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 125 毫秒
1.
Dihydroxybenzoic acid (DHB) analogues substituted at the 5-position can act as UV matrices at a wavelength of 337 nm, even when their absorption maxima are shifted past this particular wavelength.Modification of a matrix with a chiral ligand γ-(3-carboxy-4-hydroxy-anilide) (GCA) allowed it to differentiate between chiral isomers of tryptophan and also gave different intensities for glucose isomers, including structural dimers of glucose (cellobiose and maltose).An analogue that had one free hydroxyl group at the 2-OH position and a modification at the 5-position (MY10) gave protonated substance P (SP, analyte) peaks, similar in intensity to the not derivatised parent 2,5-dihydroxybenzoic acid indicating that the 5-OH position is not an important structural component.Another analogue that resembled a ‘dendrimeric’ structure of DHB (M552), also acted as a matrix, although its absorption maxima was at 552 nm suggesting the possibility of it being used at other wavelengths in addition to 337 nm.The DHB radical was complexed to a nitrone ‘spin-trap’. On complexion, the peptide (SP) peak intensity decreased. Addition of either radical initiators, such as 2,2-azobis(iso-butyronitrile) AIBN and tert-butylperoxide, or other radicals such as 2,2,6,6-tetramethyl-1-piperidinyloxy free radical (TEMPO) gave rise to higher analyte peak intensities for [SP+Na]+.It is thought that the DHB neutral radical is an intermediary in the protonation of the analyte. The photo-fragments of DHB, specifically the m/z 137 species, may also take part in proton transfer since possible mass analogues (hydroquinone, (deoxy)benzoin) can lead to analyte enhancement. Stabilization of or an increase in the matrix radical can also lead to analyte signal enhancement.  相似文献   

2.
The mass spectra of peptides obtained with different matrices were compared using a matrix-assisted laser desorption/ionization (MALDI) ion source and a multi-turn time-of-flight (TOF) mass spectrometer, MULTUM-IMG, which has been developed at Osaka University. Two types of solid matrices, alpha-cyano-4-hydroxycinnamic acid (CHCA) and 2,5-dihydroxybenzoic acid (DHB), and a liquid matrix made from a mixture of 3-aminoquinoline and CHCA were used. When measuring the peak signal intensity of human angiotensin II [M+H]+ from a fixed sample position, the liquid matrix produced a stable signal over 1000 laser shots, while the signal obtained with CHCA and DHB decayed after about 300 and 100 shots, respectively. Significant differences in the mass resolving power were not observed between the spectra obtained with the three matrices. Signal peak areas were measured as a function of the cycle number in a multi-turn ion trajectory, i.e., the total flight time over a millisecond time scale. For both [M+H]+ of human angiotensin II and bovine insulin, the decay of the signal peak area was the most significant with CHCA, while that measured with DHB was the smallest. The results of the mean initial ion velocity measurements suggested that the extent of metastable decomposition of the analyte ions increased in order of DHB, the liquid matrix, and CHCA, which is consistent with the difference in the decay of the signal peak area as the total flight time increased.  相似文献   

3.
A specially constructed split sample probe was used to unequivocally demonstrate that gas-phase cationization occurs within the desorption plume during a matrix-assisted laser desorption/ionization experiment. Two separate samples were prepared for analysis: on side A, a mixture of poly(ethylene glycol) (PEG) 1500 analyte and 2,5-dihydroxybenzoic acid (DHB) matrix, and on side B a mixture of DHB matrix and lithium hydroxide (LiOH), the cationization reagent. Analysis of the data showed that when the ionization laser was focused on the split (so that both sides were illuminated), Li(+)-cationized PEG peaks were observed. Since the PEG analyte did not come into contact with Li(+) in either the solution or solid phase, the only possibility for the observed cationization was a reaction in the gas phase. Due to the difficulty in completely removing the adventitious cations (Na(+) and K(+)) present in DHB and on sample surfaces, gas-phase cationization could not be demonstrated to be either the only or most important mechanism operating in the MALDI experiment.  相似文献   

