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 93 毫秒
1.
119Sn chemical shifts, δ(119Sn), relative to Me4Sn in five- and six-coordinate organotin chelates were measured by means of FT NMR spectroscopy. 119Sn resonances were found to lie between ca. ?90 and ?330 ppm in the five-coordinate compounds and between ca. ?125 and ?515 ppm in the six-coordinate derivatives. thus δ(119Sn) moves upfield by 60–150 ppm with a change of the coordination number of tin from four to five and by 130–200 ppm from five to six. the δ(119Sn) values were shifted depending on the nature of chelating ligands and this shift was discussed in terms of the bonding between the ligand and tin. Replacement of methyl groups attached to tin by phenyl groups in five- and six-coordinate compounds induces upfield shifts in δ(119Sn) parallel to those found in four-coordinate organotin halides.  相似文献   

2.
Twenty new compounds of the form Ph3GeCHArCH2COOSnR3 (R = n-Bu, cyclohexyl; Ar = substituted phenyl) have been synthesized. Their structures were characterized by IR and 119Sn and 1H NMR spectroscopy. The compounds are five-coordinated carboxylate bridged polymers when R = n– Bu; when R = cyclohexyl (Cy) they are four-coordinate. 119Sn NMR measurements of chemical shift for the two series of compounds have shown that there is a good linear relationship for the chemical shift of 119Sn NMR between the tributyltin and tricyclohexyltin propionates, viz. δ119Sn(Bu3Sn) = 1.0474 δ 119Sn(Cy3Sn) + 95.8076, n = 5, r = 0.993. The structure of one compound was determined by X-ray diffraction. It exists as a monomeric four-coordinated species in a distorted tetrahedronal geometry.  相似文献   

3.
The 13C and 119Sn NMR spectra of some tribenzyltin(IV) compounds and their complexes in coordinating and non-coordinating solvents have been studied. The δ(119Sn) chemical shifts and coupling constants 1J(119Sn, 13C) clearly depend on the coordination number of the central tin atom and the geometry of its coordination polyhedra. Approximate ranges of the characteristic values of both the NMR parameters were determined for various configurational types of tribenzyltin compound. The 13C and 119Sn NMR parameters found are indicative of a distinct interaction between the polarized σ(SnC) bond and adjacent π-electron system of the aromatic ring(s).  相似文献   

4.
The complexes [Rh(X)(H)(SnPh3)(PPh3)(L)] (X = NCBPh3 (a), N(CN)2 (b), NCS (c), NCO (d), N3 (e); L = 1‐methylimidazole) ( 1 ) show systematic changes in δ(119Sn), δ(103Rh), J(119Sn–1H) and J(119Sn–103Rh) that are related to the electron‐donating properties of X. As X becomes more electron‐rich, δ(103Rh), J(119Sn–1H) and J(119Sn–103Rh) increase and δ119Sn) decreases. The related complexes trans‐[Rh(X)(H)(SnPh3)(PPh3)2(L)] (X = N(CN)2, NCO; L = 4‐carboxymethylpyridine (x), pyridine (y) and 4‐dimethylaminopyridine (z)) ( 2 ), show a continuation of the trends in δ(119Sn) and J(119Sn–1H), but not δ(103Rh) or J(119Sn–103Rh). Data for 1 and 2 show that within certain limits of type of ligand varied (X = N‐donor, L = a pyridine) and coordination geometry, the response of δ(119Sn) and J(119Sn–1H) to changes in electron density on rhodium is largely independent of the means by which the change is effected.Copyright © 2001 John Wiley & Sons, Ltd.  相似文献   

5.
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   

6.
Tin(IV) Complexes with Tridentate Diacidic Ligands — 119Sn NMR and 119mSn Mössbauer Studies The tin(IV) chelates of tridentate diacidic azomethines of acetylacetone resp. salicylaldehyde with benzoylhydrazine, thiobenzoylhydrazine, 2-hydroxyaniline and 2-mercaptoaniline as well as with the ligands 2-(2′-hydroxy-4-methylphenyl)-6-(2″-hydroxyphenyl)pyridine, 2-(2′-hydroxyphenyl)-8-quinolinol and 2.6-diphenacylpyridine were synthesized. The compounds were characterized by IR-, UV/VIS-, MS-, 119Sn NMR and 119mSn Mössbauer spectroscopy. They exist as a mixture of geometrical isomers.  相似文献   

7.
Mössbauer emission spectra of defect119Sn atoms arising from119Sb were measured in InSb, GaSb, CdSb, ZnSb, In2Te3, CdTe, and Ag2Te labeled with119Sb or its parent119mTe. Together with the results of our previous studies, the isomer shifts of defect and normal119Sn were shown to correlate with the electronegativity of ligands from silver to iodine.  相似文献   

