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.
Oxidation of the cis isomer of the λ3-cyclotriphosphazane [EtNP(OCH2CF3)]3 with trimethylamine-N-oxide (TMNO) gives the cis isomer of trioxo-λ5-cyclotriphosphazane [EtNP(O)(OCH2CF3)]3; the trans isomer of [EtNP(O)(OCH2CF3)]3 is obtained by the treatment of a cis and trans mixture of [EtNP(OCH2CF3)]3 with aqueous H2O2. The two trioxocyclotriphosphazanes have been characterized by elemental analysis, IR, and NMR (1H, 13C, 19F, and 31P) spectroscopy. The solid state structures of both the isomers have been determined by single crystal X-ray diffraction. The six-membered P3N3 ring in both the isomers exhibits a twist-boat conformation; in the cis isomer, the trifluoroethoxy substituents lie on the same side of the ring, whereas, in the trans isomer, two trifluoroethoxy groups are on one side of the ring and the third on the other side of the ring. Crystal data for cis-[EtNP(O)(OCH2CF3)]3: monoclinic, P 21/ n , a = 13.593(3), b = 9.721(2), c = 17.539(3) Å, β = 99.49(2)°, V = 2286(1) Å3, Z = 4, and Final R = 0.047. Crystal data for trans-[EtNP(O)(OCH2CF3)]3: monoclinic, P 21/ n , a = 11.685(4), b = 15.115(5), c = 13.233(5) Å, - = 102.21(3)°, V = 2284(1) Å3, Z = 4, and Final R = 0.078.  相似文献   

2.
RhIII and IrIII complexes based on the λ3‐P,N hybrid ligand 2‐(2′‐pyridyl)‐4,6‐diphenylphosphinine ( 1 ) react selectively at the P?C double bond to chiral coordination compounds of the type [( 1 H ? OH)Cp*MCl]Cl ( 2 , 3 ), which can be deprotonated with triethylamine to eliminate HCl. By using different bases, the pKa value of the P? OH group could be estimated. Whereas [( 1 H ? O)Cp*IrCl] ( 4 ) is formed quantitatively upon treatment with NEt3, the corresponding rhodium compound [( 1 H ? O)Cp*RhCl] ( 5 ) undergoes tautomerization upon formation of the λ5σ4‐phosphinine rhodium(III) complex [( 1? OH)Cp*RhCl] ( 6 ) as confirmed by single‐crystal X‐ray diffraction. Blocking the acidic P? OH functionality in 3 by introducing a P? OCH3 substituent leads directly to the λ5σ4‐phosphinine iridium(III) complex ( 8 ) upon elimination of HCl. These new transformations in the coordination environment of RhIII and IrIII provide an easy and general access to new transition‐metal complexes containing λ5σ4‐phosphinine ligands.  相似文献   

3.
Perfluoroalkenyl phosphonates were formed along with Me3SiF using CF3CF=CF2, CF3CH=CF2, F5SCF=CF2 or F5SCH=CF2 and silylated phosphites, (R1O)2POSiMe3 (R1=Et, SiMe3). This straightforward method could be extended to perfluorobutadienes CF2=C(RF)C(RF)=CF2 (RF F=F, CF3). The formation of CF3C(=O)P(=O)(OSiMe3)2 and further reactions to yield bisphosphonates will be described. Acetylphosphonates, R2C(=O)P(=O)(OSiMe3)2 (R2=CH3, CF3) reacted with the ketimine, CH3C(=NiPr)Ph to give α-hydroxy-γ-imino phosphonates. Trifluoroacetylphenol and 2,6-bis(trifluoracetyl)-4-methyl-phenol have been proven to be versatile precursors for α-and γ-hydroxy phosphonates. Intermediates in these reactions were found to be cyclic λ5σ5P species.  相似文献   

