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 215 毫秒
1.
Since introduction of sinapinic acid (SA) and α‐cyano‐4‐hydroxycinnamic acid as matrices, successful application of matrix‐assisted laser desorption/ionization mass spectrometry started for protein/polypeptides. Both show some limitations in short peptide analysis because matrix clusters are quite abundant. Cinnamics currently used are E‐cinnamics. Here, Z‐SA as matrix for peptides is studied and compared with E‐SA and α‐cyano‐4‐hydroxycinnamic acid. Minor number of clusters is always observed in the low m/z region allowing the detection of short peptides. The results here described show that this novel matrix is a tool of choice for direct, rapid and sensitive detection of hydrophilic and hydrophobic peptides. Copyright © 2017 John Wiley & Sons, Ltd.  相似文献   

2.
Cinnamic acid derivatives, particularly α‐cyano‐4‐hydroxycinnamic acid (E‐α‐cyano‐4‐hydroxycinnamic acid or (E)‐2‐cyano‐3‐(4‐hydroxyphenyl)prop‐2‐enoate; CHCA), have been extensively used especially for protein and peptide analysis. Together with the introduction of ionic liquid MALDI matrix (ILM) started the study of applications of IL prepared with CHCA and a counter organic base (ie, aliphatic amines) in which CHCA moiety is the chromophore responsible of UV‐laser absorption. Despite the extensive studies of norharmane (9H‐pyrido[3,4‐b]indole; nHo) applications as matrix and its peculiar basic properties in the ground and electronic excited state, nHo containing ILM was never tested in MALDI‐MS experiments. This pyrido‐indole compound was introduced as MALDI matrix 22 years ago for different applications including low molecular weight (LMW) carbohydrates (neutral, acidic, and basic carbohydrates). These facts encouraged us to use it as a base, for the first time, for ILM preparation. As a rational design of new IL MALDI matrices, E‐α‐cyanocinnamic acid.nHo and E‐cinnamic acid.nHo were prepared and their properties as matrices studied. Their performance was compared with that of (a) the corresponding IL prepared with butylamine as basic component, (b) the corresponding crystalline E‐α‐cyanocinnamic and E‐cinnamic acid, and (c) the classical crystalline matrices (2,5‐dihydroxybenzoic acid, DHB; nHo) used in the analysis of neutral/sulfated carbohydrates. The IL DHB.nHo was tested, too. Herein, we demonstrate the outstanding performance for the IL CHCA.nHo for LMW carbohydrate in positive and negative ion mode (linear and reflectron modes). Sulfated oligosaccharides were detected in negative ion mode, and although the dissociation of sulfate groups was not completely suppressed the relative intensity (RI) of [M ? Na]? peak was quite high. Additionally, to better understand the quite different performance of each IL tested as matrix, the physical and morphological properties in solid state were studied (optical image; MS image).  相似文献   

3.
The thermal stability of several commonly used crystalline matrix‐assisted ultraviolet laser desorption/ionization mass spectrometry (UV‐MALDI‐MS) matrices, 2,5‐dihydroxybenzoic acid (gentisic acid; GA), 2,4,6‐trihydroxyacetophenone (THA), α‐cyano‐4‐hydroxycinnamic acid (CHC), 3,5‐dimethoxy‐4‐hydroxycinnamic acid (sinapinic acid; SA), 9H‐pirido[3,4‐b]indole (nor‐harmane; nor‐Ho), 1‐methyl‐9H‐pirido[3,4‐b]indole (harmane; Ho), perchlorate of nor‐harmanonium ([nor‐Ho + H]+) and perchlorate of harmanonium ([Ho + H]+) was studied by heating them at their melting point and characterizing the remaining material by using different MS techniques [electron ionization mass spectrometry (EI‐MS), ultraviolet laserdesorption/ionization‐time‐of‐flight‐mass spectrometry (UV‐LDI‐TOF‐MS) and electrospray ionization‐time‐of‐flight‐mass spectrometry (ESI‐TOF‐MS)] as well as by thin layer chromatography analysis (TLC), electronic spectroscopy (UV‐absorption, fluorescence emission and excitation spectroscopy) and 1H nuclear magnetic resonance spectroscopy (1H‐NMR). In general, all compounds, except for CHC and SA, remained unchanged after fusion. CHC showed loss of CO2, yielding the trans‐/cis‐4‐hydroxyphenylacrilonitrile mixture. This mixture was unambiguously characterized by MS and 1H‐NMR spectroscopy, and its sublimation capability was demonstrated. These results explain the well‐known cluster formation, fading (vanishing) and further recovering of CHC when used as a matrix in UV‐MALDI‐MS. Commercial SA (SA 98%; trans‐SA/cis‐SA 5 : 1) showed mainly cis‐ to‐trans thermal isomerization and, with very poor yield, loss of CO2, yielding (3′,5′‐dimethoxy‐4′‐hydroxyphenyl)‐1‐ethene as the decarboxilated product. These thermal conversions would not drastically affect its behavior as a UV‐MALDI matrix as happens in the case of CHC. Complementary studies of the photochemical stability of these matrices in solid state were also conducted. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

