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 437 毫秒
1.
Methylation is an essential metabolic process in the biological systems, and it is significant for several biological reactions in living organisms. Methylated compounds are known to be involved in most of the bodily functions, and some of them serve as biomarkers. Theoretically, all α‐amino acids can be methylated, and it is possible to encounter them in most animal/plant samples. But the analytical data, especially the mass spectral data, are available only for a few of the methylated amino acids. Thus, it is essential to generate mass spectral data and to develop mass spectrometry methods for the identification of all possible methylated amino acids for future metabolomic studies. In this study, all N‐methyl and N,N‐dimethyl amino acids were synthesized by the methylation of α‐amino acids and characterized by a GC‐MS method. The methylated amino acids were derivatized with ethyl chloroformate and analyzed by GC‐MS under EI and methane/CI conditions. The EI mass spectra of ethyl chloroformate derivatives of N‐methyl ( 1–18 ) and N,N‐dimethyl amino acids ( 19–35 ) showed abundant [M‐COOC2H5]+ ions. The fragment ions due to loss of C2H4, CO2, (CO2 + C2H4) from [M‐COOC2H5]+ were of structure indicative for 1–18 . The EI spectra of 19–35 showed less number of fragment ions when compared with those of 1–18 . The side chain group (R) caused specific fragment ions characteristic to its structure. The methane/CI spectra of the studied compounds showed [M + H]+ ions to substantiate their molecular weights. The detected EI fragment ions were characteristic of the structure that made easy identification of the studied compounds, including isomeric/isobaric compounds. Fragmentation patterns of the studied compounds ( 1–35 ) were confirmed by high‐resolution mass spectra data and further substantiated by the data obtained from 13C2‐labeled glycines and N‐ethoxycarbonyl methoxy esters. The method was applied to human plasma samples for the identification of amino acids and methylated amino acids. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

2.
In the solid state, 4‐methoxy‐N′‐(2,2,2‐trichloroethanimidoyl)benzene‐1‐carboximidamide, C10H10Cl3N3O, (I), N′‐(2,2,2‐trichloroethanimidoyl)benzene‐1‐carboximidamide, C9H8Cl3N3, (II), 4‐chloro‐N′‐(2,2,2‐trichloroethanimidoyl)benzene‐1‐carboximidamide, C9H7Cl4N3, (III), 4‐bromo‐N′‐(2,2,2‐trichloroethanimidoyl)benzene‐1‐carboximidamide, C9H7BrCl3N3, (IV), and 4‐trifluoromethyl‐N′‐(2,2,2‐trichloroethanimidoyl)benzene‐1‐carboximidamide, C10H7Cl3F3N3, (V), display strong intramolecular N—H...N hydrogen bonding across the chelate ring and also intramolecular N—H...Cl contacts. Additional intermolecular hydrogen bonds link the molecules into chains, double chains or sheets in all cases except for compound (V). For compound (II), there are three independent molecules per asymmetric unit.  相似文献   

3.
Smog chamber/gas chromatography techniques are used to investigate the atmospheric degradation of fluroxene, an anesthetic, through oxidation with OH and Cl radicals at 298 K and under atmospheric pressure of N2 or air. The measured rate constants (k) are: k(fluroxene+OH.)=(2.96±0.61)×10−11 and k(fluroxene+Cl.)=(1.62±0.19)×10−10 cm3 molecule−1 s−1. The only product detected after the oxidation of fluroxene with OH radicals is 2,2,2-trifluoroethyl formate (79 % and 83 % molar yield in the absence and presence of NOx, respectively). However, after oxidation with Cl radicals, the detected products are 2,2,2-trifluoroethyl formate (78 %), 2,2,2-trifluoroethyl-1-chloroacetate (5 %), and chloroacetaldehyde (4 %), in the absence of NOx, and 2,2,2-trifluoroethyl formate (93 %), 2,2,2-trifluoroethyl-1-chloroacetate (6 %), and chloroacetaldehyde (5 %), in the presence of NOx. The results indicate that, both in the absence and presence of NOx, the main fate of fluroxene is the addition of the oxidant to the double bond and, once the alkoxy radical is formed, the main decomposition pathway is by means of degradation. Moreover, it is expected that 2,2,2-trifluoroethyl formate is the only oxidation product able to actively contribute to climate change. To successfully assess the contribution of fluroxene to global warming, we measure the infrared spectra of fluroxene and 2,2,2-trifluoroethyl formate, and calculate the radiative efficiencies (REs) to be 0.27 and 0.28 W m−2 ppbv−1, respectively. In addition, the cumulative effect owing to the formation of 2,2,2-trifluoroethyl formate is investigated, and the direct, indirect, and net global-warming potentials are calculated by using the REs and lifetimes of fluroxene and 2,2,2-trifluoroethyl formate.  相似文献   

