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 15 毫秒
1.
The new mixed Sb2O-donor ligands O{(CH2)2SbR2}2 (R = Ph, 1; R = Me, 2) with flexible backbones have been prepared in good yields as air-sensitive oils from reaction of NaSbR2 with 0.5 mol equivalents of O(CH2CH2Br)2 in thf solution. The As2O-donor analogues, O{(CH2)2AsR2}2 (R = Ph, 3; R = Me, 4) were obtained similarly from LiAsPh2 or NaAsMe2, respectively and O(CH2CH2Br)2, although ligand 4 appears to be considerably less stable with respect to C-O bond fission under some conditions than the other ligands. Using O(CH2CH2Cl)2 leads only to partial substitution by the SbPh2 or AsPh2 nucleophile. These ligands behave as bidentate chelating Sb2- or As2-donors in the distorted tetrahedral [M(L-L)2]BF4 (M = Cu or Ag; L-L = 1-4) on the basis of solution 1H and 63Cu NMR spectroscopic studies, mass spectrometry and microanalyses. Crystal structures of three representative examples with Cu(I) and Ag(I) confirm the distorted tetrahedral Sb4 or As4 coordination at the metal and allow comparisons of geometric parameters. The crystallographic identification of an unexpected Cu(I)-Cu(I) complex, [Cu2{Me2As(CH2)2OH}3](BF4)2, obtained as a by-product via C-O bond fission within ligand 4 is also reported. The distorted octahedral [RhCl2(L-L)2]Cl and the distorted square planar cis-[PtCl2(L-L)] (L-L = 1 or 2) are also described. The ether O atoms are not involved in coordination to the metal ion in any of the late transition metal complexes isolated.  相似文献   

2.
3.
The reaction of HgCl2 and Te(R)CH2SiMe3 [R = CH2SiMe3 (1), Ph (2)] in ethanol yielded a mononuclear complex [HgCl2{Te(R)CH2SiMe3}2] (R = Ph, 3a; R = CH2SiMe3, 3b). The recrystallization of 3a or 3b from CH2Cl2 produced a dinuclear complex [Hg2Cl2(μ-Cl)2{Te(R)CH2SiMe3}2] (R = Ph, 4a; R = CH2SiMe3, 4b). When 3a was dissolved in CH2Cl2, the solvent quickly removed, and the solid recrystallized from EtOH, a stable ionic [HgCl{Te(Ph)CH2SiMe3}3]Cl·2EtOH (5a·2EtOH) was obtained. Crystals of [HgCl2{Te(CH2SiMe)2}]·2HgCl2·CH2Cl2 (6b·2HgCl2·CH2Cl2) were obtained from the CH2Cl2 solution of 3b upon prolonged standing. The complex formation was monitored by 125Te-, and 199Hg NMR spectroscopy, and the crystal structures of the complexes were determined by single crystal X-ray crystallography.  相似文献   

4.
Reaction between Os(SnClMe2)(κ2-S2CNMe2)(CO)(PPh3)2 and either LiSnMe3 or KSnPh3 produces the distannyl complexes, Os(SnMe2SnMe3)(κ2-S2CNMe2)(CO)(PPh3)2 (1) or Os(SnMe2SnPh3)(κ2-S2CNMe2)(CO)(PPh3)2 (3), respectively. Similarly, reaction between Os(SnClMe2)Cl(CO)2(PPh3)2 (6) and KSnPh3 produces the distannyl complex, Os(SnMe2SnPh3)Cl(CO)2(PPh3)2 (7). In the 119Sn NMR spectra of these stable osmium(II) distannyl complexes both the α-Sn and β-Sn atoms show well-resolved 119Sn-119Sn and 119Sn-117Sn coupling. Each of these three distannyl complexes can be selectively functionalised at the α-Sn atom by reaction with SnCl2Me2 giving Os(SnClMeSnMe3)(κ2-S2CNMe2)(CO)(PPh3)2 (2), Os(SnClMeSnPh3)(κ2-S2CNMe2)(CO)(PPh3)2 (4), and Os(SnClMeSnPh3)Cl(CO)2(PPh3)2 (8), respectively. Treatment of compounds 3 or 7 with iodine also cleaves one α-methyl group, selectively, to give Os(SnIMeSnPh3)(κ2-S2CNMe2)(CO)(PPh3)2 (5), or Os(SnIMeSnPh3)Cl(CO)2(PPh3)2 (9). Crystal structures for complexes 3 and 7 have been determined.  相似文献   

