首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 62 毫秒
1.
The reaction of the intermediate ketene N,Se‐hemiacetal 3 , prepared from cyanomethylene derivatives 1 by treatment with Et3N and aryl isoselenocyanates 2 , with bis‐electrophiles 6, 7, 9 , and 11 in DMF affords tetrahydro‐1H‐1,3‐selenazine (=1,3‐selenazinane) derivatives 8, 10 , and 12 in good yield (Scheme 2 and Tables 1–3). Chemical and spectroscopic evidence for the structures of the new compounds are described. The structures of 8d and 12e are established by X‐ray crystallography (Figs. 1 and 2).  相似文献   

2.
Pyrromethene–BF2 complexes (P–BF2) 7 were obtained from α-unsubstituted pyrroles 5 by acylation and condensation to give intermediate pyrromethene hydrohalides 6 followed by treatment with boron trifluoride etherate. Conversion of ethyl α-pyrrolecarboxylates 4 to α-unsubstituted pyrroles 5 was brought about by thermolysis in phosphoric acid at 160°C, or by saponification followed by decarboxylation in ethanolamine at 180°C, or as unisolated intermediates in the conversion of esters 4 to pyrromethene hydrobromides 6 by heating in a mixture of formic and hydrobromic acids. Addition of hydrogen cyanide followed by dehydrogenation by treatment with bromine converted 3,5,3′,5′-tetramethyl-4,4′-diethylpyrromethene hydrobromide 9 to 3,5,-3′,5′-tetramethyl-4,4′-diethyl-6-cyanopyrromethene hydrobromide 6bb , confirmed by the further conversion to 1,3,5,7-tetramethyl-2,6-diethyl-8-cyanopyrromethene–BF2 complex 7bb on treatment with boron trifluoride etherate. An alternation effect in the relative efficiency (RE) of laser activity in 1,3,5,7,8-pentamethyl-2,6-di-n-alkylpyrromethene–BF2 dyes depended on the number of methylene units in the n-alkyl substituent, -(CH2)nH, to give RE ≥ 100 when n = 0,2,4 and RE 65, 85 when n = 1,3. (The RE 100 was arbitrarily assigned to the dye rhodamine 6G). The absence of fluorescence and laser activity in 1,3,5,7-tetramethyl-2,6-diethyl-8-isopropylpyrromethene–BF2 complex 7p and a markedly diminished fluorescence quantum yield (Φ 0.23) and lack of laser activity in 1,3,5,7-tetramethyl-2,6-diethyl-8-cyclohexylpyrromethene–BF2 complex 7q were attributed to molecular nonplanarity brought about by the steric interference between each of the two bulky 8-substituents with the 1,7-dimethyl substituents. An atypically low RE 20 for a peralkylated dye without steric interference was observed for 1,2,6,7-bistrimethylene-3,5,8-trimethylpyrromethene–BF2 complex 7j . Comparisons with peralkylated dyes revealed a major reduction in RE 0–40 for the six dyes 7u–z lacking substitution at the 8-position. Low laser activity RE was brought about by functional group (polar) substitution in the 2,6-diphenyl derivative 7I , RE 20, and the 2,6-diacetamido derivative 7m , RE 5, of 1,3,5,7,8-pentamethylpyrromethene–BF2 complex (PMP–BF2) 7a and in 1,7-dimethoxy-2,3,5,6,8-pentamethylpyrromethene–BF2 complex 7n , RE 30. Diethyl 1,3,5,7-tetramethyl-8-cyanopyrromethene-2,6-dicarboxylate–BF2 complex, 7aa , and 1,3,5,7-tetramethyl-2,6-diethyl-8-cyanopyrromethene–BF2 complex, 7bb , offered examples of P–BF2 dyes with electron withdrawing substituents at the 8-position. The dye 7aa , λlas 617 nm, showed nearly twice the power efficiency that was obtained from rhodamine B, λlas 611 nm.  相似文献   

3.
Thermal substitution of 4‐hydroxypyridine 1 with trifluoroacetaldehyde ethyl hemiacetal (TFAE) leads to a moderate yield of 3‐(1‐hydroxy‐2,2,2‐trifluoroethyl)‐4‐hydroxypyridine 7 in the presence of a catalytic amount of anhydrous potassium carbonate. Under similar conditions, several α‐trifluoromethyl hydroxy‐pyridinemethanols 8–15 are easily prepared from 2‐ or 3‐hydroxypyridines 2–6 .  相似文献   

