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 31 毫秒
1.
A new fluorine-containing organosilicon compound, (bromodifluoromethyl)-phenyldimethylsilane (II), was synthesized by the N-bromosuccinimide (NBS) bromination of difluoromethyl)phenyldimethylsilane (I), which was prepared from phenyldimethylsilyllithium and chlorodifluoromethane. Compound II reacted with dimethyl sulfoxide to give dimethyl sulfide and phenyldimethylfluorosilane in quantitative yield. The reaction of II with nucleophiles, such as sodium ethoxide, Grignard or lithium reagents, afforded products arising from cleavage of the carbonsilicon bond. In contrast, the reaction of II with Grignard reagents in the presence of appropriate catalysts (Group VIII transition metal salts or complexes) afforded the homo-coupling product of II, 1,2-bis-(phenyldimethylsilyl)-1,1,2,2-tetrafluoroethane (IV), in excellent yield. The silver(I) salt-catalyzed reaction of II with ethylmagnesium bromide gave the cross-coupling product, (1,1-difluoropropyl)phenyldimethylsilane (V) as well as III and IV. When cuprous bromide was employed as catalyst, the reaction of II with ethylmagnesium bromide afforded 1-phenyldimethylsilyl-1-propene (VI) and 3-phenyldimethylsilyl-2-pentene (VII) as main products.  相似文献   

2.
The reaction of aroylphenylacetylenes (I) with acyl- or aroylhydrazines (II) gave ω-aroyl-acetophenone-N-acyl or N-aroylhydrazones (IV). The latter gave upon treatment with methanolic potassium hydroxide and with acetic anhydride in the presence of sodium acetate, the corresponding pyrazoles (V) and the N-acetylpyrazoles (VII and VIII), respectively. The acetylenic ketones ( 1 ) also reacted with methylhydrazine and 1,1-dimethylhydrazine to give 5-aryl-1-methyl-3-phenylpyrazoles (XII), and 1,1-dimethylhydrazine derivatives (XIII), respectively. When the latter compounds were heated with acetic anhydride, they gave the N-methylpyrazoles (XII).  相似文献   

3.
In the presence of zerovalent π-bis(benzene)chromium(0) (I), perfluoropropylene (II) was found to undergo oligomerization under very mild conditions to dimers (b.p. 46°, M+, 300) and trimers (b.p. 100–102°C, M+ - 19, 431) in the ratio of 2.5–3.0 to 1. One mol of metal complex could catalyze the conversion of 50 mol of perfluoropropylene. On the basis of 19F NMR, the structures of the dimers are III and IV in a ratio of ca. 80/20 and the trimers V and VI; VII and VIII also seemed to be present. V/VI/(VII + VIII) is 80/10/4. In benzene solution, perfluoropropylene was shown not to be catalytically oligomerized by fluoride ion (KF, CrF2 or (CH3)4NF) (nor by monovalent π-dibenzenchromium(I). A possible mechanism of the reaction was proposed.  相似文献   

4.
The square-wave voltammetric behaviour of cysteine and saccharin was studied at a static mercury drop electrode at pH 7.4 in the presence of Cu(II) ions. In the presence of excess Cu(II), cysteine exhibited three reduction peaks for Hg(SR)2 (−0.086 V), free Cu(II) (−0.190 V) and Cu(I)SR (−0.698 V), respectively. Saccharin produced a catalytic hydrogen peak at −1.762 V. In the presence of Cu(II), saccharin gave a new peak (−0.508 V), corresponding to the reduction of Cu(II)–saccharinate, which in the presence of cysteine formed a mixed ligand complex (−0.612 V), CuL2A2 (L=saccharin and A=cysteine). The peak potentials and currents of the obtained complexes were dependent on the ligand concentration and accumulation time. The stoichiometries and overall stability constants of these complexes were determined by Lingane's method (voltammetrically) and Job’s method (spectrophotometrically). The mixed ligand complex in the molar ratio 1:2:2 (log β=33.35) turned out to be very much stronger than the 1:1 Cu(I)SR (log β=21.64) and 1:2 Cu(II)–saccharinate (log β=16.68) complexes. Formation of a mixed ligand complex can be considered as a type of synergism.  相似文献   