4.
Matrix-assisted laser desorption/ionization mass spectrometry (MALDI-MS) is a powerful tool for the analysis and characterization of protein phosphorylation on the peptide level. In this study, the applicability of ionic liquid matrices (ILM) formed by combination of the crystalline MALDI matrix 2,5-dihydroxybenzoic acid (DHB) with pyridine or n-butylamine was tested for the analysis of phosphopeptides. Low ionization efficiency in both positive and negative ion mode was observed in acid-free sample preparations. Upon addition of 0.1% trifluoroacetic acid (TFA), ion formation was increased, but analogously to the situation described earlier for pure DHB, best results were obtained upon use of 1% phosphoric acid as matrix additive. The samples prepared in this way were significantly more homogeneous than preparations with pure DHB, thus avoiding the need for time-consuming search for hot spots. Other characteristics like metastable fragmentation of phosphopeptides did not differ from that observed in classical preparations. The limits of detection for synthetic phosphopeptides and singly or multiply phosphorylated peptides from tryptic digests of alpha- and beta-casein were comparable with those obtained when using pure DHB; in some cases even higher signal intensities could be observed in the ILM. The use of ILM in combination with 1% phosphoric acid as matrix additive significantly facilitates analysis of phosphopeptides by MALDI-MS.  相似文献   

5.
The behaviour of 2,5‐dihydroxybenzoic acid (2,5‐DHB) matrix under matrix‐assisted laser desorption/ionisation (MALDI) conditions was investigated, and the formation of 2,5‐DHB cluster ions, mainly dehydrated 2,5‐DHB ions, is reported. Interestingly, in the mass spectra of this compound, besides dimers and trimers, protonated tetramers, pentamers, hexamers and heptamers were also found with significant abundance. The MALDI behaviour of four flavonoids, quercetin, myricetin, luteolin and kaempferol, using 2,5‐DHB as matrix, was also investigated. The mass spectra of the flavonoids studied revealed a number of flavonoid–2,5‐DHB cluster ions (mainly with the dehydrated 2,5‐DHB). The number of clusters formed is dependent on the structure of the analyte. For luteolin and kaempferol, in particular, evidence was found for the formation of cluster ions involving retro Diels Alder fragments and intact flavonoids molecules, as well as the corresponding protonated retro Diels Alder fragments with dehydrated DHB molecules. All ion compositions were attributed taking into account high accuracy mass measurements and tandem mass spectrometry experiments. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

6.
The protocol and various matrices were examined for quantification of biomolecules in both the low ca. 1200 amu and mid mass 6000-12000 amu ranges using an internal standard. Comparative studies of different matrices of MALDI quantitative analysis showed that the best accuracy and standard curve linearity were obtained for two matrices: (a) 2,5-dihydroxybenzoic acid (DHB) combined with a comatrix of fucose and 5-methoxysalicylic acid (MSA) and (b) ferulic acid/fucose. In the low mass range, the quantitative limit was in the 30 fmol range and in the mid mass range the quantitative limit was in the 250 fmol range. Linear response was observed over 2-3 decades of analyte concentration. The relative error of the standard curve slope was 1.3-1.8% with correlation coefficients of 0.996-0.998.The main problem for quantitative measurement was suppression of the signal of the less concentrated component (analyte or internal standard) by the more concentrated component. The effect was identified with saturation of the matrix by the analyte. The threshold of matrix saturation was found to be in the range of 1/(3000-5000) analyte/matrix molar ratio. To avoid matrix saturation the (analyte+internal standard) to matrix molar ratio should be below this threshold. Thus the internal standard concentration should be as low as possible.DHB/MSA/fucose and ferulic acid/fucose matrices demonstrated good accuracy and linearity for standard curves even when the internal standard had chemical properties different from the analyte. However, use of an internal standard with different chemical properties requires highly stable instrumental parameters as well as constant (analyte+internal standard)/matrix molar ratio for all samples.  相似文献   

