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 109 毫秒
1.
As a part of continuing interest in the zip reaction, we present the results on a carbon ring-enlargement reaction of activated ketones with a CN group as a charge stabilizer. Two series of (1-cyano-2-oxocy-cloalkyl)alkanoates were prepared from 8- and 12-membered cyano-ketones 1 and 2 , respectively, namely the propanoates 3 and 4 , the butanoates 6 , 8 and 9 as well as the pentanoates 12 and 15 . While treatment with t-BuOK of the former two homologous esters resulted in both ring enlargement and competitive transesterification, the pentanoates 12 and 15 afforded mostly the diastereoisomeric mixtures of bicyclic alcohols 20a – c and 31a , b , respectively, which remained intact on further exposure to base. It was shown that – apart from the base used (t-BuOK) vs. Li(i-Pr)2N – the distribution of products was greatly influenced by the ring size of substrates. This is further illustrated by treatment of ketones 34 and 35 with t-BuOK. While the former rearranged smoothly to diketone 36 , no reaction at all took place with the latter. The behavior of the substrates is discussed in terms of steric and energetic reasons.  相似文献   

2.
The racemic spirosesquiterpenes β-acorenol ( 1 ), β-acoradiene ( 2 ), acorenone-B ( 3 ) and acorenone ( 4 ) (Scheme 2) have been synthesized in a simple, flexible and highly stereoselective manner from the ester 5 . The key step (Schemes 3 and 4), an intramolecular thermal ene reaction of the 1,6-diene 6 , proceeded with 100% endo-selectivity to give the separable and interconvertible epimers 7a and 7b . Transformation of the ‘trans’-ester 7a to (±)- 1 and (±)- 2 via the enone 9 (Scheme 5) involved either a thermal retro-ene reaction 10 → 12 or, alternatively, an acid-catalysed elimination 11 → 13 + 14 followed by conversion to the 2-propanols 16 and 17 and their reduction with sodium in ammonia into 1 which was then dehydrated to 2 . The conversion of the ‘cis’-ester 7b to either 3 (Scheme 6) or 4 (Scheme 7) was accomplished by transforming firstly the carbethoxy group to an isopropyl group via 7b → 18 → 19 → 20 , oxidation of 20 to 21 , then alkylative 1,2-enone transposition 21 → 22 → 23 → 3 . By regioselective hydroboration and oxidation, the same precursor 20 gave a single ketone 25 which was subjected to the regioselective sulfenylation-alkylation-desulfenylation sequence 25 → 26 → 27 → 4 .  相似文献   

3.
Addition of various amines to the 3,3‐bis(trifluoromethyl)acrylamides 10a and 10b gave the tripeptides 11a – 11f , mostly as mixtures of epimers (Scheme 3). The crystalline tripeptide 11f 2 was found to be the N‐terminal (2‐hydroxyethoxy)‐substituted (R,S,S)‐ester HOCH2CH2O‐D ‐Val(F6)‐MeLeu‐Ala‐OtBu by X‐ray crystallography. The C‐terminal‐protected tripeptide 11f 2 was condensed with the N‐terminus octapeptide 2b to the depsipeptide 12a which was thermally rearranged to the undecapeptide 13a (Scheme 4). The condensation of the epimeric tripeptide 11f 1 with the octapeptide 2b gave the undecapeptide 13b directly. The undecapeptides 13a and 13b were fully deprotected and cyclized to the [5‐[4,4,4,4′,4′,4′‐hexafluoro‐N‐(2‐hydroxyethoxy)‐D ‐valine]]‐ and [5‐[4,4,4,4′,4′,4′‐hexafluoro‐N‐(2‐hydroxyethoxy)‐L ‐valine]]cyclosporins 14a and 14b , respectively (Scheme 5). Rate differences observed for the thermal rearrangements of 12a to 13a and of 12b to 13b are discussed.  相似文献   

4.
Formal Total Synthesis of (±)-Isocomen by Application of the α-Alkinon Cyclization A total synthesis of the racemic form of the sesquiterpene isocomene ( A ) was accomplished by application of the cyclopentenone anellation B→D (Scheme 1) which includes the α-alkynone cyclization C→D , a gas-phase flow thermolytic process. Starting with the known product 2 (Scheme 3) of the anellation B→D , the elaboration of ring C of A proceeded in 9 steps to the α-alkynone 16 (Scheme 5) which was cyclized at 540° selectively to give the angularly fused triquinane 4 (77%). A two-step procedure then led to 5 (Scheme 6), a last but one intermediate in a known total synthesis of (±)- A . The conversion of 16 to 4 also demonstrated the compatibility of an acetoxy function with the anellation sequence B→D .  相似文献   

