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.
Summary Eleven compounds have been prepared by azeotropic destillation of water from toluene solutions of bis(tri-n-butyltin)oxide and N-acetyl amino acids. All derivatives are white solids.119Sn-NMR-spectra of the tri-n-butyltin compounds have been studied in coordinating and non coordinating solvents. The chemical shifts and the coupling constants1 J(119Sn,13C) depend significantly on the coordination number of the tin atom and on the properties of the substituents. The data for the compounds are discussed in comparison with those for other tri-n-butyltin compounds.
  相似文献   

2.
19F relaxation times of benzyl fluoride in acetone-d6 and in methanol-d4 were measured and extrapolated at infinitely dilute solution. The fluorine relaxes through intramolecular dipole-dipole (DDa) and spin-rotation (SR) mechanisms in acetone-d6 ,and through DDa,SR and inter-molecular dipole-dipole mechanisms in methanol-d4. The DDa contribution can be recalculated from the overall and internal reorientational motions through 2D measurements on the same solutions.The separation of the different contributions are consistent with those made on the proton relaxation times and correlates more closely with poor solvation of benzylfluoride in acetone-d6 and with greater solvation in methanol-d4.  相似文献   

3.
The 1:1 and 2:1 formulations of the free radical initiated copolymers of methyl methacrylate (MMA) and tri-n-butyltin methacrylate (TBTM), and the homopolymer, poly(TBTM), are characterized by 13C- and 119C-NMR structural analyses were performed on the tributyltin-free hydrolyzate, a copolymer of MMA and methacrcylic acid (MAA). Configurational sequencing at the triad level is performed using the α-methyl region of the 13C-NMR spectrum. The probability of isotactic (meso) dyad placement at 80°C in the homopolymer (0.19) is determined to be significantly less than the probabilities observed for the copolymers (0.23–0.24). Random compositional sequencing is established for the copolymers through a comparison of the carbonyl regions of the 13C-NMR spectra of the hydrolyzates with the carbonyl regions in published spectra of structurally characterized copolymers of MMA and MAA. The 119Sn chemical shift and the tin-carbon J coupling for the polymers are dependent on the solvent employed. This dependence is attributed to electron donor or acceptor interactions between the solvent and the strong Sn? O dipole. The tin-containing copolymers exhibit multiple 119Sn resonances, which appear related to compositional sequencing.  相似文献   

4.
Some five- and six-coordinated di- and tri-n-butyl tin(IV) complexes of the type Bu2SnL, Bu2SnL2 and Bu3SnL (where L is the anion of a monofunctional bidentate or bifunctional tridentate Schiff base) have been synthesized and characterised on the basis of microanalyses, molecular weight determinations, IR, NMR (1H, 13C, 119Sn) and 119Sn Mössbauer spectroscopy. These complexes are highly active towards bacteria.  相似文献   

5.
6.
The di- and trialkyltin(IV) complexes of composition R2SnCl2−x (OAr), and n-Bu3Sn(OAr) (R = n-Bu and Me; x = 1 and 2; OAr = OC6H3Bu t -2-Me-4) have been synthesized by the reactions of di-n-butyl and dimethyltin dichlorides and tri-n-butyltin(IV) chloride with 2-tert-butyl-4-methylphenol and triethylamine in tetrahydrofuran. The reaction of triphenyltin chloride with trimethylsilyl-2-t-butyl-4-methylphenoxide in the same solvent however, gives a complex of composition Ph3Sn(OAr). The complexes have been characterized by microanalyses, molar conductance measurements, molecular weight determinations and IR and 1H, 13C and 119Sn NMR and mass spectral studies. Thermal behaviour of the complexes has been studied by TGA and DTA techniques. From the non-isothermal TG data, the kinetic and thermodynamic parameters have been calculated employing Coats-Redfern equation and the mechanism of decomposition has been computed using non-isothermal kinetic method. Thermal investigations on the blends of poly(methylmethacrylate). PMMA, with organotin(IV)-2-tert-butyl-4-methylphenoxides have shown increased thermal stability compared to pure PMMA suggesting thereby their potential as additives towards PMMA.  相似文献   

7.
Some five- and six-coordinated di and tri-n-butyl tin(IV) semi- and thio-semi carbazates have been synthesized. The characterization of these complexes, by IR, NMR (1H, 13C, 119Sn), 119Sn), 119Sn Mössbauer and Mass spectroscopies along with X-ray diffraction, reveals that complexes of biionic ligands of the type Bu2Sn L″ are five-coordinated having trigonal bipyramidal geometry. However, complexes of monoionic ligands of the type Bu2SnL′2 are six-coordinated in a distorted cis-octahedral geometry and Bu3SnL′ are five-coordinated with a trigonal bipyramidal structure. X-ray structural studies on the compound Bu2Sn(O.C6H4.CH:N.N.CS.NH2), show that it crystallizes in a monoclinic lattice with a = 16.90 Å, b = 9.71 Å, c = 8.60 Å, and β = 103°45′.  相似文献   