5.
Metallation of 1,1-dibutyl-1-stannacyclohexadiene-2,5 (I) with lithiumamides yield the lithium compound II, from which the trimethylsilyl-, germyl-, -stannyl- and the bromoethyl-substituted stannacyclohexadienes III, IV, V and VI are obtained. The bis(trimethylsilyl- and -germyl) substituted stannacyclohexadienes VIII and X have been synthesized starting from III and IV, respectively. Arsabenzene (XII) is formed in good yields by treating arsenic trichloride with III, IV and V. 4-Trimethylsilyl-1-arsabenzene (XIII), 4-trimethylgermyl1-arsabenzene (XIV) and 4-(2-chloroethyl)-1-arsabenzene (XV) can be prepared by treating VIII, X and VI respectively with arsenic trichloride, 1H NMR, IR, UV and mass spectral data of the new compounds are described.  相似文献   

6.
A synthesis is proposed of 4,8-dimethyldecanal (VIII) — a pheromone of the flour beetlesTribolium confusum andT. castaneum. By heating 71.2 g of 4-methyltetrahydropyran (I), 83.2 g of AcBr and 1.57 g of ZnCl2 (45°C), then 120°C, 2 h), 1-acetoxy-5-bromo-3-methylpentane (II) was obtained. The hydrolysis of 19.8 g of (II) (MeOH-H2O, TsOH, 20°C, 15 h) gave 5-bromo-3-methylpentan-1-ol (III). From 18.1 g of (III) and 38.9 ml of 2,3-dihydropyran (Et2O, TsOH, 20°C, 20 h) was obtained the 2-THPL ester of (III), (IV), which was converted into 3-methyloct-7-en-1-ol (V) by the treatment of the corresponding Grignard reagent with allyl bromide (THF, CuI-bi-2-pyridyl, 2°C, 4 h, Ar). The interaction of 1.42 g of (V) with Et3Al (hexane, 20°C, Cp2ZrCl2, Ar) gave 3,7-dimethylnonan-1-ol (VI), boiling which with 48% HBr in the presence of concentrated H2SO4 gave 1-bromo-3,7-dimethylnonane (VII) which was then converted into the desired (VIII) by the reaction of the corresponding Grignard reagent with DMFA (0–2°C, 1 h; 20°C, 2 h; Ar). The characteristics of the compounds — yield (%), nD (°C): (I), 79, 1.4340 (22); (III) 89, 1.4660 (23); (IV), 82, 1.4739 (23); (V), 85, —; (VI), 90, 1.4483 (20); (VII), 88, 1.4409 (22); (VIII), 88, 1.4589 (22). Details of the IR and PMR spectra of compounds (II)–(VII) are given.Institute of Chemistry, Bashkir Scientific Center, Urals Branch, USSR Academy of Sciences, Ufa. Translated from Khimiya Prirodnykh Soedinenii, No. 2, pp. 272–276, March–April, 1989.  相似文献   

7.
Three bis(dimethylamino)silane monomers have been polymerized with 1,1'-bis-(hydroxymethyl)ferrocene to give ferrocene-containing polyoxysilanes I and II. They were bis(dimethylamino)dimethylsilane (III), bis(dimethylamino)diphenylsilane (IV), and 1,4-bis(N,N-dimethylaminodimethylsilyl)benzene (V). Mixing of the diol and III or IV at O°C followed by heating resulted in polymerization to higher molecular weights than when the monomers were initially mixed at higher temperatures. At higher temperatures the formation of monomeric cyclic products seriously competed with polymerization, and the five atom bridged derivative, 3-sila-2,4-dioxa-3,3diphenyl[5]ferrocenophane (VI) was isolated in good yield. The use of silane V, where cyclization is not expected to compete, led to higher polymer yields and molecular weights. The polymers were low melting and I (R = C6H5) could be cast into films and weak fibers were drawn from its melt. The polymers were sensitive to hydrolytic decomposition; those containing Si-CH3 linkages were completely hydrolyzed in refluxing THF-H2O (10:1) in 1 hr. The polymers were characterized by viscosity studies, gel-permeation chromatography, and infrared and NMR spectroscopy.  相似文献   

