or SP-C bound to SP-A at acidic pH as demonstrated by co-migration during centrifugation. However, the binding of acidic but not neutral vesicles to SP-A led to 1) a striking change in the CD spectra of the protein, which was interpreted as a decrease of the level of SP-A self-aggregation, and 2) a protection of the protein from endoproteinase Glu-C degradation at pH 4.5. SP-A massively aggregated acidic vesicles but poorly aggregated neutral vesicles at acidic pH. Aggregation of dipalmitoylphosphatidylcholine (DPPC) vesicles either with or without SP-B and/or SP-C strongly depended on pH, being progressively decreased as the pH was reduced and markedly increased when pH was shifted back to 7.0. At the pH of lamellar bodies, SP-A-induced aggregation of DPPC vesicles containing SP-B or a mixture of SP-B and SP-C was very low, although SP-A bound to these vesicles. These results indicate that 1) DPPC binding and DPPC aggregation are different phenomena that probably have different SP-A structural requirements and 2) aggregation of membranes induced by SP-A at acidic pH is critically dependent on the presence of acidic phospholipids, which affect protein structure, probably preventing the formation of large aggregates of protein.  相似文献   

6.
The structural characteristics of Azotobacter vinelandii apoflavodoxin II have been determined using multidimensional NMR spectroscopy. Apoflavodoxin has a stable, well-ordered core but its flavin binding region is flexible. The local stability of apoflavodoxin was probed using hydrogen/deuterium exchange measurements. The existence of an apoflavodoxin equilibrium folding intermediate is inferred from the non-coincidence of CD and fluorescence unfolding curves obtained for the guanidinium hydrochloride induced unfolding of apoflavodoxin. We suggest that the structured part of the putative intermediate is composed of the elements of secondary structure which have the slowest exchanging amide protons in the native protein. These elements are strands beta1, beta3, beta4 and beta5a and helices alpha4 and alpha5. We propose that it is a general feature of flavodoxins that the stable nucleus resides in the C-terminal part of these proteins. The results on flavodoxin are compared with those on two sequentially unrelated proteins sharing the flavodoxin-like fold: Che Y and cutinase. It is shown that the stable nucleus is found in different parts of the flavodoxin-like topology.  相似文献   

7.
The yeast nascent polypeptide-associated complex (NAC) is encoded by two genes, EGD1 and EGD2, and is associated with cytoplasmic ribosomes. Yeast mutants lacking NAC (Deltaegd2) are viable but suffer slight defects in the targeting of nascent polypeptides to several locations including the endoplasmic reticulum and mitochondria. If both NAC and Mft52p are missing from yeast cells, inefficient targeting of mitochondrial precursor proteins leads to defects in both mitochondrial function and morphology. We suggest that NAC provides a ribosomal environment for nascent mitochondrial targeting sequences to achieve secondary structure, thereby enhancing the efficiency of protein targeting.  相似文献   

8.
BACKGROUND:. The ribosomal protein L22 is one of five proteins necessary for the formation of an early folding intermediate of the 23S rRNA. L22 has been found on the cytoplasmic side of the 50S ribosomal subunit. It can also be labeled by an erythromycin derivative bound close to the peptidyl-transfer center at the interface side of the 50S subunit, and the amino acid sequence of an erythromycin-resistant mutant is known. Knowing the structure of the protein may resolve this apparent conflict regarding the location of L22 on the ribosome. RESULTS:. The structure of Thermus thermophilus L22 was solved using X-ray crystallography. L22 consists of a small alpha+beta domain and a protruding beta hairpin that is 30 A long. A large part of the surface area of the protein has the potential to be involved in interactions with rRNA. A structural similarity to other RNA-binding proteins is found, possibly indicating a common evolutionary origin. CONCLUSIONS:. The extensive surface area of L22 has the characteristics of an RNA-binding protein, consistent with its role in the folding of the 23S rRNA. The erythromycin-resistance conferring mutation is located in the protruding beta hairpin that is postulated to be important in L22-rRNA interactions. This region of the protein might be at the erythromycin-binding site close to the peptidyl transferase center, whereas the opposite end may be exposed to the cytoplasm.  相似文献   

