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 15 毫秒
1.
β-Cryptoxanthin ( 1 ) was acetylated and then epoxidized with monoperoxyphthalic acid. After hydrolysis, repeated chromatography, and crystallization, (3S,5R,6S)-5,6-epoxy-β-cryptoxanthin ( 3 ), (3S,5S,6R)-5,6-epoxy-β-cryptoxanthin ( 4 ), (3R,5′R,6′R)-5′,6′-epoxy-β-cryptoxanthin ( 5 ), (3S,5R,6S,5′R,6′S)-5,6:5′,6′-diepoxy-β-cryp-toxanthin ( 6 ), and (3S,5S,6R,5′S,6′R)-5,6:5′,6′-diepoxy-β-cryptoxanthin ( 7 ) were isolated as main products and characterized by their UV/VIS, CD, 1H- and 13C-NMR, and mass spectra. The comparison of the carotenoid isolated from yellow, tomato-shaped paprika (Capsicum annuum var. lycopersiciforme flavum) with 3–5 strongly supports the structure of 3 for the natural product.  相似文献   

2.
Valuable substrates for the synthesis of natural products , compounds 3 (R1=alkyl, aryl, alkoxy; R2, R3=alkyl) are formed from β,γ-unsaturated α-keto esters 1 and vinyl ethers 2 by the title reaction [Eq. (1)]. Copper(II ) bisoxazolines act as catalysts, and in many cases enantiomeric excesses higher than 99.5 % are achieved.  相似文献   

3.
In the presence of titanium(IV) tetraethoxide ((EtO)4Ti), menthyl arylglyoxylates are prepared by transesterification of ethyl arylglyoxylates and natural (−)‐(1R,2S,5R)‐menthol. Using menthyl as a chiral auxiliary, the corresponding novel (R)‐menthyl 2‐aryl‐2‐hydroxybutanoates are synthesized by the addition of Et2Zn with menthyl arylglyoxylates. The structures of the products are characterized by IR and 1H‐ and 13C‐NMR spectroscopy, mass spectrometry, and elemental analysis. The diastereoselectivities are analyzed by HPLC. The addition reactions are completed with good yields and high diastereoisomeric excess (de up to 95%), and, after hydrolysis, the (R)‐2‐aryl‐2‐hydroxybutanoic acids are obtained with high optical purities.  相似文献   

4.
Starting from d‐mannose, d‐galactose and d‐glucosamine hydrochloride, two disaccharide blocks were synthesized. Schmidt's inverse addition technique of trichloroacetimidate was utilized for the construction of a disaccharide with a β‐mannosidic linkage in good yield. The other disaccharide had a methyl 4,6‐(R)‐pyruvate ester. The two disaccharides in the appropriate form were then allowed to react in the presence of N‐iodosuccinimide (NIS) and trifluoromethanesulfonic acid (TfOH) to give the desired tetrasaccharide derivative, 2‐(trimethylsilyl)ethyl 2‐acetamido‐3,4,6‐tri‐O‐benzoyl‐2‐deoxy‐β‐d‐glucopyranosyl‐(1→3)‐2‐O‐benzoyl‐4,6‐O‐[(R)‐1‐methoxycarbonylethylidene]‐β‐d‐galactopyranosyl‐(l→4)‐2,3,6‐tri‐O‐benzyl‐β‐d‐mannopyranosyl‐(1→4)‐2,6‐di‐O‐benzyl‐3‐O‐(4‐methoxybenzyl)‐β‐d‐galactopyranoside.  相似文献   

