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.
The design, synthesis, and characterization of a folded high-affinity metal-binding peptide is described. Based on the previously described folded peptide NTH-18, in which an alpha-helix was constrained through two disulfide bonds to a C-terminal extension of noncanonical secondary structure, a peptide (1) was designed to contain two histidine residues in positions 3 and 7. Air oxidation of 1 led to the formation of peptide 2, which contained two intramolecular disulfide bonds. The presence of the two histidines significantly destabilized the alpha-helical structure of 2 when compared to NTH-18. However, CD spectroscopy revealed that the addition of certain transition metal ions allowed the reformation of a stable alpha-helix. CD, NMR, and EPR spectroscopy as well as MALDI-TOF mass spectrometry indicated that 2 bound to Cu2+ to form a 1:1 complex via the imidazoles of the two histidine side chains. A glycine displacement assay revealed a dissociation constant for this complex of 5 nM at pH 8, which is the lowest reported value for a designed Cu2+-binding peptide. This peptide displayed more than 100-fold selectivity for Cu2+ over Zn2+, Ni2+, and Co2+. The 1.05 A crystal structure of the Cu(II)-complex of 2 revealed a square-pyramidal coordination geometry and confirmed that 2 bound to copper in an alpha-helical conformation via its two histidine side chains. The high affinity metal binding of peptide 2 demonstrates that metals can be used for the selective nucleation of alpha-helices.  相似文献   

2.
MS/MS is indispensable for the amino acid sequencing of peptides. However, its use is limited for peptides containing disulfide bonds. We have applied the reducing properties of 1,5-diaminonaphthalene (1,5-DAN) as a MALDI matrix to amino acid sequencing and disulfide bond mapping of human urotensin II possessing one disulfide bond, and human guanylin possessing two disulfide bonds. 1,5-DAN was used in the same manner as the usual MALDI matrices without any pre-treatment of the peptide, and MS/MS was performed using a matrix-assisted laser desorption/ionization quadrupole ion trap time-of-flight mass spectrometer (MALDI QIT TOFMS). The results demonstrated that MS/MS of the molecular ions reduced by 1,5-DAN provided a series of significant b-/y-product ions. All 11 amino acid residues of urotensin II were identified using 1,5-DAN, while only 5 out of 11 residues were identified using 2,5-dihydroxybenzoic acid (DHB); similarly 11 out of 15 amino acid residues of guanylin were identified using 1,5-DAN, while only three were identified using DHB. In addition, comparison of the theoretical and measured values of the mass differences between corresponding MS/MS product ions using 1,5-DAN and DHB narrowed down the possible disulfide bond arrangement candidates. Consequently, 1,5-DAN as a reductive matrix facilitates rapid amino acid sequencing and disulfide mapping for peptides containing disulfide bonds.  相似文献   

3.
The potentiometric and spectroscopic (EPR, UV-Vis, CD) data have shown that the chicken prion hexa-repeat (Ac-His-Asn-Pro-Gly-Tyr-Pro-NH(2)) is a very specific ligand for Cu(2+) ions. The His imidazole is an anchoring binding site, then the adjacent amide nitrogen coordinates as a second donor. The presence of Pro at position 3 induces binding of phenolate oxygen as a third donor atom. The tridentate coordination dominates around physiological pH. Similar to human octapeptide fragments, chicken tandem repeats exhibit a cooperative effect in binding Cu(2+) ions, although chicken peptides are much less effective in metal ion coordination.  相似文献   

4.
Cd2+和Ni2+在粉煤灰上的吸附特性   总被引:2,自引:1,他引:1  
考察了粉煤灰对Cd2+和Ni2+的单组分吸附和双组分吸附性能。结果表明,粉煤灰可有效吸附水溶液中的Cd2+和Ni2+,去除率随pH升高而增加。吸附约60min后趋于平衡。粉煤灰对Ni2+的吸附容量高于Cd2+。单组分吸附平衡符合Freundlich模型和Redlich Peterson (R P)模型。双组分吸附时,Ni2+和Cd2+之间存在明显的竞争吸附效应;随干扰离子浓度升高,竞争吸附效应增强。不同模型拟合结果表明,双组分吸附平衡符合Freundlich竞争吸附模型。脱附实验表明,Cd2+比Ni2+易于脱附;0.1mol/L HCl、0.1mol/L HNO3 和0.05mol/L H2SO4的脱附效果接近,对Cd2+脱附率>60%,对Ni2+脱附率>35%。  相似文献   