9.
The success rates reported for secondary structural class prediction with different methods are contradictory. On one side, the problem of recognizing the secondary structural class of a protein knowing only its amino acid composition appears completely solved by simply applying jury decision with an elliptically scaled distance function. Chou and coworkers repeatedly (see Crit. Rev. Biochem. Mol. Biol. 30:275-349, 1995) published prediction accuracies near 100%. On the other hand, traditional secondary structure prediction techniques achieve success rates of about 70% for the secondary structural state per residue and about 75% for structural class only with extensive input information (full sequence of the query protein, its amino acid composition and length, multiple alignments with homologous sequences). In this article, we resolve the paradox and consider (1) the question of the secondary structural class definition, (2) the role of the representativity of the test set of protein tertiary structure for the current state of the Protein Data Bank (PDB); and (3) we estimate the real impact of amino acid composition on secondary structural class. We formulate three objective criteria for a reasonable definition of secondary structural classes and show that only the criterion of Nakashima et al. (J. Biochem. 99:153-162, 1986) complies with all of them. Only this definition matches the distribution of secondary structural content in representative PDB subsets, whereas other criteria leave many proteins (up to 65% of all PDB entries) simply unassigned. We review critically specialized secondary-structural class prediction methods, especially those of Chou and coworkers, which claim almost 100% accuracy using only amino acid composition, and resolve the paradox that these prediction accuracies are better than those from secondary structure predictions from multiple alignments. We show (i) that these techniques rely on a preselection of test sets which removes irregular proteins and other proteins without any class assignment (about 35% of all PDB entries); and (ii) that even for preselected representative test sets, the success rate drops to 60% and lower for a 4-type classification (alpha, beta, alpha + beta, alpha/beta). The prediction accuracies fall to about 50% if the secondary structural class definition of Nakashima et al. is applied and only few irregular proteins are preselected and removed from automatically generated, representative subsets of the PDB. We have applied two new vector decomposition methods for secondary structural content prediction from amino acid composition alone, with and without consideration of amino acid compositional coupling in the learning set of tertiary structures respectively, to the problem of class prediction and achieve about 60% correct assignment among four classes (alpha, beta, mixed, irregular) as well as single sequence-based secondary structure prediction methods like GORIII and COMBI. Our results demonstrate that 60% correctness is the upper limit for a 4-type class prediction from amino acid composition alone for an unknown query protein and that consideration of compositional coupling does not improve the prediction success. The prediction program SSCP offering secondary structural class assignment for query compositions and sequences has been made available as a World Wide Web and E-mail service.  相似文献   

10.
The structure of purified P-glycoprotein functionally reconstituted into liposomes was investigated by attenuated total reflection Fourier transform infrared spectroscopy. A quantitative evaluation of the secondary structure and a kinetic of 2H/H exchange of the P-glycoprotein were performed both in the presence and in the absence of MgATP, MgATP-verapamil, and MgADP. This approach was previously shown to be a useful tool to detect tertiary structure changes resulting from the interaction between a protein and its specific ligands, as established for the Neurospora crassa H+-ATPase. 2H/H exchange measurements provided evidence that a large fraction of the P-glycoprotein is poorly accessible to the aqueous medium. Addition of MgATP induced an increased accessibility to the solvent of a population of amino acids, while addition of MgATP-verapamil resulted in a subtraction of a part of the protein from access to the aqueous solvent. No significant changes were observed upon addition of MgADP or verapamil alone. The secondary structure of P-glycoprotein was not affected by addition of ligands. The variations observed in the 2H/H exchange rate when P-glycoprotein interacted with the above ligands therefore represented tertiary structure changes. Fluorescence quenching experiments confirmed that MgATP-induced changes are to be found in the tertiary structure of the enzyme.  相似文献   

11.
The plasma protein beta2-glycoprotein I (beta2-GPI) is a major target of autoantibodies in patients with the antiphospholipid syndrome. To understand the physiological function of beta2-GPI and its potential role in the pathophysiology of the antiphospholipid syndrome, the binding of beta2-GPI to phospholipid membranes was characterized. The interaction of beta2-GPI with unilamellar vesicles containing varying amounts of acidic phospholipids with phosphatidylcholine (PC) was measured at equilibrium via relative light scattering. Analysis of binding isotherms gave apparent Kd values ranging from approximately 5.0 to 0.5 microM over a range of 5-20 mol % anionic phospholipid. Inhibition of binding by increasing ionic strength and Ca2+ ions suggests that binding is primarily electrostatic. These data indicate that beta2-GPI binding to membranes with physiological anionic phospholipid content is relatively weak in comparison to plasma coagulation proteins, suggesting that beta2-GPI does not function as a physiological anticoagulant based on its phospholipid-binding properties.  相似文献   

12.
Protein L12, together with the P0/P1/P2 protein complex, forms the protein moiety of the GTPase domain in the eukaryotic ribosome. In Saccharomyces cerevisiae protein L12 is encoded by a duplicated gene, rpL12A and rpL12B. Inactivation of both copies has been performed and confirmed by Southern and Western analyses. The resulting strains are viable but grow very slowly. Growth rate is recovered upon transformation with an intact copy of the L12 gene. Ribosomes from the disrupted strain lack protein L12 but are able to carry out translation in vitro at about one fourth of the control rate. The L12-deficient ribosomes have also a defective stalk containing standard amounts of the 12-kDa acidic proteins P1beta and P2alpha, but proteins P1alpha and P2beta are drastically reduced. Moreover, the affinity of P0 is reduced in the defective ribosomes. Footprinting of the 26 S rRNA GTPase domain indicates that protein L12 protects in different extent residues G1235, G1242, A1262, A1270, and A1272 from chemical modification. The results in this report indicate that protein L12 is not essential for cell viability but has a relevant role in the structure and stability of the eukaryotic ribosomal stalk.  相似文献   