5.
Three new 1,2-dihydro-1-oxabenz[a]azulen-2-one derivatives, 1a (R1=H, R2=Me), 1b (R1=H, R2=Ph), and 1c (R1=COOEt, R2=Me), have been synthesized by the reaction of 2-hydroxyazulene (2a) and its 1-ethoxycarbonyl derivative 2b with ethyl acetoacetate (3a) or ethyl benzoylacetate (3b) in the presence of aluminum chloride. To our knowledge, these are the first examples of this type of compound, although the yield of the products is low in some cases. Their electronic properties were studied in detail utilizing the analyses of 1,2-dihydro-1-oxabenz[a]azulen-2-one derivative 1a by the spectroscopic and voltammetric analyses. The analyses revealed that the fused α-pyrone system lowers both the HOMO and the LUMO energies, relative to those of parent azulene (10), but has much pronounced effect on the LUMO, consequently, leading to decrease in HOMO–LUMO gap, compared with those of 10. These results should be attracted to the development of amphoteric redox materials. Reactivity toward electrophilic reagents was also examined by bromination and Vilsmeier–Haack formylation reactions of 1a. To evaluate the scope of the reaction products we have examined Sonogashira cross-coupling reaction of the bromination products with trimethylsilylacetylene and conversion of the formylation product to dibromoolefin by the reaction with phosphorous ylide prepared with CBr4 and Ph3P. Effective extension of the π-electron system in the ethynyl products has been revealed by the spectroscopic analysis. These reaction products would be attracted to the application as a terminal group for electronic applications.  相似文献   

6.
The enantioselective syntheses of the compounds with the assigned structures of the helibisabonols A and B have been accomplished. Using an enzymatic desymmetrization of the σ-symmetrical diol (route a) and a diastereoselective conjugate addition of the methyl to the enone with a chiral auxiliary (route b) we constructed the key tertiary stereogenic center at the benzylic position (C7) and then used an asymmetric dihydroxylation for assembling the C10 stereogenic center. In addition, possible diastereoisomers of the natural products were prepared and detailed comparisons of the 1H and 13C NMR spectra were conducted. As a result, the structure originally assigned to helibisabonol A may be revised to (7R,10R)-1. In the case of helibisabonol B, the (7R,10R)-2 would be reasonable based on a comparison of the NMR data and the biogenetic parallelism with helibisabonol A.  相似文献   

7.
Kinetic and equilibrium NMR studies of briarane diterpenes isolated from the pennatulacean coral Funiculina quadrangularis showed that funicolide A ( 1 ), funicolide D ( 5 ), and brianthein W ( 6 ), with R2 = Hβ, undergo slow flipping by rotation of the C(1)? C(2)? C(3)? C(4) dihedral angle, giving rise to two observable conformers in a 4:96 population ratio, both having an axial AcO? C(14) and C(16) pointing ‘downwards’, but differing for pseudoaxial or pseudoequatorial position of R1O, respectively. δ(C) for the minor conformers could be quickly assigned by an original emulation methodology. Similar studies revealed that funicolide B ( 2 ), 7-epifunicolide A ( 4 ), funicolide E ( 7 ), and unnatural epibrianthein W ( 8 ) undergo similar motions, where, however, the nature of the α-positioned substituent R2 determines which conformer predominates: axial R1O for R2 = Hβ ( 4 and 8 ) or equatorial R1O for R2 α-OH, ( 2 and 7 ). In contrast, funicolide C ( 3 ) proved to undergo slow conformational motions that involve also the cyclohexene ring, resulting in two observable conformers characterized by either an equatorial AcO? C(14) and trans-diaxial C(2)/C(9) or an axial AcO? C(14) and trans-diequatorial C(2)/C(9) in a 9:1 population ratio, respectively. These observations, and molecular-mechanics calculations for briaranes known to exhibit broad NMR signals, lead to general views on the conformational preferences of diterpenes of this class.  相似文献   