4.
A series of ruthenium carbene catalysts containing 2‐sulfidophenolate bidentate ligand with an ortho‐substituent next to the oxygen atom were synthesized. The molecular structure of ruthenium carbene complex containing 2‐isopropyl‐6‐sulfidophenolate ligand was confirmed through single crystal X‐ray diffraction. An oxygen atom can be found in the opposite position of the N‐heterocyclic carbene (NHC) based on the steric hindrance and strong trans‐effects of the NHC ligand. The ruthenium carbene catalyst can catalyze ring‐opening metathesis polymerization (ROMP) reaction of norbornene with high activity and Z‐selectivity and cross metathesis (CM) reactions of terminal alkenes with (Z)‐but‐2‐ene‐1,4‐diol to give Z‐olefin products (Z/E ratios, 70:30–89:11) in low yields (13%–38%). When AlCl3 was added into the CM reactions, yields (51%–88%) were considerably improved and process becomes highly selective for E‐olefin products (E/Z ratios, 79:21–96:4). Similar to other ruthenium carbene catalysts, these new complexes can tolerate different functional groups.  相似文献   

5.
In a previous study on matrix‐assisted laser desorption ionization (MALDI) of peptides using α‐cyano‐4‐hydroxycinnamic acid (CHCA) as a matrix, we found that the patterns of single‐shot spectra obtained under different experimental conditions became similar upon temperature selection. In this paper, we report that absolute ion abundances are also similar in temperature‐selected MALDI spectra, even when laser fluence is varied. The result that has been obtained using CHCA and 2,5‐dihydroxybenzoic acid as matrices is in disagreement with the hypothesis of laser‐induced ionization of matrix as the mechanism for primary ion formation in MALDI. We also report that the total number of ions in such a spectrum is unaffected by the identity, concentration and number of analytes, i.e. it is the same as that in the spectrum of pure matrix. We propose that the generation of gas‐phase ions in MALDI can be explained in terms of two thermal reactions, i.e. the autoprotolysis of matrix molecules and the matrix‐to‐analyte proton transfer, both of which are in quasi‐equilibrium in the early matrix plume. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

6.
A designed bis(dithienyl) dicyanoethene‐based, strictly E/Z photoswitch (4TCE) operates through state‐selective (E and Z isomer) photoactivation with visible light. The E and Z isomers of 4TCE exhibit remarkably different spectroscopic characteristics, including a large separation (70 nm) in their absorption maxima (λmax) and a 2.5‐fold increase in molar extinction coefficient from cis to trans. The energetically stable trans form can be completely converted to the cis form within minutes when exposed to white light, whereas the reverse isomerization occurs readily upon irradiation by blue light (λ<480 nm) or completely by thermal conversion at elevated temperatures. These features together with excellent thermal stability and photostability of both isomers make this new E/Z photoswitch a promising building block for photoswitchable materials that operate without the need for UV light.  相似文献   