13.
Papain exhibits the characteristics of molten globule under acidic conditions as seen by circular dichroism, fluorescence and ANS binding. Between pH 2.0-2.5 the protein exhibits substantial secondary structure as indicated by far-UV CD spectrum but loses the persistent tertiary interactions of the native state. Enhanced binding of ANS to the state at pH 2.0 in relation to the native and unfolded states at neutral pH indicates a considerable exposure of aromatic side chains. Temperature and guanidine hydrochloride induced unfolding of papain in this state is noncooperative and the transition curves are biphasic in nature. As papain molecule consists of two domains, the results suggest that the domains unfold independently and sequentially.  相似文献   

14.
Acidophilic microorganisms grow optimally at pH values between 1-4. They have adapted to the acid condition by maintaining their cytoplasmic pH at a value close to neutrality. Hence, only those (macro)-molecules, which face the acid medium, have had to adapt to this extreme condition. Literature data show that several exoproteins from thermoacidophilic prokaryotes are characterized by a low charge density. It is proposed that this property contributes to the stability of these proteins both below and above the pKa-values of their glutamate and aspartate residues. As an example of an acidophilic protein, the alpha-amylase from the Gram-positive Alicyclobacillus acidocaldarius ATCC27009 was studied. The enzyme is thermoacidophilic, with optima of temperature and pH of 75 degrees C and pH 3, respectively. The nucleotide sequence of the cloned gene (8) indicates that the alpha-amylase belongs to a large family of starch-degrading enzymes with a characteristic catalytic (beta alpha)8-domain. Three essential and probably catalytic acidic residues have been conserved, suggesting that the acidophilic alpha-amylase degrades starch with essentially the same mechanism as do its neutrophilic relatives. Still, the acidophilic protein contains three exchanges in residues uniformally or almost uniformally conserved among all members of the enzyme family. In order to test whether these exchanges contribute to the acidic pH optimum, the alpha-amylase gene was expressed in Escherichia coli. Sonication of the enzyme-producing cells released alpha-amylase activity associated with a 140 kDa protein. The optima of temperature and pH for the protein produced in E. coli were similar to those of the native enzyme. Experiments are underway in which it is tested which residues contribute to the acid pH optimum of the alpha-amylase.  相似文献   

15.
The DNA methyltransferases, M.HhaI and M.TaqI, and catechol O-methyl-transferase (COMT) catalyze the transfer of a methyl group from the cofactor S-adenosyl-L-methionine (AdoMet) to carbon-5 of cytosine, to nitrogen-6 of adenine, and to a hydroxyl group of catechol, respectively. The catalytic domains of the bilobal proteins, M.HhaI and M.TaqI, and the entire single domain of COMT have similar folding with an alpha/beta structure containing a mixed central beta-sheet. The functional residues are located in equivalent regions at the carboxyl ends of the parallel beta-strands. The cofactor binding sites are almost identical and the essential catalytic amino acids coincide. The comparable protein folding and the existence of equivalent amino acids in similar secondary and tertiary positions indicate that many (if not all) AdoMet-dependent methyltransferases have a common catalytic domain structure. This permits tertiary structure prediction of other DNA, RNA, protein, and small-molecule AdoMet-dependent methyltransferases from their amino acid sequences.  相似文献   

16.
A flavodoxin from Azotobacter vinelandii is chosen as a model system to study the folding of alpha/beta doubly wound proteins. The guanidinium hydrochloride induced unfolding of apoflavodoxin is demonstrated to be reversible. Apoflavodoxin thus can fold in the absence of the FMN cofactor. The unfolding curves obtained for wild-type, C69A and C69S apoflavodoxin as monitored by circular dichroism and fluorescence spectroscopy do not coincide. Apoflavodoxin unfolding occurs therefore not via a simple two-state mechanism. The experimental data can be described by a three-state mechanism of apoflavodoxin equilibrium unfolding in which a relatively stable intermediate is involved. The intermediate species lacks the characteristic tertiary structure of native apoflavodoxin as deduced from fluorescence spectroscopy, but has significant secondary structure as inferred from circular dichroism spectroscopy. Both spectroscopic techniques show that thermally-induced unfolding of apoflavodoxin also proceeds through formation of a similar molten globule-like species. Thermal unfolding of apoflavodoxin is accompanied by anomalous circular dichroism characteristics: the negative ellipticity at 222 nM increases in the transition zone of unfolding. This effect is most likely attributable to changes in tertiary interactions of aromatic side chains upon protein unfolding. From the presented results and hydrogen/deuterium exchange data, a model for the equilibrium unfolding of apoflavodoxin is presented.  相似文献   