4.
In 2,2,2‐trichloro‐N,N′‐bis(4‐methoxyphenyl)ethane‐1,1‐diamine, C16H17Cl3N2O2, molecules are linked into helical chains by N—H...O hydrogen bonds. Molecules of 2,2,2‐trichloro‐N,N′‐bis(4‐chlorophenyl)ethane‐1,1‐diamine, C14H11Cl5N2, are connected into a three‐dimensional framework by two independent Cl...Cl interactions and one C—H...Cl hydrogen bond.  相似文献   

5.
The molecules of 2,2,2‐trichloro‐N,N′‐diphenylethane‐1,1‐diamine, C14H13Cl3N2, are linked into (040) sheets by a combination of C—H...Cl and C—H...π(arene) hydrogen bonds. In 2,2,2‐trichloro‐N,N′‐bis(4‐methylphenyl)ethane‐1,1‐diamine, C16H17Cl3N2, the molecules are linked into C(7) chains by two independent C—H...Cl hydrogen bonds and one Cl...Cl contact.  相似文献   

6.
A path to new synthons for application in crystal engineering is the replacement of a strong hydrogen‐bond acceptor, like a C=O group, with a weaker acceptor, like a C=S group, in doubly or triply hydrogen‐bonded synthons. For instance, if the C=O group at the 2‐position of barbituric acid is changed into a C=S group, 2‐thiobarbituric acid is obtained. Each of the compounds comprises two ADA hydrogen‐bonding sites (D = donor and A = acceptor). We report the results of cocrystallization experiments of barbituric acid and 2‐thiobarbituric acid, respectively, with 2,4‐diaminopyrimidine, which contains a complementary DAD hydrogen‐bonding site and is therefore capable of forming an ADA/DAD synthon with barbituric acid and 2‐thiobarbituric acid. In addition, pure 2,4‐diaminopyrimidine was crystallized in order to study its preferred hydrogen‐bonding motifs. The experiments yielded one ansolvate of 2,4‐diaminopyrimidine (pyrimidine‐2,4‐diamine, DAPY), C4H6N4, (I), three solvates of DAPY, namely 2,4‐diaminopyrimidine–1,4‐dioxane (2/1), 2C4H6N4·C4H8O2, (II), 2,4‐diaminopyrimidine–N,N‐dimethylacetamide (1/1), C4H6N4·C4H9NO, (III), and 2,4‐diaminopyrimidine–1‐methylpyrrolidin‐2‐one (1/1), C4H6N4·C5H9NO, (IV), one salt of barbituric acid, viz. 2,4‐diaminopyrimidinium barbiturate (barbiturate is 2,4,6‐trioxopyrimidin‐5‐ide), C4H7N4+·C4H3N2O3, (V), and two solvated salts of 2‐thiobarbituric acid, viz. 2,4‐diaminopyrimidinium 2‐thiobarbiturate–N,N‐dimethylformamide (1/2) (2‐thiobarbiturate is 4,6‐dioxo‐2‐sulfanylidenepyrimidin‐5‐ide), C4H7N4+·C4H3N2O2S·2C3H7NO, (VI), and 2,4‐diaminopyrimidinium 2‐thiobarbiturate–N,N‐dimethylacetamide (1/2), C4H7N4+·C4H3N2O2S·2C4H9NO, (VII). The ADA/DAD synthon was succesfully formed in the salt of barbituric acid, i.e. (V), as well as in the salts of 2‐thiobarbituric acid, i.e. (VI) and (VII). In the crystal structures of 2,4‐diaminopyrimidine, i.e. (I)–(IV), R22(8) N—H…N hydrogen‐bond motifs are preferred and, in two structures, additional R32(8) patterns were observed.  相似文献   