4.
A new procedure for the cyclization of 2-indolecarbohydrazones (5) to 1,2,3,4-tetrahydro-4-oxo-5H-pyridazino[4,5-b]indoles (6) and for the cyclization of 3-indolecarbohydrazones (7) to 1-oxo-1,2,3,4-tetrahydro-5H-pyridazino[4,5-b]indoles (8 and 9) is described. The hydrazones (5 or 7) were treated with an acyl halide (acetyl or benzoyl chlorides) and triethylamine in ethyl acetate of chloroform as solvents to give the compounds 6 (20–70%) from the compounds 5 , and the compounds 8 (20–60%) from the compounds 7 . Through refluxing with ethanol-hydrochloric acid the compounds 8a-8f selectively separate the acetyl group on N5 to give the respective compounds, 9a-9f. The ir and 1H-nmr spectra of all the compounds 5, 6, 7, 8 and 9 and the uv, mass and 13C-nmr spectra of the compounds 7h, 7i, 8h and 8i are discussed.  相似文献   

5.
β-Functionalized Hydrazines from N-Phthalimidoaziridines and their Hydrogenolytic N,N-Cleavage to Amines The three N-phthalimido-aziridines 1–3 were reacted with phenol, thiophenol, aniline, p-toluenesulfonic acid, and H2O in selected combinations. These nucleophiles opened the 3-membered ring to yield the N-phthalimidoamines 4a–d, 5a–d, 6a–c , and 6e ; all these products (except the carbinol 6e ) carry an aryl-substituted functional group on the C-atom vicinal to the N-substituent. Hydrazinolysis of 4, 5, 6a–c , and 6e afforded the β-functionalized hydrazines 7, 8, 9a–c , and 9e . The reducing medium Raney-Ni/N2H4 transformed 4, 5, 6a–c , and 6e to the β-functionalized amines 10, 11, 12a–c , and 12e . By a study with the hydrazide 6a and the hydrazine 9a , it was shown that the N,N-cleavage is a catalytic hydrogenolysis by H2 generated from N2H4 with Raney-Ni and that it does not take place on the hydrazide 6 , but rather on the hydrazine 9 , generated as intermediate from 6 with N2H4. Spectroscopic data confirmed that the conversions of 1–3 to 4–6 occurred exclusively with inversion and that the resulting configurations remained fully intact during the transformations of 4, 5 , and 6 (via 7, 8 , and 9 ) to 10, 11 , and 12 , respectively.  相似文献   

6.
The reactions of 3, 6-diphenyl-1, 2, 4, 5-tetrazine 1 and 3, 6-di(2-pyridyl)-1, 2, 4, 5-tetrazine 2 with the enamines 3a-d derived from morpholine and the 5-,6-,7- and 8-membered cyclic ketones have been investigated. A number of pyridazine derivatives 4–7 most of which are new have been reported. Moreover, a novel procedure for the aromatization of pyridazines 5a-d to the corresponding pyridazine 7b-d via oxidative elimination using hydrogen peroxide is described. The structures of products 4–7 were confirmed by spectral methods and elemental analysis.  相似文献   

7.
Title compounds were obtained starting from the key imidazole intermediate, 5‐amino‐1‐phenyl‐methyl‐2‐mercapto‐1H‐imidazole‐4‐carboxylic acid amide 5 , readily derived from the base catalyzed rearrangement of a thiazole, 5‐amino‐2‐phenylmethylaminothiazole‐4‐carboxylic acid amide 4 . Alkylation of the thiol function on 5 with phenylmethyl and allylic chlorides gave compounds 6 and 7 respectively. Cyclization of 6 with a variety of esters afforded 8‐phenylmethylthiohypoxanthines, 8–11 . Similarly, 7 was cyclized to 8‐allylthiohypoxanthines, 20–21 . Compound 5 was also cyclized, but formed 8‐mercaptohypox‐anthines, 22–24 . Alkylation of 8‐mercaptohypoxanthines afforded 8‐alkylthiohypoxanthines, 8, 9,25 and 26 (see Scheme 2). Chlorination of 9–11 afforded 16–18 ; adenine 19 was derived from 16 . Oxidation of hypox‐anthines 8–11 with m‐chloroperbenzoic acid gave the corresponding 8‐phenylmethylsulfonyl derivatives 12 ‐ 15 . These derivatives proved resistant to nucleophilic displacement reactions with primary amines.  相似文献   