8.
Two new one‐dimensional CuII coordination polymers (CPs) containing the C2h‐symmetric terphenyl‐based dicarboxylate linker 1,1′:4′,1′′‐terphenyl‐3,3′‐dicarboxylate (3,3′‐TPDC), namely catena‐poly[[bis(dimethylamine‐κN)copper(II)]‐μ‐1,1′:4′,1′′‐terphenyl‐3,3′‐dicarboxylato‐κ4O,O′:O′′:O′′′] monohydrate], {[Cu(C20H12O4)(C2H7N)2]·H2O}n, (I), and catena‐poly[[aquabis(dimethylamine‐κN)copper(II)]‐μ‐1,1′:4′,1′′‐terphenyl‐3,3′‐dicarboxylato‐κ2O3:O3′] monohydrate], {[Cu(C20H12O4)(C2H7N)2(H2O)]·H2O}n, (II), were both obtained from two different methods of preparation: one reaction was performed in the presence of 1,4‐diazabicyclo[2.2.2]octane (DABCO) as a potential pillar ligand and the other was carried out in the absence of the DABCO pillar. Both reactions afforded crystals of different colours, i.e. violet plates for (I) and blue needles for (II), both of which were analysed by X‐ray crystallography. The 3,3′‐TPDC bridging ligands coordinate the CuII ions in asymmetric chelating modes in (I) and in monodenate binding modes in (II), forming one‐dimensional chains in each case. Both coordination polymers contain two coordinated dimethylamine ligands in mutually trans positions, and there is an additional aqua ligand in (II). The solvent water molecules are involved in hydrogen bonds between the one‐dimensional coordination polymer chains, forming a two‐dimensional network in (I) and a three‐dimensional network in (II).  相似文献   

9.
Decacarbonyldimanganese(0) (I) gives photochemically with 1,1-dimethylallene (II) at 248 K a mixture of three mononuclear and two dinuclear complexes, which are separated by HPLC. The main product of the reaction is the bridged octacarbonyl-μ-η2:2-1,1-dimethylallene-dimanganese(0)(V). As side products, a mixture of the isomeric tetracarbonyl-η3-3-methyl-2-buten-1-yl-manganese (III) and tetracarbonyl-η3-2-methyl-2-buten-1-yl-manganese (IV) is obtained. In addition the novel electron deficient compounds tricarbonyl-η3:CH-2-methyl-E-3-(3(8), 6-m-menthadien-6-yl)-2-buten-1-yl-manganese (VI) and hexacarbonyl-μ-η3:CH:3:CH-2,3,4,5-tetramethyl-2,5-hexadien-1,4-diyl-dimanganese (VII) are isolated. VI and VII add under ambient conditions reversibly carbon monoxide and form the moderately stable tetracarbonyl-η3-2-methyl-E-3-(3(8), 6-m-menthadien-6-yl)-2-buten-1-yl-manganese (VIII) and the instable hepta- and octacarbonyl-μ-η3:3-2,3,4,5-tetramethyl-2,5-hexadien-1,4-diyl-dimanganese (IX, X). The mixture of the isomers III and IV, and the complexes V – VIII were characterized by C and H elemental analyses and by NMR and IR spectroscopy. The molecular structures of VI and VII were determined by X-ray structure analyses.  相似文献   

10.
The potassium hydroxide-induced (Stevens) rearrangement of 1,3,4-trimethyl-1-(3,4,5-trimethoxybenzyl)-1,2,5,6-tetrahydropyridinium chloride (I) gives the desired 1,3,4-trimethyl-2-(3,4,5-trimethoxybenzyl)-1,2,5,6-tetrahydropyridine (III) and the Hofmann elimination product, N-methyl-N-(3,4,5-trimethoxybenzyl)-2,3-dimethyl-2,4-pentadienamine (II). In the presence of ethereal phenyllithium, the salt I undergoes rearrangement giving the expected tetrahydropyridine III in about 17% yield and four other products, N-(3,4,5-trimethoxybenzyl)methylamine (VI), 1,3,4-trimethyl-2-(6-methyl-2,3,4-trimethoxyphenyl)-1,2,5,6-tetrahydropyridine (IV), 1,3,3-trirnethyl-2-(3,4,5-trimethoxyphenyl)-4-rnethylenepiperidine (V) and 1,3,4-trimethyl-4-(3,4,5-trimethoxybenzyl)-1,4,5,6-tetrahydropyridine (VII), the latter being the 1,4-Stevens rearrangement product which cyclizes easily to β-2′,3′,4′-trimethoxy-2,5,9-trimethyl-7,8-benzomorphan (VIII). Their structures have been proved both by analytical and spectral data. A possible route for VIII and its stereochemical aspects are discussed.  相似文献   