7.
The Z and E isomers of 3‐[4‐(dimethylamino)phenyl]‐2‐(2,4,6‐tribromophenyl)acrylonitrile, C17H13Br3N2, ( 1 ), were obtained simultaneously by a Knoevenagel condensation between 4‐(dimethylamino)benzaldehyde and 2‐(2,4,6‐tribromophenyl)acetonitrile, and were investigated by X‐ray diffraction and density functional theory (DFT) quantum‐chemical calculations. The (Z)‐( 1 ) isomer is monoclinic (space group P21/n, Z′ = 1), whereas the (E)‐( 1 ) isomer is triclinic (space group P, Z′ = 2). The two crystallographically‐independent molecules of (E)‐( 1 ) adopt similar geometries. The corresponding bond lengths and angles in the two isomers of ( 1 ) are very similar. The difference in the calculated total energies of isolated molecules of (Z)‐( 1 ) and (E)‐( 1 ) with DFT‐optimized geometries is ∼4.47 kJ mol−1, with the minimum value corresponding to the Z isomer. The crystal structure of (Z)‐( 1 ) reveals strong intermolecular nonvalent Br…N [3.100 (2) and 3.216 (3) Å] interactions which link the molecules into layers parallel to (10). In contrast, molecules of (E)‐( 1 ) in the crystal are bound to each other by strong nonvalent Br…Br [3.5556 (10) Å] and weak Br…N [3.433 (4) Å] interactions, forming chains propagating along [110]. The crystal packing of (Z)‐( 1 ) is denser than that of (E)‐( 1 ), implying that the crystal structure realized for (Z)‐( 1 ) is more stable than that for (E)‐( 1 ).  相似文献   

8.
New data on sample preparation and matrix selection for the fast screening of androgenic anabolic steroids (AAS) by matrix‐assisted laser desorption/ionisation time‐of‐flight mass spectrometry (MALDI‐TOF‐MS) is presented. The rapid screening of 15 steroids included in the World Anti‐Doping Agency (WADA) prohibited list using MALDI was evaluated. Nine organic and two inorganic matrices were assessed in order to determine the best matrix for steroid identification in terms of ionisation yield and interference by characteristic matrix ions. The best results were achieved for the organic matrices 2‐(4‐hydroxyphenylazo)benzoic acid (HABA) and trans‐3‐indoleacrylic acid (IAA). Good signals for all the steroids studied were obtained for concentrations as low as 0.010 and 0.050 µg/mL on the MALDI sample plate for the HABA and IAA matrices, respectively. For these two matrices, the sensitivity achieved by MALDI is comparable with the sensitivity achieved by gas chromatography/mass spectrometry (GC/MS), which is the conventional technique used for AAS detection. Furthermore, the accuracy and precision obtained with MALDI are very good, since an internal mass calibration is performed with the matrix ions. For the inorganic matrices, laser fluences higher than those used with organic matrices are required to obtain good MALDI signals. When inorganic matrices were used in combination with glycerol as a dispersing agent, an important reduction of the background noise was observed. Urine samples spiked with the study compounds were processed by solid‐phase extraction (SPE) and the screening was consistently positive. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

9.
A novel three‐phase hollow fiber liquid‐phase microextraction was developed based on reverse micelle as extraction solvent and acceptor phase, and compared with conventional two‐phase hollow fiber liquid‐phase microextraction. Both procedures were used in the extraction and concentration of four cinnamic acids (caffeic acid, p‐hydroxycinnamic acid, ferulic acid, and cinnamic acid) in traditional Chinese medicines prior to high‐performance liquid chromatography analysis. Parameters affecting the two procedures were investigated and optimized to obtain the optimum enrichment factors. The mechanism of the developed procedure was explored and elucidated by comparison with conventional two‐phase hollow fiber liquid‐phase microextraction. Under the optimized conditions, the analytes’ enrichment factors were between 50 and 118 for the proposed procedure, and 31–96 for conventional two‐phase mode. Satisfactory linear ranges (r2 ≥ 0.99), detection limits (0.1–0.6 ng/mL), precisions (<9.2%), and accuracies (recoveries: 80–123.1%) were observed for the two procedures. The results showed that the enrichment capacity of the proposed procedure for the cinnamic acids is better than that of conventional two‐phase procedure, and both are eco‐friendly, simple, and effective for the enrichment and detection of cinnamic acids in traditional Chinese medicines.  相似文献   

