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 656 毫秒
1.
The mixed pincer palladacycles (Me2NCH2(Cl)CCCH2CH2Y‐κNCY)PdCl ( 1 , Y=PPh2; 2 , OPPh2) and (t‐BuSCH2CH2CC(Cl)(o‐NC5H4)‐κSCN)PdCl 3 have been obtained in high yields by chloropalladation of heterosusbstituted alkynes Me2NCH2CCCH2CH2PPh2, Me2NCH2CCCH2CH2OPPh2 and t‐BuSCH2CH2CC(o‐NC5H4), respectively. The molecular structures of 1 and 3 have been ascertained by means of X‐ray diffraction analysis. The catalytic properties of these mixed donor group pincer‐type palladacycles have been evaluated in the arylation of olefins (Heck reaction). The pincer palladacycle 1 is highly active for the coupling of aryl iodides and aryl bromides with n‐butyl acrylate. In contrast it is only moderately active for the coupling of aryl chlorides substituted with electron‐withdrawing groups and inactive for the coupling of electron neutral and electron deactivated aryl chlorides.  相似文献   

2.
The reaction of the Cu(II) bis N,O‐chelate‐complexes of L‐2,4‐diaminobutyric acid, L‐ornithine and L‐lysine {Cu[H2N–CH(COO)(CH2)nNH3]2}2+(Cl)2 (n = 2–4) with terephthaloyl dichloride or isophthaloyl dichloride gives the polymeric complexes {‐OC–C6H4–CO–NH–(CH2)n–CH(nh2)(COO)Cu(OOC)(NH2)CH–CH2)n–NH‐}x 1 – 5 . From these the metal can be removed by precipitation of Cu(II) with H2S. The liberated ω,ω′‐N,N′‐diterephthaloyl (or iso‐phthaloyl)‐diaminoacids 6 – 10 react with [Ru(cymene)Cl2]2, [Ru(C6Me6)Cl2]2, [Cp*RhCl2]2 or [Cp*IrCl2]2 to the ligand bridged bis‐amino acidate complexes [Ln(Cl)M–(OOC)(NH2)CH–(CH2)nNH–CO]2–C6H4 11 – 14 .  相似文献   

3.
A large variety of conjugated dienones R1R2CCHCHC(R3)C(O)R4 and diene‐diones R1R2CCHCHC{C(O)R3}C(O)R4 have been synthesized in high yields by reacting terminal propargylic alcohols HCCCR1R2(OH) with enolizable ketones R3CH2C(O)R4 and β‐dicarbonyl compounds R3C(O)CH2C(O)R4, respectively. The process, which is catalyzed by the 16e3‐allyl)‐ruthenium(II ) complex [Ru(η3‐2‐C3H4Me)(CO)(dppf)] [SbF6] associated with CF3CO2H, involves the initial isomerization of the propargylic alcohol into the corresponding α,β‐unsaturated aldehyde R1R2CCHCHO (Meyer–Schuster rearrangement) and subsequent aldol‐type condensation.  相似文献   

4.
The reaction of [Ir(μ‐Cl)(COD)]2 with various fluorous derivatives of triphenylphosphane containing a para‐, meta‐, or ortho‐(1H,1H‐perfluoroalkoxy)‐substituted fluorous phosphane P(C6H4‐ORf)3 (Rf=CH2C7F15, CH2CH2CH2C8F17) and CO (1 atm) gives the corresponding trans‐[Ir(μ‐Cl)(CO){P(C6H4ORf)3}2]. The IR νCO values of these complexes give some information on the donor/acceptor properties of the phosphanes. These fluorous derivatives of triphenylphosphane, as well as a phosphane bearing two (1H,1H‐perfluoroalkyloxy) chains at the 3,5‐positions, were used in association with [Rh(μ‐Cl)(COD)]2 or [Rh(COD)2]PF6 in the reduction of methyl cinnamate, 2‐cyclohexen‐1‐one, cinnamaldehyde, and methyl α‐acetamidocinnamate in a two‐phase system D‐100/ethanol under 1 bar hydrogen at room temperature. Some differences in catalytic activity were observed in the reduction of methyl cinnamate, the most active catalyst being the rhodium complex containing the phosphane with the p‐fluorous ponytail. Recycling of the fluorous catalyst was possible, particularly using the p‐substituted phosphane, where no significant loss of catalyst or activity was observed, and generally with very low leaching of rhodium or phosphane in the organic phase.  相似文献   

