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.
《Inorganica chimica acta》1988,143(2):161-167
Complexes formed from the reaction of palladium(II) and platinum(II) halides with (p-EtO·C6H4)Te(CH2)nTe(C6H4OEt-p) (Ln, n = 6, 7, 8, 9, 10) are reported together with data for some mercury(II) complexes, [HgLnCl2] which are used for comparative purposes. The compounds [MLnX2] (M  Pd, Pt; n = 7, 8 (Pt only), 9, 10; X  Cl, Br) have molecular weights in molten naphthalene which fluctuate about the monomer value. [ML6X2] (M  Pd, Pt; X  Cl, Br) are totally insoluble and are believed to be polymeric. The palladium(II) complexes have trans dichloro groups whereas the platinum compounds show cis dichloro groups in the solid state.13C NMR spectra are valuable to confirm the coordination of the ligand; the methylene resonance of the TeCH2 group undergoes a 19–20 ppm downfield shift on coordination. 125Te NMR spectra of the Pd(II) and Pt(II) complexes show two broad resonances the chemical shifts of which imply the presence of cis and trans isomers in CDCl3 solution. A more detailed variable temperature high field study of [PtL8Cl2] (125Te and 195Pt NMR) reveals a complex solution chemistry involving at least two cis and two trans species. The temperature range over which the solution is stable (−10 to 70 °C) is insufficient to allow a totally unambiguous interpretation but a model based on monomer ⇍ dimer equilibria provides a self consistent interpretation.  相似文献   

2.
《Inorganica chimica acta》1986,115(2):207-209
The reaction of [Au(CH2)2PPh2]2 with excess CHBr3 in benzene initially gives [Au(CH2)2PPh2]2− (CHBr2)Br. This observation establishes that halomethanes, CHyX4−y (y=3,2,1,0; X=Cl, Br, I), react with [Au(CH2)2PPh2]2 to initially give Au(II) adducts of the general form [Au(CH2)2PPh2]2−(CHyX3−y)X (y=3,2,1,0) via oxidative addition across the carbon-halogen bond. The order of reactivity inversely follows the order of carbon-halogen bond dissociation energies of haloalkanes. Methyl chloride is the only halomethane of the series that does not give a Au(II) adduct under similar reaction conditions.  相似文献   

3.
《Inorganica chimica acta》1986,115(2):121-128
The interaction between Cp2MoH2 (Cp=η5- C5H5) and SnMenCl4−n (n=0−3) proceeds in aprotic solvent with the elimination of HCl and the formation of heterometallic complexes of the composition Cp2Mo(H)SnMenCl3−n (n=0−3) and Cp2Mo(SnMe2Cl)2 which contains an MoSn σ-bond. It has been found that in all studied compounds the length of this bond is 0.20–0.30 Å less than the sum of the covalent radii of the Mo and Sn atoms.Based on analysis of the geometry of the Mo and Sn environment, the high values of the isomeric shifts (IS) in the Mössbauer spectra, the constants of the spin-spin interactions (SSI) J3Cp-Sn and J2HMoSn, and the considerably decreased values of the J2Me-sn constants in 1H NMR spectra, it was concluded that the decrease in the interatomic distance Mo-Sn is due to the high s-character of this bond. It is suggested that this effect, which is most pronounced in wedge-like complexes, is brought about by changing the orbital hybridization type of the tin atom from sp3 to s + 3p. This can explain the shorter interatomic distance M-Sn in heterometallic complexes of other types.  相似文献   

4.
《Inorganica chimica acta》2006,359(5):1531-1540
A series of phosphine-linked oligomers of oxo-centered triruthenium-acetate clusters have been prepared by the reaction of [Ru3O(OAc)6(py)2(CH3OH)](PF6) (1) with di- or poly-phosphine. They have been characterized by elemental analysis, ESI-MS spectrometry, UV–Vis, IR, and 31P NMR spectroscopy, and cyclic and differential pulse voltammetry. The structures of diphosphine-linked dimeric compounds 4 and 7 were determined by X-ray crystallography. As revealed by redox wave splitting, weak to moderate electronic communication is operative between triruthenium clusters across bridging di- or poly-phosphine. With increase of the methylene number in Ph2P(CH2)nPPh2, electronic communication decreases rapidly in diphosphine-linked dimeric complexes [{Ru3O(OAc)6(py)2}2{μ-Ph2P(CH2)nPPh2}]2+ (n = 1–5).  相似文献   