8.
Tin isotopes were fractionated by the liquid-liquid extraction technique with a crown ether, dicyclohexano-18-crown-6. The isotopic ratios of mSn/120Sn (m: 116, 117, 118, 119, 122 and 124) were measured by multi-collector inductively coupled plasma spectrometry (MC-ICP-MS) on a Nu Plasma 500 with a precision better than 0.05 permil amu−1 on each isotopic ratio. Odd atomic mass isotopes (117Sn and 119Sn) showed depletions compared to the even atomic mass isotopes (116Sn, 118Sn, 122Sn and 124Sn). We show that this odd-even staggering property originates from the nuclear field shift effect. The contribution of the nuclear field shift effect to the observed isotope enrichment factor was estimated to be ∼35%.  相似文献   

9.
Silicon-29, carbon-13 and oxygen-17 NMR data are reported for the methylethoxysilanes, MenSi(OEt)4?n with n = C to 3. The values of 1J(SiC) vary greatly, in a manner which appears to be inconsistent with an effect of s-character variation alone. The chemical shifts are discussed in terms of possible π-bonding. Silicon-29 and carbon-13 spin-lattice relaxation times and nuclear Overhauser effects are also reported and discussed.  相似文献   

10.
The 13C and 119Sn NMR spectra of 33 organotin compounds of the type RSnMenCl3 ? n and related types are discussed. The substituent effects of the groups SnMe3, SnMe2Cl, SnMeCl2 and SnCl3 (and of some related groups) on the carbon chemical shifts in the alkyl group R have been determined; the SnMe3 group causes a small upfield shift of the carbon attached to it, while the other groups cause downfield shifts. The shifts show a monotonic change on replacing methyl groups in Me3Sn by chlorine atoms. The effects on carbons further removed from the tin atom are discussed. Variation in R causes little change in nJ(Sn? C) or δ(119Sn).  相似文献   

11.
The small negative magnetogyric ratio (γ) of the 15N nucleus decreases the efficiency of 15N? 1H dipole-dipole relaxation to about 25% of that for an analogous 13C nucleus. This may lead to greater competition from other relaxation mechanisms in 15N n.m.r. and consequent partial or total quenching of the negative nuclear Overhauser effect (NOE). In unfavorable circumstances nulling of the 15N resonance can occur. Previous 15N relaxation studies have examined isotopically enriched, low molecular weight compounds. The present study examines several small to intermediate size organic compounds containing nitrogen at natural isotopic abundance. In contrast to some of the earlier studies, 15N? 1H dipolar relaxation was found to be dominant for protonated nitrogen atoms, even for two tertiary nitrogens (the tertiary amine nitrogen in 1,2,3,4,6,7,12,12b-octahydroindolo[2,3-a] quinolizine and the oxime nitrogen in 3-methyl-2-pentanone ketoxime). The magnitude of the NOE and the moderate value of T1 indicate effective dipolar relaxation from neighboring but not directly bonded protons in these cases. Nitro groups were found, as expected, to have predominant contributions from non-dipolar mechanisms, and in one case (2-methyl-2-nitro-1, 3-propanediol) signal nulling (NOE of η = ?1) was observed. The effect of paramagnetic impurities was demonstrated for ethanolamine, which contains a basic nitrogen. In this case T1DD(15N? 1H) = 4·3 s; added Ni(acac)2 at 1 × 10?4 Molar reduced the 15N T1 to 0·065 s and consequently the NOE to η = 0.  相似文献   

12.
Abstract  Cholesterol complexes with tri-n-butyl phosphate, tri-n-octylamine, N,N-dimethylacetamide, and cyclohexanone in benzene and toluene solutions were studied using conventional IR spectroscopy. The spectra were recorded in the region of fundamental OH stretching (3,700–3,100 cm−1) at 298 K. The experimental spectra were resolved into bands corresponding to the cholesterol monomer and particular oligomeric and complex species. The formation constants of complexes were determined from the-least squares plots of the linearized expressions of Bjerrum’s formation function. The stoichiometry of complexes was also identified in this way. The identification of the particular resolved bands was performed from their location, and from the dependence of their intensity on the cholesterol monomer and free base concentration. Graphical Abstract     相似文献   