8.
A new family of optically active cyclophane receptors for the complexation of mono‐ and disaccharides in competitive protic solvent mixtures is described. Macrocycles (−)‐(R,R,R,R)‐ 1 – 4 feature preorganized binding cavities formed by four 1,1′‐binaphthalene‐2,2′‐diyl phosphate moieties bridged in the 3,3′‐positions by acetylenic or phenylacetylenic spacers. The four phosphodiester groups converge towards the binding cavity and provide efficient bidentate ionic H‐bond acceptor sites (Fig. 2). Benzyloxy groups in the 7,7′‐positions of the 1,1′‐binaphthalene moieties ensure solubility of the nanometer‐sized receptors and prevent undesirable aggregation. The construction of the macrocyclic framework of the four cyclophanes takes advantage of Pd0‐catalyzed aryl—acetylene cross‐coupling by the Sonogashira protocol, and oxidative acetylenic homo‐coupling methodology (Schemes 2 and 8 – 10). Several cleft‐type receptors featuring one 1,1′‐binaphthalene‐2,2′‐diyl phosphate moiety were also prepared (Schemes 1, 6, and 7). An undesired side reaction encountered during the synthesis of the target compounds was the formation of naptho[b]furan rings from 3‐ethynylnaphthalene‐2‐ol derivatives, proceeding via 5‐endo‐dig cyclization (Schemes 35). Computer‐assisted molecular modeling indicated that the macrocycles prefer nonplanar puckered, cyclobutane‐type conformations (Figs. 7 and 8). According to these calculations, receptor (−)‐(R,R,R,R)‐ 1 has, on average, a square binding site, which is complementary in size to one monosaccharide. The three other cyclophanes (−)‐(R,R,R,R)‐ 2 – 4 feature, on average, wider rectangular cavities, providing a good fit to one disaccharide, while being too large for the complexation of one monosaccharide. This substrate selectivity was fully confirmed in 1H‐NMR binding titrations. The chiroptical properties of the cyclophanes and their nonmacrocyclic precursors were investigated by circular dichroism (CD) spectroscopy. The CD spectra of the acyclic precursors showed a large dependence from the number of 1,1′‐binaphthalene moieties (Fig. 9), and those of the cyclophanes were remarkably influenced by the nature of the functional groups lining the macrocyclic cavity (Fig. 11). Profound differences were also observed between the CD spectra of linear and macrocyclic tetrakis(1,1′‐binaphthalene) scaffolds, which feature very different molecular shapes (Fig. 10). In 1H‐NMR binding titrations with mono‐ and disaccharides (Fig. 13), concentration ranges were chosen to favor 1 : 1 host−guest binding. This stoichiometry was experimentally established by the curve‐fitting analysis of the titration data and by Job plots. The titration data demonstrate conclusively that the strength of carbohydrate recognition is enhanced with an increasing number of bidentate ionic host−guest H‐bonds (Table 1) in the complex formed. As a result of the formation of these highly stable H‐bonds, carbohydrate complexation in competitive protic solvent mixtures becomes more favorable. Thus, cleft‐type receptors (−)‐(R)‐ 7 and (−)‐(R)‐ 38 with one phosphodiester moiety form weak 1 : 1 complexes only in CD3CN. In contrast, macrocycle (−)‐(R,R,R,R)‐ 1 with four phosphodiester groups undergoes stable inclusion complexation with monosaccharides in CD3CN containing 2% CD3OD. With their larger number of H‐bonding sites, disaccharide substrates bind even more strongly to the four phosphodiester groups lining the cavity of (−)‐(R,R,R,R)‐ 2 and complexation becomes efficient in CD3CN containing 12% CD3OD. Finally, the introduction of two additional methyl ester residues further enhances the receptor capacity of (−)‐(R,R,R,R)‐ 3 , and efficient disaccharide complexation occurs already in CD3CN containing 20% CD3OD.  相似文献   

9.
A new disaccharide block for OSW-1 natural steroidal antitumor agent was described. Regioisomeric 2- and 3-O-p-methoxybenzoyl derivatives of phenyl 1-thio-??-d-xylopyranoside and phenyl 2-O-acetyl-1-thio-??-l-arabinopyranoside derivatives blocked at positions 3 and 4 by R3Si groups were synthesized with a view to use them in the preparation of OSW-1 analogs modified at the disaccharide fragment.  相似文献   

10.
The triazenols 4-R1? C6H4? N?N? N(OH)? R2 ( 1 ), oxidized with t-BuO radicals, produced nitroxide radicals R1? C6H4? N(O?)? N?N(R2) +O? ( 5 ). The suggested radical structure was confirmed by 15N-labeling. The reaction of triazenols 1 with PbO2 proceeded under N2 elimination, in which case nitroxides R1? C6H4? N(R2)? O?( 2 ) were observed as the final radical products. The intermediate R1? C6H radicals were identified by spin-trapping.  相似文献   