5.
Soluble conjugated polymers were obtained in the presence of Pd(II), Pt(II) and Rh(I) complexes from monosubstituted acetylene 3‐dimethylamino‐1‐propyne (H? C≡CCH2N(CH3)2, 1 ) and the corresponding hydrochloride (H? C≡CCH2N(CH3)2·HCl, 2 ) and hydrobromide (H? C≡CCH2N(CH3)2·HBr, 3 ) derivatives. A series of reactions were performed to achieve the optimization of the polymerization conditions. The highest yields were found for polymers synthesized using Pd(II) bisacetylides specially prepared, i.e. trans‐[Pd(PPh3)2(C≡CCH2N(CH3)2)2], trans‐[Pd(PPh3)2(C≡CCH2N(CH3)2)2HCl] and trans‐[Pd(PPh3)2(C≡CCH2N(CH3)2)2HBr], respectively. The dimension and size distribution of the polymers were investigated using dynamic light scattering. Polymers containing quaternary ammonium groups showed evidence of a hydrodynamic radius of about 300 nm if prepared with the Rh(I) catalyst and of 160 nm if prepared with the Pd(II) catalysts. Polymers obtained from 1 showed smaller hydrodynamic radius compared to polymers obtained from 2 and 3 , regardless the polymerization catalyst. The ionic polymeric materials were soluble in organic solvents and, more interestingly, in water. The formation of nanoparticles with pearl‐like morphology was achieved using a recently developed osmosis‐based method, with dimensions varying from 60 nm up to micrometres. Copyright © 2011 Society of Chemical Industry  相似文献   

6.
Iridium(III) complexes of the type [Ir(η5‐C5Me5)Cl2{Ph2PCH2CH2CH2S(O)xPh‐κP}] (x=0–2; 1 – 3 ) and [Ir(η5‐C5Me5)Cl{Ph2PCH2CH2CH2S(O)xPh‐κPS}][PF6] (x=0–1; 4 and 5 ) with 3‐(diphenylphosphino)propyl phenyl sulfide, sulfoxide, and sulfone ligands Ph2PCH2CH2CH2S(O)xPh were designed, synthesized, and characterized fully, including X‐ray diffraction analyses for complexes 3 and 4 . In vitro studies against human thyroid carcinoma (8505C), submandibular carcinoma (A253), breast adenocarcinoma (MCF‐7), colon adenocarcinoma (SW480), and melanoma (518A2) cell lines provided evidence for the high biological potential of the neutral and cationic iridium(III) complexes. Neutral iridium(III) complex 5 proved to be the most active, with IC50 values up to about 0.1 μM , representing activities of up to one order of magnitude higher than cisplatin. Using 8505C cells, apoptosis was shown to be the main mechanism through which complex 5 exerts its tumoricidal action. The described iridium(III) complexes represent potential leads in the search for novel metal‐based anticancer agents.  相似文献   

7.
Copolymerization of ethylene with 1‐octadecene was studied using [η51‐C5Me4‐4‐R1‐6‐R‐C6H2O]TiCl2 [R1 = tBu (1), H (2, 3, 4); R = tBu (1, 2), Me (3), Ph (4)] as catalysts in the presence of Al(i‐Bu)3 and [Ph3C][B(C6F5)4]. The effect of the concentration of comonomer in the feed and Al/Ti molar ratio on the catalytic activity and molecular weight of the resultant copolymer were investigated. The substituents on the phenyl ring of the ligand affect considerably both the catalytic activity and comonomer incorporation. The 1 /Al(i‐Bu)3/[Ph3C][B(C6F5)4] catalyst system exhibits the highest catalytic activity and produces copolymers with the highest molecular weight, while the 2 /Al(i‐Bu)3/[Ph3C][B(C6F5)4] catalyst system gives copolymers with the highest comonomer incorporation under similar conditions. © 2010 Wiley Periodicals, Inc. J Appl Polym Sci, 2011  相似文献   

