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.
Tetraphenylbismuth(V) derivatives of the general formula Ph4BiX [X = OSO2C6H4Me-4, OC6H2(NO2)3-2,4,6, OC6H2(NO2-4)(Br2-2,6), OSO2C6H3(OH)(COOH)] react with methyl acrylate in the presence of palladium dichloride (1:3:0.04 molar ratio) in acetonitrile at 20°C to form the cross-coupling products, methyl cinnamate (0.17–0.54 mol mol?1 starting bismuth compound) and methylhydrocinnamate (0.10–0.73 mol mol?1), diphenyl (0.06–0.80 mol mol?1), and benzene (0.02–0.36 mol mol?1). The highest C-phenylating activity is shown by Ph4BiOSO2C6H4Me-4. The mechanisms with the palladium-catalyzed cross-coupling reactions are suggested.  相似文献   

2.
The reactions of tetraphenylbismuthonium and -stibonium salts Ph4EX (E = Bi, Sb; X = I, OSO2 (C6H3(CH3)2-2,5), OSO2C6H3(OH-4)(COOH-3)) with bismuth triiodide in acetone afford complexes [Ph4Bi]+[PhBi(C5H5N)I3]-, [(Ph4BiO)2S(O){2,5-(CH3)2C6H3S(O)} [Ph2Bi2I6]2–, [Ph4Sb [Bi4I16]4-·2(CH3)2C=O, and [Ph4Sb] 3+ + [Bi5I18]3-, whose structural units, according to the X-ray diffraction data, are tetraphenylbismuthonium (-stibonium) cations and mono-, di-, tetra-, and pentanuclear anions, respectively.  相似文献   

3.
A new catalytic reaction of the competing phenylation and hydrophenylation in air of methyl acrylate with tetraphenylantimony chloride in the presence of PdCl2 (0.04 mol per 1 mol of organometallic compound) in acetonitrile at 50°C for 6 h was studied. The yields of methyl cynnamate and methyl hydrocynnamate were 0.73 and 0.27 mol mol?1 respectively. The products ratio obtained depends slightly on the process duration, the Ph4SbCl and methyl acrylate ratio, and the structure of Pd salt [PdCl2, Pd(OAc)2, Li2PdCl4], but significantly on the nature of a solvent (MeCN > DMF > THF). The use of Ph4SbCl instead of Ph4SbBr leads to decrease in the yield of methyl hydrocynnamate to 0.04 mol mol?1. In the reactions of Ph4SbX (X = F, I, OAc, O2CEt) the product is not formed at all.  相似文献   

4.
The formation heats ΔHofPhICl2(c) = 21.7 kJ mol?1 and ΔHofpFC6H4ICl2(c) = ?173.0 kJ mol?1 were determined by reaction of iodides with excess chlorine in methyl cyanide. These heats are used to show:(i) that thermal or photo-decompositions of aryl iodine dichlorides are kinetically controlled,(ii) the greater halogenating ability of aryl iodine dichlorides than Ph3ACl2 (A = P, As, Sb) compounds, and(iii) the weakening of halogenating ability with phenylation.  相似文献   

5.
The kinetics of the formation of the titanium‐peroxide [TiO2+2] complex from the reaction of Ti(IV)OSO4 with hydrogen peroxide and the hydrolysis of hydroxymethyl hydroperoxide (HMHP) were examined to determine whether Ti(IV)OSO4 could be used to distinguish between hydrogen peroxide and HMHP in mixed solutions. Stopped‐flow analysis coupled to UV‐vis spectroscopy was used to examine the reaction kinetics at various temperatures. The molar absorptivity (ε) of the [TiO2+2] complex was found to be 679.5 ± 20.8 L mol?1 cm?1 at 405 nm. The reaction between hydrogen peroxide and Ti(IV)OSO4 was first order with respect to both Ti(IV)OSO4 and H2O2 with a rate constant of 5.70 ± 0.18 × 104 M?1 s?1 at 25°C, and an activation energy, Ea = 40.5 ± 1.9 kJ mol?1. The rate constant for the hydrolysis of HMHP was 4.3 × 10?3 s?1 at pH 8.5. Since the rate of complex formation between Ti(IV)OSO4 and hydrogen peroxide is much faster than the rate of hydrolysis of HMHP, the Ti(IV)OSO4 reaction coupled to time‐dependent UV‐vis spectroscopic measurements can be used to distinguish between hydrogen peroxide and HMHP in solution. © 2007 Wiley Periodicals, Inc. Int J Chem Kinet 39: 457–461, 2007  相似文献   