5.
Mass spectra of the volatile derivatives of short peptides were studied by the photoionisation method with the use of monochromatic photons. The dependence of the intensity of the main peaks on the photon energy was analysed from 7·5 to 13·0 eV. The data obtain reveal the influence of the chemical structure of amino acid residues on the relative probability of the decomposition of peptide molecular ions at the CH? CO and CO? NH bonds, resulting in the formation of positively charged aldimine and amino acid N-terminal fragments, respectively. These data may be used as a basis for the application of the photoionisation technique to mass spectrometric sequential studies in peptides. In peptides containing residues of aliphatic amino acid the decomposition results mainly in the formation of aldimine ions, the stability of which increase with the increase of the alkyl chain size. In peptides containing residues of aromatic amino acids the decomposition is usually observed leading to formation of the amino acid ions. Ionisation potentials, as well as photoionisation efficiency curves and appearance potentials were determined for characteristic ions. Comparison was made of the values of the appearance potentials of different fragments formed upon decomposition of molecular ions. It has been shown that for peptides containing aliphatic amino acid moieties the appearance potentials of aldimine fragments are always lower than those inherent to peptides containing aromatic amino acid residues. The larger the size of an aliphatic chain, the lower is the energy of formation of these fragments. For all the compounds studied, including the peptides containing aromatic amino acid residues, the appearance potentials of the aldimine ions did not exceed those of the amino acid ions. These data indicate that, contrary to the experiments with electron-impact with energy of about 70 eV, upon ionisation with photons with energy from 7·5 to 13·0 eV, the aldimine fragments appear directly due to primary decomposition of molecular ions, independent of the formation of the amino acid fragments. The photoionisation efficiency curves for peptides containing different types of amino acid residues facilitate the choice of an optimal photon energy providing unequivocal information on the amino acid sequence in the peptide under study.  相似文献   

6.
A theoretical method for studying the local lattice structure of Ni2+ ions in (NiF6)(4-) coordination complex is presented. Using the ligand-field model, the formulas relating the microscopic spin Hamiltonian parameters with the crystal structure parameters are derived. Based on the theoretical formulas, the 45 x 45 complete energy matrices for d8 (d2) configuration ions in a tetragonal ligand-field are constructed. By diagonalizing the complete energy matrices, the local distortion structure parameters (R perpendicular and R || ) of Ni2+ ions in K2ZnF4:Ni2+ system have been investigated. The theoretical results are accorded well with the experimental values. Moreover, to understand the detailed physical and chemical properties of the fluoroperovskite crystals, the theoretical values of the g factor of K2ZnF4:Ni2+ system at 78 and 290 K are reported first.  相似文献   

7.
铜转运蛋白(CTR1)不仅参与铜的细胞摄取,而且在其它重金属离子的摄取过程中也发挥重要作用. 本文采用紫外-可见(UV-Vis)光谱,核磁共振(NMR)和质谱(MS)的方法,研究了人源CTR1 (hCTR1)的C端金属结合域(C8)与Ag+和Hg2+的相互作用. 研究表明,Ag+和Hg2+都能与C8结合,但二者与C8的结合机制明显不同. 每个C8分子可以结合两个Ag+离子,但一个Hg2+却可以与两个C8形成桥联. 此外,Ag+离子与C8的配位是一个中等速度的交换过程,而Hg2+离子则为快速交换过程. C8的半胱氨酸残基是两种离子的重要结合位点,同时组氨酸残基也参与两种金属离子的配位,其中Ag+优先结合组氨酸,而Hg2+则对半胱氨酸的结合具有显著的优势. 虽然HCH基序对C8 与金属配位至关重要,一些远端的其它氨基酸也可以参与C8 与银离子的配位,这可能与CTR1 在摄取Ag+过程中的金属转移机制相关. 这些结果为理解hCTR1 蛋白摄取重金属离子的作用机制提供了必要的信息.  相似文献   

