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 375 毫秒
1.
A peptide peak at m/z 1634 in the mass spectrum of tryptically digested cytochrome c has been ambiguously assigned to either a peptide IFVQKCAQCHTVEK or a peptide CAQCHTVEK combined with a heme group (CAQCHTVEK + heme (Fe(III))). A comprehensive investigation was performed to clearly identify the origin of the peak. Tryptic digests of cytochrome c were analyzed by matrix‐assisted laser desorption/ionization time‐of‐flight mass spectrometry (MALDI‐TOF MS), liquid chromatography‐tandem MS (LC‐MS/MS), LC‐ultraviolet (LC‐UV), and MALDI Fourier transform‐ion cyclotron resonance (FT‐ICR) MS. The use of instruments with extremely high mass accuracy revealed the mass difference between the IFVQKCAQCHTVEK and the (CAQCHTVEK + heme (Fe(III))) ions. Fragmentation of the peptide associated with the unknown peak yielded a heme ion and other fragment ions originating from a (CAQCHTVEK + heme (Fe(III))) ion. Furthermore, an absorption peak at 395 nm confirmed the presence of a heme group in the unknown peptide. High mass accuracy analyses of MS and MS/MS spectra, in addition to three‐dimensional UV contour mapping, showed that the peak at m/z 1634 is due to a (CAQCHTVEK + heme (Fe(III))) ion and not from protonated IFVQKCAQCHTVEK. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

2.
《中国化学会会志》2017,64(3):282-288
Based on two published UV absorbing molecules 2‐(5′‐tert‐butyl‐2′‐hydroxyl)benzotriazole (Tinuvin‐PS ) and hydroxybenzophenone (HBP ), both having characteristic absorption capability, a refined UV absorber, 2‐(3′‐benzoyl‐2′‐hydroxy‐5′‐tert‐butylphenyl)benzotriazole, has been synthesized. This new UV absorber exhibits a wider absorption range and better thermal stability compared to either Tinuvin‐PS and HBP and a higher absorption efficiency in the UV ‐C range. Theoretical studies reveal that the UV ‐C absorption arises from the electron density transfer from the Tinuvin‐PS moiety to the HBP moiety and the enhanced absorption efficiency is ascribed to the presence of the carbonyl's π system.  相似文献   

3.
In this work the effect in secondary ion mass spectrometry (SIMS) of several frequently used matrix‐assisted laser desorption/ionisation (MALDI) matrices on the secondary ion intensities of low molecular weight (m/z 400–800) organic dyes and a pharmaceutical is tested. Matrix (10?1 M) and analyte (10?2 M) solutions were made in methanol. Mixtures with several concentration ratios were prepared from these solutions and spincoated on Si substrates prior to time‐of‐flight (TOF)‐SIMS analysis. In some cases the presence of the MALDI matrices caused a considerable increase in the positive secondary (protonated) molecular ion signals. Enhancements of a factor of 20 and more were recorded. Generally, of the matrices used, 2,5‐dihydroxybenzoic acid and 2,4,6‐trihydroxyacetophenone brought about the highest intensity increases. It was also shown that matrix‐enhanced (ME‐)SIMS is capable of lowering the detection limits for molecule ions. However, the enhancement effect is strongly influenced by the analyte/matrix combination and its concentration ratio. As a result, finding an optimal analyte/matrix mixture can be a very time‐consuming process. Mostly, the presence of the matrices causes changes in the relative ion intensities in the TOF‐S‐SIMS spectra. Compared to the spectra recorded from samples without matrices, only a few additional peaks, such as signals that originate directly from the applied matrix or adduct ions, are observed in the mass spectra. Sometimes molecule ions and some characteristic fragments at high m/z values, that cannot be recorded without matrix, do appear in the spectrum when a matrix is present. In the negative mode no enhancement effect is observed on applying the studied MALDI matrices. The results obtained from samples treated with MALDI matrices are also compared to SIMS results for the same samples after Ag and Au metallisation (MetA‐SIMS). For three of the four tested compounds Au MetA‐SIMS resulted in higher ion yields than ME‐SIMS. For both techniques possible mechanisms that can account for the enhancement effect are proposed. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