5.
It is shown that 4- or 8-[(E)-styryl]-substituted azulenes can easily be prepared from 4- or 8-methylazulenes in the presence of potassium tert-butoxide (t-BuOK) with the corresponding benzaldehydes in tetrahydrofuran (THF) at −5 to 25° (see Schemes 1 and 2). 6-(tert-Butyl)-4,8-dimethylazulene ( 5 ) with both Me groups in reactive positions leads to the formation of a mixture of the mono- and distyryl-substituted azulenes 6 and 7 , respectively (Scheme 3). Vilsmeier formylation of 6 results in the formation of 3 : 2 mixture of the azulene-carbaldehydes 8a and 8b , which can be separated by chromatography on silica gel. Reduction of 8a and 8b with NaBH4 in trifluoroacetic acid (TFA)/CH2Cl2 gives the 1-methyl forms 9a and 9b , respectively, in good yields (Scheme 4). The latter two azulenes are not separable on silica gel.  相似文献   

6.
Synthetic attempts towards fully conjugated polymers 9 with pentafulvene-diyl structural units are described. Cationic polymerization of pentafulvenes 1a (R = X = Me) and 1b (R = X = MeS) nearly quantitatively gives polymers 8a and 8b with typical Mn and Mw values of 38800 and 53750, respectively, for 8a , and 12000 and 35900, respectively, for 8b . Key step of the conversion 8a → 9a (Scheme 6) is a quantitative bromination 8a → 32a , the structure of 32a being confirmed by analytical data as well as by spectroscopic comparison with model compound 23 . Best results in view of two-fold the HBr elimination 32a → 9a are obtained with Et3N, but so far elimination has not been complete. Synthetic sequences are optimized with model compound 21 (Scheme 4). Here again, bromination 21 → 23 is quantitative, while two-fold HBr elimination 23 → 22 with Et3N proceeds in 51% yield. Dibromide 23 easily undergoes HBr elimination followed by a Br shift to give bromide 29 . Contrary to cationic polymerization, anionic polymerization of simple pentafulvenes 1 to 2 (which would be attractive in view of the conjugated polymers 3 ) is not successful: For pentafulvene 1b (R = X = MeS), the main reaction is Diels-Alder-type dimerization 1b → 15b (Scheme 2), even under anionic conditions.  相似文献   

7.
A mechanistic model is presented for the base‐catalyzed intramolecular cyclization of polycyclic unsaturated alcohols of type A to ethers D (Scheme 1). The alkoxide anion B is formed first in a fast acid‐base equilibrium. For the subsequent reaction to D , a carbanion‐like transition state C is proposed. This mechanism is in full agreement with our results regarding the influence of substituents on the regioselectivity and the rate of cyclization. We studied the effect of alkyl substituents in allylic position (alkylated endocylic olefinic alcohols 1 – 3 ) and, especially, at the exocyclic double bond ( 12 – 15 ). The fastest cyclization (krel=1) is 12 → 16 , which proceeds via a primary carbanion‐like transition state ( E : R1=R2=H). The corresponding processes 13 → 17 and 14 → 17 are characterized by a less‐stable secondary carbanion‐like transtition state ( E : R1=Me, R2=H, or vice versa) and are slower by a factor of 104. The slowest reaction (krel ca. 10−6) is the cyclization 15 → 18 via a tertiary carbanion‐like transition state ( E : R1=R2=Me).  相似文献   

8.
The antibiotic myxovirescine M2 was synthesized from seven building blocks ( 1 – 7 , Scheme 1), with the following chiral starting materials being employed: (S)-malic acid, (+)-D -ribonolactone, (S)-2-(hydroxymethyl)butanoate, and (2R,4S)-5-hydroxy-2,4-dimethylpenLanoate. Three new nucleophilic reagents, 8 – 10 , for C-C bond formation have been used. The key steps of the synthesis are: a Suzuki coupling between an alkyl borane and a vinyl bromide ( 4 + 12e → 13 ), a Julia olefinalion ( 14 + 17 → 18 ), and a Yamaguchi macrolactonizalion to form the 28-membered lactone ( 18 → 19 ), This extremely convergenl synthetic approach will allow the preparation of a number of the 31 known myxovirescine molecules.  相似文献   