8.
Matrix-assisted laser desorption/ionization (MALDI) mass spectrometry was used to study peptide-peptide interaction. The interaction was seen when 6-aza-2-thiothymine was used as a matrix (pH 5.4), but was disrupted with a more acidic matrix, alpha-cyano-4-hydroxycinnamic acid (pH 2.0). In the present study, we show that dynorphin, an opioid peptide, and five of its fragments that contain two adjacent basic residues (Arg6-Arg7), all interact noncovalently with peptides that contain two to five adjacent acidic residues (Asp or Glu). Two other nonrelated peptides containing two (Arg6-Arg7) or three (Arg1-Lys2-Arg3) adjacent basic amino acid residues were studied and exhibited the same behavior. However, peptides containing adjacent Lys or His did not form noncovalent complexes with acidic peptides. The noncovalent bonding was sufficiently stable that digestion with trypsin only cleaved Arg and Lys residues that were not involved in hydrogen bonding with the acidic residues. In an equimolar mixture of dynorphin, dynorphin fragments (containing the motif RR), and an acidic peptide (minigastrin), the acidic peptide preferentially complexed with dynorphin. If the concentration of minigastrin was increased 10 fold, noncovalent interaction was seen with dynorphin and all its fragments containing the motif RR. In the absence of dynorphin, minigastrin formed noncovalent complexes with all dynorphin fragments. These findings suggest that conformation, equilibrium, and concentration do play a role in the occurrence of peptide-peptide interaction. Observations from this study include: (1) ionic bonds were not disrupted by enzymatic digests, (2) conformation and concentration influenced complex formation, and (3) the complex did not form with fragments of dynorphin or unrelated peptides that did not contain the motifs RR or RKR, nor with a fragment of dynorphin where Arg7 was mutated to a phenylalanine residue. These findings strongly suggest that peptide-peptide interaction does occur, and can be studied by MALDI if near physiologic pH is maintained.  相似文献   

9.
The synthesis, protonation behavior, and Cu2+ and Zn2+ coordination chemistry of the novel bibrachial aza lariat ether (naphthalen-1-ylmethyl)[2-(20-[2-[(naphthalen-1-ylmethyl)amino]ethyl]-3,6,9,17,20,23,29,30-octaazatricyclo[23.3.1.1*11,15*]triaconta-1(29),11(30),12,14,25,27-hexaen-6-yl)ethyl]amine (L) are discussed. The macrocycle, which has two aminoethyl naphthyl moieties symmetrically appended to a 2:2 azapyridinophane structure, displays, in the pH range 2-11, six protonation steps that correspond to the protonation of the secondary amino groups. Steady-state fluorescence measurements show emissions due to the monomer and to the excimer formed between the two naphthalene fragments of the macrocycle. The time-resolved fluorescence data, obtained by the time-correlated single photon counting technique, show that a significant percentage of excimer is preformed as ground-state dimers. The ligand L forms with the metal ions Cu2+ and Zn2+ mono- and dinuclear complexes in aqueous solution. The influence of metal coordination in the fluorescence emission of L is analyzed. The acid-base, coordination capabilities, and emissive behavior of L are compared with those presented by its synthetic precursor L1, which has a tripodal tris(2-aminoethyl)amine structure functionalized at one of its terminal amino groups with a naphthyl moiety.  相似文献   

10.
Multiply protonated ions of disulfide-intact and -reduced peptides were generated by electrospray ionization and studied by Fourier transform ion cyclotron resonance mass spectrometry. The effects of disulfide bonds on gas-phase deprotonation reactions and hydrogen/deuterium (H/D) exchange were investigated. Insight into conformations was gained from molecular dynamics calculations. For ions from three small peptides containing 9–14 amino acid residues, H/D exchange is more sensitive to changes in conformation than deprotonation. However, with both gas-phase reactions the more diffuse forms of the peptides (as determined by molecular modeling) react more readily. The effects of disulfide cleavage on the conformations and on the reactions were found to depend upon the sequence of the peptide. For [M + 3H]3+ of TGF-α (34–43), reduction of the disulfide linkage leads to a greatly extended structure and a dramatic increase in the rate and extent of H/D exchange. In contrast, [M + 2H]2+ of Arg8 -vasopressin becomes slightly more compact upon cleavage of the disulfide bond; these reduced ions are slower to react. For [M + 3H]3+ of somatostatin-14, reduction of the disulfide bond has little effect on conformation or gas-phase reactivity. Overall, these results indicate that there is no general rule on how cleavage of a disulfide bond will effect a peptide ion’s gas-phase reactivity.  相似文献   