4.
Monosaccharide C‐glycoside ketones have been synthesized by aqueous‐based Knoevenagel condensation of isotopically labeled and unlabeled aldoses with cyclic diketones, 5,5‐dimethyl‐1,3‐cyclohexanedione (dimedone) and 1,3‐cyclohexanedione (1,3‐CHD). The reaction products and their corresponding acetylated analogs produce characteristic molecular adduct ions by matrix‐assisted laser desorption/ionization mass spectrometry (MALDI‐TOF MS). Analysis of the peracetylated C‐glycosides by electron ionization (EI) gas chromatography/mass spectrometry (GC/MS) revealed diagnostic fragment ions that have been used to deduce the EI fragmentation pathways and the structure of each C‐glycoside ketone. Characteristic gluco‐ and ribo‐specific ions were observed at m/z 350 and 278, respectively. Ions common to both carbohydrate fragmentation pathways were observed at m/z 193 and 169 for the dimedone‐C‐glycosides, and m/z 165 and 141 for the 1,3‐CHD‐C‐glycosides. Ions with m/z 169 and 141 retain the anomeric carbon (carbon‐1) of the original sugar, while m/z 193 and 165 are shown to retain carbons‐1, 2, and 3. Published in 2009 by John Wiley & Sons, Ltd.  相似文献   

5.
Matrix‐assisted laser desorption/ionization in‐source decay (MALDI‐ISD) has been used for characterization of a phosphorylated peptides and proteins because labile phosphate group is not lost during the MALDI‐ISD process. The conventional MALDI‐ISD is initiated by the hydrogen transfer from reducing matrix molecules to peptide backbone, leading to c′‐ and z′‐series ions. In contrast, when an oxidizing chemical 5‐nitrosalicylic acid (5‐NSA) is served as the MALDI‐ISD matrix, a‐ and x‐series ions are specifically generated by hydrogen abstraction from peptide backbone to matrix molecule. The 5‐NSA provides useful complementary information to the conventional MALDI‐ISD for the analysis of amino acid sequencing and site localization of phosphorylation in peptides. The MALDI‐ISD with reducing and oxidizing matrix could be a useful method for the de novo peptide sequencing. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

6.
Rhizoma Atractylodes Macrocephala (RAM) is an important traditional Chinese medicinal herb that is used for treatment of dyspepsia and anorexia. The active ingredients, atractylenolide I (AO‐I) and atractylenolide III (AO‐III), were identified by direct‐injection ion trap‐mass spectrometry (IT‐MS) for collecting MSn spectra. The major fragment ions of AO‐I and AO‐III were confirmed by MSn both in negative ion mode and in positive ion mode. The possible main cleavage pathway of fragment ions was studied. The determinations of AO‐I and AO‐III were accomplished by liquid chromatography (LC) with UV and MS. The analytes provided good signals corresponding to the protonated molecular ions [M + H]+ and product ions. The precursor ions and product ions for quantification of AO‐III and AO‐I were m/z 249 → 231 and m/z 233 → 215, respectively, using selected ion monitoring by LC‐IT‐MS. Two methods were evaluated for a number of validation characteristics (repeatability, limit of detection, calibration range, and recovery). MS provides a high selectivity and sensitivity for determination of AO‐III and AO‐I in positive mode. After optimization of the methods, separation, identification and quantification of the two components in RAM were comprehensively tested by HPLC with UV and MS. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

7.
A method to measure CsI cluster ions ((CsI)nCs+, (CsI)nI?) from CsI samples in matrix‐assisted laser desorption/ionization mass spectrometry (MALDI‐MS) was developed with a 2‐[(2E)‐3‐(4‐tert‐butylphenyl)‐2‐methylprop‐2‐enylidene] malononitrile (DCTB) matrix and additives. Solvent combinations in which the CsI and DCTB solutions were miscible were effective in detecting CsI cluster ions at a mass range of over m/z 2000 and are associated with a characteristic spread of DCTB within the CsI/DCTB mixture. The addition of saccharides or sugar alcohols to the CsI/DCTB mixture improved the DCTB distribution and widened the mass distribution of CsI cluster ions up to m/z 10 000 in the linear mode. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