5.
Chiral “P-N-P” ligands, (C20H12O2)PN(R)PY2 [R = CHMe2, Y = C6H5 (1), OC6H5 (2), OC6H4-4-Me (3), OC6H4-4-OMe (4) or OC6H4-4-tBu (5)] bearing the axially chiral 1,1′-binaphthyl-2,2′-dioxy moiety have been synthesised. Palladium allyl chemistry of two of these chiral ligands (1 and 2) has been investigated. The structures of isomeric η3-allyl palladium complexes, (R′ = Me or Ph; Y = C6H5 or OC6H5) have been elucidated by high field two-dimensional NMR spectroscopy. The solid state structure of [Pd(η3-1,3-Ph2-C3H3){κ2-(racemic)-(C20H12O2)PN(CHMe2)PPh2}](PF6) has been determined by X-ray crystallography. Preliminary investigations show that the diphosphazanes, 1 and 2 function as efficient auxiliary ligands for catalytic allylic alkylation but give rise to only moderate levels of enantiomeric excess.  相似文献   

6.
A series of novel arylantimony(V) triphenylgermanylpropionates with the formula (Ph3GeCHR1CHR2CO2)nSbAr(5−n) (R1=H, Ph; R2=H, CH3; n=1, 2) were synthesized and characterized by elemental analysis, IR, 1H-NMR, 13C-NMR and mass spectroscopy. The crystal structures of Ph3GeCH(Ph)CH2CO2SbPh4 and [Ph3GeCH2CH(CH3)CO2]2Sb(4-ClC6H4)3 were determined by X-ray diffraction. The in vitro antitumor activities of some selected compounds against five cancer cells are reported.  相似文献   

7.
Novel half-sandwich [C9H5(SiMe3)2]ZrCl3 (3) and sandwich [C9H5(SiMe3)2](C5Me4R)ZrCl2 (R = CH3 (1), CH2CH2NMe2 (2)) complexes were prepared and characterized. The reduction of 2 by Mg in THF lead to (η5-C9H5(SiMe3)2)[η52(C,N)-C5Me4CH2CH2N(Me)CH2]ZrH (7). The structure of 7 was proved by NMR spectroscopy data. Hydrolysis of 2 resulted in the binuclear complex ([C5Me4CH2CH2NMe2]ZrCl2)2O (6). The crystal structures of 1 and 6 were established by X-ray diffraction analysis.  相似文献   

8.
Trimethylindium reacted with phenyl- and tert-butylhydrazine by the release of methane and the formation of the corresponding dimethylindium hydrazides (1 and 2, respectively). Both products form dimers and possess four-membered In2N2 heterocycles with two exocyclic N-N bonds in their molecular cores. Interestingly, one compound (1) crystallizes with centrosymmetric molecules in which the N-N bonds are located on different sides of the In2N2 ring (C2h), while both N-N bonds are on the same side in 2 (C2v). In contrast, the reaction of tri(tert-butyl)indium with tert-butylhydrazine yielded a quite unexpected product. Partial decomposition occurred, and in a low yield the adduct of tribenzylindium with the unchanged tert-butylhydrazine was isolated. In a remarkable reaction, the trialkylindium derivative did not react with the relatively acidic hydrazine, but by the release of the corresponding alkane with the solvent toluene.  相似文献   

9.
Reaction of Cl3SiR or (EtO)3SiR with [PW11O39]7− affords the disubstituted hybrid anions [PW11O39(SiR)2O]3−. These species have been characterized by IR spectroscopy in the solid state and by multinuclear NMR (1H, 29Si, 31P and 183W) and cyclic voltammetry in solution. The hydrosilylation of [PW11O39(Si-CHCH2)2O]3− has been achieved with Et3SiH and PhSiMe2H. These are the first examples of hydrosilylation on a hybrid tungstophosphate core. The chromogenic behaviour of hybrid species has been demonstrated in solution.  相似文献   

10.
The novel silicon-, germanium- and tin-containing imido alkyl complexes of tungsten of the type (ArN)2W(CH2EMe3)2 (; E = Si (1), Ge (2), Sn (3)) have been prepared by the reactions of (ArN)2WCl2(dme) (dme = 1,2-dimethoxyethane) with heteroelement-containing alkyllithium or Grignard reagents Me3ECH2Li (E = Si, Ge), Me3ECH2MgCl (E = Ge, Sn). The title compounds were isolated in high yields as crystalline solids and characterized by elemental analysis, IR, 1H, 13C, 29Si and 119Sn NMR spectroscopy and X-ray diffraction studies. The geometry of the W atoms in the compounds can be described as a distorted tetrahedron.  相似文献   

11.
The reactions of dimethyl-, diethyl- and dibutyltin(IV) oxides with pyridoxine (PN) in toluene/ethanol led to the formation of compounds [SnR2(PN-2H)] which were characterized by EI and FAB mass spectrometry and by IR, Raman, Mössbauer and 1H, 13C and 119Sn NMR spectroscopy. The structures of [SnEt2(PN-2H)] · CH3OH, [SnBu2(PN-2H)] and [SnEt2(PN-2H)(DMSO)] were determined by X-ray diffractometry. The first two contain dimeric [SnR2(PN-2H)]2 units in which two bridging-chelating pyridoxinate anions link the Sn atoms, while in [SnEt2(PN-2H)(DMSO)] the DMSO coordinates to the tin atom via its O atom in a similar dimeric unit.  相似文献   