11.
All four stereoisomers of (E)-vitamin Kb i. e. (21E, 7R, 111R)-l (= 1a), (21E, 71 R, l11S)-1 (= 1b), (21E, 71 S, 111S) 1 ( = 1c), and (21E, 71S, 111R)-l ( = Id), have been synthesized in a state of high chemical and stereoisomeric purity. The synthesis of stereoisomers lb-d relied on the use of the optically active Cf1* and C*10-building blocks (R)- or (S)-4-(benzyloxy)-3-methylbutanal ((R)- or (S)-2) and (R)- or (S)-citronellal ((R)- or (S)- 3 ) which had been secured by the Rh1-catalyzed allylamine-to-enamine isomerization technology. For the synthesis of the natural (E)-vitamin-K1 stereoisomer 1a , a new route starting from natural phylol was developed, based on an O-alkylation/rearrangement procedure. A HPLC method was developed which separates with remarkable efficiency all four stereoisomers of (E)- as well as three out of the four stereoisomers of (Z)-vitamin K1 on optically active poly(trityl methacrylate) as the chiral stationary phase supported on Nucleosil. By this method, the stereoisomeric content of the stereoisomers 1b-d synthesized was shown to be in the range of 96-98 %, while the natural isomer 1a was configurationally uniform. The biological activity of the four (E)-vitamin-K1 stereoisomers was determined by means of the curative prothrombin time test with vitamin-K-depleted chicks. A high precision of the results was obtained with the recently introduced up-and-down organization of the test and the statistical evaluation according to an estimation procedure. With the natural (E)-vitamin-K1 stereoisomer 1a as standard (set at 1. 0), activities of 0. 93, 1. 19, and 0. 99 were found for stereoisomers 1b, 1c , and 1d , respectively. Within the confidence limits, these activity ratios can be regarded as identical, A very similar efficacy was obtained by comparison of (E, all-rac )-vitamin K1 ((21E, RS, 11′ RS)- 1 ; equimolar mixture of the four stereoisomers 1a-d) with the natural (E)-vitamin-K1 stereoisomer 1a ). A synergistic effect was not detectable, as was the case with the eight α-tocopheryl-acetate stereoisomers.  相似文献   

12.
The sex pheromone of the endoparasitoid insect Xenos peckii (Strepsiptera: Xenidae) was recently identified as (7E,11E)‐3,5,9,11‐tetramethyl‐7,11‐tridecadienal. Herein we report the asymmetric synthesis of three candidate stereostructures for this pheromone using a synthetic strategy that relies on an sp3–sp2 Suzuki–Miyaura coupling to construct the correctly configured C7‐alkene function. Comparison of 1H NMR spectra derived from the candidate stereostructures to that of the natural sex pheromone indicated a relative configuration of (3R*,5S*,9R*). Chiral gas chromatographic (GC) analyses of these compounds supported an assignment of (3R,5S,9R) for the natural product. Furthermore, in a 16‐replicate field experiment, traps baited with the synthetic (3R,5S,9R)‐enantiomer alone or in combination with the (3S,5R,9S)‐enantiomer captured 23 and 18 X. peckii males, respectively (mean±SE: 1.4±0.33 and 1.1±0.39), whereas traps baited with the synthetic (3S,5R,9S)‐enantiomer or a solvent control yielded no captures of males. These strong field trapping data, in combination with spectroscopic and chiral GC data, unambiguously demonstrate that (3R,5S,9R,7E,11E)‐3,5,9,11‐tetramethyl‐7,11‐tridecadienal is the X. peckii sex pheromone.  相似文献   

13.
From cultures of a Phoma species (strain S 298) the hitherto unknown metabolite deoxaphomine has been isolated. On the basis of the spectral data, structure 1 of the (7S, 16R, 20R)-7,20-dihydroxy-16-methyl-10-phenyl-[13]cytochalasa-6(12),13t,21t-triene-1,23-dione is assigned to deoxaphomine. This structure is confirmed by the chemical degradation of 1 , yielding the products 4 and 6 which are identical with derivatives of phomine ( 2 )((7S,16R,20R)-7,20-dihydroxy-16-methyl-10-phenyl-24-oxa-[14]cytochalasa-6(12), 13t,21t-triene-1,23-dione) and cytohalasin D ( 8 ) ((7S,16S,18R,21R)-21-acetoxy-7,18-dihydroxy-16,18-dimethyl-10-phenyl-[11]sytochalasa-6(12),13t,19t-triene-1,17-dione). Deoxaphomine ( 1 ) is a potential biogenetic precursor of the 24-oxa-[14]cytochalasans. Preliminary results of the biological activity of deoxaphomine are reported.  相似文献   