5.
The inhibitory effect of a phenyl group in quaternary ammonium compounds on thiamine uptake in isolated rat hepatocytes was investigated. The phenyltrimethylammonium ion was a more potent inhibitor than the tetramethylammonium ion, while the dibenzyldimethylammonium ion was the most potent inhibitor of thiamine uptake among those compounds examined. A kinetic study showed that this compound was a competitive inhibitor. The cetyltrimethylammonium ion was a less effective inhibitor than the benzyltrimethylammonium ion, and the palmitoylcholine ion was a weak inhibitor. These results indicate that the lipophilicity of a quaternary ammonium compound is not always correlated with its affinity for thiamine-carrier binding, but the presence of a phenyl group plays a significant role in affinity. The inhibitory effect of the series of (CH3)3N+(CH2)nC6H5 (n = 0−6) compounds on thiamine uptake in isolated rat hepatocytes was studied. The maximal inhibitory activity occurred at n = 5. These results suggest that the phenyl group in a quaternary ammonium compound has a specific interaction with the thiamine-binding site in rat liver plasma membrane.  相似文献   

6.
Reactions of alkanolamines [R1R2NXOH; R1 = H, CH3, C2H5; R2 = H, CH3, C2H5 and X = -CH2CH2-, -CH2CH2CH2-, -CH2CHCH3, -C6H4CH2CH2-] with aluminium isopropoxide in different molar ratios (1 to 3) yield compounds of the type Al(OPri)3?n(OXNR1R2)n, where ‘n’ can be 1, 2 and 3. Most of the derivatives are distillable liquids, soluble in common organic solvents and susceptible to hydrolysis even by atmospheric moisture. The new derivatives are characterized by elemental analysis, IR and 1H NMR spectra. Molecular weight measurements of Al(OPri)3?n(OXNR1R2)n reveal them to be tetrameric in nature.  相似文献   

7.
A study of the reactions of Mo(CO)5(P(OCH2- CMe2CH20)X) (X = C1, Br) with a variety of nucleophiles of the type HER (E = NH, O, S; R = H, alkyl, aryl) is reported. The mechanism of these reactions is shown to be SN2, and the significantly slower rates of reactions of n-propylamine with the above complexes relative to the rate of reaction of n-propylamine with Mo(CO)5(Ph2PC1) is discussed.The 13C, 17O, 31P and 95Mo NMR data and infrared data for these complexes are presented. Good correlations between chemical shifts of the cis and trans CO13C and trans CO170 NMR resonances, CO infrared stretching force constants and the magnitudes of 1JMop and 2JPC (trans CO) are observed and the reasons for these correlations are discussed.The correlations between the chemical shifts of NMR-active nuclei in the Mo(CO)5(P(OCH2CMe2- CH2O)ER) complexes with the chemical shifts of similar nuclei in the Mo(CO)5(Ph2PER) complexes fange from excellent to very poor. This indicates that the effects of-the P-substituents on the chemical shifts of the NMR-active nuclei in these complexes are not additive.  相似文献   

8.
《Inorganica chimica acta》2006,359(9):2721-2727
Salts of the new complexes [PW11O39{Rh2(O2CR)2}]5−; R = Prn (1), CH2Cl (2), CH2OH (3), o-C6H4OH (4), p-C6H4OH (5), and [XW11O39{Rh2(p-O2CC6H4OH)2}]6−; X = Si (6), Ge (7) have been prepared in good yield and characterized by elemental analysis, multinuclear NMR spectroscopy, and structural crystallography of the cesium salts of anions 1 and 5 with chloride ions in axial positions of the dirhodium moiety. The incorporation of the dirhodium group into the Keggin structure significantly increases the hydrolytic kinetic stability of the polyoxotungstates at pH 7–8.5, a result that has implications for the use of such complexes for imaging and for phase determination in structural studies of large biomolecules. Based on NMR studies, the axial sites of the dirhodium moiety of the polytungstate anions are accessible to ligation by molecules such as cysteine, methionine, and isonicotinic acid.  相似文献   

9.
《Inorganica chimica acta》1987,134(2):275-278
Silicon-29 (δ 29Si) NMR chemical shifts are reported for the first time for a number of tris(phenyldimethylsilyl)methyl silicon compounds (disilylated derivatives), (PhMe2SiA)3CαL, where LSiBR1R2R3, where R varies widely in electronegativity. δ 29SiB for these series exhibited to some extent good correlations with the electronegativities of the groups bonded to silicon (particularly, δ 29SiBMe2X, X H, OH, OCH3, COCH3 and Cl). Substitution with electronegative atoms shifts the chemical shift of silicon to high field.The 13C NMR spectra of these compounds have been recorded and assigned. The chemical shifts of the α-carbon (Cα) resonances are shown to depend on the type of substituent on the silicon-B, thus 13Cα exhibit downfield shifts when X=oxygen ligand. The 13C phenyl resonances have been measured and show the same order of o, m and p signals, viz. δ ortho (downfield)>δ parameta.The variation of 29Si-1H coupling constants with the electronegativity of X was studied.  相似文献   