8.
The syntheses of N-(trans-2-iodocyclohexyl)- 1–4 , N-(2-iodo-3,3-dimethylbutyl) 5 , and N-(2-iodo-1,1-diphenylethyl)ureas 6, 7 and the cyclization of 6 and 7 into 2-amino-2-oxazoline derivatives 8, 9 are reported. The structures of prepared compounds are based on analytical and spectroscopic data.  相似文献   

9.
Photocyclization reactions were carried out on ethyl 2-(8-oxo-5,6,7,8-tetrahydro-1-naphthyloxy)acetates 1a-e and ethyl 2-(5-oxo-6,7,8,9-tetrahydro-5H-benzocyclohepten-4-yloxy)acetates 2a-e in acetonitrile. Irradiation of 1a-e gave naphtho[1,8-bc]furanols 3a-e and naphtho[1,8-bc]furans 4a-e in 33–83% yields and ethyl acrylates 5b-d were produced in 3–25% yields during irradiation of 1b-d . On the other hand, 2a-e afforded cyclohepta[ad|benzofuranols 6a-e and cyclohepta[ad]benzofurans 7a-e in 44–87% yields. Ethyl acrylates 8b-d were also produced in 7–43% yields from irradiation of 2b-d . Substituent effects on photocyclization and reaction pathways are discussed.  相似文献   

10.
Methyl 2-(benzyloxycarbonyl)aimno-3-dimemylaminopropenoate ( 2 ) was prepared from methyl N-(benzyloxycarbonyl)glycinate ( 1 ) and t-butoxybis(dimethylamino)methane, and used as a reagent for preparation of substituted 3-(benzyloxycarbonyl)amino-4H-quinolizin-4-ones 5 and 6 , ?2H-pyran-2-ones 17–19 , ?2H-1-benzopyran-2-ones 28–31 , and -naphthopyrans 32–35 , ?2H-pyrano[3,2-c]pyridine-2,5-dione 46 , -pyrano-[4,3-b]pyran-2,5-dione 47 , -pyrano[3,2-c]benzopyran-2,5-dione 48 , -pyrano[2,3-c]pyrazol-6-ones 49 and 50 , -pyrano[2,3-d]pyrirnidin-7-ones 51 and 52 derivatives. In the reaction of 2 with 1,3-diketones trisubsti tuted pyrroles 14–16 were formed. Selective removal of benzyloxycarbonyl group was achieved by cat alytic transfer hydrogenation with Pd/C in the presence of cyclohexene to afford free 3-amino compounds 7 , 8 , 20 , 36–38 and 53–57 in yields better than 80%.  相似文献   

11.
The protected disaccharide 44 , a precursor for the synthesis of allosamidin, was prepared from the glycosyl acceptor 8 and the donors 26–28 , best yields being obtained with the trichloroacetimidate 28 (Scheme 6). Glycosidation of 8 or of 32 by the triacetylated, less reactive donors 38–40 gave the disaccharides 46 and 45 , respectively, in lower yields (Scheme 7). Regioselective glycosidation of the diol 35 by the donors 38–40 gave 42 , the axial, intramolecularly H-bonded OH? C(3) group reacting exclusively (Scheme 5). The glycosyl acceptor 8 was prepared from 9 by reductive opening of the dioxolane ring (Scheme 3). The donors 26–28 were prepared from the same precursor 9 via the hemiacetal 25 . To obtain 9 , the known 10 was de-N-acetylated (→ 18 ), treated with phthalic anhydride (→ 19 ), and benzylated, leading to 9 and 23 (Schemes 2 and 3). Saponification of 23 , followed by acetylation also gave 9 . Depending upon the conditions, acetylation of 19 yielded a mixture of 20 and 21 or exclusively 20 . Deacetylation of 20 led to the hydroxyphthalamide 22 . De-N-acetylation of the 3-O-benzylated β-D -glycosides 11 and 15 , which were both obtained from 10 , was very sluggish and accompanied by partial reduction of the O-allyl to an O-propyl group (Scheme 2). The β-D -glycoside 30 behaved very similarly to 11 and 15 . Reductive ring opening of 31 , derived from 29 , yielded the 3-O-acetylated acceptor 32 , while the analogous reaction of the β-D -anomer 20 was accompanied by a rapid 3-O→4-O acyl migration (→ 34 ; Scheme 4). Reductive ring opening of 21 gave the diol 35 . The triacetylated donors 38–40 were obtained from 20 by debenzylidenation, acetylation (→ 36 ), and deallylation (→ 37 ), followed by either acetylation (→ 38 ), treatment with Me3SiSEt (→ 39 ), or Cl3CCN (→ 40 ).  相似文献   