8.
Matrix‐assisted laser desorption/ionization in‐source decay (MALDI‐ISD) induces N–Cα bond cleavage via hydrogen transfer from the matrix to the peptide backbone, which produces a c′/z? fragment pair. Subsequently, the z? generates z′ and [z + matrix] fragments via further radical reactions because of the low stability of the z?. In the present study, we investigated MALDI‐ISD of a cyclic peptide. The N–Cα bond cleavage in the cyclic peptide by MALDI‐ISD produced the hydrogen‐abundant peptide radical [M + 2H]+? with a radical site on the α‐carbon atom, which then reacted with the matrix to give [M + 3H]+ and [M + H + matrix]+. For 1,5‐diaminonaphthalene (1,5‐DAN) adducts with z fragments, post‐source decay of [M + H + 1,5‐DAN]+ generated from the cyclic peptide showed predominant loss of an amino acid with 1,5‐DAN. Additionally, MALDI‐ISD with Fourier transform‐ion cyclotron resonance mass spectrometry allowed for the detection of both [M + 3H]+ and [M + H]+ with two 13C atoms. These results strongly suggested that [M + 3H]+ and [M + H + 1,5‐DAN]+ were formed by N–Cα bond cleavage with further radical reactions. As a consequence, the cleavage efficiency of the N–Cα bond during MALDI‐ISD could be estimated by the ratio of the intensity of [M + H]+ and [M + 3H]+ in the Fourier transform‐ion cyclotron resonance spectrum. Because the reduction efficiency of a matrix for the cyclic peptide cyclo(Arg‐Gly‐Asp‐D‐Phe‐Val) was correlated to its tendency to cleave the N–Cα bond in linear peptides, the present method could allow the evaluation of the efficiency of N–Cα bond cleavage for MALDI matrix development. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

9.
IgA nephropathy (IgAN) is one of the most common types of glomerulonephritis worldwide and is diagnosed only with a renal biopsy. The purpose of the present studies was to identify the potential biomarkers for the non‐invasive diagnosis of IgAN. The combination of a magnetic bead separation system with matrix‐assisted laser desorption ionization time‐of‐flight mass spectrometry (MALDI‐TOF MS) was used to analyze urinary peptides of IgAN patients, other glomerulopathy patients, and healthy controls. ClinProTools v.2.0 software was also applied to establish a diagnostic model for IgAN. Our results demonstrated that 11 features had optimal discriminatory performance (p <0.00001). Among these features, the peptide with m/z 1913.14 was identified as a fragment of uromodulin. Receiver operating characteristic (ROC) analysis for m/z 1913.14 showed that the area under the curve (AUC) was 0.998 for distinguishing IgAN versus healthy controls, and 0.815 for distinguishing IgAN versus other glomerulopathy. Analysis of urine peptides patterns by the magnetic bead separation system and MALDI‐TOF‐MS was a non‐invasive diagnostic tool. We conclude that the urinary peptide with m/z 1913.14, which was identified as a uromodulin fragment, may be used as a biomarker for the non‐invasive diagnosis of IgAN clinically. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

10.
Unambiguous identification of mycotoxin‐producing fungal species as Fusarium is of great relevance to agriculture and the food‐producing industry as well as in medicine. Protein profiles of intact fungal spores, such as Penicillium, Aspergillus and Trichoderma, derived from matrix‐assisted laser desorption/ionization time‐of‐flight mass spectrometry (MALDI‐TOF MS) were shown to provide a rapid and straightforward method for species identification and characterization. In this study, we applied this approach to five different Fusarium spp. strains which are known to affect the growth of different grain plants. To obtain a suitable MALDI matrix system and sample preparation method, thin‐layer, dried‐droplet and sandwich methods and several MALDI matrices, namely CHCA, DHB, FA, SA and THAP dissolved in various solvent mixtures (organic solvents such as ACN, MeOH, EtOH and iPrOH and for the aqueous phase water and 0.1% TFA), were evaluated in terms of mass spectrometric pattern and signal intensities. The most significant peptide/protein profiles were obtained with 10 mg ferulic acid (FA) in 1 mL ACN/0.1% TFA (7:3, v/v) used as matrix system. Mixing the spores with the matrix solution directly on the MALDI target (dried‐droplet technique) resulted in an evenly distributed spores/matrix crystal layer, yielding highly reproducible peptide/protein profiles from the spore surfaces. Numerous abundant ions throughout the investigated m/z range (m/z 1500–15 000) could be detected. Differences in the obtained mass spectral patterns allowed the differentiation of spores of various Fusarium species. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