11.
The catalysis of isoprene hydrosilylation with different silanes HSiR3″ (R″ = Et, Ome, OEt), in the presence of rhodium(I) or ruthenium(II) complexes with unsaturated N-containing controlling ligands as co-catalysts, occurs under mild conditions and gives with high selectivity (Z)-2-methyl-1-silyl-2-butenes (VI), 2-methyl-4-silyl-2-butenes (VIII), 2-methyl-4-silyl-1-butenes (IX) or the 3-methyl-4-silyl-1-butenes (X). Thus four out of five possible isomers of the Si addition to a terminal sp2-C atom can be obtained as the main products of catalysis (51–87%) by changing the metal and the controlling ligand. Chiral X is obtained for the first time via catalysis. Full 1H NMR assignment is given for compounds VIII–X.  相似文献   

12.
The Schiff base ligands I–V, made by condensing either 2-acetylpyridine (I), 8-quinolinecarboxaldehyde (II and III), or o-methylthiobenzaldehyde (IV and V) with either N,N′-dimethyl-1,3-diaminopropane (I, II, and IV), 2-aminomethylpyridine (III), or 2-(2-aminoethyl)-pyridine (V), give ionic PtIVMe3 complexes containing tridentate NNN- or SNN-bonded ligands. With PtMe3Br ligand V gives a neutral complex XI in which it is coordinated only via the two N atoms. A monomeric PtIVMe3 salicyladiminate complex results on treating the dimeric trimethylplatinum(IV) salicylaldehyde complex with the bidentate amine H2N (CH2)3NMe2. The complexes have been fully characterised by 1H NMR spectroscopy.  相似文献   

13.
The reaction of 1,3-diaryl-2-propene-1-ones I with arylacetamides II, in the presence of sodium ethoxide under reflux, for two hours, gave the corresponding 3,4,6-triaryl-3,4-dihydro-2(1H)-pyridones IV. However, when the reaction of these ketones was carried out in the presence of sodium hydride, they gave the corresponding 3,4,6-triaryl-2(1H)-pyridones VI or a mixture of IV and VI. When 1,3-diaryl-2-propyne-1-ones V were reacted with arylacetamides, in the presence of sodium hydride, they yielded the corresponding 2-pyridones VI. Treatment of compounds IV with selenium produced the corresponding 2-pyridones VI. Acetylation of the latter compounds gave the corresponding 2-acetyl derivatives VIII. The structure of all products was confirmed by chemical and spectroscopic evidence, and the mechanism of the reactions was discussed.  相似文献   

14.
Treatment of pyrene and some its derivatives with Cu(II) tetrafluoroborate or perchlorate in CH3CN cleanly led to the formation of 1,1′-dipyrenyls. The other polycyclic hydrocarbons (anthracene, perylene) under the same conditions provide cation-radicals. 1,1′-Dipyrenyl strongly differs from pyrene in the chemical behavior: it does not undergo either formylation by Vilsmeier reaction, neither acylation (AcCl, ZnCl2) or nitration (by pyridinium nitrate in boiling pyridine whereas the pyrene is easily involved into these reactions.  相似文献   