11.
Polyamic acids containing biquinolyl (biQ) fragments in the polymer backbone are capable of coordinating with NiII ions to form two types of coordination units: with one or two biQ ligands in the Ni coordination environment. The relative number of [Ni(biQ)2]2+ coordination units in the polymer chain increases with an increase in the concentration of the biQ fragments in solution and with a decrease in the conformational rigidity of the polymer. The synthesis of the nickel complexes with the polymer ligands can be carried out using both Ni salts and the electrochemical dissolution of the Ni anode.  相似文献   

12.
The widespread occurrence of the neutral loss of one to six amino acid residues as neutral fragments from doubly protonated tryptic peptides is documented for 23 peptides with individual sequences. Neutral loss of amino acids from the N-terminus of doubly charged tryptic peptides results in doubly charged y-ions, forming a ladder-like series with the ions [M + 2H](2+) = y(max) (2+), y(max - 1) (2+), y(max - 2) (2+), etc. An internal residue such as histidine, proline, lysine or arginine appears to favor this type of fragmentation, although it was sometimes also observed for peptides without this structure. For doubly protonated non-tryptic peptides with one of these residues at or near the N-terminus, we observed neutral loss from the C-terminus, resulting in a doubly charged b-type ion ladder. The analyses were performed by Q-TOF tandem mass spectrometry, facilitating the recognition of neutral loss ladders by their 2+ charge state and the conversion of the observed mass differences into reliable sequence information. It is shown that the neutral loss of amino acid residues requires low collision offset values, a simple mechanistic explanation based on established fragmentation rules is proposed and the utility of this neutral loss fragmentation pathway as an additional source for dependable peptide sequence information is documented.  相似文献   

13.
Peptides with biological functions often contain disulfide bridges connecting two cysteine residues. In an attempt to screen biological fluids for peptides containing cysteine residues, we have developed a sensitive and specific method to label cysteines selectively and detect the resulting molecular mass shift by differential mass spectrometry. First, reduction of disulfide bridges and carboxyamidomethylation of free thiols is adjusted to quantitatively achieve cysteine alkylation for complex peptide extracts. In a second step, matrix-assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-TOFMS) before and after chemical derivatization is performed, followed by differential analysis to determine shifted peaks; shifted peaks belong to cysteine-containing peptides, other peaks remain unchanged. The number of cysteines can then be determined by the resulting molecular mass shift. Free, reduced cysteines are shifted by 57 u, two oxidized cysteines involved in disulfide bridges (cystine) result in a shift to higher mass per disulfide bridge of 116 u. Disulfide bridges connecting different amino acid chains like insulin break up during reduction. In this case, two peaks with lower molecular masses result from a single one in the unmodified sample. With this technique, we were able to identify cysteine-containing peptides and short fragments of proteins present in human blood filtrate.  相似文献   

14.
The cation-pi interaction, a noncovalent interaction of electrostatic nature between a cation and an electron-rich pi system, is increasingly recognized as an important force that influences the structures and functions of molecules including proteins. Unlike other metal cations, the transition metal cation Cu2+ is not regarded to take part in a cation-pi interaction because Cu2+ tends to oxidize the pi electron system, in particular that of Trp, and to introduce covalency in the metal-pi electron interaction. This paper reports the first spectral evidence for the cation-pi interaction between Cu2+ and Trp. The Cu2+ ion bound to the amino N-terminal Cu2+/Ni2+ binding motif composed of three amino acid residues interacts with the indole ring of the fourth Trp residue in a noncovalent manner. The Cu2+-Trp interaction produces a distinct negative band at 223 nm in circular dichroism (CD), which disappears upon mutation or depletion of the Trp residue or upon replacement of the Cu2+ ion by Ni2+. In UV absorption, a pair of negative/positive intensity changes is generated at 222/231 nm by the Cu2+-Trp interaction, being consistent with the previous observations on the indole ring interacting with K+ or a cationic His imidazole ring. The negative CD band around 223 nm is characteristic of the Cu2+-Trp pair and may be useful as a marker of the Cu2+-Trp cation-pi interaction. Coordination of negatively charged ligands to Cu2+ is suggested to be important for the cation to be involved in a cation-pi interaction.  相似文献   