4.
Abstract

The System CF3I/Me3P is re-investigated and Me2PCF3, Me4P+γ, (CF3)2PMe3, Me3PI2, [Me3(CF3)P]+γ are found as products. Using CF3Br/P(NEt2)3 the phosphines R1 2PCF3 and R1P(CF3)2 (e.g. R1 = Me, iPr, NEt2) can be obtained which are precursors either for phosphoranes (e.g. 1,2λ5σ5-oxaphosphetanes) or phosphonium salts (e.g. [R1 2(Me)PCF3]+X? or [R1(Me)P(CF3)2X?]. The latter are deprotonated to furnish methylene phosphoranes R1 2(CH2=)PCF3 or R1(CH2=)P(CF3)2, reactive synthons. From CF3Br/P(NEt2)3/P(OPh)3 the phosphine P(CF3)3 is available, which turned out to be a potent electrophile. Amido phospites ROP(NEt2)2 and halides R2X (R2=CCl2CF3, X=Cl; R2=CF=CFCF3, X=F; R2=C6F5, X=Br, I; R2=C(CF3)3, X=Br; R2=SCF3, X=CF3) undergo an ARBUZOV reaction.  相似文献   

5.
Syntheses of Oxovanadium(V) Halide Complexes Stabilized with Tripodal Oxygen Ligands LR = [η5‐(C5H5)Co{PR2(O)}3], R = OMe, OEt The sodium salts of the tripodal oxygen ligands LR = [η5‐(C5H5)Co{PR2(O)}3] (R = OMe, OEt) react with the oxovanadium halides V(O)F3 and V(O)Cl3 to yield deep red compounds of the type [V(O)X2LR]. Halide exchange reactions with [V(O)Cl2LOMe] und [V(O)F2LOMe] aiming at the preparation of the analogous bromide complex [V(O)Br2LOMe] led to the isomer [VO(LOMe)2][V(O)Br4]. The crystal structure of [V(O)Cl2LOMe] has been determined by single crystal x‐ray diffraction. The compound crystallizes in the monoclinic space group P21/n with a = 9.6332(8), b = 15.0312(11) and c = 15.3742(12)Å, β = 100.181(8)°. The coordination around vanadium is distorted octahedral.  相似文献   

6.
The reactions of pentachloro(2′, 2′, 2′-triphenylphosphazen-1′-yl)cyclotriphosphazene, N3P3Cl5(NPPh3), with primary and secondary amines have been investigated using diethyl ether, methyl cyanid or benzene as the solvent. The structures of the products obtained, N3P3Cl5minus;nRn(NPPh3) [n = 1, R = NHMe, NHBut, NMe2, NC5H10, NEt2; n = 2, R = NMe2, NC5H10, NEt2; n = 3, R = NMe2, NHBut; n = 5, R = NMe2] are elucidated by 1H and 31P NMR spectroscopy. The ? NPPh3 substituent exerts a pronounced geminal directing influence on incoming secondary amino nucleophiles; compounds containing a ≡ PCl(NPPh3) group are not formed at the bis and subsequent stages of chlorine replacement. The reactions that involve primary amines follow the pattern established for the analogous reactions of hexachlorocyclotriphosphazene. The effect of solvent and possible mechanism(s) are discussed.  相似文献   

7.
Synthesis of Fluoro-λ5-monophosphazenes and Fluoro-1,3-diaza-2λ5,4λ5-diphosphetidines by Means of the Staudinger Reaction 35 Tetrafluoro- and 2 difluorodiaza-diphosphetidines as well as 4 difluoro- and 30 monofluoro-λ5-monophosphazenes were prepared by the Staudinger reaction between tervalent phosphorus fluorides, RnPF3?n (n = 1, 2; R = R2N, (CH2)5N, O(CH2)4N, RO, (CH2O)2, alkyl, aryl) and phenylazides, X? C6H4N3 (X = H, 4-CH3, 4-Cl, 4-Br, 4-NO2, 3-NO2). PF3 does not react with phenylazide The influence of substituents on the structure of the reaction products is discussed. Kinetic measurements allowed to determine the constants λPI of the substituents (CH2)5N, O(CH2)4N and R(C6H5)N (R = CH3, C2H5, n-C4H9).  相似文献   

