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.
A series of the extremely crowded triarylpnitogens, tris(2,4,6-triisopropylphenyl)phosphine (1), arsine (2), stibine (3), and bismuthine (4), were synthesized by the reaction of 2,4,6-triisopropylphenylcopper(I) with the corresponding pnictogen trichlorides. Introducion of the three bulky aryl groups resulted in the unusual structures and redox properties, which were studied by X-ray crystallography and cyclic voltammetry. The triarylpnictogens 1, 2, and 3 had extremely large bond angles around pnictogen atoms (1: 111.5 degrees , 2: 109.2 degrees , 3: 106.7 degrees ), and not only 1 and 2, but also stibine 3 displayed a reversible redox wave in the cyclic voltammograms at very low potentials (1: 0.16 V, 2: 0.50 V, 3: 0.57 V vs Ag/Ag+), which suggests considerable stability of the corresponding cation radicals.  相似文献   

2.
3.
4.
The new tris(ferrocenylamine) ditertiary phosphine 1,1′-{FcCH2N(CH2PPh2)CH25-C5H4)}2Fe [Fc = (η5-C5H5)Fe(η5-C5H4)] has been prepared along with two coordination complexes. All compounds have been characterised by a combination of spectroscopic and analytical methods. The single crystal X-ray structure of the pentametallic Ru2Fe3 complex 5 has been determined.  相似文献   

5.
The phosphine oxide complexes [GaX3(Me3PO)] and [(GaX3)2{μ-o-C6H4(CH2P(O)Ph2)2}] have been prepared and characterised by microanalysis, IR and multinuclear NMR (1H, 13C{1H}, 31P{1H} and 71Ga) spectroscopy. The structures of [GaCl3(Me3PO)], [(GaBr3)2{μ-o-C6H4(CH2P(O)Ph2)2}] and of the ionic product [GaI2(Me3PO)2][GaI4] have been determined and show that the Lewis acidity of the gallium halides towards phosphinoyl ligands diminishes as the halogen becomes heavier. The [GaX3(Ph3E)] (X = Cl, Br or I; E = P or As) and [(GaX3)2{μ-o-C6H4(CH2PPh2)2}] (X = Br or I) have been prepared and their structural and spectroscopic properties compared with those of the phosphinoyl complexes. The results, and competitive solution NMR studies, show that Ga(III) binds the hard R3PO in preference to the softer phosphine or arsine ligands. Hydrolysis of gallium(III) phosphines is shown to lead to [R3PH][GaX4], but in contrast to some other p-block halides, GaX3 do not promote air-oxidation of R3P to R3PO.  相似文献   

6.
Treatment of [LOEtTi(OTf)3] (, OTf = triflate) with S-binapO2 (binap = 2,2′-bis(diphenylphosphinoyl)-1,1′-binaphthyl) afforded the terminal hydroxo complex [LOEtTi(S-binapO2)(OH)][OTf]2 (1). Treatment of [LOEtTi(OTf)3] with K(tpip) (tpip = [N(Ph2PO)2]) afforded [LOEtTi(tpip)(OTf)][OTf] (2) that reacted with CsOH to give [LOEtTi(tpip)(OH)][OTf] (3). The structures of 1 and 2 have been determined.  相似文献   

7.
Ultra-branched mixed tetradentate tripodal phosphines and phosphine chalcogenides have been synthesized by the exhaustive regioselective addition of secondary phosphines, phosphine sulfides and phosphine selenides to available tris(4-vinylbenzyl)phosphine oxide under free-radical conditions (UV irradiation or AIBN) in good to excellent yields.  相似文献   

8.
Palladium(II) complexes with a tetradentate pseudo-tripodal ligand having two phosphino groups and two phosphine sulfide or selenide groups, pp3X2 (pp3 = tris[2-(diphenylphosphino)ethyl]phosphine, X = S (1) or Se (2)), were prepared from [PdCl(pp3)]Cl. Both of these phosphine chalcogenide complexes 1 and 2 showed rapid equilibrium between the five-coordinate [PdCl(pp3X2)]Cl with two bound phosphine chalcogenide groups and four-coordinate [PdCl2(pp3X2)] with two dissociated pendant ones in chloroform. The thermodynamic parameters for the reaction, [PdCl(pp3X2)]+ + Cl?[PdCl2(pp3X2)], were obtained by low-temperature 31P NMR as follows: K298 = 3.7 × 103 and 5.4 × 102 mol−1, ΔH° = 11.3 ± 0.3 and 13.4 ± 0.4 kJ mol−1, and ΔS° = 106 ± 2 and 97 ± 2 J mol−1 K−1 for 1 and 2, respectively. The rate for the geometrical change at 246.7 K for 1 was appreciably faster than that for 2. These thermodynamic and kinetic results indicate that the phosphine selenide Se atoms can stabilize the five-coordinate structure by effective π-back donation from Pd(II) compared with the phosphine sulfide S atoms. Difference in retention of the catalytic activity for Suzuki coupling, 2 > 1 > [PdCl(pp3 or p3)]Cl, was explained by difference in the π-accepting ability that stabilizes the catalytically active Pd(0) species. Considering the rapid dissociation-coordination equilibrium of the phosphine chalcogenide groups on Pd(II), it is probable that the oxidative addition and the subsequent transmetallation of the Pd(II) species are hardly blocked by the phosphine chalcogenide groups.  相似文献   