11.
The Escherichia coli single‐stranded DNA binding protein (SSB) selectively binds single‐stranded (ss) DNA and participates in the process of DNA replication, recombination and repair. Different binding modes have previously been observed in SSB?ssDNA complexes, due to the four potential binding sites of SSB. Here, chemical cross‐linking, combined with high‐mass matrix‐assisted laser desorption/ionization (MALDI) mass spectrometry (MS), is used to determine the stoichiometry of the SSB?ssDNA complex. SSB forms a stable homotetramer in solution, but only the monomeric species (m/z 19 100) can be detected with standard MALDI‐MS. With chemical cross‐linking, the quaternary structure of SSB is conserved, and the tetramer (m/z 79 500) was observed. We found that ssDNA also functions as a stabilizer to conserve the quaternary structure of SSB, as evidenced by the detection of a SSB?ssDNA complex at m/z 94 200 even in the absence of chemical cross‐linking. The stability of the SSB?ssDNA complex with MALDI strongly depends on the length and strand of oligonucleotides and the stoichiometry of the SSB?ssDNA complex, which could be attributed to electrostatic interactions that are enhanced in the gas phase. The key factor affecting the stoichiometry of the SSB?ssDNA complex is how ssDNA binds to SSB, rather than the protein‐to‐DNA ratio. This further suggests that detection of the complex by MALDI is a result of specific binding, and not due to non‐specific aggregation in the MALDI plume. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

12.
Data‐independent acquisition (DIA)‐based proteomics has become increasingly complicated in recent years because of the vast number of workflows described, coupled with a lack of studies indicating a rational framework for selecting effective settings to use. To address this issue and provide a resource for the proteomics community, we compared 12 DIA methods that assay tryptic peptides using various mass‐isolation windows. Our findings indicate that the most sensitive single injection LC‐DIA method uses 6 m/z isolation windows to analyze the densely populated tryptic peptide range from 450 to 730 m/z, which allowed quantification of 4465 Escherichia coli peptides. In contrast, using the sequential windowed acquisition of all theoretical fragment‐ions (SWATH) approach with 26 m/z isolation windows across the entire 400–1200 m/z range, allowed quantification of only 3309 peptides. This reduced sensitivity with 26 m/z windows is caused by an increase in co‐eluting compounds with similar precursor values detected in the same tandem MS spectra, which lowers the signal‐to‐noise of peptide fragment‐ion chromatograms and reduces the amount of low abundance peptides that can be quantified from 410 to 920 m/z. Above 920 m/z, more peptides were quantified with 26 m/z windows because of substantial peptide 13C isotope distributions that parse peptide ions into separate isolation windows. Because reproducible quantification has been a long‐standing aim of quantitative proteomics, and is a so‐called trait of DIA, we sought to determine whether precursor‐level chromatograms used in some methods rather than their fragment‐level counterparts have similar precision. Our data show that extracted fragment‐ion chromatograms are the reason DIA provides superior reproducibility. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

13.
A new fragmentation process was proposed to interpret the characteristic product ion at m/z 130 of protonated arginine. The α‐amino group was dissociated from protonated arginine and then combined with the (M + H‐NH3) fragment to form an ion‐neutral complex which further generated a hydroxyl‐amino exchange intermediate compound through an ion‐molecule reaction. This intermediate compound was synthesized from argininamide through a diazo reaction, and then the reaction mixture was analyzed using liquid chromatography combined with mass spectrometry (LC‐MS). The collision‐induced dissociation experiments under the same conditions revealed that this intermediate compound produced the characteristic product ion at m/z 130 as well as protonated arginine, and in addition, density functional theory calculations were performed to confirm simultaneous loss of NH3 and CO from this intermediate to give the m/z 130 ion.  相似文献   