8.
Abstract

The reactions of a variety of electrophiles with the N-silyl-P-trifluoroethoxyphosphoranimine anion Me3Sin°P(Me)(OCH2CF3)CH? 2 (1a), prepared by the deprotonation of the dimethyl precursor Me3SiN[dbnd]P(OCH2CF3)Me2 (1) with n-BuLi in Et2O at-78°C, were studied. Thus, treatment of 1a with alkyl halides, ethyl chloroformate, or bromine afforded the new N-silylphosphoranimine derivatives Me3SiN[dbnd]P(Me)(OCH2CF3)CH2R [2: R = Me, 3: R = CH2Ph, 4: R = CH[sbnd]CH2, 5: R = C(O)OEt, and 6: R = Br]. In another series, when 1a was allowed to react with various carbonyl compounds, 1,2-addition of the anion to the carbonyl group was observed. Quenching with Me3SiCl gave the O-silylated products Me3SiN[dbnd]P(Me)(OCH2CF3)CH2°C(OSiMe3)R1R2 [7: R 1 = R 2 = Me; 8: R 1 = Me, R 2 = Ph; 9: R1 = Me, R 2 = CH[sbnd]CH2; and 10: R 1 = H, R 2 = Ph]. Compounds 2–10 were obtained as distillable, thermally stable liquids and were characterized by NMR spectroscopy (1H, 13C, and 31P) and elemental analysis.  相似文献   

9.
Copolymerizations of ethylene with α‐olefins (i.e., 1‐hexene, 1‐octene, allylbenzene, and 4‐phenyl‐1‐butene) using the bis(β‐enaminoketonato) titanium complexes [(Ph)NC(R2)CHC(R1)O]2TiCl2 ( 1a : R1 = CF3, R2 = CH3; 1b : R1 = Ph, R2 = CF3; and 1c : R1 = t‐Bu, R2 = CF3), activated with modified methylaluminoxane as a cocatalyst, have been investigated. The catalyst activity, comonomer incorporation, and molecular weight, and molecular weight distribution of the polymers produced can be controlled over a wide range by the variation of the catalyst structure, α‐olefin, and reaction parameters such as the comonomer feed concentration. The substituents R1 and R2 of the ligands affect considerably both the catalyst activity and comonomer incorporation. Precatalyst 1a exhibits high catalytic activity and produces high‐molecular‐weight copolymers with high α‐olefin insertion. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 6323–6330, 2005  相似文献   

10.
Ethylene–propylene copolymerization, using [(Ph)NC(R2)CHC(R1)O]2TiCl2 (R1 = CF3, Ph, or t‐Bu; R2 = CH3 or CF3) titanium complexes activated with modified methylaluminoxane as a cocatalyst, was investigated. High‐molecular‐weight ethylene–propylene copolymers with relatively narrow molecular weight distributions and a broad range of chemical compositions were obtained. Substituents R1 and R2 influenced the copolymerization behavior, including the copolymerization activity, methylene sequence distribution, molecular weight, and polydispersity. With small steric hindrance at R1 and R2, one complex (R1 = CF3; R2 = CH3) displayed high catalytic activity and produced copolymers with high propylene incorporation but low molecular weight. The microstructures of the copolymers were analyzed with 13C NMR to determine the methylene sequence distribution and number‐average sequence lengths of uninterrupted methylene carbons. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 5846–5854, 2006  相似文献   

11.
A series of heteroligated (salicylaldiminato)(β‐enaminoketonato)titanium complexes [3‐But‐2‐OC6H3CH = N(C6F5)] [PhN = C(R1)CHC(R2)O]TiCl2 [ 3a : R1 = CF3, R2 = tBu; 3b : R1 = Me, R2 = CF3; 3c : R1 = CF3, R2 = Ph; 3d : R1 = CF3, R2 = C6H4Ph(p ); 3e : R1 = CF3, R2 = C6H4Ph(o ); 3f : R = CF3, R2 = C6H4Cl(p ); 3g : R1 = CF3; R2 = C6H3Cl2(2,5); 3h : R1 = CF3, R2 = C6H4Me(p )] were investigated as catalysts for ethylene (co)polymerization. In the presence of modified methylaluminoxane as a cocatalyst, these complexes showed activities about 50%–1000% and 10%–100% higher than their corresponding bis(β‐enaminoketonato) titanium complexes for ethylene homo‐ and ethylene/1‐hexene copolymerization, respectively. They produced high or moderate molecular weight copolymers with 1‐hexene incorporations about 10%–200% higher than their homoligated counterpart pentafluorinated FI‐Ti complex. Among them, complex 3b displayed the highest activity [2.06 × 106 g/molTi?h], affording copolymers with the highest 1‐hexene incorporations of 34.8 mol% under mild conditions. Moreover, catalyst 3h with electron‐donating group not only exhibited much higher 1‐hexene incorporations (9.0 mol% vs. 3.2 mol%) than pentafluorinated FI‐Ti complex but also generated copolymers with similar narrow molecular weight distributions (M w/M n = 1.20–1.26). When the 1‐hexene concentration in the feed was about 2.0 mol/L and the hexene incorporation of resultant polymer was about 9.0 mol%, a quasi‐living copolymerization behavior could be achieved. 1H and 13C NMR spectroscopic analysis of their resulting copolymers demonstrated the possible copolymerization mechanism, which was related with the chain initiation, monomer insertion style, chain transfer and termination during the polymerization process. © 2017 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2017 , 55 , 2787–2797  相似文献   