8.
On Polystannanes. III. 1,2-Dichloro-tetramethyl-distannane. Forming a Sn? Sn-connected Helical Double Chain Structure [(…?SnMe2Cl…?SnMe2? Cl…?)]2 The crystal structure of the title compound has been determined at ?160°C and refined to R = 0.071 (bond lengths Sn? Sn 277.0(2), Sn? Cl 244.2(3) and 244.8(3), Sn? C 214(2) pm). Intermolecular Sn…?Cl connection (324.0(3) and 329.2(3) pm) results in a double chain structure. 119Sn-NMR spectra in CH2Cl2 and acetone exhibit a movable temperature dependent coordination of acetone at the distannane (1J(119Sn? 119Sn) 8000 to 9000 Hz; appr. 5000 Hz in CH2Cl2).  相似文献   

9.
Abstract

The 119Sn NMR spectra of several sugar-tin derivatives were recorded. The geometric and steric isomers of all of the organotin derivatives studied were easily differentiated by 119Sn NMR. The appropriate 119Sn resonances are: ca - 50 ppm for trans and ?60 ppm for cis vinyltin derivatives (1-3), ca 16 ppm for allyltins 4-6, and ca ?32 ppm for tin-carbinols 9 and 11. When the hydroxyl group in carbinol 9 was converted to an O-acetyl group, the chemical shift of 119Sn was shifted to ?22 ppm.  相似文献   

10.
Diorganotin complexes of monoisopropyl and monomethyl nadiate, succinate, and phthalate were synthesized and characterized by elemental analysis, FT-IR, 1H NMR, 13C NMR, and 119Sn NMR spectroscopic techniques. The spectroscopic investigation demonstrated that carboxylate is bidentate in the diorganotin complexes. On the basis of 1 J(119Sn–13C) and 2 J(119Sn–1H) values, C–Sn–C bond angles were also calculated. The newly synthesized complexes were also screened for their antibacterial activities against Gram-positive and Gram-negative pathogenic strains of bacteria.  相似文献   

11.
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   

12.
In recent years, various protocols on preparing Lewis acidic Sn‐β zeolite hydrothermally and postsynthetically have been reported. However, very little is known about the effects of different synthesis protocols on the Sn(IV) speciation in the final material. Even the effects of individual synthesis parameters within a certain preparation method have not been studied systematically. Here, we demonstrate that hydrothermally synthesized Sn‐β zeolites prepared via very similar recipes show significantly different 119Sn‐NMR spectra, suggesting different Sn site speciation. Among postsynthetically prepared Sn‐β zeolites, less variation in the resulting 119Sn‐NMR spectra have been observed, indicating a more reproducible synthesis procedure compared to hydrothermal synthesis in fluoride media. This work highlights the importance of 119Sn‐NMR measurements to elucidate the precise local geometry of the Sn heteroatoms in Sn‐β, and the need to quantify the number of reactive Sn sites on each sample that participate in a given catalytic reaction, in order to accurately compare materials prepared by different routes.  相似文献   

13.
Adsorption behaviour of trace elements, In(III), Sn(IV), Sb(V) and Te(IV) on activated carbon and graphite powder was studied. Adsorption characteristics of the ions enabled the separation of In(III)–Sn(IV), Sn(IV)–Sb(V) and Sb(V)–Te(IV) pairs. Applications to practical separation, milking of113mIn from113Sn, removal of tin impurity from119Sb, and milking of119Sb from119mTe, are presented.  相似文献   

14.
The synthesis and 119Sn NMR characteristics of new five-coordinate tris(trichlorostannato) complexes of RhI, IrI and PtII are reported. The RhI and IrI complexes are complex dianions of the form (PPN)2[M(SnCl3)3L2] where L can be CO, CN (cyclohexyl) or L2, a diolefin such as 1,5-COD or NBD (norbornadiene). The anionic platinum complexes (PPN)[Pt(SnCl3)3L2] contain similar L ligands. A number of neutral monotrichlorostannato complexes of type [M(SnCl3)L4] including [Ir(SnCl3)(NBD)(1,5-COD)] have been prepared and characterized. Their δ(119Sn), δ(13C), δ(195Pt) as well as 1J(103Rh, 119Sn), 1J(195Pt, 119Sn), 2J(119Sn, 117Sn) and 2J(119Sn, 13C) data are given. A trans influence series, based on 1J(195Pt, 119Sn), reveals the following sequence: H? > PR3 > AsR3 > SnCl3? > olefin > Cl?.  相似文献   

15.
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   

16.
Dipole-dipole relaxation via non-bonded protons is an important relaxation mechanism for119Sn in tri-n-propyltin and tri-n -butyltin compounds. This causes a negative nuclear Overhauser effect, arising from the negative magnetogyric ratio, which in some cases nulls the signal. The relative contributions from the spin-rotation and dipole-dipole mechanisms vary: larger molecules have lower spin-rotation and higher dipolar relaxation rates. The practical significance of large nuclear Overhauser enhancement factors in recording 119Sn spectra and the relation of the dipole-dipole contribution to the molecular motion and of the spin-rotation contribution to the absolute shift scale for 119Sn are discussed.  相似文献   