15.
We report two fluorophores with open coordination sites for specific sensing of Ni(2+) and Cu(2+) ions via a simple synthetic route. The fluorescence activity was completely quenched on coordination of the metal ions with the phenanthroline ring present in the fluorophore as is clearly evident from the photophysical studies.  相似文献   

16.
The complexes formed between chromium(III) and synthetic acidic peptides were studied by sustained off-resonance irradiation collision-induced dissociation (SORI-CID) in a Fourier transform ion-cyclotron resonance (FT-ICR) mass spectrometer equipped with electrospray ionization (ESI). Neutral peptides and peptides containing one, two, and multiple acidic residues were studied. Formation of [M + Cr-2H]+ occurred for all peptides. Three noteworthy features were found in the CID spectra of [M + Cr-2H]+. The first is that fewer fragment ions were produced from [M + Cr-2H]+ than from [M + H]+. The reason may be that multiple coordination between chromium(III) and carboxylate or carbonyl groups hinders the production of fragment ions by continuing to bind pieces of the peptide to chromium(III) after cleavage of bonds within the peptide. The second feature is loss of CO from [M + Cr-2H]+ and [y(n) + Cr-H]+. A mechanism involving coordination of chromium(III) with carboxylate groups is proposed to rationalize elimination of CO. The third feature is that chromium(III) is retained in all fragment ions, indicating strong binding of the metal ion to the peptides. The complex [M + 2Cr-5H]+ is formed as the peptide chain length and number of acidic residues increases. Longer peptides have more sites to coordinate with chromium(III) and more conformational flexibility. In addition, formation of [M + Cr-2H]+ from AGGAAAA-OCH(3), which has no carboxylic acid groups, suggests that chromium(III) can coordinate with sites on the peptide backbone, albeit in low abundance. In the negative mode, [M + Cr-4H](-) was only found for peptides containing four or more carboxylic acid groups. This is consistent with deprotonated carboxylic acid groups being involved in chromium(III) coordination and with chromium existing in the 3 + state in the gas-phase ions.  相似文献   

17.

A novel fragmentation of metastable peptide [M + H]+ ions is described. Loss of the C-terminal amino acid residue is accomqanied by retention of one of the carboxyl oxygens, as judged by 18O-labeling. The retained 8O label is located at the new C-terminus. Sequential mass spectrometric analyses indicate that the structure of the first-generation product ion is indistinguishable from that of the [M + H]+ ion of the peptide with one fewer amino acid residues. Thus, for example, the metastable decompositions of ions of m/z 904 are similar whether they correspond to des-Arg9-bradykinin [M + H]+ ions or to fragments derived from bradykinin [M + H]+ ions. No corresponding rearrangements have been observed for peptides with C-terminal amide or ester functions. The mechanism of this fragmentation may be considered to be analogous to that previously suggested for fragmentations of [M + alkali metal cation]+ ions. For the examples of bradykinin and related peptides, the rearrangement is strongly promoted when arginine is the amino acid residue lost. The same fragmentation is also favored by the presence of an arginine residue at or near the N-terminus. The strong influence of peptide amino acid composition, including residues remote from the C-terminus, on the prevalence of this fragmentation suggests mechanistic complexities that require further elucidation.