14.
Unusual behaviour has been noted for allyl methyl ether (1) chemically ionized in a high-pressure ion source. Tandem mass spectrometry indicates the formation of methylcyclopentadienyl and methoxy-1-propenylcarbenium ions (d, m/z 81 and e, m/z 85). The origin of these unexpected ions has been elucidated using conventional and Fourier transform ion cyclotron resonance experiments: primary fragment ions derived from protonated 1 (allyl ions a and methoxymethyl cations b) generate collision complexes with neutral 1, giving rise to the ions d and e, respectively, after methanol elimination.  相似文献   

15.
Analytical Chemistry Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee, USA Ultraviolet matrix-assisted laser desorption can be used to enhance formation of [M + H]+, [M + Na]+, and [M + K)+ ions from small peptides for Fourier transform mass spectrometry (FTMS). In accord with laser desorption (LD) time-of-flight experiments, matrices such as nicotinic acid and 2-pyrazinecarboxylic acid exhibit strong enhancement effects (i.e., formation of abundant protonated and cationized molecules for the analyte with virtually no fragment ions) for 266 nm LD/FTMS, whereas pyrazinedicarboxylic acid provides no matrix enhancement at this wavelength. Both sinapinic acid and coumarin-120 provide strong matrix enhancement effects for the 355-nm LD of peptides. For the small peptides examined in this study, no significant differences in the abundance of fragment ions were observed between the 266- and 355-nm wavelengths. Matrix-assisted LD/FTMS is useful for the generation and characterization of ions corresponding to protonated and cationized molecules from virtually all biological compounds with molecular weights up to 2000. The lack of observation of biological ions with m/ z > 2500 may be related to inefficient trapping of these laser-desorbed ions or instrumental detection limitations of FTMS and is under further investigation.  相似文献   

16.
During the process of exploring aqueous piperazine chemistry under simulated flue‐gas scrubbing conditions, positive‐ion electrospray ionisation mass spectrometric (ESI‐MS) analyses of the resulting reaction mixtures in a triple quadrupole system revealed the presence of peaks at m/z 116 and 145, the putative N‐nitroso derivatives of piperazine. Confirmation of the presence of these species in the reaction mixtures was achieved using collision‐induced dissociation experiments. A purchased standard, together with in‐house synthesised N‐nitrosopiperazine standards (including N‐nitroso derivatives derived from deuterium‐labelled precursor materials), were used for this purpose. Across a small range of collision energies, large fluctuations in the abundance of the two major product ions of protonated N‐nitrosopiperazine, m/z 86 and 85, were observed. Using B3LYP/6‐311 + +G(d,p) computations, the potential energy surface was determined for loss of NO and [H,N,O]. At an activation energy slightly in excess of 1 eV, intramolecular isomerisation precedes loss of NO (m/z 86) via a 4,1 H‐shift, and at activation energies between 2.1–2.3 eV, consecutive loss of NO and atomic hydrogen competes with the direct loss of nitrosyl hydride (m/z 85). It is recommended that any multiple reaction monitoring method for quantifying N‐nitrosopiperazines at low collision energies use the sum of both transitions (m/z 116 ← 85, m/z 116 ← 86) to avoid errors that could be introduced by subtle changes in ES source conditions or collision voltages. This approach is adopted in an HPLC/MS/MS method used to monitor the degradation of N‐nitrosopiperazine exposed to (i) broad‐band UV light and (ii) heat typical of an amine regeneration (stripper) tower. The results reveal that aqueous N‐nitrosopiperazine is thermally stable at 150°C but will degrade slowly upon exposure to UV light. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