17.
This paper presents the method of double labelling in the study of the kinetics of homogeneous isotope exchange reactions. This method was tested by the determination of the Sn(II)−Sn(IV) exchange rate in hydrochloric acid medium. The system was labelled by the tracer119mSn [initially in the Sn(IV) state]; when the isotope equilibrium was established, Sn(IV) was again labelled by tracer113Sn. The separation of Sn(II) and Sn(IV) in the given time of exchange was performed by the extraction of Sn(IV)-hydroxyquinolate into chloroform. The specific activities of the separated components were determined from the ratio of113Sn and119mSn activities. The exchange rate was calculated from the time dependence of specific activities. The advantage and possibilities of the method of double labelling in the study of isotope exchange are discussed.  相似文献   

18.
Bis(tributyltin) oxide or trimethyltin hydroxide react with carbon dioxide to afford the bis(trialkyltin) carbonates, (R3SnO)2CO; 119Sn NMR (in the case of R = Bu) or 119mSn Mössbauer spectroscopy show that these compounds contain 4- and 5-coordinate tin atom sites.  相似文献   

19.
The 119Sn nuclei of hexamethylditin, formed during the photochemical reaction of trimethyltin hydride with d1-t-butyl peroxide or dibenzyl ketone, or during the thermal decomposition of azodusobutyronitrile with trimethyltin hydride, exhibit CIDNP. The nuclear polarisation is built up in radical pairs Me3Sn··SnMe31. The full CKO theory has to be used for explaining the net effect in the main 119Sn signals of the hexamethylditin. The high field approximation is not valid because of the high value of the 119Sn hyperfine splitting in trimethylstannyl radicals. The multiplet effect in the 117Sn satellites is interpreted in terms of the high field treatment. A negative sign is found for a117Sn(Me3Sn) and a119Sn(Me3Sn). 119Sn-CIDNP also appears in benzyltrimethyltin during photolysis of dibenzylketone with trumethyltin hydride. It is concluded from 1H-CIDNP investigations that nuclear polarisations built up in radical pairs containing both stannyl radicals and others are not observed in hexamethylditin. A positive sign is found for
.  相似文献   

20.
195Pt, 119Sn and 31P NMR characteristics of the complexes trans-[Pt(SnCl3)(carbon ligand)(PEt3)2] (1a-1e) are reported, (carbon ligand = CH3 (1a), CH2Ph (1b), COPh (1c), C6Cl5 (1d), C6Cl4Y (e); Y = meta- and para-NO2, CF3, Br, H, CH3, OCH3, or Pt(SnCl3)(PEt3)2. The values of 1J(195Pt, 119Sn) vary from 2376 to 11895 Hz with the COPh ligand having the smallest and the C6Cl5 ligand the largest value, making a total range for this coupling constant, when the dimer syn-trans-[PtCl(SnCl3)(PEt3)]2 is included, of ca. 33000 Hz. In the meta- and para-substituted phenyl complexes 1J(195Pt, 119Sn) (a) is greater for electron-withdrawing substituents, (b) varies more for the meta-substituted derivatives (5634 to 7906 Hz) than for the para analogues (6088 to 7644 Hz) and (c) has the lowest values when the Pt(SnCl3)(PEt3)2 group is the meta- or para-substituent. The direction of the change in 1J(195Pt, 119Sn) is opposite to that found for 1J(195Pt, 119P). For the aryl complexes linear correlations are observed between δ(119Sn), 1J(195Pt, 119Sn), 1J(195Pt, 31P), 1J(119Sn, 31P) and the Hammett substituent constant σn. δ(119Sn) and 1J(195Pt, 119Sn) are related linearly to v(Pt-H) in the complexes trans-[PtH(C6H4Y)(PEt3)2]; δ(119Sn) and δ(1H) (hydride) are also linearly related. Based on 1J(195Pt, 119Sn), the acyl ligand is suggested to have a very large NMR trans influence. The differences in the NMR parameters for (1a-e) are rationalized in terms of differing σ- and π-bonding abilities of the carbon ligands.The structure of 1c has been determined by crystallographic methods. The complex has a slightly distorted square planar geometry with trans-PEt3 ligands. Relevant bond lengths (Å) and bond angles (°) are: PtSn, 2.634(1), PtP, 2.324(4) and 2.329(4), PtC, 2.05(1); PPtP, 170.7(6), SnPtC, 173.0(3), SnPtP, 92.1(1), 91.7(1), PPtC, 88.8(4) and 88.3(4). The PtSn bond separation is the longest yet observed for square-planar platinum trichlorostannate complexes, and would be consistent with a large crystallographic trans influence of the benzoyl ligand. The PtSn bond separation is shown to correlate with 1J(195Pt, 119Sn).  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