12.
Enantiomerically pure triflones R1CH(R2)SO2CF3 have been synthesized starting from the corresponding chiral alcohols via thiols and trifluoromethylsulfanes. Key steps of the syntheses of the sulfanes are the photochemical trifluoromethylation of the thiols with CF3Hal (Hal=halide) or substitution of alkoxyphosphinediamines with CF3SSCF3. The deprotonation of RCH(Me)SO2CF3 (R=CH2Ph, iHex) with nBuLi with the formation of salts [RC(Me)? SO2CF3]Li and their electrophilic capture both occurred with high enantioselectivities. Displacement of the SO2CF3 group of (S)‐MeOCH2C(Me)(CH2Ph)SO2CF3 (95 % ee) by an ethyl group through the reaction with AlEt3 gave alkane MeOCH2C(Me)(CH2Ph)Et of 96 % ee. Racemization of salts [R1C(R2)SO2CF3]Li follows first‐order kinetics and is mainly an enthalpic process with small negative activation entropy as revealed by polarimetry and dynamic NMR (DNMR) spectroscopy. This is in accordance with a Cα? S bond rotation as the rate‐determining step. Lithium α‐(S)‐trifluoromethyl‐ and α‐(S)‐nonafluorobutylsulfonyl carbanion salts have a much higher racemization barrier than the corresponding α‐(S)‐tert‐butylsulfonyl carbanion salts. Whereas [PhCH2C(Me)SO2tBu]Li/DMPU (DMPU = dimethylpropylurea) has a half‐life of racemization at ?105 °C of 2.4 h, that of [PhCH2C(Me)SO2CF3]Li at ?78 °C is 30 d. DNMR spectroscopy of amides (PhCH2)2NSO2CF3 and (PhCH2)N(Ph)SO2CF3 gave N? S rotational barriers that seem to be distinctly higher than those of nonfluorinated sulfonamides. NMR spectroscopy of [PhCH2C(Ph)SO2R]M (M=Li, K, NBu4; R=CF3, tBu) shows for both salts a confinement of the negative charge mainly to the Cα atom and a significant benzylic stabilization that is weaker in the trifluoromethylsulfonyl carbanion. According to crystal structure analyses, the carbanions of salts {[PhCH2C(Ph)SO2CF3]Li? L }2 ( L =2 THF, tetramethylethylenediamine (TMEDA)) and [PhCH2C(Ph)SO2CF3]NBu4 have the typical chiral Cα? S conformation of α‐sulfonyl carbanions, planar Cα atoms, and short Cα? S bonds. Ab initio calculations of [MeC(Ph)SO2tBu]? and [MeC(Ph)SO2CF3]? showed for the fluorinated carbanion stronger nC→σ* and nO→σ* interactions and a weaker benzylic stabilization. According to natural bond orbital (NBO) calculations of [R1C(R2)SO2R]? (R=tBu, CF3) the nC→σ*S? R interaction is much stronger for R=CF3. Ab initio calculations gave for [MeC(Ph)SO2tBu]Li ? 2 Me2O an O,Li,Cα contact ion pair (CIP) and for [MeC(Ph)SO2CF3]Li ? 2 Me2O an O,Li,O CIP. According to cryoscopy, [PhCH2C(Ph)SO2CF3]Li, [iHexC(Me)SO2CF3]Li, and [PhCH2C(Ph)SO2CF3]NBu4 predominantly form monomers in tetrahydrofuran (THF) at ?108 °C. The NMR spectroscopic data of salts [R1(R2)SO2R3]Li (R3=tBu, CF3) indicate that the dominating monomeric CIPs are devoid of Cα? Li bonds.  相似文献   