9.
The reaction of 2,2-dimethyl-5-(1,2-epoxypropyl)cyclohexanone ( 7 ) with t-BuOK in DMSO furnished a small amount of 5-(1-hydroxyprop-2-enyl)-2,2-dimethylcyclohexanone ( 12 ) and the 4 unexpected products 13–16 which contain one to three additional C-atoms (Scheme 2). The relative configuration of the major product 1-(4′,4′-dimethyl-2′,3′-dimethylidenecyclohexyl)propane-1,2-diol ( 15 ) was shown to be 1RS, 2RS,1′SR via NOE measurements performed on a derivative thereof. A crossover experiment in DMSO/[13C2]DMSO 1:1 as solvent showed that the two additional C-atoms of this product originate from a single molecule of DMSO (Scheme 5). A tentative mechanistic scheme, consistent with all experimental observations, is proposed which involves a [2,3]-sigmatropic rearrangement of an (allylsulfinyl)methanide to a sulfenic acid as one of the key steps ( V → 24 , Scheme 8). We corroborated part of this hypothetic scheme by taking recourse to a model compound (7-(methylsulfinyl)-p-mentha-1,8-diene ( 32/33 ), readily prepared in two steps from perilla alcohol ( 30 )), which reacted as predicted by the proposed mechanism (Schemes 9 and 10).  相似文献   

10.
Parent fulvenes and fulvalenes are thermally unstable cross‐conjugated olefins for which low‐temperature syntheses are indispensable. In this review 5 syntheses (in the temperature range between ?100 and ?10°) are discussed: 1. Reaction of sodium cyclopentadienide with 1‐acetoxy‐1‐chloroalkanes or 1‐acetoxy‐1‐bromoalkanes ( 26 ) gives acetoxy‐alkyl‐cyclopentadienes ( 27 ) which are easily converted to pentafulvenes ( 2 ) by low‐temperature HOAc‐elimination with NEt3. This synthesis has been applied to parent pentafulvene ( 2a ), heptafulvene ( 3a ), nonafulvene ( 4a ) and sesquifulvalene ( 19a ) (Schemes 811). 2. Based on a nearly quantitative oxidative coupling of cyclononatetraenide ( 8 ) to give dihydrononafulvalene ( 38 ) (Scheme 10), a general synthetic plan for fulvalenes has been outlined (Scheme 11) and applied to the synthesis of pentafulvalene ( 12 ), nonapentafulvalene ( 16 ) and nonafulvalene ( 14 ). Several applications of oxidative couplings of Hückel anions are discussed (Schemes 20 and 21). 3. Trifunctional cyclopropanes 67 (in most cases 1,1‐dibromo‐2‐X‐cyclopropanes) are attractive precursors of parent triafulvene ( 1a ) and calicene ( 17 ) (Scheme 18). Contrary to classical procedures they are transformed into nucleophiles ( 67 → 68 ) by halogen‐lithium exchange, methylation ( 68 → 69 ) and HBr‐elimination to give 1‐methylidene‐2‐X‐cyclopropanes of type 71 . By subsequent HX‐elimination triafulvene ( 1a ) has been synthesized and trapped as a [4+2]‐cycloadduct 73 (Scheme 20). Furthermore, calicene precursors 77 are available by using cyclopentenone as an electrophilic cyclopentadiene equivalent. 4. Similarly, 1‐lithio‐1‐bromo‐2‐X‐cyclopropanes 68 are directly transformed into triafulvalene precursors 81 (Scheme 26) by a novel CuCl2‐catalyzed oxidative coupling. 5. In view of the synthesis of parent triafulvene ( 1a ), triafulvalene ( 11 ) and calicene ( 17 ), retro‐Diels? Alder reactions of stable precursors – prepared by low‐temperature reactions (described in chapters 3 and 4 ) – have been explored.  相似文献   