10.
To assign the observed vibrationsl modes in the resonance Raman spectrum of the retinylidene chromophore of rhodopsin, we have studied chemically modified retinals. The series of analogs investigated are the n-butyl retinals substituted at C9 and C13. The results obtained for the 11-cis isomer have clearly assigned the CCH3 vibrational frequencies observed in the spectrum of the retinylidene chromophore. The data show that the C(9)CH3 stretching vibration can be assigned to the vibrational mode observed in the 1017 cm?1 region, and the vibration detected at 997 cm?1 can be assigned to the C(13CH3 vibration. The C(5)CH3 stretching mode does not contribute to the vibrations observed in this region. The splitting in the C(n)CH3 (n = 9, 13) vibration is characteristic of the 11-cis conformation. The results on the modified retinals do not support the hypothesis that the splitting arises from equilibrium mixtures of 11-cis, 12-s-cis and 11-cis, 12-s-trans in solution. Thus, this splitting cannot be used to determine whether the chromophore in rhodopsin is in a 12-s-cis or 12-s-trans conformation. However, our results demonstrate that there are other vibrational modes in the spectra which are sensitive to this conformational equilibrium and we use the presence of a strong ~ 1271 cm?1 mode in bovine and squid rhodopsin spectra as an indication that the chromophore in these pigments is 11-cis, 12-s-trans.  相似文献   

11.
The nonenzymatic reaction of ethanol-derived CH3CHO with tissue constituents continues to be of interest as a potential mechanism underlying the toxicity of alcohol. The current study has focused on the spontaneous condensation of CH3CHO with H4folate under physiological conditions (38 °C, pH 7.0, I = 0.25 M). Computer analysis of uv spectral changes with increasing CH3CHO concentrations demonstrated the presence of at least two different adducts. The observed equilibrium constant (Kobs) for the formation of the first adduct is 91 ± 2 m?1 (121 ± 2 m?1 at 25 °C), a value which is unaffected by variations in ionic strength (0.06–1.0 m) or by free [Mg2+] up to 5 mm. The NMR spectrum is compatible with the structure: 5,10-CH3CH-H4folate analogous to the naturally occurring 5,10-CH2-H4folate. The formation of the latter compound from HCHO and H4folate, however, is much more favorable under the same conditions [Kobs = 3.0 ± 0.2 × 104 M?1 (38 °C), 3.6 ± 0.1 × 104 M?1 (25 °C)]. At the levels of CH3CHO which accumulate during ethanol metabolism in vivo only a small fraction of the H4folate will exist as the CH3CHO derivative, yet it may ultimately be the ratio of free CH3CHO to free HCHO in tissue which determines the physiological importance of the CH3CHO adduct. Other adduct(s) of CH3CHO with H4folate are observed at very high levels of CH3CHO but are unlikely to be of physiological significance.  相似文献   