8.
The oxorhenium(V) chelates [ReOCl(N,O‐L)(PPh3)] [N,O‐L=(OCH2CH2)N(CH2CH2OH)(CH2COO) ( 2 ), (OCH2CH2)N(CH2COO)(CH2COOCH3) ( 3 )] and [ReOCl2(N,O‐L)(PPh3)] [N,O‐L=C5H4N(COO‐2) ( 4 ) C5H3N(COOCH3‐2)(COO‐6) ( 5 )] have been prepared by reaction of [ReOCl3(PPh3)2] ( 1 ), in refluxing methanol, with N,N‐bis(2‐hydroxyethyl)glycine [bicine; N(CH2CH2OH)2(CH2COOH)], N‐(2‐hydroxyethyl)iminodiacetic acid [N(CH2CH2OH)(CH2COOH)2], picolinic acid [NC5H4(COOH‐2)] or 2,6‐pyridinedicarboxylic acid [NC5H3(COOH‐2,6)2], respectively, with ligand esterification in the cases of 3 and 5 . All these complexes have been characterized by IR and multinuclear NMR spectroscopy, FAB+‐MS, elemental and X‐ray diffraction structural analyses. They act as catalysts, in a single‐pot process, for the carboxylation of ethane by CO, in the presence of potassium peroxodisulfate K2S2O8, in trifluoroacetic acid (TFA), to give propionic and acetic acids, in a remarkable yield (up to ca. 30%) and under relatively mild conditions, with some advantages over the industrial processes. The picolinate complex 4 provides the most active catalyst and the carboxylation also occurs, although much less efficiently, by the TFA solvent in the absence of CO. The selectivity can be controlled by the ethane and CO pressures, propionic acid being the dominant product for pressures about ca. 7 and 4 atm, respectively (catalyst 4 ), whereas lower pressures lead mainly to acetic acid in lower yields. These reactions constitute an unprecedented use of Re complexes as catalysts in alkane functionalization.  相似文献   

9.
An N‐propargylamide monomer, CH?CCH2NHCOC(CH3)2CH2CH3 (monomer 9), was polymerized in the presence of (nbd)Rh+B?(C6H5)4 (nbd represents norbornadiene) in CH2Cl2, CHCl3, tetrahydrofuran or dimethylformamide, to provide polymers with moderate number‐average molecular weights (Mn = 8700–12 100 g mol?1) in high yields (≥92%). The resulting poly(N‐propargylamide) (polymer 9) dissolves almost completely in CHCl3 (>95%). According to the UV‐visible spectra, measured at various temperatures, polymer 9 forms relatively stable helices over a wide temperature range (35–65 °C). Moreover, it exhibits reversible conformational transitions from an ordered helix to a random coil. On copolymerization of monomer 9 with CH?CCH2NHCO(CH2)3CH3 (monomer 4) or CH?CCH2NHCO(CH2)7CH3 (monomer 8), the solubility of polymer 9 improves noticeably. All the copolymers form helices under the experimental conditions. From the viewpoint of monomers 4 and 8, copolymerization with monomer 9 is favorable in terms of the copolymers forming helices. These findings reveal that the helical content and thermodynamic stability of the helices formed in the copolymers are likely to be controlled by selecting a suitable comonomer and by adjusting the composition of the copolymer. Copyright © 2007 Society of Chemical Industry  相似文献   

10.
The hexamethylbenzene ruthenium(II) dimer [{RuCl(μ‐Cl)(η6‐C6Me6)}]2 (5 mol%), tested among a series of ruthenium(II) and ruthenium(IV) complexes, represents an efficient precatalyst source for the dimerization of terminal arylalkynes ArCCH [Ar=C6H5, 3,4,5‐(OMe)3C6H2, 4‐MeOC6H4, 2‐MeOC6H4, 4‐MeC6H4, 2,4,5‐Me3C6H2, 4‐BrC6H4, 4‐ClC6H4, 4‐FC6H4, 4‐HC(O)C6H4, 4‐CH2CHC6H4, 3‐NCC6H4, 4‐O2NC6H4, 4‐EtO2C‐(CH2)3OC6H4, 4‐HO(CH2CH2O)3C6H4, 3‐HO(CH2CH2O)3‐C6H4] in acetic acid/water mixture (1:1, v/v). The reactions proceed for 24 h at room temperature under heterogeneous conditions and afford the dimeric enyne derivatives (E)‐Ar CHCH CC Ar in high yields and stereoselectivity. The preformed acetato complex [RuCl(η6‐C6Me6)(κ2‐OAc)] catalyzes the dimerization of phenylacetylene under analogous conditions, with rapid substrate conversion. The presence of cosolvents of acetic acid different from water reduces dramatically the efficiency and selectivity of the reaction. The aqueous medium facilitates the activation stage of the precatalyst by assisting the splitting of the ruthenium dimer. The addition or generation in situ of acetate salts results in shorter reactions times (0.5–3 h) and excellent yields, due to the rapid formation of active acetato complexes. Circumstantial evidence indicates that the π‐bound alkyne molecule is activated by intramolecular proton abstraction. This is currently the most efficient, E‐selective and wide‐scope catalytic system for the alkyne dimerization reaction in protic aqueous media.  相似文献   