11.
The reactions of 1,4,5-trisubstituted imidazole 3-oxides 1a – k with cyclobutanethiones 5a , b in CHCl3 at room temperature give imidazole-2(3H)-thiones 9a – k in high yield. The second product formed in this reaction is 2,2,4,4-tetramethylcyclobutane-1,3-dione ( 6a ; Scheme 2). Similar reactions occur with 1 and adamantanethione ( 5c ) as thiocarbonyl compound, as well as with 1,2,4-triazole-4-oxide derivative 10 and 5a (Scheme 3). A reaction mechanism by a two-step formation of the formal cycloadduct of type 7 via zwitterion 16 is proposed in Scheme 5. Spontaneous decomposition of 7 yields the products of this novel sulfur-transfer reaction. The starting imidazole 3-oxides are conveniently prepared by heating a mixture of 1,3,5-trisubstituted hexahydro-1,3,5-triazines 3 and α-(hydroxyimino) ketones 2 in EtOH (cf. Scheme 1). As demonstrated in the case of 9d , a `one-pot' procedure allows the preparation of 9 without isolation of the imidazole 3-oxides 1 . The reaction of 1c with thioketene 12 leads to a mixture of four products (Scheme 4). The minor products, 9c and the ketene 15 , result from an analogous sulfur-transfer reaction (Path a in Scheme 5), whereas the parent imidazole 14 and thiiranone 13 are the products of an oxygen-transfer reaction (Path b in Scheme 5).  相似文献   

12.
To demonstrate the relevance of the kinetic acidity of individual OH groups for the regioselectivity of glycosylation by glycosylidene carbenes, we compared the glycosylation by 1 of the known triol 2 with the glycosylation of the diol D - 3 and the fluorodiol L - 4 . Deoxygenation with Bu3SnH of the phenoxythiocarbonyl derivative of 5 (Scheme 1) or the carbonothioate 6 gave the racemic alcohol (±)- 7 . The enantiomers were separated via the allophanates 9a and 9b , and desilylated to the deoxydiols D - and L - 3 , respectively. The assignment of their absolute configuration is based upon the CD spectra of the bis(4-bromobenzoates) D - and L - 10 . The (+)-(R)-1-phenylethylcarbamates 13a and 13b (Scheme 2) were prepared from the fluoroinositol (±)- 11 via (±)- 4 and the silyl ether (±)- 12 and separated by chromatography. The absolute configuration of 13a was established by X-ray analysis. Decarbamoylation of 13a ( → L - 12 ) and desilylation afforded the fluorodiol L - 4 . The H-bonds of D - 3 and L - 4 in chlorinated solvents and in dioxane were studied by IR and 1H-NMR spectroscopy (Fig. 2). In both diols, HO? C(2) forms an intramolecular, bifurcated H-bond. There is an intramolecular H-bond between HO? C(6) and F in solutions of L - 4 in CH2Cl2, but not in 1,4-dioxane; the solubility of L - 4 in CH2Cl2 is too low to permit a meaningful glycosidation in this solvent. Glycosidation of D - 3 in dioxane by the carbene derived from 1 (Scheme 3) followed by acetylation gave predominantly the pseudodisaccharides 18/19 (38%), derived from glycosidation of the axial OH group besides the pseudodisaccharides 16 / 17 (13%) and the epoxides 20 / 21 (7%), derived from protonation of the carbene by the equatorial OH group. Similarly, the reaction of L - 4 with 1 (Scheme 4) led to the pseudodisaccharides 28 / 29 (46%) and 26 / 27 (14%), derived from deprotonation of the axial and equatorial OH groups, respectively. Formation of the epoxides involved deprotonation of the intramolecularly H-bonded tautomer, followed by intramolecular alkylation, elimination, and substitution (Scheme 4). The regio- and diastereoselectivities of the glycosidation correlate with the H-bonds in the starting diols.  相似文献   