13.
Bis(trifluoromethyl)phosphines RP(CF3)2 (R = Me, NEt2) were methylated by MeOSO2CF3, yielding the respective phosphonium salts [RP(CF3)2Me]+ and CF3SO3. Deprotonation using MeNP(NEt2)3 led to the phosphorus ylides RP(CF3)2CH2, stable in solution at ambient temperature, which could be converted into 1,2λ5σ5‐oxaphosphetanes by adding hexafluoroacetone. © 2002 Wiley Periodicals, Inc. Heteroatom Chem 13:650–653, 2002; Published online in Wiley InterScience (www.interscience.wiley.com). DOI 10.1002/hc.10061  相似文献   

14.
F2P(NEt2)3 – Difluorophosphanes as Versatile Fluoridation Agent F2P(NEt2)3 ( 1 ), prepared by the reaction of P(NEt2)3 with SF4 (3:2) in Et2O at –78°C, acts as fluoridation agent on Lewis acids such as AlMe3, GaMe3, or [Mes3V(THF)] by transfer of a F ion. The products [FP(NEt2)3][Me3MF] [M = Al ( 2 ), Ga ( 3 )] and [FP(NEt2)3][Mes3VF] ( 4 ) were characterized by NMR, IR and MS techniques. With 1 and 4 an X‐ray structure determination could be performed. According to them 1 consists of trigonal‐bipyramidal λ5‐phosphane molecules with an occupation of the axial positions by F ligands. In 4 the centers of the cations [FP(NEt2)3]+ and of the anions [Mes3VF] are distorted terahedrally coordinated.  相似文献   

15.
On Sn[OCH(CF3)2]2 and Sn(OCH2CF3)2 (n = 1, 2) The sulfoxylates S[OCH(R)CF3]2, 1 and 2 and the disulfides S2[OCH(R)CF3]2, 5 and 6 (R = CF3, H) are obtained by reacting SCl2 or S2Cl2, respectively, and the lithium alcoxides LiOCH(R)CF3. Chlorine and compound 2 give ClS(O)OCH2F3 and CF3CH2Cl, whereas the sulfur-sulfur bound is cleaved in 5 and 6 furnishing SCI2, 1 and 2 , respectively. The 19F n.m.r. spectrum of 5 and the 1H n.m.r. spectrum of 6 are interpreted in terms of hindered rotation about the sulfur-sulfur axis.  相似文献   

16.
Trifluoromethylated phosphines R2PCF3 (R = NEt2, Me, 1Pr) were methylated by CH4OSO2CF3, yielding the corresponding phosphonium salts [R2P(CF3)CH3] [F3CSO3]. Treatment with LiN(SiMe3)2 at −80°C furnished the phosphorus ylides R2P(CF3) = CH2 that could be trapped by use of hexafluoroacetone with formation of stable 1,2λ5σ5-oxaphosphetanes. The single-crystal X-ray structure determination of one of these oxaphosphetanes showed a distorted trigonal bipyramid at phosphorus with the P-CF3 group in an axial position. © 1996 John Wiley & Sons, Inc.  相似文献   

17.
The introduction of the organosilicon substituent into the α‐position of an amino group results in cardinal change of the amine reactivity irrespective of the coordination state of silicon. Amines R2NCH2SiX3 [R = Me, Et, PhCH2, CH2SiX3; SiX3 = SiMe3, Si(OEt)3, Si(OCH2CH2)3N] easily react with AgNO3, to give the corresponding ammonium salts (R2NH+ CH2SiX3)·NO3?. At the same time, Ag(I) is reduced to Ag(0). The interaction of N‐methyl‐N,N‐bis(silatranylmethyl)amine with AgNO3 has been investigated by EPR spectroscopy. It was proven that the reaction involved a single electron transfer stage with the formation of cation radical of this amine. A mechanism of the reaction is proposed. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