7.
Intracluster proton transfer from the matrix-assisted laser desorption/ionization matrix 2,5-dihydroxybenzoic acid (DHB) to the peptide valyl-prolyl-leucine has been investigated as a function of excitation laser wavelength and power. Ionization laser power studies at 308 nm indicate that cluster ionization occurs with a two-photon dependence, whereas matrix-to-analyte proton transfer and cluster dissociation requires an additional photon. At 266 nm, two-photon absorption leads to both cluster ionization and cluster dissociation/proton transfer. A consideration of these results clearly indicates that analyte protonation occurs following ionization of the cluster to produce a radical cation matrix/analyte cluster. Mass spectral features also indicate that mixed DHB/peptide cluster ionization can occur via two-photon ionization at wavelengths as long as 355 nm. These results suggest a reduction in the ionization potential of larger mixed DHB/peptide clusters of greater than 1 eV. The reduced ionization potential seen in these clusters suggests that radical cation initiated proton transfer remains a viable mechanism for analyte protonation in matrix-assisted laser desorption/ionization at these longer wavelengths.  相似文献   

8.
Bashir S  Mutter R  Derrick PJ 《The Analyst》2003,128(12):1452-1457
Dihydroxybenzoic acid was modified to three analogues (M2, M4 and M6). The analogues exhibited specific properties that resulted in enhancement of analyte signal intensity with or without addition of iodine compared to the underivatized parent. Addition of iodine to M2, an ester of dihydroxybenzoic acid that had a terminal double bond in the alkyl chain, resulted in peak intensities comparable to the parent, indicating that iodine interaction across the double bond resulted in enhancement although the exact mechanism is not fully understood. No enhancement on addition of iodine was observed for M4, which had a long alkyl chain that contained no double bonds. The alkyl chain allowed micelle formation in solution, which in turn allowed more uniform analyte-to-matrix mixing. The final analogue combined the long alkyl chain of M4 with the double bond of M2 and exhibited either similar peak intensities to that of dihydroxybenzoic acid or better. Micelle formation in solution was examined using spectroscopy and in the solid by reflective microscopy. The standard deviation from spot to spot was considerably lower relative to dihydroxybenzoic acid (RSD 3.4%vs. 14.2%). Unlike dihydroxybenzoic acid, the novel matrix M6 was able to yield characteristic peaks for analytes such as ubiquitin.  相似文献   

9.
Negative-ion matrix-assisted laser desorption/ionization mass spectra of sialylated glycoconjugates were acquired employing 2,5-dihydroxybenzoic acid (DHB) in conjunction with spermine as a co-matrix. The addition of spermine to DHB permitted an improved crystal formation as well as a higher analyte solubility. Moreover, DHB/spermine appears to minimize alkali adduct formation, thus allowing the sample analysis without desalting. The combined matrix permitted the analysis of complex sialylated and sialylated/fucosylated structures down to the femtomole range. The ability to use such a matrix also facilitates determination of the sialic acid linkages (in combination with a specific enzyme cleavage). The matrix also appears suitable for studies on gangliosides.  相似文献   

10.
The protocol and various matrices were examined for quantification of biomolecules in both the low ca. 1200 amu and mid mass 6000–12000 amu ranges using an internal standard. Comparative studies of different matrices of MALDI quantitative analysis showed that the best accuracy and standard curve linearity were obtained for two matrices: (a) 2,5-dihydroxybenzoic acid (DHB) combined with a comatrix of fucose and 5-methoxysalicylic acid (MSA) and (b) ferulic acid/fucose. In the low mass range, the quantitative limit was in the 30 fmol range and in the mid mass range the quantitative limit was in the 250 fmol range. Linear response was observed over 2–3 decades of analyte concentration. The relative error of the standard curve slope was 1.3–1.8% with correlation coefficients of 0.996–0.998.The main problem for quantitative measurement was suppression of the signal of the less concentrated component (analyte or internal standard) by the more concentrated component. The effect was identified with saturation of the matrix by the analyte. The threshold of matrix saturation was found to be in the range of 1/(3000–5000) analyte/matrix molar ratio. To avoid matrix saturation the (analyte+internal standard) to matrix molar ratio should be below this threshold. Thus the internal standard concentration should be as low as possible.DHB/MSA/fucose and ferulic acid/fucose matrices demonstrated good accuracy and linearity for standard curves even when the internal standard had chemical properties different from the analyte. However, use of an internal standard with different chemical properties requires highly stable instrumental parameters as well as constant (analyte+internal standard)/matrix molar ratio for all samples.  相似文献   