13.
Attempted Synthesis of Calicene from Trisubstitued Cyclopropanes and Cyclopentenone The Li carbenoids 4 , prepared by treatment of substituted 1,1-dihalocyclopropanes with BuLi, are reacted with cyclopent-2-enone under thermodynamic and kinetic control (Scheme 1). In general, the latter procedure gives better yields of cyclopropylcyclopentenols 5a – e , but the reaction seems to be controlled mainly by the steric and electronic properties of the substituent Y. So, with 4b and 4e , the main reaction is the attack of the carbenoid at C(1) of cyclopent-2-enone, while 4a (Y = PhS) predominantly deprotonates the ketone (Scheme 4). Whereas 5d and 5e can easily be converted to the dihydrocalicenes 6d and 6e (Scheme 6), the attempted elimination of H2O from 5a – c leads to the rearranged products 13 – 2 due to the opening of the cyclopropane ring (Scheme 5). Finally, the generation of the parent compound 2 from the silylated precursor 6d is attempted: treatment with MeO? gives the addition products 18A/18B , while the reaction with Br2 provides 19 by a bromination/dehydrobromination sequence (Scheme 7).  相似文献   

14.
Three new dammarane monodesmosides, named notoginsenosides Ft1 ( 1 ), Ft2 ( 2 ), and Ft3 ( 3 ), together with three known ginsenosides, were obtained from a mild acidic hydrolysis of the saponins from notoginseng (Panax notoginseng (Burk .) F. H. Chen ) leaves. Their structures were elucidated to be (3β,12β,20R)‐12,20‐dihydroxydammar‐24‐en‐3‐yl O‐β‐D ‐xylopyranosyl‐(1 → 2)‐O‐β‐D ‐glucopyranosyl‐(1 → 2)‐β‐D ‐glucopyranoside ( 1 ), (3β,12β)‐12,20,25‐trihydroxydammaran‐3‐yl O‐β‐D ‐xylopyranosyl‐(1 → 2)‐O‐β‐D ‐glucopyranosyl‐(1 → 2)‐β‐D ‐glucopyranoside ( 2 ), and (3β,12β,24ξ)‐12,20,24‐trihydroxydammar‐25‐en‐3‐yl O‐β‐D ‐xylopyranosyl‐(1 → 2)‐O‐β‐D ‐glucopyranosyl‐(1 → 2)‐β‐D ‐glucopyranoside ( 3 ), by means of spectroscopic evidences. The known ginsenosides Rh2 and Rg3 4 – 6 were obtained as the major products from this acidic deglycosylation.  相似文献   

15.
The [1,1′‐biisoquinoline]‐4,4′‐diol ( 4a ), which was obtained as hydrochloride 4a ?2 HCl in two steps starting from the methoxymethyl (MOM)‐protected 1‐chloroisoquinoline 8 (Scheme 3), opens access to further O‐functionalized biisoquinoline derivatives. Compound 4a ?2 HCl was esterified with 4‐(hexadecyloxy)benzoyl chloride ( 5b ) to give the corresponding diester 3b (Scheme 4), which could not be obtained by Ni‐mediated homocoupling of 6b (Scheme 2). The ether derivative 2b was accessible in good yield by reaction of 4a ?2 HCl with the respective alkyl bromide 9 under the conditions of Williamson etherification (Scheme 4). Slightly modified conditions were applied to the esterification of 4a ?2 HCl with galloyl chlorides 10a – h as well as etherification of 4a ?2 HCl with 6‐bromohexyl tris(alkyloxy)benzoates 11b , d – h and [(6‐bromohexyl)oxy]‐substituted pentakis(alkyloxy)triphenylenes 14a – c (Scheme 5). Despite the bulky substituents, the respective target 1,1′‐biisoquinolines 12, 13 , and 15 were isolated in 14–86% yield (Table).  相似文献   

16.
Heterodiamantanes and Structurally Related Compounds. Part III. The Pentacyclic C11-Diethers 5, 13-Dioxapentacyclo [6.5.0.02,6.03,12.04,9]tridecane, 4, 13-Dioxapentacyclo [6.4.1.02,7.03,10.05,9]tridecane, and 3, 10-Dioxapentacyclo [7.3.1.02,7.04,12.06,11]tridecane In connection with the studies on heterodiamantanes and structurally related compounds the three novel pentacyclic diethers 3 – 5 were prepared starting from the cyclopentadienone dimer 6 . All four compounds have as common features a central carbocyclic 6-membered ring with four axial alkyl substituents and two oxygen functions in 1, 4 position. The required eleventh C-atom was introduced by dichlorocarbene addition either to 6 ( → 7 ) (Scheme 2) or to 29 ( → 28 ) (Scheme 4). Diether 3 was obtained by reduction of 26 (Scheme 2), a suitable precursor prepared either by intramolecular addition ( 24 → 26 ; Scheme 2) or substitution ( 30 → 26 , 31 → 26 ; Scheme 4), as well as by direct substitution ( 44 → 3 , 42 → 3 ; Scheme 5). Diether 4 was the product of a direct substitution ( 39 → 4 , 36 → 4 ; Scheme 5). The synthesis of diether 5 was achieved from the addition product 51 (resulting from the alcohols 47 and 48 ; Scheme 6). Diether 4 is the thermodynamically least stable of the three diethers 3 – 5 . It was easily isomerized to 5 on treatment with concentrated sulfuric acid in benzene whereas 3 and 5 remained unchanged under these conditions.  相似文献   