7.
Novel methods for the determination of inorganic oxyanions by electrospray (ES) ionization mass spectrometry have been developed using dehydration reactions between oxyanions and carboxylic acids at the ES interface. Twelve oxyanions (VO3?, CrO42?, MoO42?, WO42?, BO33?, SiO32?, SiO44?, AsO44?, AsO2?, SeO42?, SeO32? and NO2?), out of 16 tested, reacted with at least one of four aminopolycarboxylic acids, i.e. iminodiacetic acid (IDA), nitrilotriacetic acid (NTA), trans‐1,2‐diaminocyclohexane‐N,N,N′,N′‐tetraacetic acid and triethylenetetramine‐N,N,N′,N″,N′″,N′″‐hexaacetic acid, at the ES interface to produce the dehydration products that gave intense mass ion responses, sufficient for trace analysis. As examples, trace determinations of CrVI and silica in water samples were achieved after online ion exchange chromatography, where the dehydration product of CrO42? and NTA (m/z 290) and that of SiO44? and IDA (m/z 192) were measured. The limits of detection of the respective methods were 17 nM (0.83 ng Cr/ml) for CrVI and 0.17 μM (4.8 ng Si/mL) for SiO44?. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

8.
Summary: Carboxylic acids were efficiently activated with N,N′‐carbonyldiimidazole (CDI) and applied for the acylation of cellulose under homogeneous conditions using dimethyl sulfoxide (DMSO)/tetrabutylammonium fluoride trihydrate (TBAF) as solvent. The simple and elegant method is a very mild and easily applicable tool for the synthesis of pure aliphatic, alicyclic, bulky, and unsaturated cellulose esters with degrees of substitution of up to 1.9. Products are soluble in organic solvents, e.g., DMSO or N,N‐dimethylformamide (DMF). The cellulose esters were characterized by elemental analysis, FT‐IR, 1H and 13C NMR spectroscopy and show no impurities or substructures resulting from side reactions.

The esterification of cellulose using carboxylic acids activated in situ with N,N′‐carbonyldiimidazole.  相似文献   


9.
Two D‐π‐A‐type 2,2,2‐trifluoroacetophenone derivatives, namely, 4′‐(4‐( N,N‐diphenyl)amino‐phenyl)‐phenyl‐2,2,2‐trifluoroacetophenone (PI‐Ben) and 4′‐(4‐(7‐(N,N‐diphenylamino)‐9,9‐dimethyl‐9H‐fluoren‐2‐yl)‐phenyl‐2,2,2‐trifluoroacetophenone (PI‐Flu), are developed as high‐performance photoinitiators combined with an amine or an iodonium salt for both the free‐radical polymerization of acrylates and the cationic polymerization of epoxides and vinyl ether upon exposure to near‐UV and visible light‐emitting diodes (LEDs; e.g., 365, 385, 405, and 450 nm). The photochemical mechanisms are investigated by UV‐Vis spectra, molecular‐orbital calculations, fluorescence, cyclic voltammetry, photolysis, and electron‐spin‐resonance spin‐trapping techniques. Compared with 2,2,2‐trifluoroacetophenone, both photoinitiators exhibit larger redshift of the absorption spectra and higher molar‐extinction coefficients. PI‐Ben and PI‐Flu themselves can produce free radicals to initiate the polymerization of acrylate without any added hydrogen donor. These novel D‐π‐A type trifluoroacetophenone‐based photoinitiating systems exhibit good efficiencies (acrylate conversion = 48%–66%; epoxide conversion = 85%–95%; LEDs at 365–450 nm exposure) even in low‐concentration initiators (0.5%, w/w) and very low curing light intensities (1–2 mW cm?2). © 2016 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2016 , 54, 1945–1954  相似文献   