12.
Arylpiperazines, XC6H4N(CH2CH2)2NH, are readily alkylated to give the N-alkylpiperazines of the type XC6H4N(CH2CH2)2N(CH2)nNH2. The amine functions of these derivatives are in turn easily subjected to mono- or dialkylation to provide potentially tridentate ligands of the types XC6H4N(CH2CH2)2N(CH2)nN(H)(CH2Y) and XC6H4N(CH2CH2)2N(CH2)nN(CH2Y)(CH2Z), respectively. The latter class of dialkylated derivatives may be symmetrically (Y=Z) or unsymmetrically (Y ≠ Z) substituted. The donor groups Y and Z of this study include pyridine, imidazole, methyl-imidazole, thiazole, carboxylate and thiolate.The reactions of these ligands with [NEt4]2[Re(CO)3Br3] yield complexes of the type [Re(CO)3{(YCH2)N(H)(CH2)n(H)xN(CH2CH2)2N(H)yC6H4X}]n and [Re(CO)3{(ZCH2)(YCH2)N(CH2)n(H)xN(CH2CH2)2N(H)yC6H4X}]n where the molecular charge n (0, +1, or +2) depends on the nature of the donor groups Y and Z (whether neutral or anionic or a combination of neutral and anionic) and on the degree of protonation of the piperazine unit (x=0 or 1; y=0 or 1). This variety of tridentate chelators provides complexes with fac-{Re(CO)3N3}, {Re(CO)3N2O}, {Re(CO)3NO2}, {Re(CO)3N2S} and {Re(CO)3NS2} coordination geometries. The structures of the model compound [Re(CO)3{(CH3N2C3H2CH2)N(H)CH2CH2-piperidine}]Br · H2O, [Re(CO)3{(CH3N2C3H2CH2)N(H)CH2CH2-Fphenpip}]Br, [Re(CO)3{(NC5H4CH2)N(H)CH2CH2-Fphenpip}]Br, [Re(CO)3{(O2CCH2)2NCH2CH2CH2-CH3OphenpipH}] · xCH3OH (x≈0.875), [Re(CO)3{(NC5H4CH2)2NCH2CH2CH2-CH3OphenpipH}]Br2 · 2CH2Cl2 · H2O and [Re(CO)3{(CH3N2C3H2CH2)(O2CCH2)NCH2CH2CH2-CH3OphenpipH2}]BrCl · 1.5CH3OH · H2O are discussed (phenpip: phenylpiperazine, -C6H4N(CH2CH2)2N-).  相似文献   

13.
The reaction of 8-thioguanosine (8-thioGuoH2 with methylmercury(II) has been shown to give rise to 1:1 (neutral and cationic), 1:2 (neutral and cationic), and 1:3 (cationic) complexes of the type [CH3Hg(8-thioGuoH)], [(CH3Hg(8-thioGuoH2)]NO3, [(CH3Hg)2(8-thioGuo)], [(CH3Hg)2(8-thioGuoH)]NO3 and [(CH3Hg)3(8-thioGuo)]NO3, depending upon the reactant stoichiometry and pH. 1H NMR, 13C NMR, and IR, as well as analytical data were used to characterize the complexes. Coupling of methylmercury(II)-protons to mercury-199 has been observed in all compounds. The magnitude of the coupling, 2J(1H-199Hg), is strongly dependent on the nature of the ligand bonded to the methylmercury(II) group and correlates with the 13C chemical shifts of coordinated CH3Hg(II) at the different binding sites.  相似文献   

14.
《Inorganica chimica acta》1986,111(2):163-166
Tri- and di-organosilicon O,O-alkylenedithiophosphates, R4−nSi[S2PO2G]n (where R = Ph, Me, G = −C(CH3)2·C(CH3)2−, −CH2C(CH3)2CH2−, −CH-CH3CH2C(CH3)2−, n = 1,2) were synthesized by treatment of organosilicon(IV) chlorides with ammonium O,O-alkylenedithiophosphates in benzene. The compounds are volatile, yellow oily liquids, miscible with common organic solvents and monomeric in refluxing benzene. Like dialkyldithiophosphate derivatives of organosilicon(IV), these cyclic chain derivatives appear to be tetrahedral, the ligand behaving as unidentate.  相似文献   

15.
The gem-dialkyl effect has been investigated in the reactions of cyclotriphosphazene, N3P3Cl61, with various 2,2′-derivatives of 1,3-propandiol, CXY(CH2OH)2, in either THF or DCM to form spiro (6-membered) and ansa (8-membered ring) derivatives. The reactions were made with a number of symmetrically-substituted (X = Y, methyl, ethyl, n-butyl and a malonate ester) and unsymmetrically-substituted (X ≠ Y, methyl/H, phenyl/H, methyl/n-propyl, ethyl/n-butyl and Br/NO2) 1,3-propandiols. The products were analysed by 1H and 31P NMR spectroscopy and some of the spiro and ansa derivatives were also characterized by X-ray crystallography. Reactions of 1 with unsymmetrically-substituted 1,3-propandiols results in the formation of two structural isomers of ansa-substituted compounds, both isomers (endo and exo) have been structurally-characterized by X-ray crystallography for the ethyl/n-butyl derivative. It is found that the regioselectivity of the reaction is changed when the base is changed. The relative proportions of spiro and ansa compounds formed under different reaction conditions were quantified by 31P NMR measurements of the reaction mixtures. The results were rationalised mainly in terms of the electronic effect of the substituents, whereas the steric effect has a secondary role in the formation of both spiro and ansa compounds.  相似文献   