15.
Ten vinylhydroquinone and one vinyl resorcinol derivatives are compared, particularly with respect to NMR spectra and copolymerizability with styrene. They are vinylhydroquinone dimethyl ether (I), vinyl-O,O′-bis(1-ethoxyethyl)hydroquinone (II), vinylhydroquinone di(2-pentyl)ether (III), 4-vinyl resorcinol bismethoxymethyl ether (IV), 2-vinyl-5-methylhydroquinone dimethyl ether (V), 2-vinyl-5-methyl-O,O′-bis(1-ethoxyethyl)hydroquinone (VI), 2-vinyl-6-methylhydroquinone dimethyl ether (VII), 2-vinyl-5-tert-butylhydroquinone dimethyl ether (VIII), 2-vinyl-5-chlorohydroquinone dimethyl ether (IX), 2-vinyl-3,6-dimethylhydroquinone dimethyl ether (X), and 2-vinyl-3,5,6-trimethylhydroquinone dimethyl ether (XI). All the vinyl protons have almost the same coupling constants. Though subtle distinctions are found among all the spectra, they can in general be put into two groups on the basis of the chemical shifts. Let the hydrogen on carbon-1 of the vinyl group be A, the hydrogen cis to A be B the hydrogen trans to A be C, then in the first group, (I) through (IX), the chemical shifts (τ) are (A) 3.02 ± 0.08, (C) 4.41 ± 0.05, and (B) 4.87 ± 0.07, and in the second group, (X) and (XI), they are (A) 3.30 ± 0.03, (C) 4.49 ± 0.01, and (B) 4.59 ± 0.03. It is supposed that in (X) and (XI) the vinyl group is out of the plane of the ring, because of the two ortho substituents, and this conformation is reflected in the NMR data. Ultraviolet spectra are consonant with this interpretation, since the λmax of (X) and (XI) correspond closely with those of nonvinyl reference compounds, while those of (II), (V), and (VIII) are shifted to longer wavelengths. When these compounds are copolymerized separately with styrene, the behaviors are classifiable into the following three groups, where r1 and r2 are monomer reactivity ratios with styrene as the first monomer: (i) r1 < 1 and r2 < 1 for compounds (II) and (III) and the reference compound O,O′-dibenzoylvinylhydroquinone, (ii) r1 < 1 and r2 > 1 for compounds (I), (V), (VII), (VIII), (IX), and (iii) r1 > 1 and r2 = 0 for compounds (X) and (XI). These behaviors are correlated with the effect of electronegativity of groups on the stability of the radical at the growing end of the chain and with the simultaneous effects of steric hindrance.  相似文献   

16.
From an extract of Laurelia novae-zelandiae A. CUNN . the aporphine alkaloids (?)-pukateine (I), (?)-pukateine methyl ether (II), (?)-roemerine (IV), (?)-mecambroline (V), (+)-boldine (VII), (+)-isoboldine (VIII), (+)-laurolitsine (IX), and the proaporphine alkaloid (+)-stepharine (X) were isolated. Compounds II and V were up to now not described as natural alkaloids. These and the alkaloids IV, VII, VIII, IX and X are new for L. novae-zelandiae.  相似文献   

17.
(±)-15,19-Dimethyltritriacontane (II) — a component of the pheromone of the stable fly — has been obtained by a five-stage synthesis from dimethylcyclooctadiene (I). The coupling of 1,1-dimethoxy-4-methyl-8-oxonon-4Z-ene [the product of the ozonolysis of (I)] with n-C13H27CH=PPh3 (THF; ?30°, 2 h; 25°, 15 h; Ar) gave 1,1-dimethoxy-4,8-dimethyldocosa-4Z,8Z(E)-diene (III). The hydrolysis of (III) (TsOH·Py, H2O-Ac, boiling, 4 h) gave the corresponding aldehyde (IV). The condensation of (IV) with n-C10H21CH=PPh3 (THF; ?60° to ?30°C, 2 h, 25°C, 15 h) led to 15,19-dimethyltritriaconta-11Z(E),15Z,19Z(E)-triene (V), the exhaustive hydrogenation of which (ethanol, H2, 5% Pd/C, 25°C) gave (II). The substance, the yield in %, and Rf values are given, respectively: (II), 95, 0.92; (III), 29, 0.74; (IV), 80, 0.72; (V) 50, 0.8. The IR and PMR spectra of compounds (II)–(V) and the mass spectra of (II) and (III) are given.  相似文献   