10.
Antonios Kolocouris 《Tetrahedron》2009,65(45):9428-9435
Dynamic NMR spectroscopy and ab initio correlated calculations revealed that the attachment of a spiroadamantane entity at the C-2 position of N-methylpyrrolidine or N-methylpiperidine induces a severe steric crowding around nitrogen, which changes the conformational space of the heterocycle resulting in: (a) the complete destabilization of the N-Me(eq) conformer in spiranic structures; in contrast the N-Me(eq) conformer corresponds to the global minimum in N-methylpyrrolidine or N-methylpiperidine. The spiroadamantane structure raises the energy of the equatorial conformer because of the severe van der Waals repulsion between the N-Me(eq) group and adamantane C-H bonds. (b) The interconversion between the only populated enantiomeric N-Me(ax) conformers ax→[eq]→ax′; the interconversion eq→ax between N-Me(eq) and N-Me(ax) conformers, which are both populated, is observed in N-methylpyrrolidine or N-methylpiperidine. (c) The raising of ring and nitrogen inversion barriers ax→ts by ∼4-6 kcal mol−1. The dynamic NMR study provides evidence that the most important process required for the enantiomerization between the axial N-Me conformers in spiropiperidine 4 and spiropyrrolidine 5 are different, i.e., a nitrogen inversion in 5 (9.10 kcal mol−1) and a ring inversion in 4 (15.2 kcal mol−1). While an enantiomerization interconverts N-Me axial conformers in spiropiperidine 5 and spiropyrrolidine 4, substitution of the pyrrolidine ring of 5 with a C-Me group effects a diastereomerization between two N-Me axial conformers and reduces effectively the nitrogen inversion barrier according to the protonation experiments and the calculations. In general, all the calculations levels used, i.e., the MM3, B3LYP/6-31+G∗∗ and MP2/6-311++G∗∗//B3LYP/6-31+G∗∗, predict correctly the different stability of the local minima; however only MP2/6-311++G∗∗//B3LYP/6-31+G∗∗ was found to be reliable for the calculation of the nitrogen inversion barriers.  相似文献   

11.
An sp 2 /sp 3 get‐together : A novel and efficient method can be used to synthesize 3,3‐disubstitued oxindoles by the direct intramolecular oxidative coupling of an aryl C? H and a C? H center (see scheme; DMF=N,N‐dimethylformamide).

  相似文献   


12.
In order to examine the preferred hydrogen‐bonding pattern of various uracil derivatives, namely 5‐(hydroxymethyl)uracil, 5‐carboxyuracil and 5‐carboxy‐2‐thiouracil, and for a conformational study, crystallization experiments yielded eight different structures: 5‐(hydroxymethyl)uracil, C5H6N2O3, (I), 5‐carboxyuracil–N,N‐dimethylformamide (1/1), C5H4N2O4·C3H7NO, (II), 5‐carboxyuracil–dimethyl sulfoxide (1/1), C5H4N2O4·C2H6OS, (III), 5‐carboxyuracil–N,N‐dimethylacetamide (1/1), C5H4N2O4·C4H9NO, (IV), 5‐carboxy‐2‐thiouracil–N,N‐dimethylformamide (1/1), C5H4N2O3S·C3H7NO, (V), 5‐carboxy‐2‐thiouracil–dimethyl sulfoxide (1/1), C5H4N2O3S·C2H6OS, (VI), 5‐carboxy‐2‐thiouracil–1,4‐dioxane (2/3), 2C5H4N2O3S·3C6H12O3, (VII), and 5‐carboxy‐2‐thiouracil, C10H8N4O6S2, (VIII). While the six solvated structures, i.e. (II)–(VII), contain intramolecular S(6) O—H…O hydrogen‐bond motifs between the carboxy and carbonyl groups, the usually favoured R22(8) pattern between two carboxy groups is formed in the solvent‐free structure, i.e. (VIII). Further R22(8) hydrogen‐bond motifs involving either two N—H…O or two N—H…S hydrogen bonds were observed in three crystal structures, namely (I), (IV) and (VIII). In all eight structures, the residue at the ring 5‐position shows a coplanar arrangement with respect to the pyrimidine ring which is in agreement with a search of the Cambridge Structural Database for six‐membered cyclic compounds containing a carboxy group. The search confirmed that coplanarity between the carboxy group and the cyclic residue is strongly favoured.  相似文献   