17.
The phenyl‐substituted conduritol B 8 was prepared in racemic form in a five‐step sequence starting from 2‐phenyl‐1,4‐benzoquinone ( 10 ) (Scheme 1). The reaction mechanism of the key step 12b → 13 is discussed (Scheme 2).  相似文献   

18.
The 6‐aminopyrazin‐2(1H)‐one, when incorporated as a pyrimidine‐base analog into an oligonucleotide chain, presents a H‐bond donor? donor? acceptor pattern to a complementary DNA or RNA strand. When paired with the corresponding acceptor? acceptor? donor purine in oligonucleotides, the heterocycle selectively contributes to the stability of the duplex, presumably by forming a base pair of Watson? Crick geometry joined by a nonstandard H‐bonding pattern, expanding the genetic alphabet. Reported here is a short, high yielding, β‐D ‐selective synthesis of a 6‐aminopyrazin‐2(1H)‐one nucleoside via the glycine riboside derivative 28 . The key steps include a Wittig? Horner reaction of an appropriately protected ribose derivative (Scheme 10, 19 → 21 ) followed by a Michael‐like ring closure (Scheme 12, 30 → 1a and 32 → 1b ). Thus, a variety of pyrazine nucleosides (Scheme 13) including the target 6‐aminopyrazin‐2(1H)‐one riboside 1a , and its 5‐methyl derivative 1b , 6‐amino‐5‐methylpyrazin‐2(1H)‐one riboside, are obtained.  相似文献   

19.
Photochemical behaviors of the pyrazinone derivatives 5,6,7,8-tetrahydroquinoxalin-2(1H)-ones 1a – c and 1,5,6,7,8,9-hexahydro-2H-cyclohepta[b]pyrazin-2-one 1d were investigated. Dye-sensitized photo-oxygenation of 1a-c gave the 1:1 adducts 5a – c of the corresponding 3,8a-epidioxy-3,5,6,7,8,8a-hexahydroquinoxalin-2(1H)-one 4 and H2O, whereas 1d gave 3,9a-epidioxy-1,3,5,6,7,8,9,9a-octahydro-2H-cyclohepta[b]pyrazin-2-one 4d (Scheme 2). The different kind of products was interpreted as being the result of the ring strain and steric hindrance of endoperoxides produced from 1a – d with singlet oxygen. Irradiation of 1a – b in the presence of alkenes gave tricyclic azetidine derivatives 9 by [2 + 2] cycloaddition of the C?N bond of 1 to the alkene.  相似文献   

20.
[(1,3-Dioxolan-2-ylidene)methyl]phosphonates and -phosphinates as [simple] Synthons in Heterocyclic Synthesis The readily available [(1,3-dioxolane-2-ylidene)methyl]phosphonates and -phosphinates 2a–f (Scheme 1) can be transformed with amines to aliphatic ketene N,O-and N,N-acetales (see Scheme 2, 2a → 3–7 ). Alkanediamines yield with 2a–f the imidazolidines 8a–f and the hexahydropyrimidines 9a–d (Scheme 3). the oxazolidine derivatives 10a–e and the thiazolidine 11 are accessible under special reaction conditions starting from 2a, b (Scheme 4). Hydrazines react with the CN-group-containing ketene O,O-acetals 2a–c to the pyrazoles 12a–g , whereof 12a, d, e can be cyclized to pyrazolo[1,5-a]pyrimidines 13a–d (Scheme 5). Amidines as starting materials transform 2a–c in an analogous way to the pyrimidine derivatives 14a–c (Scheme 6).  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