12.
Two homologous series of quinoline-containing liquid crystalline compounds were synthesised. Preparation of these compounds was completed in a short two-step reaction. Fair to good two-step overall yields of 63%–68%?and 50%–60%?were obtained respectively for the liquid crystalline compounds of 2-(4-alkylphenyl)-6-methylquinolines (nPQMe, n?=?4–8) and 2-(4-alkoxyphenyl)-6-methylquinolines (mOPQMe, m?=?3–7). Spectral analyses were in accordance with the expected structures. Polarising optical microscopy showed both series of compounds only display a nematic phase. Their thermotropic behaviours were further confirmed by differential scanning calorimetry.  相似文献   

13.
Two series of difunctional proton acceptors, stilbazole derivatives 4a – c and 6a – c with different spacers, oligo(methylene) and oligo(ethylene glycol), respectively, were synthesized. Hydrogen‐bonded polymeric complexes 4 / 7 and 6 / 7 and trimeric complexes 4 / 82 and 6 / 82 ( 7 and 8 are aromatic dicarboxylic acids and monocarboxylic acids, respectively) were prepared, and their liquid crystallinity was examined. The effects of the spacer in 4 , 6 , and 7 and the terminal group in 8 on the thermal behaviors of hydrogen‐bonded complexes were investigated using differential scanning calorimetry and polarizing optical microscopy. X‐ray diffraction at elevated temperatures was used to verify liquid crystal phases. The study showed that the phase transition temperatures for isotropic to nematic (TI–Ns) of polymeric complexes 4 / 7 and 6 / 7 in general decreased with the increase in length of spacer in the corresponding proton donors 7 . The increase in length of the proton acceptors 6 led to a drop of TI–Ns of the corresponding complexes 6 / 7 ; however, the TI–Ns of complexes 4 / 7 did not show any correlation with the spacer length in 4 . In contrast, the increase in length of the terminal group in 8 resulted in a slight decrease in TI–Ns of trimeric complexes 4 / 82 , but had a negligible effect on the TI–Ns of 6 / 82 because of the presence of the more flexible spacer in the proton acceptors 6 . © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 4731–4743, 2005  相似文献   

14.
1,3-Dipolar cycloadditions of diazoalkanes to pyridazin-3(2H)-ones 1–7 and pyridazin-3(2H)-thiones 8 and 9 are regioselective producing 3H -pyrazolo[3,4-d]pyridazin-4(5H)-ones 15–19, 27–29 and 34–38 as the major products. In some instances, the isomeric 3H-pyrazolo[3,4-d]pyridazin-7(6H)-ones, such as 20 and 23 were isolated as the minor products. From 3 and 6 the primary 3a,7a-dihydro cycloadducts 25 and 26 , and rearranged 1,2-dihydro intermediate 31 were isolated. From 10 and 1-diazoindane the isomeric exo- and endospiro products 39 and 40 were formed.  相似文献   