14.
Existing data on the self-reactions of tertiary peroxy radicals RO2 has been reanalyzed and corrected to deduce Arrhenius parameters for both termination and nontermination paths. For R = t-Butyl, these are logkt(M?1sec?1) = 7.1 - (7.0/θ) and logknt(M?1sec?1) = 9.4 - (9.0/θ), respectively, different from those recommended by other authors. The higher magnitudes observed for termination processes of tertiary peroxy radicals like those of cumyl and 1,1-diphenylethyl have been discussed in terms of a much greater cage recombination of cumyloxy radicals as contrasted with t-butoxy radicals. It is shown that for benzyl peroxy radicals, the R—O bond dissociation energy is sufficiently low (18–20 kcal) that reversible dissociation into R˙ + O2 opens a competing second-order path to fast recombination R˙ + RO → ROOR. This path is probably not important for cumyl peroxy radicals under usual experimental conditions but can become important for 1,1-diphenyl ethyl peroxy radicals at (O2) < 10?3M. At very low RO concentrations (<10?5M), in the absence of added O2, an apparent first-order disappearance of RO can occur reflecting the rate determining breaking of the cumyl—O bond followed by the second step above. The thermochemistry of RO is used to show that the reaction of R2O4 → 2RO + O2 must be concerted and cannot proceed via RO which is too unstable and cannot form even from RO˙ + O2.  相似文献   

15.
王长生  高坤  张艳  刘阳 《化学学报》2005,63(15):1383-1390
使用B3LYP方法研究了发生在有机钼化合物R3R4Mo(≡CH)(CHR1R2)和R3R4Mo(=CH2)(=CR1R2)之间的α-氢转移反应, 探讨了R1, R2, R3和R4位置上不同取代基对α-氢转移反应势垒和产物稳定性的影响. 研究发现, 金属钼有机化合物中, 发生α-氢转移的碳原子在过渡态中采用sp2杂化. R1和R2位置上取代基对α-氢转移反应势垒的影响取决于取代基对过渡态中碳原子的未参与sp2杂化的pz轨道上单电子的离域作用. 当R1, R2位置是甲基时, 由于碳原子的pz轨道与甲基的一个C—H键轨道间存在强的超共轭效应, 从而可以较大程度地降低α-氢转移反应的势垒. 研究结果还表明, 当R3, R4位置为SiH3时的反应势垒较低. 所以当R1和R2 位置为Me, R3和R4位置为SiH3时, 反应势垒最低. 第一个甲基取代R1或R2位置的H时, 反应势垒降低很大; 第二个甲基继续取代时, 反应势垒的降低约为第一个甲基的一半. 第一个SiH3取代R3或R4位置的甲基时, 反应势垒降低较大; 第二个SiH3继续取代时, 反应势垒的降低小于第一个SiH3的一半. 对反应物和产物的相对稳定性的研究表明, 第一个甲基和第二个甲基对产物的相对能量的降低几乎相同; 第一个SiH3降低产物的相对能量, 但是第二个SiH3使产物的相对能量升高, 从而抵消了第一个SiH3对产物的稳定作用.  相似文献   

16.
(Z)-1,6-Anhydro-3-deoxy-4-methylsulfanyl-3-[(methylsulfanyl)methylene]-β-D-erythro-hexopyranos-2-ulose (1) reacted with diethyl malonate, 1,3-diketones, N-aryl-3-oxobutyramides and dialkyl 3-oxoglutarate, respectively, in the presence of potassium carbonate and crown ether to yield diethyl 2-(1,6-anhydro-4-methylsulfanyl—D-arabino-hex-2-ulopyranos-3-ylmethylene) malonate (2), 1-{(1R,2S,8S,9R)-2-hydroxy-4-methyl-8-methylthio-3,11,12- trioxatricyclo7.2.1.02,7dodeca-4,6-dien-5-yl} ethanone (3), (1R,2S,12S,13R)-2-hydroxy-12-methylthio-3,15,16-trioxatetracyclo[11.2.1. 02,11. 04,9] hexadeca- 4(9),10-dien-8-one (4), (1R,8S,9R)-5-acetyl-3-aryl-8-methylthio-11,12-dioxa- 3-azatricyclo-[7.2.1.02,7]dodeca-2(7),5-dien-4-ones (5,6) and dialkyl (1R,8S,-9R)-4-hydroxy-8-methylthio-11,12-dioxatricyclo[7.2.1.02,7]dodeca-2(7),3,5-triene-3,5-dicarboxylates (7,8), respectively.  相似文献   