9.
A comparison of the tris(trimethylsilyl)silyl I and tris(trimethylsilyl)germyl II radical reactivity is provided. Their formation as well as their reactivity encountered in a large variety of chemical processes (addition to double bond, halogen abstraction, peroxyl radical formation…) is examined by laser flash photolysis, quantum mechanical calculations and electron spin resonance (ESR) experiments. The starting compound (TMS)3GeH is more reactive than (TMS)3SiH toward the t-butoxyl, the t-butylperoxyl and the phosphinoyl radicals. A similar behavior is noted for an aromatic ketone triplet state. II exhibits a lower absolute electronegativity: accordingly, the addition to electron rich alkenes is less efficient than for I. Radical II is also found less reactive for both the peroxylation and the halogen abstraction reactions. The rearrangement of is slower than for ; this is related to the respective exothermicity of the processes.  相似文献   

10.
11.
12.
The proton and carbon-13 NMR spectra of solutions in CDCl3 of dimethyl(phenylthio)arsine and tris(phenylthio)arsine where investigated. All resolved experimental resonances were assigned and the coupling constants determined by means of an iterative procedure employing the LAO-COON3 program. The spectra were simulated using Gaussian—Lorentzian lineshapes. The protons in the 2,6- and 3,5-positions were found to be equivalent. The resonances appear in the sequence δ2,6 ⪢ δ3,5 ⪢ δ4 and δ2,6 ⪢ δ4 ⪢ δ3,5 for dimethyl(phenylthio)arsine and tris(phenylthio)arsine, respectively. Whereas the protons in the 2,6-positions resonate at the same frequency (7.44 ppm) in the two compounds, the shifts for the 3,5- and 4-protons of tris(phenylthio)arsine are downfield from those of dimethyl(phenylthio)arsine. These differences are explained in terms of the hypothesis that the conformation of the rigid tris(phenylthio)arsine molecule does not allow overlap between the sulfur atomholding the lone electron pair and the π-ring orbital to the same extent as in dimethyl(phenylthio)arsine; in which the phenyl group can oscillate about an equilibrium position characterized by maximal ps-π-ring overlap.  相似文献   

13.
Molecular mechanics (MM2) calculations were performed on 54 conformations of 18 phosphines (PH3; PH3−nRn, where n = 1,…3, and R = Me and Et, n = 1 or 2 and R =iPr, and n = 1 and R =tBu, PMe2Et, PMeEt2, and PPhMe2, and PPh2R where R = Me, Et, iPr, tBu and Ph). The results are compared to those previously obtained from MINDO/3 and MNDO calculations, and to experimental data. Single conformer cone angles and weighted average cone angles were calculated from MM2 optimized geometries employing Tolman's general definition, and they are compared to Tolman's values, MINDO/3 results, and T.L. Brown's ER values. Of the cone angle definitions used, the weighted average values are suggested as the best single representation of phosphine ligand sizes. The steric parameters (cone angle and ER values) alone, and in conjunction with electronic parameters, are correlated with experimental data.  相似文献   

14.
The synthesis of a series of (fluoroalkyl)phosphine complexes of nickel is reported. Treatment of (cod)2Ni with dfepe (dfepe=(C2F5)2PCH2CH2P(C2F5)2) yields (dfepe)Ni(cod) (1), which has been structurally characterized. Treatment of 1 with CO or bipy results in the formation of (dfepe)Ni(CO)2 (2) and (dfepe)Ni(bipy) (3), respectively. Addition of excess dfepe to 1 results in incomplete cod displacement to form (dfepe)2Ni (4). The homoleptic complex 4 may be independently prepared in high yield by reduction of (acac)2Ni with (iBu)3Al in the presence of butadiene and excess dfepe. Solvation of (dfepe)Ni(cod) in acetonitrile gives a new complex tentatively identified as (dfepe)Ni(MeCN)2 (6), whereas dissolution of (dfepe)2Ni in acetonitrile leads to a mixture of 6 and the partial displacement product (dfepe)(η1-dfepe)Ni(MeCN) (5). In contrast to (R3P)4Ni(0) phosphine and phosphite complexes, which undergo protonation by strong anhydrous acids such as HCl, H2SO4 and CF3CO2H to give (R3P)4Ni(H)+ products, Treatment of (dfepe)2Ni with neat CF3CO2H or excess HOTf in dichloromethane gave no spectroscopic evidence for (dfepe)2Ni(H)+. Exposure for extended periods leads to dfepe loss and decomposition to Ni(II) products. The synthesis of the first cobalt complex of dfepe, (dfepe)Co(CO)2H, is also reported.  相似文献   

15.
16.
17.
The reaction of 1,1,1-tris(chloromethyl)propane with diphenylphosphine under phase-transfer conditions afforded 1,1,1-tris(diphenylphosphinomethyl)propane, whose oxidation gave a previously unknown representative of trigonal tris(phosphine oxides), viz., stable 1,1,1-tris(diphenylphosphorylmethyl)propane. Its analogs, viz., bis(diphenylphosphoryl)diphenylphosphinomethane and tris(diphenylphosphoryl)methane, are unstable in air and decompose with the cleavage of the P-C bond.Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 9, pp. 1926–1929, September, 2004.  相似文献   

18.
19.
20.
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