10.
The performance of a matrix‐assisted laser desorption/ionization (MALDI) ionic liquid matrix (ILM) consisting of α‐cyano‐4‐hydroxycinnamic acid (CHCA) and aniline (ANI) was evaluated to assess whether it could offer possible advantages over conventional matrices. Ultraviolet (UV), Fourier transform infrared (FTIR), nuclear magnetic resonance (NMR) and laser desorption/ionization mass spectrometry (LDI‐MS) experiments were carried out with the aim of confirming the structure of the ANI‐CHCA ILM. Different model analytes such as amino acids, peptides, proteins, lipids, phospholipids, synthetic polymers, and sugars were tested. Mass spectra with similar or improved signal‐to‐noise (S/N) ratio (compared to CHCA) were invariably obtained demonstrating the potential of this ILM as a general purpose matrix. Furthermore, protein identification by peptide mass fingerprinting (PMF) and database search was facilitated compared to CHCA since higher scores and increased sequence coverage were observed. Finally, a complex lipid mixture (i.e. a raw extract of a milk sample) analysed by MALDI‐MS showed improved S/N ratio, a reduced chemical noise and a limited formation of matrix‐clusters. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

11.
Hydroboration of internal alkynes with N‐heterocyclic carbene boranes (NHC‐boranes) occurs to provide stable NHC (E)‐alkenylboranes upon thermolysis in the presence of di‐tert‐butyl peroxide. The E isomer results from an unusual trans‐hydroboration, and the E/Z selectivity is typically high (90:10 or greater). Evidence suggests that this hydroboration occurs by a radical‐chain reaction involving addition of an NHC‐boryl radical to an alkyne to give a β‐NHC‐borylalkenyl radical. Ensuing hydrogen abstraction from the starting NHC‐borane provides the product and returns the starting NHC‐boryl radical. Experiments suggest that the observed trans‐selectivity results from kinetic control in the hydrogen‐transfer reaction.  相似文献   

12.
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   

13.
Aerosol matrix-assisted laser desorption ionization (MALDI) with a reflection time-of-flight mass spectrometer was used to study fragmentation of vitamin B12. Six MALDI matrices were used: 2,5-di-hydroxy benzoic acid (gentisic acid), 4-nitroaniline, 3,5-dimethoxy-4-hydroxy cinnamic acid (sinapic acid), 3,4-di-hydroxy cinnamic acid (caffeic acid), trans-4-hydroxy-3-methoxy cinnamic acid (ferulic acid), and α-cyano-4-hydroxy cinnamic acid (4-HCCA). Mass spectra were obtained with a 355-nm pulsed Nd:YAG laser at irradiances between 0. 1 and 5 GW/cm2 (between 3- and 150-mJ pulse energy). Loss of CN was a major product of prompt ion source fragmentation and the ratio of fragmented to intact analyte was found to be strongly dependent on matrix and weakly dependent on laser irradiance. Additionally, free cobalt ions and cobalt ions bound to small methanol clusters were observed in the mass spectra. The cobalt removal from the corrin ring of vitamin B12 results from direct photon absorption by vitamin B12, but is enhanced by the presence of matrix.  相似文献   

14.
The single‐crystal‐to‐single‐crystal (SCSC) E?Z photoisomerization of TA in the supramolecular solid CECR‐[Zn(TA)2(bpy)]·H2O (CECR = C‐ethylcalixresorcinarene, HTA = tiglic acid, and bpy = 2,2′‐bipyridine, is induced by 458 nm light, indicating a red‐shift of the photo‐active wavelength on introduction of the aromatic bpy ligand compared with the previously studied reaction of CECR‐[Zn(TA)2(H2O)2] 4H2O. Theoretical calculations show that the initial excitation involves the bipyridyl ligand, which acts as an intramolecular photosensitizer for the isomerization process. The reaction is topotactic and illustrated by photodifference maps.  相似文献   