13.
Satellites corresponding to metal-proton coupling constants through two and four bonds are observed in PMR spectra of Pb, Sn and Hg allenic derivatives. The relative signs of these coupling constants are deduced from analysis of the satellite spectra: 2J(X? H) and 4J(X? H) are of opposite signs for X = 207Pb, 119Sn, 117Sn and of same sign for X = 199Hg. Probable absolute signs of reduced coupling constants are discussed in relation to published data: 2K(X? C? H) is probably positive for X = 207Pb, 119Sn, 117Sn and 199Hg. 4K(X? C?C?C? H) is probably negative for X = 207Pb, 119Sn, 117Sn and positive for X = 199Hg.  相似文献   

14.
Rate constants for the tri-n-butyltin radical ( Sn · ) induced decomposition of a number of peroxides have been measured in benzene at 10°C. The values range from ~100 M?1 sec?1 for di-t-butyl peroxide to 2.6 × 107 M?1 sec?1 for di-t-butyl diperoxyisophthalate. The majority of the peroxides, including diethyl peroxide, diacetyl peroxide, and t-butyl peracetate, have rate constants of ~105 M?1 sec?1. It is shown that di-n-alkyl disulfides are ten times as reactive toward Sn · as di-n-alkyl peroxides, although the exothermicities of these reactions are ~15 and ~39 kcal/mole, respectively. The enhanced reactivity of the disulfides is attributed to the easier formation of an intermediate or transition state with 9 electrons around sulfur, compared with an analogous species with 9 electrons around oxygen. The following bond strengths (kcal/mole) have been estimated: D[ Sn ? OR] = 77; D[ Sn ? H] = 82; D[ Sn ? SR] = 83; and D[ Sn ? OC(O)R] = 86, where R = alkyl. Rate constants for reaction of Sn · with some benzyl esters have also been measured. It has been found that t-butoxy radicals can add to benzene and abstract hydrogen from benzene at ambient temperatures.  相似文献   

15.
The new stannide Li2AuSn2 was prepared by reaction of the elements in a sealed tantalum tube in a resistance furnace at 970 K followed by annealing at 720 K for five days. Li2AuSn2 was investigated by X‐ray diffraction on powders and single crystals and the structure was refined from single‐crystal data: Z=4, I41/amd, a=455.60(7), c=1957.4(4) pm, wR2=0.0681, 278 F2 values, 10 parameters. The gold atoms display a slightly distorted tetrahedral tin coordination with Au? Sn distances of 273 pm. These tetrahedra are condensed through common corners leading to the formation of two‐dimensional AuSn4/2 layers. The latter are connected in the third dimension through Sn? Sn bonds (296 pm). The lithium atoms fill distorted hexagonal channels formed by the three‐dimensional [AuSn2] network. Modestly small 7Li Knight shifts are measured by solid‐state NMR spectroscopy that are consistent with a nearly complete state of lithium ionization. The noncubic local symmetry at the tin site is reflected by a nuclear electric quadrupolar splitting in the 119Sn Mössbauer spectra and a small chemical shift anisotropy evident from 119Sn solid‐state NMR spectroscopy. Variable‐temperature static 7Li solid‐state NMR spectra reveal motional narrowing effects at temperatures above 200 K, revealing lithium atomic mobility on the kHz time scale. Detailed lineshape as well as temperature‐dependent spin lattice relaxation time measurements indicate an activation energy of lithium motion of 27 kJ mol?1.  相似文献   

16.
Complexes [Me2SnL2 ( I ), Me3SnL ( II ), Et2SnL2 ( III ), n‐Bu2SnL2 ( IV ), n‐Bu3SnL ( V ), n‐Oct2SnL2 ( VI )], where L is (E)‐3‐furanyl‐2‐phenyl‐2‐propenoate, have been synthesized and structurally characterized by vibrational and NMR (1H, 13C and 119Sn) spectroscopic techniques in combination with mass spectrometric and elemental analyses. The IR data indicate that in both the di‐ and triorganotin(IV) carboxylates the ligand moiety COO acts as a bidentate group in the solid state. The 119Sn NMR spectroscopic data, 1J[119Sn,13C] and 2J[119Sn, 1H], coupling constants show a four‐coordinated environment around the tin atom in triorganotin(IV) and five‐coordinated in diorganotin(IV) carboxylates in noncoordinating solvents. The complexes have been screened against bacteria, fungi, and brine‐shrimp larvae to assess their biological activity. © 2008 Wiley Periodicals, Inc. Heteroatom Chem 19:612–620, 2008; Published online in Wiley InterScience ( www.interscience.wiley.com ). DOI 10.1002/hc.20488  相似文献   