15.
The synthesis of 8-azaguanine N9-, N8-, and N7-(2′-deoxyribonucleosides) 1–3 , related to 2′-deoxyguanosine ( 4 ), is described. Glycosylation of the anion of 5-amino-7-methoxy-3H-1,2,3-triazolo[4,5-d]pyrimidine ( 5 ) with 2-deoxy-3,5-di-O-(4-toluoyl)-α-D -erythro-pentofuranosyl chloride ( 6 ) afforded the regioisomeric glycosylation products 7a/7b, 8a/8b , and 9 (Scheme 1) which were detoluoylated to give 10a, 10b, 11a, 11b , and 12a . The anomeric configuration as well as the position of glycosylation were determined by combination of UV, 13C-NMR, and 1H-NMR NOE-difference spectroscopy. The 2-amino-8-aza-2′-deoxyadenosine ( 13 ), obtained from 7a , was deaminated by adenosine deaminase to yield 8-aza-2′-deoxyguanosine ( 1 ), whereas the N7- and N8-regioisomers were no substrates of the enzyme. The N-glycosylic bond of compound 1 (0.1 N HCl) is ca. 10 times more stable than that of 2′-deoxyguanosine ( 4 ).  相似文献   

16.
In this work, six coumarins, including two new ones, 8-(3-hydroxy-2,2-dimethylpropyl)-7-methoxy-2H-chromen-2-one (2) and 5-[(7′,8′-dihydroxy-3′,8′-dimethyl-2-nonadienyl)oxy] psoralen (4), as well as four known ones, 5-[(6′,7′-dihydroxy-3′,7′-dimethyl-2-octenyl) oxy] psoralen (1), marmin (3), epoxybergamottin (5), and aurapten (6) were successfully separated from the crude extract of pummelo (Citrus maxima (Burm.) Merr. Cv. Shatian Yu) peel by high-speed countercurrent chromatography in a single run with petroleum-ether–ethyl acetate–methanol–water (4:6:6:4, v/v). The structures of these six coumarins were elucidated by ESI-MS, extensive 1D and 2D NMR spectroscopy.  相似文献   

17.
Synthesis of Ketoses by Chain Elongation of 1-Deoxy-1-nitroaldoses. Nucleophilic Additions and Solvolysis of Nitro Ethers A method for the preparation of chain elongated uloses based upon the base-catalyzed addition of 1-deoxy-1-nitroaldoses to aldehydes and Michael acceptors and subsequent solvolytic replacement of the nitro group by a hydroxy group is described. Thus, addition of 1 , 3 and 9 to formaldehyde, followed by solvolysis gave the chain elongated ulose derivatives 2 , 8 and 10 (63–76%), respectively. The configuration at the anomeric center of the addition products was deduced from 13 C – NMR . spectra and mutarotation. In the case of 3 , the primary addition products 4 and 6 were isolated and acetylated to 5 and 7 . The nitro derivatives 4 – 7 do not follow Hudson's rule of isorotation. Addition of 1 to benzaldehyde (44%) and to nonanal (74%) preceded with a small degree of diastereoselectivity to give 15a / 15b , and 11 / 12 , respectively. The configuration of the secondary hydroxyl group of 12 was determined by correlation with methyl 2-hydroxydecanoate ( 14 ). Addition of 1 to the galacroaldehyde 16 gave a single compound 17 (78%). The structure of this dodecosulose was determined by X-ray crystallography. Solvolysis of the acetylation product 18 in formamide gave the hemiacetal 19 (69%). Michael addition of 1 to acrylonitrile, methyl vinyl ketone and cyclohexenone under solvolytic conditions gave the hemiacetals 27 , 30 and 31a , b (49%, 71% and 76%, respectively). Under non-solvolytic conditions (Bu4NF), 1 reacted with acrylonitrile, and crotononitrile to give the anomeric nitro ethers 23 and 24 (67%) and 25 and 26 (84%). respectively. Similarly. 3 added to acrylonitrile to give 28 and 29 (55%, 4:1). This reaction appears to proceed under kinetic control. Addition of 1 to ethyl propiolate and solvolysis yielded the unsaturated spirolactone 32 (50%) and the hemiacetal 33 (17%). Hydrogenation of 32 gave the saturated spirolactone 34 (100%) which was also obtained from 1 and methyl acrylate (63%). Addition of 1 to dimethylmaleate gave the unsaturated ester 35 (48%).  相似文献   