17.
Using the murine teratocarcinoma cell line F9 we investigated the influence of serum stimulation and cisplatin treatment on the p53, CK2, MDM2 levels. Both treatments led to an increase of p53, though with different kinetics; the other proteins investigated were not affected. We present direct evidence by immunoprecipitation for an association of protein kinase CK2 holoenzyme (alpha2beta2), p53, and the ribosomal protein L5. The results suggest complexes between the CK2 holoenzyme and p53 but also p53/CKbeta complexes. Furthermore we provide evidence for the existence of high molecular mass complexes of CK2 in vivo. This is the first evidence that, under physiological conditions, protein kinase CK2 does not exist solely as a heterotetramer, but predominantly in association with other proteins.  相似文献   

18.
Signaling across integrins is regulated by interaction of these receptors with cytoskeletal proteins and signaling molecules. To identify molecules interacting with the cytoplasmic domain of the beta3-integrin subunit (glycoprotein IIIa), a placental cDNA library was screened in the yeast two-hybrid system. Two identical clones coding for a 96-amino acid sequence were identified. This sequence was 100% identical to a sequence in skelemin, a protein identified previously in skeletal muscle. Skelemin is a member of a superfamily of cytoskeletal proteins that contain fibronectin-type III-like motifs and immunoglobulin C2-like motifs and that regulate the organization of myosin filaments in muscle. The amino acid residues in the isolated clones encompassed C2 motifs 4 and 5 of skelemin. A recombinant skelemin protein consisting of C2 motifs 3-7 interacted with beta1- and beta3-integrin cytoplasmic domains expressed as glutathione S-transferase (GST) fusion proteins, but not with GST-beta2-integrin cytoplasmic tail or GST alone. The skelemin-binding region was in the membrane proximal cytoplasmic domains of the integrins. Full-length skelemin interacted with integrin in intact cells as demonstrated by the colocalization of hemagglutinin-tagged skelemin in Chinese hamster ovary (CHO) cells containing alphaIIbbeta3-integrin and by the finding that microinjection of C2 motif 4 of skelemin into C2C12 mouse myoblast cells caused spread cells to round up. A skelemin-like protein was detected in CHO cells, endothelial cells, and platelets, and this protein colocalized with beta1- and beta3-integrins in CHO cells. This study suggests the presence of a skelemin-like protein in non-muscle cells and provides evidence that it may be involved in linking integrins to the cytoskeleton.  相似文献   

19.
Hen egg-white lysozyme dissolved in glycerol containing 1% water was studied by using CD and amide proton exchange monitored by two-dimensional 1H NMR. The far- and near-UV CD spectra of the protein showed that the secondary and tertiary structures of lysozyme in glycerol were similar to those in water. Thermal melting of lysozyme in glycerol followed by CD spectral changes indicated unfolding of the tertiary structure with a Tm of 76.0 +/- 0.2 degreesC and no appreciable loss of the secondary structure up to 85 degreesC. This is in contrast to the coincident denaturation of both tertiary and secondary structures with Tm values of 74.8 +/- 0.4 degreesC and 74.3 +/- 0.7 degreesC, respectively, under analogous conditions in water. Quenched amide proton exchange experiments revealed a greater structural protection of amide protons in glycerol than in water for a majority of the slowly exchanging protons. The results point to a highly ordered, native-like structure of lysozyme in glycerol, with the stability exceeding that in water.  相似文献   

20.
The conformational stability of Plasmodium falciparum triosephosphate isomerase (TIMWT) enzyme has been investigated in urea and guanidinium chloride (GdmCl) solutions using circular dichroism, fluorescence, and size-exclusion chromatography. The dimeric enzyme is remarkably stable in urea solutions. It retains considerable secondary, tertiary, and quaternary structure even in 8 M urea. In contrast, the unfolding transition is complete by 2.4 M GdmCl. Although the secondary as well as the tertiary interactions melt before the perturbation of the quaternary structure, these studies imply that the dissociation of the dimer into monomers ultimately leads to the collapse of the structure, suggesting that the interfacial interactions play a major role in determining multimeric protein stability. The Cm(urea)/Cm(GdmCl) ratio (where Cm is the concentration of the denaturant required at the transition midpoint) is unusually high for triosephosphate isomerase as compared to other monomeric and dimeric proteins. A disulfide cross-linked mutant protein (Y74C) engineered to form two disulfide cross-links across the interface (13-74') and (13'-74) is dramatically destablized in urea. The unfolding transition is complete by 6 M urea and involves a novel mechanism of dimer dissociation through intramolecular thiol-disulfide exchange.  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