16.
The reaction of the ruthenium complexes RuCl2(PPh3)3, RuCl2(PPh3)4, RuCl2(PMe3)4, RuCl2(Me2SO)4, or RuBr2(PPh3)3 with the tripod tetrakis(tertiary) phosphine P(CH2CH2CH2PMe2)3 gave the compounds cis-RuCl2 [P(CH2CH2CH2PMe2)3] (1) and cis-RuBr2[P(CH2CH2CH2PMe2)3] (2). The coordination geometry of 1 and 2 was derived from the ABX2 type 31P NMR patterns of the complexes, as well as from an X-ray structure determination for the chloride 1. Crystals of 1 were found to be monoclinic, space group P21/n (Z = 4), with a = 942.0(3), b = 1446.2(4), c = 1680(1) pm, and β = 104.99(4)°. Anisotropic refinement of the structure converged at R = 0.040 and Rw = 0.034 (3318 data). Selected bond lengths are (in pm): RuP(CH2−)Me2 (trans-atom P), 235.8(1) and 239.3(1); RuP(CH2−)Me2 (trans-atom Cl), 227.9(1); RuP(CH2−)3, 225.3(1); RuCl (trans-group P(CH2−)3), 252.1(1); and RuCl (trans-group P(CH2)Me2), 250.5(1). Reaction of 1 with LiAlH4 yielded the hydro derivatives cis-Ru(H)Cl[P(CH2CH2CH2PMe2)3] (3) and cis-RuH2[P(CH2CH2CH2PMe2)3] (4), which were characterized by IR and 1H and 31p NMR spectroscopy.  相似文献   

17.
A series of mono- and bis-amide scandium and yttrium compounds incorporating the furyl-substituted disilazide ligand, [N{SiMe2R}2] {i} (where R = 2-methylfuryl) have been synthesized. The compounds Sc{i}Cl2 (1), Sc{i}(CH2SiMe3)2 (2) and Sc{i}(OAr)2 (3) were made from suitable scandium starting materials employing either a salt metathesis protocol with Li{i} or via protonolysis of Sc-C bonds by the neutral amine H{i}. The thermally unstable bis-alkyl yttrium compound, ‘Y{i}(CH2SiMe3)2 was isolated as the bis-THF adduct (4) and the bis-aryloxide Y{i}(OAr)2 (5) was synthesized by elimination of LiOAr from Y(OAr)3. The bis-amide complex Y{i}2Cl (6) and conversion to a rare example of an yttrium benzyl compound Y{i}2(CH2Ph) (7) are described. The yttrium cation, [Y{i}2]+, was synthesized by benzyl abstraction from 7 using B(C6F5)3. Structural characterization of representative examples show variation in the coordination modes for amide ligand {i}, differing primarily in the number of furyl groups that coordinate to the metal, with examples in which zero, one or two M-Ofuryl bonds are present. Preliminary investigation in two areas of catalysis are presented.  相似文献   

18.
《Inorganica chimica acta》1988,144(1):129-134
Several new or partially described complexes of uranium(IV) of the series Cp 4−n U(BH 4) n [Cp=η 5- C 5H 5, η 5-C 5H 4CH 3 or η 5-C 5H 4Si(CH 3) 3] are reported, mainly in order to systematically compare their physical, chemical and spectroscopic properties. X-ray data of the crystal structure of Cp 3UBH 4 are also reported.  相似文献   

19.
A new high yielding synthesis of the seven-coordinate complexes [MI2(CO)3{Ph2P(CH2)nPPh2}] (M = Mo and W; n = 1–6) is described. The procedure involves reacting the complexes [MI2(CO)3(NCMe)2] in CH2Cl2 with an equimolar amount of the bidentate phosphorus ligand. The low temperature (−70 °C) 13C NMR spectra of the complexes [Wl2(CO)3{Ph2P(CH2)nPPh2}] (n = 3 and 5) indicates that the geometry is capped octahedral with a carbonyl ligand in the unique capping position.  相似文献   

20.
Monodentate and chelating phosphines with long alkyl chains, incorporating ethoxy- or chlorosilane functions for immobilizations, have been synthesized and fully characterized. The new compounds (EtO)3Si(CH2)xPPh2, Cl2Si(CH2CH2PPh2)2, and (EtO)2Si[(CH2)xPPh2]2 (x = 7, 11) could be prepared in high yields from cheap starting materials, and they have been characterized by multinuclear NMR spectroscopy and X-ray crystallography. The phosphines have been immobilized on silica in a well-defined manner, and the modified silicas have been studied by 31P and 29Si solid-state NMR of the dry materials and of the suspensions.  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