17.
Phosphorus clusters Pn (n = 1–89) are easily formed from red phosphorus by laser desorption ionization (LDI) and they cover a range of up to approx. m/z 3000 in both positive and negative ion mode. The clusters are singly charged and the spectra are simple because phosphorus is monoisotopic. The mass spectra can be measured with an acceptable resolution and intensity. The use of positively charged Pn clusters for calibration in mass spectrometry was examined and it was demonstrated that in external calibration a standard deviation of ±0.04 m/z units can be achieved even when using a common commercial matrix‐assisted laser desorption/ionization time‐of‐flight (MALDI‐TOF) instrument. When used as internal standards the Pn clusters react with some analytes – C60 and C70 fullerenes and cucurbituril[8], for example. It was also found that red phosphorus is a suitable MALDI matrix for peptides and proteins, illustrated by the examples of a Calmix mixture of bradykinin, angiotensin, renin, adrenocorticotropic hormone ACTH fragment 18‐359 and insulin, and of insulin alone. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

18.
A MALDI source is interfaced to a modified LTQ Orbitrap XL instrument. This work gives insight into the MALDI source design and shows results obtained with the MALDI source coupled to an accurate mass, high-resolution hybrid mass spectrometer. MALDI-produced ions and fragment ions thereof produced in the mass spectrometer may be analyzed and detected by the Orbitrap analyzer at a maximum mass resolution of 100,000 (FWHM) at m/z 400 with high mass accuracy. An accuracy of ≤2 ppm is achieved by internal mass calibration using lock mass functionality; using external mass calibration, an accuracy of ≤3 ppm is routinely obtained. External mass calibration of the hybrid mass spectrometer is performed using a standard calibration mixture of different peptides and matrix components. The instrumental capabilities are demonstrated for analytical methodologies such as Protein ID using Peptide Mass Fingerprint (PMF) and MS/MS analyses of small molecule samples. Stability of mass accuracy and signal-to-noise ratio for low samples loads (on plates) are demonstrated as well as the experimental dynamic range using α-cyano-4-hydroxy cinnamic acid (CHCA) matrix.  相似文献   

19.
Matrix‐assisted laser desorption/ionization in‐source decay (MALDI‐ISD) is initiated by hydrogen transfer from matrix molecules to the carbonyl oxygen of peptide backbone with subsequent radical‐induced cleavage leading to c′/z? fragments pair. MALDI‐ISD is a very powerful method to obtain long sequence tags from proteins or to do de novo sequencing of peptides. Besides classical fragmentation, MALDI‐ISD also shows specific fragments for which the mechanism of formation enlightened the MALDI‐ISD process. In this study, the MALDI‐ISD mechanism is reviewed, and a specific mechanism is studied in details: the N‐terminal side of Cys residue (Xxx‐Cys) is described to promote the generation of c′ and w fragments in MALDI‐ISD. Our data suggest that for sequences containing Xxx‐Cys motifs, the N–Cα bond cleavage occurs following the hydrogen attachment to the thiol group of Cys side‐chain. The c?/w fragments pair is formed by side‐chain loss of the Cys residue with subsequent radical‐induced cleavage at the N–Cα bond located at the left side (N‐terminal direction) of the Cys residue. This fragmentation pathway preferentially occurs at free Cys residue and is suppressed when the cysteines are involved in disulfide bonds. Hydrogen attachment to alkylated Cys residues using iodoacetamide gives free Cys residue by the loss of ?CH2CONH2 radical. The presence of alkylated Cys residue also suppress the formation of c?/w fragments pair via the (Cβ)‐centered radical, whereas w fragment is still observed as intense signal. In this case, the z? fragment formed by hydrogen attachment of carbonyl oxygen followed side‐chain loss at alkylated Cys leads to a w fragment. Hydrogen attachment on peptide backbone and side‐chain of Cys residue occurs therefore competitively during MALDI‐ISD process. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

20.
The methane and isobutane chemical ionization mass spectra of alicyclic substituted 2-aryl-1,3-dithianes were examined by gas chromatography mass spectrometry. The protonated molecular ion was found to be of low abundance in the methane spectra, while a protonated cyclic sulfide cation (m/z 107) appeared as the base peak. A protonated molecular ion was the base peak when isobutane was used as the reagent gas. Electron impact mass spectra displayed weak molecular ions and were characterized by the m/z 106 fragment.  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