  相似文献   

18.
The possible use of divalent alkaline-earth metal ions, including Mg2+, Ca2+, Sr2+, and Ba2+, as charge carrier for electron capture dissociation of peptides was investigated. Model peptides of RGGGVGGGR and NGGGWGGGN were used to simplify the interpretation of spectral information. It was demonstrated that useful electron capture dissociation (ECD) tandem mass spectra of these metalated peptides could be generated. Interestingly, peptides metalated with different alkaline-earth metal ions generated very similar ECD tandem mass spectra. Metalated c-ions and z-ions were the predominant fragment ions. Only Mg2+-metalated peptides gave somewhat different results. Some nonmetalated c-ions were observed from ECD of [RGGGVGGGR + Mg]2+ but not from [NGGGWGGGN + Mg]2+. Together with some ab initio calculations, it was established that the bound metal ions might activate the acidity of the amide hydrogen. With the presence of high proton affinity moiety, such as N-terminal amino group and/or side chain of the arginine residues, the metalated peptide ions could exist predominantly in their zwitterion forms, in which one or two backbone amide group(s) was deprotonated and the high proton affinity functional group(s) was protonated. It was believed that electron capture leads primarily to the reduction of the mobile proton rather than the metal ions. With this zwitterion model, the formation of nonmetalated c-fragments and the generation of similar ECD spectra for peptides metalated with various alkaline-earth metal ions could readily to be explained. Another interesting observation in the ECD mass spectra of metalated peptides is related to the enhanced formation of the minor ECD products, i.e., (c - 1)(+*) and (z + 1)+ ions. Together with ab initio calculations using a truncated peptide model, various possible reaction mechanisms for the formation of these minor ECD products were evaluated. It was concluded that hydrogen transfer between the initiated formed c and z(.) species plays an important role in the formation (c - 1)(+*) and (z + 1)+ ions. Although peptides metalated with these metal ions do not have better ECD efficiency compared to the multiply-protonated peptides, it provides practical accessibility of ECD methods to analyze small peptides with no basic amino acid residues.  相似文献   

19.
The prion protein (PrP(C)) is a copper binding cell surface glycoprotein which when misfolded causes transmissible spongiform encephalopathies. The cooperative binding of Cu2+ to an unstructured octarepeat sequence within PrP(C) causes profound folding of this region. The use of NMR to determine the solution structure of the octarepeat region of PrP with Cu2+ bound has been hampered by the paramagnetic nature of the Cu2+ ions. Using NMR we have investigated the binding of candidate diamagnetic replacement ions, to the octarepeat region of PrP. We show that Pd2+ forms diamagnetic complexes with the peptides HGGG, HGGGW and QPHGGGWGQ with 1:1 stoichiometry. The 1H NMR spectra indicate that these peptides are in slow-exchange between free and bound Pd2+ on the chemical-shift time-scale. We demonstrate that the Pd-peptide complex forms slowly with a time taken to reach half-maximal signal of 3 hours. Other candidate metal ions, Ni2+, Pt2+ and Au3+, were investigated but only the Pd2+ complexes gave resolvable 1H NMR spectra. We have determined the solution structure of the QPHGGGWGQ-Pd 1:1 complex using 71 NOE distance restraints. A backbone RMSD of 0.30 A was observed over residues 3 to 7 in the final ensemble. The co-ordinating ligands consist of the histidine imidazole side chain N epsilon, the amide N of the second and third glycines with possibly H2O as the fourth ligand. The co-ordination geometry differs markedly from that of the HGGGW-Cu crystal structure. This survey of potential replacement metal ions to Cu2+ provides insight into the metal specificity and co-ordination chemistry of the metal bound octarepeats.  相似文献   

20.
Abstract

Besides being of interest in photochemistry, photoinduced electron transfer (PET) is a process largely used in the design of fluorescent ion sensing molecules. One of the simplest systems is based on fluorescent aromatic groups linked to amino groups and proposed as possible fluorescent transition metal ion chemosensor [1]. In this case, the fluorescence of the fluorophore “ligths on” when the amino group is complexed. On the other hand, in the absence of metal ions, the fluorescence is quenched by a PET originating from the nitrogen lone electron pairs [2]. We prepared a new fluorescent chemosensor, abbreviated as Ant-NH-O-O-NH-Ant (shown in Fig. 1) in which the intramolecular PET is expected to be efficient. The chemosensor consists of a metal-binding dioxodiamino unit linked to two light-emitting anthracene fragments. This type of supramolecules when irradiated in methanol solution (conc. 1.89—10?5 M.) at 368 nm displays a characteristic fluorescence spectrum for anthracene group with the most intensive band at 415 nm [Fig. 2(a)]. The emission is slightly enhanced upon coordination of such metal ions as Ni2+ and Zn2+ by the ligand fragment of the Ant-NH-O-O-NH-Ant molecule [Fig. 2(b) and (d)]. However, much higher intensity of emission is observed in the case of Cu2+ complex [see Fig 2(c)]. The fluorescence enhancement is presumably due to suppression of photoinduced fluorophore-to-metal electron-transfer mechanism.  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