12.
A straightforward method of synthesis of heteroleptic tin (II) alkoxides stabilized by one intramolecular coordination bond was developed. Addition of one equivalent of dimethylamino ethanol to diamide Sn(N(SiMe3)2)2 (5) yields alkoxy-amido derivative Sn(OCH2CH2NMe2)(N(SiMe3)2) (2). Further addition of alcohol leads to corresponding heteroleptic dialkoxides Sn(OCH2CH2NMe2)(OR) (R = Me (6), Et (7), iPr (8), tBu (9), Ph (10)). Catalytic activity of tin (II) compounds in polyurethane formation was tested.  相似文献   

13.
The metal-metal bonds of the title compounds have been investigated with the help of energy decomposition analysis at the DFT/TZ2P level. In good agreement with experiment, computations yield Hg-Hg bond distance in [H3SiHg-HgSiH3] of 2.706 Å and Zn-Zn bond distance in [(η5-C5Me5)Zn-Zn(η5-C5Me5)] of 2.281 Å. The Cd-Cd bond distances are longer than the Hg-Hg bond distances. Bond dissociation energies (-BDE) for Zn-Zn bonds in zincocene −70.6 kcal/mol in [(η5-C5H5)2Zn2] and −70.3 kcal/mol in [(η5-C5Me5)2Zn2] are greater amongst the compounds under study. In addition, [(η5-C5H5)2M2] is found to have a binding energy slightly larger than those in [(η5-C5Me5)2M2]. The trend of the M-M bond dissociation energy for the substituents R shows for metals the order GeH3 < SiH3 < CH3 < C5Me5 < C5H5. Electrostatic forces between the metals are always attractive and they are strong (−75.8 to −110.5 kcal/mol). The results demonstrate clearly that the atomic partial charges cannot be taken as a measure of the electrostatic interactions between the atoms. The orbital interaction (covalent bonding) ΔEorb is always smaller than the electrostatic attraction ΔEelstat. The M-M bonding in [RM-M-R] (R = CH3, SiH3, GeH3, C5H5, C5Me5; M = Zn, Cd, Hg) has more than half ionic character (56-64%). The values of Pauli repulsions, ΔEPauli, electrostatic interactions, ΔEelstat, and orbital interactions, ΔEelstat are larger for mercury compounds as compared to zinc and cadmium.  相似文献   

14.
Mismatched molecular 1:1 complexes of C10F8 with catenated chalcogen-nitrogen compounds C6H5-X-NSN-SiMe3 (X = S, Se) were prepared and characterized by X-ray crystallography. The complexes provide examples of structurally non-rigid polyheteroatom molecules involved in non-covalent arene-polyfluoroarene π-stacking interactions. In going from homocrystals to the co-crystals, the molecular Z, E configuration of the catenated compounds changes from noticeably non-planar to perfectly planar, i.e. C10F8 acts as “molecular iron”. On the other hand, C10H8 does not produce complexes with C6F5-X-NSN-SiMe3 (X = S, Se).  相似文献   

15.
Nine organotin esters, Me2SnL21, Me3SnL 2, n-Bu2SnL23, n-Bu3SnL 4, Ph3SnL 5, (PhCH2)2SnL26, [(Me2SnL)2O]27, Et2SnL28 and n-Oct2SnL29, of (E)-3-(3-fluorophenyl)-2-(4-chlorophenyl)-2-propenoic acid, HL have been synthesized and characterized by elemental analysis, IR, Multinuclear NMR (1H, 13C and 119Sn) and mass spectrometry. The geometry around the tin atom has been deduced and compared both in solution and solid states. The crystal structure of compound 5 has been determined by X-ray single crystal analysis, which shows a tetrahedral geometry around the tin atom with space group . These compounds have also been screened for bactericidal, fungicidal activities and cytotoxicity data.  相似文献   

16.
A series of three new trithioether compounds containing fluorinated phenyl moieties, 1,3,5-(CH2SRf)3-2,4,6-(CH3)3C6, Rf = C6F5 (1), 4-HC6F4 (2), or 2-FC6H4 (3), were prepared by treatment of 1,3,5-(CH2Br)3-2,4,6-(CH3)3C6 with the corresponding Pb(SC6F5)2 or NaSRf. The new structures were verified by elemental analyses, IR, 1H NMR, 19F NMR spectroscopies, and mass spectra. The single crystal X-ray diffraction studies of 1-3 show a similar cis,trans,trans-conformation for the three fluorophenylthiomethyl groups attached to the central benzene ring with all dihedral angles between planes of central ring and external rings close to 0°, giving flat molecules. Comparing, 1-3 with closely related tripodal molecules built-up on 2,4,6-trimethylbenzene, arrangement of one SR group respect to others seems to be defined by the nature of the R substituent. Then in the case of 1-3, a parallel arrangement of rings is favored over an orthogonal one, which would bring the ortho-F atoms close to H atoms of the methylene groups.  相似文献   