11.
Diethyl-2,2′-biquinoline-4,4′-dicarboxylate (debq) reacts with pentacarbonylbromorhenium in toluene to give fac-[Re(CO)3(debq)Br]; subsequent reaction with AgOTf in MeCN affords the cationic complex fac-[Re(CO)3(debq)(MeCN)]OTf. X-ray crystallographic data on fac-[Re(CO)3(debq)Br] show that the debq ligand is coordinated in a bidentate fashion at an angle of 32° from the equatorial plane of the complex: DFT calculations confirm this as the lowest energy conformation. fac-[Re(CO)3(debq)Br] and fac-[Re(CO)3(debq)(MeCN)]OTf both possess 1MLCT absorptions in the visible region, the former most red-shifted to ca. 465 nm. Both complexes demonstrated unusual luminescent properties: in solution, emission appears to be dominated by ligand-centred processes, whereas only 3MLCT emission, tailing into the near-IR, was observed in the solid state at 671 (τ = 59 ns) and 711 nm (τ = 19 ns) for fac-[Re(CO)3(debq)(MeCN)]OTf and fac-[Re(CO)3(debq)Br], respectively. However, the debq ligand is not labile in these complexes, which are robust to competing ligands and coordinating solvents.  相似文献   

12.
Block copolymers having a pendant trichlorogermyl group as a part of polyamide segment? (CO? R′? CO? NH? Ar? NH? )xCO? R′? CO? and polydimethylsiloxane of general formula [(? CO? R′? CO? HN? Ar? NH)x? CO? R′? CO? NH(CH2)3SiO(CH3)2 ((CH3)2SiO)ySi(CH3)2(CH2)3 NH? ]n (where R′ = CH2CH(GeCl3), CH(CH3)CH(GeCl3), CH(GeCl3)CH(CH3); Ar = C6H4, (? C6H3? CH3)2, (? C6H3? OCH3)2, 2,5‐(CH3)2? C6H2, C6H4? O? C6H4) were prepared by a polycondensation reaction and characterized using CHN and Ge analysis, Fourier transform infrared (FTIR) and 1H NMR spectroscopy, thermogravimetric analysis (TGA) and molecular weight determination. They have a lamellar structure with weight‐average molecular weight in the range 1.21 × 105–4.79 × 105 g mol?1. These copolymers display two glass transition temperatures and have an average decomposition temperature of 489 °C. TGA, FTIR and gas chromatography/mass spectrometry studies indicate that degradation of these block copolymers results in carbon monoxide, oligomeric siloxanes and polyamide fragments. They are thermally stable due to the hydrogen bonded interlinked chains of polyamide, while they absorb water due to the presence of Ge? Cl bonding. Copyright © 2010 Society of Chemical Industry  相似文献   

13.
An asymmetric 3‐oxa‐pentamethylene bridged dinuclear titanocenium complex (CpTiCl2)25‐η5‐C9H6(CH2CH2OCH2CH2)C5H4) ( 1 ) has been prepared by treating two equivalents of CpTiCl3 with the corresponding dilithium salts of the ligand C9H7(CH2CH2OCH2 CH2)C5H5. The complex 1 was characterized by 1H‐, 13C‐NMR, and elemental analysis. Homogenous ethylene polymerization catalyzed using complex 1 has been conducted in the presence of methylaluminoxane (MAO). The influences ofreaction parameters, such as [MAO]/[Cat] molar ratio, catalyst concentration, ethylene pressure, temperature, and time have been studied in detail. The results show that the catalytic activity and the molecular weight (MW) of polyethylene produced by 1 /MAO decrease gradually with increasing the catalyst concentration or polymerization temperature. The most important feature of this catalytic system is the molecular weight distribution (MWD) of polyethylene reaching 12.4, which is higher than using common mononuclear metallocenes, as well as asymmetric dinuclear titanocene complexes like [(CpTiCl2)25‐η5‐C9H6(CH2)nC5H4)] (n = 3, MWD = 7.31; n = 4, MWD = 6.91). The melting point of polyethylene is higher than 135°C, indicating highly linear and highly crystalline polymers. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  相似文献   