17.
Recently we presented the dynamics of 13CO2 molecules sorbed in silicone rubber (PDMS) ascertained from spin relaxation experiments. Results of a similar investigation for 13CO2 sorbed in polyisobutene (PIB) are presented in this report. The spin-lattice and spin-spin relaxation times as well as nuclear Overhauser enhancements (NOE) were determined as a function of temperature and Larmor frequency. The relaxation mechanisms found to be important for 13CO2/PIB system are intermolecular dipole-dipole relaxation and chemical shift anisotropy with a minor contribution from spin rotation relaxation. We have determined the parameters which characterize correlation times for 13CO2 collisional motion, rotational motion, and translational motions in the PIB. The self-diffusion coefficient of 5.15 × 10?8 cm2/s obtained from the nuclear magnetic resonance (NMR) data is close to the literature value of the mutual diffusion coefficient of CO2 in PIB at 300 K obtained from permeability measurements. In contrast to the case of CO2/PDMS in which a broad distribution (characterized by a fractional exponential correlation function of the Williams-Watts type with α = 0.58) is observed, a sharp distribution with a fractional exponent, α, of 0.99 is found for the CO2/PIB system. Instead of assuming an Arrhenius type temperature dependence, we used a Williams-Landel-Ferry type temperature dependence and found it to be better suited to describe the behavior of this system. PIB is a densely packed “strong” chain polymer which responds gradually to the temperature variation and gas sorption. In contrast PDMS is a relatively loosely packed “fragile” polymer with a propensity to exhibit rapid dynamic responses to the temperature change and gas sorption. © 1993 John Wiley & Sons, Inc.  相似文献   

18.
Zwitterionic amphiphiles of the general formula H(CH2)y + N(CH3)2(CH2)n PO2C6H 5 , where the number of intercharge methylenesn is varied, were studied in dilute aqueous solution. Their critical micellar concentrations show a peculiar variation withn, first increasing asn varies from 1 to 4 and then slowly decreasing as methylenes are added up to 10. This behavior is interpreted as being the consequence of two opposite contributions. The first is the classical CMC lowering due to the increase of hydrophobic character with the total number of methylene groups in the surfactant molecule. The second contribution is the increase in the dipole moment of the zwitterionic headgroup withn, leading to stronger dipole-dipole repulsions between headgroups at the micellar surface. Experimental results suggest that the dipole moment does not increase linearly withn because of the polymethylene chain flexibility. This is supported by13C NMR relaxation experiments.  相似文献   

19.
Chemical shifts δ(13C), δ(119Sn) and coupling constants J(119Sn13C) for alkynylstannanes of the type R4-nSn(CCR′)n (n = 1–4) are reported. The values of 1J(119Sn13C) and 2J(119SnC13C) depend upon the nature of the substituent R′. 1J(119Sn13C) in Sn(CCCH3)4 is 1168 Hz, much larger than a value predicted in the literature of ca. 700 Hz. The comparison of δ(119Sn) for (CH3)2Sn(CCR′)2 and 1,1,4,4-tetramethyl-1-stannacyclohexadi-2,5-ene suggests that the δ(119Sn) of alkynylstannanes are determined only to a small extent by the diamagnetic anisotropic effect of the CC-triple bond.  相似文献   

20.
A number of alkyltin(IV) paratoluenesulfonates, RnSn(OSO2C6H4CH3‐4)4?n (n = 2, 3; R = C2H5, n‐C3H7, n‐C4H9), have been prepared and IR spectra and solution NMR (1H, 13C, 119Sn) are reported for these compounds, including (n‐C4H9)2Sn(OSO2X)2 (X = CH3 and CF3), the NMR spectra of which have not been reported previously. From the chemical shift δ(119Sn) and the coupling constants 1J(13C, 119Sn) and 2J(1H, 119Sn), the coordination of the tin atom and the geometry of its coordination sphere in solutions of these compounds is suggested. IR spectra of the compounds are very similar to that observed for the paratoluenesulfonate anion in its sodium salt. The studies indicate that diorganotin(IV) paratoluenesulfonates, and the previously reported compounds (n‐C4H9)2Sn(OSO2X)2 (X = CH3 and CF3), contain bridging SO3X groups that yield polymeric structures with hexacoordination around tin and contain non‐linear C? Sn? C bonds. In triorganotin(IV) sulfonates, pentacoordination for tin with a planar SnC3 skeleton and bidentate bridging paratoluenesulfonate anionic groups are suggested by IR and NMR spectral studies. The X‐ray structure shows [(n‐C4H9)2Sn(OSO2C6H4CH3‐4)2·2H2O] to be monomeric containing six‐coordinate tin and crystallizes from methanol–chloroform in monoclinic space group C2/c. The Sn? O (paratoluenesulfonate) bond distance (2.26(2) Å) is indicative of a relatively high degree of ionic character in the metal–anion bonds. Copyright © 2003 John Wiley & Sons, Ltd.  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