15.
Due to the characteristic absorption peaks in the IR region, various molecules can be used as a matrix for infrared matrix‐assisted laser desorption/ionization (IR‐MALDI). Especially in the 6–7 µm‐band IR region, solvents used as the mobile phase for liquid chromatography have absorption peaks that correspond to their functional groups, such as O–H, CO, and CH3. Additionally, atmospheric pressure (AP) IR‐MALDI, which is applicable to liquid‐state samples, is a promising technique to directly analyze untreated samples. Herein we perform AP‐IR‐MALDI mass spectrometry of a peptide, angiotensin II, using a mid‐IR tunable laser with a tunable wavelength range of 5.50–10.00 µm and several different matrices. The wavelength dependences of the ion signal intensity of [M + H]+ of the peptide are measured using a conventional solid matrix, α‐cyano‐4‐hydroxycinnamic acid (CHCA) and a liquid matrix composed of CHCA and 3‐aminoquinoline. Other than the O–H stretching and bending vibration modes, the characteristic absorption peaks are useful for AP‐IR‐MALDI. Peptide ions are also observed from an aqueous solution of the peptide without an additional matrix, and the highest peak intensity of [M + H]+ is at 6.00 µm, which is somewhat shorter than the absorption peak wavelength of liquid water corresponding to the O–H bending vibration mode. Moreover, long‐lasting and stable ion signals are obtained from the aqueous solution. AP‐IR‐MALDI using a 6–7 µm‐band IR tunable laser and solvents as the matrix may provide a novel on‐line interface between liquid chromatography and mass spectrometry. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

16.
Asymmetric oxyallylation reactions and ring‐closing metathesis have been used to synthesize compound 3 , a key advanced intermediate used in the total synthesis of eleutherobin reported by Danishefsky and co‐workers. The aldehyde 6 , which is readily prepared from commercially available R‐(?)‐carvone in six steps in 30 % overall yield on multigram quantities, was converted into the diene 5 utilizing two stereoselective titanium‐mediated Hafner–Duthaler oxyallylation reactions. The reactions gave the desired products ( 8 and 12 ) in high yields (73 and 83 %, respectively) as single diastereoisomers, with the allylic alcohol already protected as the p‐methoxyphenyl (PMP) ether, which previous work has demonstrated actually aids ring‐closing metathesis compared to other protective groups and the corresponding free alcohol. Cyclization under forcing conditions, using Grubbs' second‐generation catalyst 13 , gave the ten‐membered carbocycle (E)‐ 14 in 64 % yield. This result is in sharp contrast to similar, but less functionalized, dienes, which have all undergone cyclization to give the Z stereoisomers exclusively. A detailed investigation of this unusual cyclization stereochemistry by computational methods has shown that the E isomer of the ten‐membered carbocycle is indeed less thermodynamically stable than the corresponding Z isomer. In fact, the selectivity is believed to be due to the dense functionality around the ruthenacyclobutane intermediate that favors the trans‐ruthenacycle, which ultimately leads to the less stable E isomer of the ten‐membered carbocycle under kinetic control. During the final synthetic manipulations the double bond of enedione (E)‐ 16 isomerized to the more thermodynamically stable enedione (Z)‐ 4 , giving access to the advanced key‐intermediate 3 , which was spectroscopically and analytically identical to the data reported by Danishefsky and co‐workers, and thereby completing the formal synthesis of eleutherobin.  相似文献   