11.
D-Glucose and 19 glucose derivatives were investigated by positive and negative ion matrix assisted laser desorption/ionization time-of-flight mass spectrometry using 2,5-dihydroxybenzoic acid (DHB) as the matrix. The set of substrates includes oligomers of amylose and cellulose, native alpha-, beta-, and gamma-cyclodextrin, and chemically modified beta- and gamma-cyclodextrins. These analytes were chosen to modulate molecular weight, polarity, and capability of establishing noncovalent interactions with guest molecules. In the negative-ion mode, the DHB matrix gave rise to charged multicomponent adducts of type [M + DHB - H]- (M = oligosaccharide) selectively for those analytes matching the following conditions: (i) underivatized chemical structure and (ii) number of glucose units > or = 4. In the positive-ion polarity, only some amylose and cellulose derivatives and methylated beta-cyclodextrins provided small amount of cationized adducts with the matrix of type [M + DHB + X]+ (X = Na or K), along with ubiquitous [M + X]+ ions. The results are discussed by taking into account analyte-matrix association phenomena, such as hydrogen bond and inclusion phenomena, as a function of the molecular structure of the analyte. The conclusions derived by mass spectrometric data are compared with the X-ray diffraction data obtained on a single crystal of the 1:1 alpha-cyclodextrin - DHB noncovalent adduct.  相似文献   

12.
Selected benzoic acid derivatives and related substances were used as additives to 2,5-dihydroxybenzoic acid (2,5DHB) and the performance of the mixtures in matrix-assisted laser desorption/ionization mass spectrometry was investigated. Using benzoic acid derivatives substituted at position 2 and/or 5 or related substances as a co-matrix in the 1–10% range with 2,5DHB results in improved ion yields and signal-to-noise ratio of analyte molecules, especially for the high-mass range. The enhanced performance is prominent for 2-hydroxy-5-methoxybenzoic acid and exists for both proteins and oligosaccharides. It is suggested that the improvement is caused by a disorder in the 2,5DHB crystal lattice allowing ‘softer’ desorption. Charge transfer from matrix ions to additive molecules at the expense of analyte ionization gives a simple explanation for the deteriorating effects of some tested additives.  相似文献   

13.
Time‐of‐flight secondary ion mass spectrometry (ToF‐SIMS) can be used to image biological samples with nanometer‐scale resolution, albeit with the drawback that it often cannot detect large molecular signals. One way to increase secondary ion molecular yield is to chemically modify the surface in the so‐called matrix‐enhanced SIMS (ME‐SIMS) approach, which is based on embedding analyte molecules in low‐weight organic matrices. In this study, a solvent‐free sample preparation technique was employed using sublimation/deposition for coating a mouse brain section with a thin layer of a 2,5‐dihydroxybenzoic acid (DHB) matrix. Using this preparation technique, signal enhancements of up to a factor of 18 could be detected. It was found that the matrix layer thickness plays an important role in the efficiency of yield enhancement. Also, a complex influence of the matrix layer on various signals was observed. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

14.
The molecular structure of lipids in whole bacteria cells was characterized in detail by using two different and complementarily direct analyses; thermally assisted hydrolysis and methylation-gas chromatography (THM-GC) and matrix-assisted laser desorption/ionization mass spectrometry (MALDI-MS) combined with on-probe sample pretreatment. First, THM-GC in the presence of tetramethylammonium hydroxide (TMAH) was applied to compositional analysis of the fatty acid components of lipids in whole bacterial cells. On the resulting chromatograms, a series of fatty acid components in bacterial lipids were clearly observed as resolved peaks of their methyl esters. The fatty acid compositions determined on the basis of the peak intensities were in good agreement with those obtained by the conventional technique involving solvent extraction of the lipids in the samples. Furthermore, MALDI-MS combined with the on-probe sample treatment, using 2,5-dihydroxybenzoic acid (DHB) and sodium iodide (NaI) as matrix and cationization reagents, respectively, was used to detect intact phospholipids directly from whole bacterial cells. The MALDI spectra thus obtained showed an array of ions generated from bacterial phospholipids, such as phosphatidylethanolamines (PEs) and phosphatidylglycerols (PGs). Finally, the bacterial lipids were comprehensively characterized in terms of the acyl groups and the molecular structures by taking both of the results obtained by THM-GC and MALDI-MS into consideration.  相似文献   