18.
Reaction of the Two-component Systems P(OR)3 ? x(NR2)x (x = 0–3)/CCl4 and P4/CCl4 with HF-Donators The combination of organylammonium fluorides and carbon tetrachloride is a good agent for oxidative fluorination of trivalent phosphorus compounds. As oxidation products [(RO)PF5]? and (RO)2P(O)F are obtained from P(OR)3, (Et2N)2P(O)F and (Et2N)2(EtO)PF2 from P(OEt)(NEt2)2 as well as (Et2N)3PF2 and [(Et2N)3PF]+ from P(NEt2)3. In the system R2NH/CCl4/Et3N · n HF P4 is fast oxidized forming [HPF5]?, R2NH · PF5 and (R2N)2P(O)F. In the case of simultaneous addition of alcohols [(RO)PF5]?, (RO)3PO and (R2N)2P(O)F are formed. The reactions are controlled by the nucleophilic power and the concentration of fluoride, the acidity of the system, and the temperature.  相似文献   

19.
Tris)(η 5-cyclopentadienyl-μ-carbonyl-iron)-μ3-nitrosyl cluster was obtained from the reaction of cyclopentadienyl dicarbonyliron dimer with nitrogen monoxide in xylene. The cluster was characterized by elemental analyses, IR, MS and 1H NMR. The crystal structure of [(η5-C5H5)(μ-CO)Fe]3(μ3-NO).C4H8O was determined by X-ray diffraction analysis. It crystallizes in the orthorhombic space group Pnma, a=9.053(2), 6=10.545(2), c=22.525(4) A, V=2150.3(7) A3, Z=4,Dc=1.68 g.cm-3; structure solution and refinement based on 1141 reflections with I > 3.0 (I) (MoKa, A=0.71073 A) converged at R=0.0540. The infrared absorption band at 1325 cm-1 of the μ3-NO in the cluster, which is red shifted, shows that μ3-NO is activated.  相似文献   

20.
Reactions of rubidium or barium salts of the ortho‐selenostannate anion, [Rb4(H2O)4][SnSe4] ( 1 ) or [Ba2(H2O)5][SnSe4] ( 2 ) with Zn(OAc)2 or ZnCl2 in aqueous solution yielded two novel compounds with different ternary Zn/Sn/Se anions, [Rb10(H2O)14.5][Zn4(μ4‐Se)2(SnSe4)4] ( 3 ) and [Ba5(H2O)32][Zn5Sn(μ3‐Se)4(SnSe4)4] ( 4 ). 1 – 4 have been determined by means of single crystal X‐ray diffraction: 1 : triclinic space group lattice dimensions at 203 K: a = 8.2582(17) Å, b = 10.634(2) Å, c = 10.922(2) Å, α = 110.16(3)°, β = 91.74(3)°, γ = 97.86(3)°, V = 888.8(3) Å3; R1 [I > 2σ(I)] = 0.0669; wR2 = 0.1619; 2 : orthorhombic space group Pnma; lattice dimensions at 203 K: a = 17.828(4) Å, b = 11.101(2) Å, c = 6.7784(14) Å, V = 1341.5(5) Å3; R1 [I > 2σ(I)] = 0.0561; wR2 = 0.1523; 3 : triclinic space group ; lattice dimension at 203 K: a = 17.431(4) Å, b = 17.459(4) Å, c = 22.730(5) Å, α = 105.82(3)°, β = 99.17(3)°, γ = 90.06(3)°, V = 6563.1(2) Å3; R1 [I > 2σ(I)] = 0.0822; wR2 = 0.1782; 4 : monoclinic space group P21/c; lattice dimensions at 203 K: a = 25.231(5) Å, b = 24.776(5) Å, c = 25.396(5) Å, β = 106.59(3)°, V = 15215.0(5) Å3; R1 [I > 2σ(I)] = 0.0767; wR2 = 0.1734. The results serve to underline the crucial role of the counterion for the type of ternary anion to be observed in the crystal. Whereas Rb+(aq) stabilizes a P1‐type Zn/Sn/Se supertetrahedron in 3 like K+, the Ba2+(aq) ions better fit to an anionic T3‐type Zn/Sn/Se cluster arrangement as do Na+ ions. It is possible to estimate a radius:charge ratio for the stabilization of the two structural motifs.  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