17.
Heat capacity measurements on the adducts of thiourea with cyclohexane, cycloheptane, cyclo-octane, and 2,2-dimethylbutane have been made over the temperature range 12 to 300 K by means of an adiabatic calorimeter. All the samples showed several regions of anomalously high heat capacity. Some of the smaller anomalies are suggestive of displacive ferroelectric transitions. The curve for the cyclo-octane adduct showed a large hump with its maximum at 240 K and having an associated molar enthalpy of c-C6H16 of 7186 J mol−1 and entropy of 31.86 J K−1 mol−1, which is here ascribed to conformational changes of the cyclo-octane; a smaller and more gradual hump in the curve for the cycloheptane adduct may be due to a similar cause.  相似文献   

18.
The facile reaction of [CpCr(CO)3]2 (Cp = η5-C5H5) (1) with one mole equivalent of 2,2′-dithiodipyridine ((C5H4NS)2(SPy)2) at ambient temperature led to the isolation of dark brown crystalline solids of CpCr(CO)22-SPy) (2) in ca. 72% yield. 2 undergoes quantitative conversion to CpCrCl21-SPyH) (3) with HCl. The reaction 1 with one mole equivalent of 2-mercaptopyrimidine (C4H3N2SHHSPym) at ambient temperature led to the isolation of reddish-brown crystalline solids of CpCr(CO)22-SPym) (4) and green solids of CpCr(CO)3H (5) in yields of ca. 42% and 46%, respectively. Reaction of 4 with HCl and subsequent workup in acetonitrile resulted in the cleavage of the thiolate ligand, giving the 15-electron chromium(III) species CpCrCl2(CH3CN) (6) and free 2-mercaptopyrimidine. The complexes 2-4 have been determined by single X-ray diffraction analysis.  相似文献   

19.
A series of mononuclear ruthenium complexes [RuCl(CO)(PMe3)3(CHCH-C6H4-R-p)] (R = H (2a), CH3 (2b), OCH3 (2c), NO2 (2d), NH2 (2e), NMe2 (2f)) has been prepared. The respective products have been characterized by elemental analyses, NMR spectrometry, and UV-Vis spectrophotometry. The structures of complexes 2c and 2d have been established by X-ray crystallography. Electrochemical studies have revealed that electron-releasing substituents facilitate monometallic ruthenium complex oxidation, and the substituent parameter values (σ) show a strong linear correlation with the anodic half-wave or oxidation peak potentials of the complexes.  相似文献   

20.
Complexes M(CCCSiMe3)(CO)2Tp′ (Tp′ = Tp [HB(pz)3], M = Mo 2, W 4; Tp′ = Tp [HB(dmpz)3], M = Mo 3) are obtained from M(CCCSiMe3)(O2CCF3)(CO)2(tmeda) (1) and K[Tp′].Reactions of 2 or 4 with AuCl(PPh3)/K2CO3 in MeOH afforded M{CCCAu(PPh3)}(CO)2Tp′ (M = Mo 5, W 6) containing C3 chains linking the Group 6 metal and gold centres.In turn, the gold complexes react with Co33-CBr)(μ-dppm)(CO)7 to give the C4-bridged {Tp(OC)2M}CCCC{Co3(μ-dppm)(CO)7} (M = Mo 7, W 8), while Mo(CBr)(CO)2Tp and Co33-C(CC)2Au(PPh3)}(μ-dppm)(CO)7 give {Tp(OC)2Mo}C(CC)2C{Co3(μ-dppm)(CO)7} (9) via a phosphine-gold(I) halide elimination reaction. The C3 complexes Tp′(OC)2MCCCRu(dppe)Cp (Tp′ = Tp, M = Mo 10, W 11; Tp′ = Tp, M = Mo 12) were obtained from 2-4 and RuCl(dppe)Cp via KF-induced metalla-desilylation reactions. Reactions between Mo(CBr)(CO)2Tp and Ru{(CC)nAu(PPh3)}(dppe)Cp (n = 2, 3) afforded {Tp(OC)2Mo}C(CC)n{Ru(dppe)Cp} (n = 2 13, 3 14), containing C5 and C7 chains, respectively. Single-crystal X-ray structure determinations of 1, 2, 7, 8, 9 and 12 are reported.  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