13.
Gas chromatographic analysis of ethyl chloroformate derivatives of samples taken from the paint layers of post-Byzantine panel paintings permitted the successful characterisation of the different binding media used in them. This paper describes an analytical study of various post-Byzantine binding media such as egg yolk and egg/oil emulsion, using gas chromatography. The characterisation of these icons’ binding media is an important task, as it contributes to our understanding of and the reconstruction of the post-Byzantine artists’ palette. It also enables us to investigate the validity of our assumptions about the influences of Venetian style on Greek icon painting techniques from the sixteenth to the early nineteenth century, which up to now have been based on information in artists’ handbooks. The methodology involves two experimental steps: (1) hydrolysis of the proteins and triglycerides in the binding media to obtain free amino acids and fatty acids, and (2) the formation of ethyl chloroformate derivatives via derivatization with ethyl chloroformate (ECF). This methodology is of considerable interest, since it permits the identifcation of the nature of the proteinaceous binders used in these works through the simultaneous derivatization and determination of amino acids and fatty acids. Advantages of this methodology include the small quantity of sample required and the minimum preparation time involved. The proteinaceous media can be determined based on the ratios of seven stable amino acids, while the type of emulsions and drying oils used can be determined from the fatty acid ratio. Electronic supplementary material Supplementary material is available in the online version of this article at and is accessible for authorized users.  相似文献   

14.
An approach towards precision NMR measurements of four‐bond deuterium isotope effects on the chemical shifts of backbone amide nitrogen nuclei in proteins is described. Three types of four‐bond 15 N deuterium isotope effects are distinguished depending on the site of proton‐to‐deuterium substitution: 4ΔN(Ni‐1D), 4ΔN(Ni+1D) and 4ΔN(Cβ,i‐1D). All the three types of isotope shifts are quantified in the (partially) deuterated protein ubiquitin. The 4ΔN(Ni+1D) and 4ΔN(Cβ,i‐1D) effects are by far the largest in magnitude and vary between 16 and 75 ppb and ?18 and 46 ppb, respectively. A semi‐quantitative correlation between experimental 4ΔN(Ni+1D) and 4ΔN(Cβ,i‐1D) values and the distances between nitrogen nuclei and the sites of 1H‐to‐D substitution is noted. The largest isotope shifts in both cases correspond to the shortest inter‐nuclear distances. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

15.
肖锋  罗宇  吕伟  汤杰 《有机化学》2010,30(2):311-313
10-羟基喜树碱首先在N,N-二甲基甲酰胺(DMF)中经N-溴代丁二酰亚胺(NBS)溴代得到9-溴-10-羟基喜树碱, 9-溴-10-羟基喜树碱和氯甲酸乙酯反应得到9-溴-10-羟基喜树碱-10,20-双乙氧基碳酸酯(6). 化合物6和烯丙基三正丁基锡通过Stille偶联反应[9]得到关键中间体7, 最后水解化合物7得到目标化合物. 通过柱层析纯化得到纯度大于99.8%, 单杂小于0.1%的吉咪替康(HPLC). 所有中间体及目标产物经1H NMR, 13C NMR, LRMS, HRMS表征确证.  相似文献   

16.
A reaction with phenol and pyrocatechol of N-(2,2,2-trichloroethylidene)arenesulfonyl-, ethoxycarbonylamides and 1-hydroxy-substituted N-(2,2,2-trichloroethyl)amides of arenesulfonic, carbamic, and acetic acids in the presence of oleum or in sulfuric acid provided the corresponding (1-amido-2,2,2-trichloroethyl)-substituted phenols. N-(2,2,2-Trichloroethylidene)-4-chlorobenzenesulfonamide reacted with salicylamide in the presence of oleum to afford 3-aminocarbonyl-4-[2,2,2-trichloro-1-(4-chlorobenzenesulfonamido)ethyl]benzene whereas the 1-hydroxy-2,2,2-trichloroethylamides of the acetic, carbamic, and arenesulfonic acids did not enter into such reactions.  相似文献   

17.
《合成通讯》2013,43(23):4013-4018
Abstract

Several N-methoxy-N-methylamides were prepared by the reaction of the corresponding carboxylic acids with N,O-dimethylhydroxylamine hydrochloride at room temperature using trichloromethyl chloroformate in the presence of triethylamine in excellent yields.  相似文献   