15.
Conditions for the detection of three odd-electron cholesterol oxidation peaks were determined and these peaks were shown to be artifacts of the matrix-assisted laser desorption time of flight (MALDI-TOF) process. Matrix choice, solvent, laser intensity and cholesterol concentration were systematically varied to characterize the conditions leading to the highest signals of the radical cation peaks, and it was found that initial cholesterol solution concentration and resultant density of solid cholesterol on the MALDI target were important parameters in determining signal intensities. It is proposed that hydroxyl radicals, generated as a result of laser irradiation of the employed 2, 5-dihydroxybenzoic acid (DHB) matrix, initiate cholesterol oxidation on the MALDI target. An attempt to induce the odd-electron oxidation peaks by means of adding an oxidizing agent succeeded using an acetonitrile solution of DHB, cholesterol, and cumene hydroperoxide. Moreover, addition of free radical scavengers reduced the abundances of some oxidation products under certain conditions. These results are consistent with the mechanism of oxidation proposed herein involving laser-induced hydroxyl radical production followed by attack on neutral cholesterol. Hydroxyl radical production upon irradiation of dithranol matrix may also be responsible for generation of the same radical peaks observed from cholesterol in dithranol by an analogous mechanism.  相似文献   

16.
Matrix-assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-TOFMS), utilizing an on-probe sample pretreatment, was applied to the rapid and direct detection of intact phospholipids from whole bacterial cells. The sample preparation procedure involved depositing growing bacterial colonies from culture dishes directly onto the MALDI probe followed by treatment of the sample spot with a 3 micro L aliquot of an aqueous 0.05 M solution of sodium iodide prior to the addition of a 2,5-dihydroxybenzoic acid (DHB) matrix solution (ca. 8 mg dissolved in 70% acetonitrile/30% H(2)O containing 0.1% of trifluoroacetic acid). The MALDI spectra obtained from whole bacteria cells showed a series of ions generated from bacterial phospholipids, such as phosphatidylethanol-amines (PEs) and phosphatidylglycerols (PGs), which were clearly observed as well-resolved peaks. The ranges of the observed total carbon numbers in two acyl groups for PEs and PGs (30-36 and 33-36, respectively) were in good agreement with those reported previously. Furthermore, the distinct discrimination of four species of the Enterobacteriaceae family cultured identically was achieved by using principal components analysis (PCA) conducted on the relative peak intensities of phospholipids observed from the MALDI spectra.  相似文献   

17.
The properties of several cinnamic acid compounds used as matrices for matrix-assisted laser desorption ionization time-of-flight mass spectrometry (MALDI-TOF MS) were investigated as standard dried droplet (DD) and vacuum sublimed preparations. The differences between both preparation methods were analyzed with regard to matrix grain size, internal ion energy, initial velocity, analyte intensity, and analyte incorporation depth. Some of the used cinnamic acid derivatives exhibit clearly reduced grain sizes as sublimed preparations compared with standard DD approaches. In these cases higher effective temperatures could be measured accompanied by increased analyte intensities, which can be explained by stronger volatilization processes caused by a hindered heat dissipation resulting in a raised analyte transfer into the gas phase. For all sublimed compounds, a strong increase of the initial ion velocity compared with DD preparations could be measured. Higher initial ion velocities correlate with a decrease in internal ion energy which might be attributed to the very uniform crystal morphology exhibited by sublimed compounds. For sublimed matrices without reduced grain size, at least slightly higher analyte intensities could be detected at raised laser fluences. Analyte accumulation in the uppermost matrix layers or the detected higher ion stability can be explanations for these results.  相似文献   