18.
Zusammenfassung Durch Einwirkung von CH2N2 auf Isopropyliden-benzylidenmalonat (I) entsteht 2-Phenylcyclopropan-1,1-dicarbonsäure-isopropylidenacylal (II) und 2-Benzylcyclopropan-1,1-dicarbonsäure-isopropylidenacylal (III). Der Konstitutionsbeweis für die Reaktionsprodukte wird erbracht, der Chemismus ihrer Bildungsreaktion diskutiert und weiters über einige Umsetzungen mit ihnen berichtet.The reaction of diazomethane with isopropylidene benzylidene malonate (I) leads to the formation of 2-phenylcyclopropane-1,1-dicarboxylic acid isopropylidene acylal (II) and 2-benzylcyclopropane-1,1-dicarboxylic acid isopropylidene acylal (III). Evidence for the structure of the reaction products II and III is given, the mechanism of their formation discussed and several of their reactions reported.Mit 1 AbbildungEinige wichtige Ergebnisse der in dieser Arbeit mitgeteilten Untersuchungen wurden bereits in zwei kurzen vorläufigen Mitt.1 veröffentlicht. Hier sollen die dort erwähnten Befunde ergänzt und das experimentelle Material nachgetragen werden.  相似文献   

19.
A CF3‐containing diamine, 1,4‐bis(4‐amino‐2‐trifluromethylphenoxy) benzene ( I ), was prepared from hydroquinone and 2‐chloro‐5‐nitrobenzotrifluoride. Imide‐containing diacids ( V a–h and VI a,b ) were prepared through the condensation reaction of amino acids, aromatic diamines, and trimellitic anhydride. Then, a series of soluble fluorinated polyamides ( VII a–h ) and poly(amide imide)s ( VIII a–h and X a,b ) were synthesized from I with various aromatic diacids ( II a–h ) and imide‐containing diacids ( V a–h and VI a,b ) via direct polycondensation with triphenyl phosphate and pyridine. The polyamides and poly(amide imide)s had inherent viscosities of 1.00–1.70 and 0.79–1.34 dL/g, respectively. All the synthesized polymers showed excellent solubility in amide‐type solvents such as N‐methyl‐2‐pyrrolidinone, N,N‐dimethylacetamide, and N‐dimethylformamide and afforded transparent and tough films via solvent casting. Polymer films of VII a–h , VIII a–h , and X a,b had tensile strengths of 91–113 MPa, elongations to break of 8–40%, and initial moduli of 2.1–2.8 GPa. The glass‐transition temperatures of the polyamides and poly(amide imide)s were 254–276 and 255–292 °C, respectively, and the imide‐containing poly(amide imide)s had better thermal stability than the polyamides. The polyamides showed higher transparency and were much lighter in color than the poly(amide imide)s, and their cutoff wave numbers were below 400 nm. In comparison with isomeric IX c – h , poly(amide imide)s VIII c–h exhibited less coloring and showed lower yellowness indices. © 2004 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 42: 3116–3129, 2004  相似文献   

20.
Treatment of niobocene carbonylhydride, Cp2Nb(CO)H (I), with PhnSnCl4?n and Et2SnCl2 in THF in the presence of Et3N leads to the respective heteronuclear complexes Cp2Nb(CO)SnRnCl3?n (R = Ph, n = 3 ÷ 1 (II–IV), R = Et, n = 2 (V)). Treatment of II with HCl in ether gives Cp2Nb(CO)SnCl3 (VI). Complex VI and its analog (MeC5H4)2Nb(CO)SnCl3 (VIII) were prepared by an alternative synthesis using direct reaction of I or (MeC5H4)2Nb(CO)H with an equimolar quantity of SnCl4 in THF in the presence of Et3N. Complex VI is also generated by insertion of SnCl2 into the NbCl bond in Cp2Nb(CO)Cl (VII). X-Ray analysis of complexes II and VIII was performed: for II, space group P21/n, a = 10.1021(21), b = 17.4633(32), c= 14.2473(29) Å, β = 95.578(16)°, Z = 4; for VIII, space group. P21/n, a= 8.9369(15), b = 13.3589(12), c = 13.9292(20) Å, β = 99.490(14)°, Z = 4. The NbSn bond in VIII (2.764(9) Å) is shorter than that in II (2.825(2) Å). In both cases the NbSn bond is significantly shorter than the sum of Nb and Sn covalent radii (1.66 + 1.40 = 3.06 Å). It is probably partly multiple in character owing to an additional interaction of the lone electron pair of the NbIII ion (d2 configuration) with the antibonding Sn orbitals. The PMR spectra of II–VI exhibit two satellites of the singlet of C5H5 protons because of HSn117 and HSn119 spin-spin coupling (SSC). The SSC constant correlates with the number of electronegative chlorine atoms on the Sn atom.  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