14.
Stereoregular trans‐arylene‐silylene‐vinylene polymers of Mw=13100–34800 and PDI=1.6–2.9 of the general formulas CH2CH [ SiMe2C6H4‐SiMe2CHCH ] ( 16, 17, 18 ) and CH2CH [ (R)CHCHC6H4CHCH ] (where R= Me2Si‐p C6H4‐ SiMe2 ,  Me2Si‐m C6H4SiMe2 and  Me2SiC6H4C6H4SiMe2 ) ( 19, 20, 21 ) have been effectively synthesized via silylative coupling (SC) homopolycondensation of bis(vinyldimethylsilyl)arenes ( 10, 12, 14 ) and cross‐polycondensation of 4‐(vinyldimethylsilyl)styrene ( 11 ) as well as cross‐copolycondensation of bis(vinyldimethylsilyl)arenes ( 10, 12 and 14 ) with 1,4‐divinylbenzene ( 9 ) catalyzed by [RuH(Cl)(CO)(PCy3)2] ( 7 ). Such highly stereoregular products cannot be synthesized via ADMET polycondensation or ring opening metathesis ROM or polyaddition of hydridosilanes to acetylenes.  相似文献   

15.
Novel facial tricarbonylrhenium(I) complexes, fac-[Re(urca)2(CO)3Cl] (1) and fac-[Re(bzta)(CO)3Cl] (2) were isolated from the coordination reactions of 5-((5-hydroxypentylimino)methyl)uracil (urca) and 2-((5-hydroxypentylimino)methyl)benzothiazole (bzta) with [Re(CO)5Cl], respectively. Spectral characterization of metal complexes 1 and 2 were supported by their X-ray crystal structures. DNA interactions were assessed via UV–Vis calf-thymus (CT)-DNA binding titrations and gel electrophoresis. Redox properties of the metal complexes were probed using voltammetry.  相似文献   

16.
Regioselective Markovnikov‐type addition of PhSH to alkynes (HC≡C‐R) has been performed using easily available nickel complexes. The non‐catalytic side reaction leading to anti‐Markovnikov products was suppressed by addition of γ‐terpinene to the catalytic system. The other side reaction leading to the bis(phenylthio)alkene was avoided by excluding phosphine and phosphite ligands from the catalytic system. It was found that catalytic amounts of Et3N significantly increased the yield and selectivity of the catalytic reaction. Under optimized conditions high product yields of 60–85% were obtained for various alkynes [R=n‐C5H11, CH2NMe2, CH2OMe, CH2SPh, C6H11(OH), (CH2)3CN]. The X‐ray structure of one of the synthesized products is reported.  相似文献   

17.
Ruthenium complexes with the formulae Ru(CO)2(PR3)2(O2CPh)2 [ 6a – h ; R=n‐Bu, p‐MeO‐C6H4, p‐Me‐C6H4, Ph, p‐Cl‐C6H4, m‐Cl‐C6H4, p‐CF3‐C6H4, m,m′‐(CF3)2C6H3] were prepared by treatment of triruthenium dodecacarbonyl [Ru3(CO)12] with the respective phosphine and benzoic acid or by the conversion of Ru(CO)3(PR3)2 ( 8e – h ) with benzoic acid. During the preparation of 8 , ruthenium hydride complexes of type Ru(CO)(PR3)3(H)2 ( 9g , h ) could be isolated as side products. The molecular structures of the newly synthesized complexes in the solid state are discussed. Compounds 6a – h were found to be highly effective catalysts in the addition of carboxylic acids to propargylic alcohols to give valuable β‐oxo esters. The catalyst screening revealed a considerably influence of the phosphine′s electronic nature on the resulting activities. The best performances were obtained with complexes 6g and 6h , featuring electron‐withdrawing phosphine ligands. Additionally, catalyst 6g is very active in the conversion of sterically demanding substrates, leading to a broad substrate scope. The catalytic preparation of simple as well as challenging substrates succeeds with catalyst 6g in yields that often exceed those of established literature systems. Furthermore, the reactions can be carried out with catalyst loadings down to 0.1 mol% and reaction temperatures down to 50 °C.