17.
The most common secondary‐ionization mechanism in positive ion matrix‐assisted laser desorption/ionization (MALDI) involves a proton transfer reaction to ionize the analyte. Peptides and proteins are molecules that have basic (and acidic) sites that make them susceptible to proton transfer. However, non‐polar, aprotic compounds that lack basic sites are more difficult to protonate, and creating charged forms of this type of analyte can pose a problem when conventional MALDI matrices are employed. In this case, forming a radical molecular ion through electron transfer is a viable alternative, and certain matrices may facilitate the process. In this work, we investigate the performance of a newly developed electron‐transfer secondary reaction matrix: 9,10‐diphenylanthracene (9,10‐DPA). The use of 9,10‐DPA as matrix for MALDI analysis has been tested using several model compounds. It appears to promote ionization through electron transfer in a highly efficient manner as compared to other potential matrices. Thermodynamic aspects of the observed electron transfers in secondary‐ionization reactions were also considered, as was the possibility for kinetically controlled/endothermic, electron‐transfer reactions in the MALDI plume. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

18.
This article proposes a simple and sensitive HPLC method with photo‐diode array detection for the analysis of organic acids, monomeric polyphenols and furanic compounds in wine samples by direct injection. The chromatographic separation of 8 organic acids, 2 furans and 22 phenolic compounds was carried out with a buffered solution (pH 2.70) and acetonitrile as mobile phases and a difunctionally bonded C18 stationary phase, Atlantis dC18 (250×4.6 mm, 5 μm) column. The elution was performed in 12 min for the organic acids and in 60 min for the phenolic compounds, including phenolic acids, stilbenes and flavonoids. Target compounds were detected at 210 nm (organic acids, flavan‐3‐ols and benzoic acids), 254 nm (ellagic acid), 280 nm (furans and cinnamic acid), 315 nm (hydroxycinnamic acids and trans‐resveratrol) and 360 nm (flavonoids). The RSD for the repeatability test (n=5) of peak area and retention times were below 3.1 and 0.3%, respectively, for phenolics and below 1.0 and 0.2% for organic acids. The RSDs expressing the reproducibility of the method were higher than for the repeatability results but all below 9.0%. Method accuracy was evaluated by the recovery results, with averaged values between 80 and 104% for polyphenols and 97–105% for organic acids. The calibration curves, obtained by triplicate injection of standard solutions, showed good linearity with regression coefficients higher than 0.9982 for polyphenols and 0.9997 for organic acids. The LOD was in the range of 0.07–0.49 mg/L for polyphenols (cinnamic and gallic acids, respectively) and 0.001–0.046 g/L for organic acids (oxalic and lactic acids, respectively). The method was successfully used to measure and assess the polyphenolic fingerprint and organic acids profile of red, white, rosé and fortified wines.  相似文献   

19.
The localization of polymeric composition in samples prepared for matrix‐assisted laser desorption/ionization (MALDI) analysis has been investigated by imaging mass spectrometry. Various matrices and solvents were used for sample spot preparation of a polybutyleneglycol (PBG 1000). It was shown that in visibly homogeneous spots, prepared using the ‘dried droplet’ method, separation between matrix and polymer takes place. Moreover, using α‐cyano‐4‐hydroxycinnamic acid (CCA) as matrix and methanol as solvent molecular mass separation of the polymer homologues in the spots was detectable. In contrast to manually spotted samples, dry spray deposition results in homogeneous layers showing no separation effects. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

20.
Bis((Z)‐5‐phenyl‐2‐phenylmethylidene‐1, 3‐dithiole‐4‐yl)monosulfane ( 6 ), a molecule consisting of two diphenyldithiafulvene units connected by a sulfur bridge, was synthesized by the selective lithiation of (Z)‐4‐phenyl‐2‐phenylmethylidene‐1, 3‐dithiole ( 7a ) at the endocyclic double bond and by subsequent reaction of the lithiated intermediate with bis(phenylsulfonyl)sulfane. Since this reaction sequence proceeded with retention of configuration, of three possible isomers (E, E, Z, E, and Z, Z) only the Z, Z form was obtained. On the basis of the X‐ray structure analysis and the NMR‐spectroscopic characterization of 6 supplemented by the NMR parameters of (E)‐ and (Z)‐4‐phenyl‐2‐phenylmethylidene‐1, 3‐dithiole, it was demonstrated that two characteristic 5J coupling constants of the proton at the exocyclic double bond indicate the configuration (Z or E) of disubstituted dithiafuvene derivatives.  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