18.
Treatment with a mixture of formic and hydrobromic acids converted ethyl 3,4-diethyl-5-methyl-pyrrole-2-carboxylate 7a to 3,3′,4,4′-tetraethyl-5,5′-dimethylpyrromethene hydrobromide 8a presumably via the condensation of α-unsubstituted and α-formylpyrrole intermediates 7c and 7e formed in situ. The corresponding 6-cyanohexaalkylpyrromethane 9a was obtained by the addition of hydrogen cyanide to the pyrromethene 8a and was oxidized with bromine to an unstable pyrromethene 10a , an intermediate converted to 1,2,6,7-tetraethyl-3,5-dimethyl-8-cyanopyrromethene–BF2 complex 3 , (PM TEDC), λlas (plastic) 613–639 nm, in a reaction with boron trifluoride etherate. Ethyl 3,4,5-trimethylpyrrole-2-carboxylate 7b was similarly converted to 1,2,3,5,6,7-hexamethyl-8-cyanopyrromethene–BF2 complex 4 , (PM HMC), λlas (plastic) 615–639 nm. Immediately after formation by a condensation between propionyl chloride and 2,4-dimethyl-3-cyanopyrrole 16 , unstable 3,3′,5,5′-tetramethyl-6-ethyl-4,4′-dicyanopyrromethene hydrochloride 17 was treated with boron trifluoride etherate to give 1,3,5,7-tetramethyl -2,6-dicyano-8-ethylpyrromethene–BF2 complex 18 , λlas (ethanol) 540–565 nm.  相似文献   

19.
The syntheses of 7-deaza-N6-methyladenine N9-(2′-deoxy-β-D -ribofuranoside) ( 2 ) as well as of 8-aza-7-deaza-N6-methyladenine N8? and N9?(2′-deoxyribofuranosides) ( 3 and 4 , resp.) are described. A 4,4′-dimeth-oxylritylation followed by phosphitylation yielded the methyl phosphoramidites 12–14 . They were employed together with the phosphoramidite of 2′-deoxy-N6v-methyladenosine ( 15 ) in automated solid-phase oligonucleotide synthesis. Alternating or palindromic oligonucleotides derived from d(A-T)6 or d(A-T-G-C-A-G-A*-T-C-T-G-C-A) but containing one methylated pyrrolo[2,3-d]pyrimidine or pyrazolo[3,4-d]pyrimidine moiety in place of a N6-methylaminopurine (A*) were synthesized. Melting experiments showed that duplex destabilization induced by a N6-Me group of 2′-deoxy-N6-methyladenosine is reversed by incorporation of 8-aza-7-deaza-2′-deoxy-N6-meihyladenosine, whereas 7-deaza-2′-deoxy-N6-methyladenostne decreased the Tm value further. Regiospecific phosphodiester hydrolysis of d(A-T-G-C-A-G-m6A-T-C-T-G1-C-A) by the endodeoxyribonuclease Dpn I, yielding d(A-T-G-C-A-G-m6A) and d(pT-C-T-G-C-A), was prevented when the residue c7m6Ad ( 2 ), c7m6z8Ad ( 3 ), or c7m6z8Ad′ ( 4 ) replaced m6Ad ( 1 ) indicating that N(7) of N6-methyladenine is a proton-acceptor site for the endodeoxyribonuclease.  相似文献   

20.
Conversion of Bicyclo [3.2.0]hept-2-en-6-one into Cyclopentadienylacetic Acid Derivatives The reaction of a mixture of 4exo-bromobicyclo [3.2.0]hept-2-en-6-one ( 2 ) and -7-one ( 3 ) with O- or N-nucleophiles yielded cyclopentadien-5′-yl-acetates 4a–f or-acetamides 4g–h . Due to their rapid isomerization, the products 4 were not isolated, but some of them were demonstrated spectroscopically or captured in situ with maleimide as 10′-substituted norbornene derivatives 7 . The formation of 4 from 2/3 involves a fragmentation of the bond between the carbonyl and the bridge-head C-atom, induced by the attacking nucleophile and the leaving Br-ion and aided by the relief of the four-membered ring strain. Some of the isomerization products of 4 , i.e. the cyclopentadiene-1′-yl- and 2′-yl-acetyl derivatives were captured with maleimide as the 1′- and 8′-substituted norbornene-derivatives 8 and 9 . Two C-nucleophiles did not induce the fragmentation: sodium acetylacetonate substituted the Br-atom and sodium (diethoxyphosphoryl)ethoxycarbonylmethide condensed with the carbonyl group of 2/3 , yielding 11/12 and 13/14 , respectively.  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-23

京公网安备 11010802026262号