18.
Ultraviolet resonance Raman spectroscopy (UVRRS), electronic absorption spectroscopy, and X-ray crystallography were used to probe the nature of the binding of 2,3-dihydroxybiphenyl (DHB) to the extradiol ring-cleavage enzyme, 2,3-dihydroxybiphenyl 1,2-dioxygenase (DHBD; EC 1.13.11.39). The lowest lying transitions in the electronic absorption spectrum of DHBD-bound DHB occurred at 299 nm, compared to 305 nm for the monoanionic DHB species in buffer. In contrast, the corresponding transitions in neutral and dianionic DHB occurred at 283 and 348 nm, respectively, indicating that DHBD-bound DHB is monoanionic. These binding-induced spectral changes, and the use of custom-designed optical fiber probes, facilitated UVRR experiments. The strongest feature of the UVRR spectrum of DHB was a Y8a-like mode around 1600 cm(-1), whose position depended strongly on the protonation state of the DHB. In the spectrum of the DHBD-bound species, this feature occurred at 1603 cm(-1), as observed in the spectrum of monoanionic DHB. Raman band shifts were observed in deuterated solvent, ruling out dianionic binding of the substrate. Thus, the electronic absorption and UVRRS data demonstrate that DHBD binds its catecholic substrate as a monoanion, definitively establishing this feature of the proposed mechanism of extradiol dioxygenases. This conclusion is supported by a crystal structure of the DHBD:DHB complex at 2.0 A resolution, which suggests that the substrate's 2-hydroxyl substituent, and not the 3-hydroxyl group, deprotonates upon binding. The structural data also show that the aromatic rings of the enzyme-bound DHB are essentially orthogonal to each other. Thus, the 6 nm blue shift of the transition for bound DHB relative to the monoanion in solution could indicate a conformational change upon binding. Catalytic roles of active site residues are proposed based on the structural data and previously proposed mechanistic schemes.  相似文献   

19.
Fragmentation of different generations of poly(amidoamine) dendrimers was explored in five common MALDI matrices: 2,5-dihydroxybenzoic acid (DHB), 4-hydroxy-3-methoxycinnamic acid (FER), α-cyano-4-hydroxycinnamic acid (ACH), 2,4,6-trihydroxyacetophenone (THAP), and 3-hydroxypicolinic acid (HPA). Of these, DHB was the softest matrix and ACH produced significant fragment intensity already at MALDI threshold, FER and THAP being in between. HPA was not a convenient matrix for dendrimers and produced a specific fragmentation pattern. Fragmentation analysis was mainly concentrated on generation G1, which contains already all essential structural elements. Dendrimers showed complicated fragmentation behavior with multiple fragmentation channels in our MALDI experiments. The relative intensities of these channels depended selectively on choice of the matrix and showed dissimilar dependence on the laser pulse energy. This was attributed to different fragmentation mechanisms, due to different protonation pathways, occurring in the same MALDI plume. The fragmentation pathways were proposed for all observed fragmentation channels. All fragmentation sites of protonated ions were found to be directly attached to the protonation sites and the fragmentation was surplus charge driven in this sense. No charge remote fragmentation channels were detected. Cationized dendrimers showed higher stability than the protonated ions.  相似文献   

20.
Application of matrix‐assisted laser‐desorption/ionization mass spectrometry (MALDI MS) to analysis and characterization of phosphopeptides in peptide mixtures may have a limitation, because of the lower ionizing efficiency of phosphopeptides than nonphosphorylated peptides in MALDI MS. In this work, a binary matrix that consists of two conventional matrices of 3‐hydroxypicolinic acid (3‐HPA) and α‐cyano‐4‐hydroxycinnamic acid (CCA) was tested for phosphopeptide analysis. 3‐HPA and CCA were found to be hot matrices, and 3‐HPA not as good as CCA and 2,5‐dihydroxybenzoic acid (DHB) for peptide analysis. However, the presence of 3‐HPA in the CCA solution with a volume ratio of 1:1 could significantly enhance ion signals for phosphopeptides in both positive‐ion and negative‐ion detection modes compared with the use of pure CCA or DHB, the most common phosphopeptide matrices. Higher signal intensities of phosphopeptides could be obtained with lower laser power using the binary matrix. Neutral loss of the phosphate group (?80 Da) and phosphoric acid (?98 Da) from the phosphorylated‐residue‐containing peptide ions with the binary matrix was decreased compared with CCA alone. In addition, since the crystal shape prepared with the binary matrix was more homogeneous than that prepared with DHB, searching for ‘sweet’ spots can be avoided. The sensitivity to detect singly or doubly phosphorylated peptides in peptide mixtures was higher than that obtained with pure CCA and as good as that obtained using DHB. We also used the binary matrix to detect the in‐solution tryptic digest of the crude casein extracted from commercially available low fat milk sample, and found six phosphopeptides to match the digestion products of casein, based on mass‐to‐charge values and LIFT TOF‐TOF spectra. Copyright © 2009 John Wiley & Sons, Ltd.  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