18.
We report a range of new transformations of the diamide–amine supported Ti?NNPh2 functional group with a variety of unsaturated substrates, along with DFT studies of the key mechanisms. Reaction of [Ti(N2Npy)(NNPh2)(py)] ( 4 , N2Npy=(2‐NC5H4)CMe(CH2NSiMe3)2; py=pyridine) with MeCN gave the dimeric species [Ti2(N2Npy)2{μ‐NC(Me)(NNPh2)}2] through a [2+2] cycloaddition process. Reaction of 4 or [Ti(N2NMe)(NNPh2)(py)] ( 5 , N2NMe=MeN(CH2CH2NSiMe3)2) with fluorinated benzonitriles gave the terminal hydrazonamide complexes [Ti(N2NR){NC(Ar)NNPh2}(py)] (R=py or Me; Ar=2,6‐C6H3F2 or C6F5). DFT studies showed that this proceeds through an overall [2+2] cycloaddition–reverse cycloaddition, resulting in net insertion of ArCN into the Ti?Nα bonds of the respective hydrazides. Reaction of 4 with a mixture of MeCN and PhCCMe gave the metallacycle [Ti(N2Npy){NC(Me)C(Ph)C(Me)NNPh2}] by sequential coupling of Ti?NNPh2 with PhCCMe and then MeCN. A related product, [Ti(N2Npy){NC(Me)C(ArF)C(H)NNPh2}], was formed by insertion of MeCN into the Ti? C bond of the isolated azatitanacyclobutene [Ti(N2Npy){N(NPh2)C(H)C(ArF)}] (ArF=3‐C6H4F). Reaction of 4 with two equivalents of B(Ar)3 (Ar=C6F5) formed the zwitterionic borate [Ti(N2Npy){η2‐N(NPh2)B(Ar)3}] by electrophilic attack at Nα. Compounds 4 and 5 reacted with tBuNC and/or XylNC (Xyl=2,6‐C6H3Me2) to give the Nα? Nβ bond cleavage products, [Ti(N2NR)(NCNR′)(NPh2)] (R=py or Me; R′=tBu or Xyl), containing metallated carbodiimide ligands. DFT studies of these reactions found an initial addition of RNC across Ti?Nα followed by Nβ coordination, and finally complete Nα transfer from the NNPh2 to the RNC fragment. Reaction of 5 with Ar′NCE (E=O, S, Se; Ar′=2,6‐C6H3iPr2) gave the [2+2] cycloaddition products [Ti(N2NMe){N(NPh2)C(NAr′)O}(py)] and [Ti(N2NMe){N(NPh2)C(NAr′)E}] (E=S or Se), which did not undergo further transformation of the Ti? N? NPh2 moiety.  相似文献   

19.
The complex [Pd(O,N,C‐L)(OAc)], in which L is a monoanionic pincer ligand derived from 2,6‐diacetylpyridine, reacts with 2‐iodobenzoic acid at room temperature to afford the very stable pair of PdIV complexes (OC‐6‐54)‐ and (OC‐6‐26)‐[Pd(O,N,C‐L)(O,C‐C6H4CO2‐2)I] (1.5:1 molar ratio, at ?55 °C). These complexes and the PdII species [Pd(O,N,C‐L)(OX)] and [Pd(O,N,C‐L′)(NCMe)]ClO4, (X=MeC(O) or ClO3, L′=another monoanionic pincer ligand derived from 2,6‐diacetylpyridine), are precatalysts for the arylation of CH2?CHR (R?CO2Me, CO2Et, Ph) using IC6H4CO2H‐2 and AgClO4. These catalytic reactions have been studied and a tentative mechanism is proposed. The presence of two PdIV complexes was detected by ESI(+)‐MS during the catalytic process. All the data obtained strongly support a PdII/PdIV catalytic cycle.  相似文献   

20.
The title compound, C34H38Cl6O14·C4H8O2·H2O, prepared by the reaction of 10‐deacetyl baccatin III with 2,2,2‐trichloroethyl chloroformate in pyridine, crystallizes via strong intermolecular hydrogen bonds and noncovalent interactions between 7,10‐bis‐O‐(2,2,2‐trichloroethoxycarbonyl)‐10‐deacetyl baccatin III (7,10‐di‐Troc‐DAB), water and ethyl acetate. A detailed comparison of the molecular conformation with those of related structures is presented.  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