  相似文献   


18.
The reaction of [RhCl(PiPr3)2] ( 1 ) with 1,4-C6H4(C≡CH)2 at 0°C leads almost quantitatively to the formation of the bis(alkyne) complex [(PiPr3)2ClRh-(HC≡C6H4-C≡CH)RhCl(PiPr3)2] (2). At elevated temperatures (THF, 60°C) it rearranges to give the isomeric bis(vinylidene) complex [(PiPr3)2ClRh-(=C=CH-C6H4-CH=C=)RhCl(PiPr3)2] (3). A one-pot synthesis of 3 is also described. Treatment of either 2 or 3 with pyridine affords the bis(alkynyl)dihydrido compound [(PiPr3)2(py)Cl(H)Rh(-C≡C-C6H4-C≡C-)Rh(H)Cl(py)(PiPr3)2] ( 4 ) in which both metal centers are octahedrally coordinated. Whereas the reaction of 2 with NaC5H5 produces the complex CsH5(PiPr3)Rh(HC≡C-C6H4-C≡CH)Rh(PiPr3)C5H5 ( 7 ), the bis(vinyl-idene) isomer C5H5(PiPr3)Rh(=C=CH-C6H4-CH=C=)Rh(PiPr3)C5Hs ( 8 ) is obtained from 4 and NaC5H5. Electrophiles preferably attack the Rh=C bonds of 8 and thus on protonation with CF3CO2H the bis(vinyl) complex C5H5(PiPr3)(CF3CO2)-Rh(Z,Z-CH=CH-C6H4-CH=CH)Rh(O2CCF3)(PiPr3)C5Hs ( Z-9 ) is formed. In acetone solution, it rearranges to give the E isomer. Reaction of 8 with sulfur affords the bis(thioketene) complex C5H5(PiPr3)Rh(≡2-C,S; η2-C,S-S=C=CH-C6H4-CH=C=S)-Rh(PiPr3)C5H5 ( 12 ), for which only one diastereomer is observed. All attempts to prepare mononuclear rhodium compounds containing the diyne HC≡C-C6H4-G≡CH or the isomeric vinylidene: C=CH-C6H4-G≡CH as ligand failed.  相似文献   

19.
Copolymerizations of ethylene with 1‐decene have been carried out by using two syndiospecific metallocenes synthesized by modifying the bridge: highly syndiospecific isopropylidene(1‐η5‐cyclopentadienyl)(1‐η5‐fluorenyl)‐dimethylzirconium (Me2C(Cp)(Flu)ZrMe2, 1 ) and less syndiospecific (1‐fluorenyl‐2‐cyclopentadienylethane)‐dimethylzirconium (Et(Cp)(Flu)ZrMe2, 2 ), in the presence of [Ph3C][B(C6F5)4] as a cocatalyst. The ethano bridged 2 compound of smaller dihedral angle showed much higher activity than 1 compound. The catalytic activities of the two compounds were enhanced about twice when a suitable amount of 1‐decene comonomer is present in the feed. The compound 1 showed better comonomer reactivity than 2 . The properties (Tm, density, and crystallinity) of copolymers seem not to be affected by the type of bridge of the metallocenes, and mainly depend on 1‐decene content in the copolymer.  相似文献   

20.
Alkinyl functionalized titanocenes of general type RC≡C-[Ti]-C≡CR and Cl-[Ti]-C≡CR {[Ti] = (η5-C5H4-SiMe3)2Ti}; R = singly bonded organic ligand} can successfully be used as organometallic π-tweezers to stabilize numerous mononuclear MX/R1 species (M = Cu, Ag, Au; X = singly bonded inorganic group; R1 = singly bonded organic ligand). The synthesis, manifold reaction chemistry as well as bonding and spectroscopy of {[Ti](C≡CR)2}MX/R1 and {[Ti] (C≡CR)(Cl)}CuX complexes is described  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