17.
Four stereoisomers of nodulisporacid A (1) were synthesized by the concise approach which includes three-component reaction and subsequent one-pot construction of the whole framework. The 1H NMR comparison of the derivatives (10-12) revealed the absolute configuration of natural 1 to be 4R,4′R,6′R.  相似文献   

18.
4,5-Dihydro-3H-naphtho[1,8-bc]furans 4 and 6 which have various substituents (R1 and R2) have been synthesized from 8-oxo-5,6,7,8-tetrahydro-1-naphthyloxyacetic acids 1 and 3 or their ethyl esters 2 . The reaction of acids 1 and 3 with sodium acetate in acetic anhydride gave a mixture of furans 4 and 6 and lactones 5 and 7 . The ratios of the products were varied according to the types of substituents (R1 and R2) in acids 1 and 3 . As the substituent R1 (R2 = hydrogen) in acids 1 was changed from hydrogen to a methyl, ethyl or isopropyl group, production of furans 4 became more difficult. However, when a phenyl group was used as the substituent, furan 4 was obtained in good yield. Similarly, as the substituent R2 (R1 = hydrogen) in acids 1 was changed from hydrogen to a methyl, ethyl or isopropyl group, furan formation was more difficult. In contrast, acids 3 which had electron-withdrawing substituents such as chlorine, bromine or a nitro group at the 4-position afforded furans 6 in good yield. 4,5-Dihydro-3H-naphtho[1,8-bc]furans 4 and 4,5-dihydro-3H-naphtho[1,8-bc]furan-2-carbocylic acids 8 were synthesized from the reaction of esters 2 and potassium hydroxide in dioxane. When the substituents R1 or R2 in esters 2 were varied from hydrogen to a methyl, ethyl or isopropyl group the total yields of furans 4 and furancarboxylic acids 8 were reduced.  相似文献   

19.
Abstract

Oligosaccharide components of glycoconjugates play a fundamental role in recognition phenomena and intercellular interactions. Disaccharide H, 2-O-(α-L-fucopyranosyl)-D-galactopyranose, is one such biologically important, L-fucose containing oligosaccharide.1 This disaccharide is found in natural antigenic determinants such as blood group determinants and tumor associated antigens.1,2  相似文献   

20.
Modular polyketide synthases such as 6‐deoxyerythronolide B synthase (DEBS) catalyze the biosynthesis of structurally complex natural products. Streptomyces coelicolor CH999/pJRJ2 harbors a plasmid encoding DEBS(KS10), a mutant form of 6‐deoxyerythronolide B synthase that is blocked in the formation of 6‐deoxyerythronolide B ( 1 , 6‐dEB) due to a mutation in the active site of the ketosynthase (KS1) domain that normally catalyzes the first polyketide chain‐elongation step of 6‐dEB biosynthesis. Administration of (2S,3R,4S)‐ and (2S,3R,4R)‐3‐hydroxy‐2,4‐dimethylhexanoic acid N‐acetylcysteamine (SNAC) thioesters (= S‐[2‐(acetylamino)ethyl] (2S,3R,4S)‐ and (2S,3R,4R)‐3‐hydroxy‐2,4‐dimethylhexanethioates) 3 and 4 in separate experiments to cultures of Streptomyces coelicolor CH999/pJRJ2 led to production of the corresponding (14S)‐ and (14R)‐14‐methyl analogues of 6‐dEB, 10 and 11 , respectively. Unexpectedly, when a 3 : 2 mixture of 4 and 3 was fed under the same conditions, exclusively branched‐chain macrolactone 11 was isolated. In similar experiments, feeding of 3 and 4 to S. coelicolor CH999/pCK16, an engineered strain harboring DEBS1+TE(KS10), resulted in formation of the branched‐chain triketide lactones 13 and 14 , while feeding of the 3 : 2 mixture of 4 and 3 gave exclusively 14 . The biochemical basis for this stereochemical discrimination was established by using purified DEBS module 2+TE to determine the steady‐state kinetic parameters for 3 and 4 , with the kcat/KM for 4 shown to be sevenfold greater than that of 3 .  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