6.
The influence of the central metal atom and acid residue in a molecule of organometallic compound Ph3MX2 (M = Bi, Sb; X = Cl, Br, OAc, O2CEt) on the direction and yield of dephenylation products in a methyl acrylate— methanol system in the presence of Cu(OAc)2 and Na2PdCl2(OAc)2 was studied at 50 °C in acetonitrile. The main product of dephenylation of the antimony compounds is methyl cinnamate (yield 0.31–1.59 moles per mole of the starting Ph3SbX2), while anisole (0.55–0.97 mol) and halobenzene (0.67–1.04 mol) are those for compounds Ph3Bi(O2CR)2 and Ph3BiHal2, respectively. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 4, pp. 655–658, April, 2006.  相似文献   

7.
Triphenylantimony was used as an efficient agent for C-phenylation of methyl acrylate in the presence of ButOOH (1 to 2 mol) and a palladium salt (PdCl2, Li2PdCl4; 0.04 mol) in AcOH at 50 °C. The yield of methyl cinnamate is two moles per mole of the starting Ph3Sb.  相似文献   

8.
1-Substituted 4,4,6-trimethyl-1H, 4H-pyrimidine-2-thiol(L), whose substituents are C6H5 ( 1 ), p-CH3OC6H4 ( 2 ), p-O2NC6H4 ( 3 ), n-C4H9 ( 4 ), and NH2 ( 5 ), react with TeCl4, ArTeCl3 (Ar = C6H5, 4-CH3OC6H4, or 4-C2H5OC6H4), Ph2TeCl2, and Ph3TeCl, resulting in TeCl3(L–H), ArTeCl2(L–H), Ph2TeCl2.L, and Ph3TeCl.L types of compounds. Elemental analyses, molecular weights, conductances in solutions, IR, and 1H and 13C NMR spectral data of these compounds suggest that, in all of them, the pyrimidine-2-thiols ligate through sulfur only. In the first two types of derivatives, the thiol form of L reacts with tellurium moieties, resulting in the liberation of HCl. On the basis of these data TeCl3(L-H) appears to exist as a dimer in which two Te atoms are probably 5-coordinated and are bridged by two Cl. Tellurium in ArTeCl2(L-H), Ph2TeCl2.L, and Ph3TeCl.L also appears to be 5-coordinated. IR data suggest that the Te–S bond in Ph2TeCl2.L is weak and, therefore, because of partial dissociation, the molecular complexes exhibit lower molecular weights in CH3CN. Ph3TeCl.L in acetonitrile behaves as a 1:1 electrolyte. Attempts to synthesize N-, S-, and Te-containing heterocycle ring compounds from TeCl3( 1/2 -H) and ArTe( 1/2 -H)Cl2 did not succeed.  相似文献   

9.
The anionic gold(I) complexes [1‐(Ph3PAu)‐closo‐1‐CB11H11]? ( 1 ), [1‐(Ph3PAu)‐closo‐1‐CB9H9]? ( 2 ), and [2‐(Ph3PAu)‐closo‐2‐CB9H9]? ( 3 ) with gold–carbon 2c–2e σ bonds have been prepared from [AuCl(PPh3)] and the respective carba‐closo‐borate dianion. The anions have been isolated as their Cs+ salts and the corresponding [Et4N]+ salts were obtained by salt metathesis reactions. The salt Cs‐ 3 isomerizes in the solid state and in solution at elevated temperatures to Cs‐ 2 with ΔHiso=(?75±5) kJ mol?1 (solid state) and ΔH=(118±10) kJ mol?1 (solution). The compounds were characterized by vibrational and multi‐NMR spectroscopies, mass spectrometry, elemental analysis, and differential scanning calorimetry. The crystal structures of [Et4N]‐ 1 , [Et4N]‐ 2 , and [Et4N]‐ 3 were determined. The bonding parameters, NMR chemical shifts, and the isomerization enthalpy of Cs‐ 3 to Cs‐ 2 are compared to theoretical data.  相似文献   

10.
Four new complexes of pentavalent bismuth are synthesized: Ph3Bi[OC6H2(Br3-2,4,6)]2, Ph3Bi[OC6H2(Cl3-2,4,6)]2, Ph4BiOC6H2(Br3-2,4,6), and Ph4BiOC6H2(Cl3-2,4,6). Tetraphenylbismuth aroxides are produced by the disproportionation reaction of ligands from pentaphenylbismuth and triphenylbismuth diaroxides in toluene or from pentaphenylbismuth and phenol. Triphenylbismuth diaroxides are synthesized from phenol, triphenylbismuth, and hydrogen peroxide taken at a molar ratio of 2 : 1 : 1, respectively, in diethyl ether. According to the X-ray diffraction data, the bismuth atom surrounding in 2,4,6-tribromophenoxytetraphenylbismuth has the configuration of a trigonal bipyramid with the aroxyl ligand in the axial position. The Bi-C and Bi-O bond lengths are 2,184, 2.190, 2.234, and 2.514 Å, respectively, and the equatorial CSbC angles are equal to 111.4°, 121.3°, and 121.3°.Translated from Koordinatsionnaya Khimiya, Vol. 30, No. 12, 2004, pp. 935–938.Original Russian Text Copyright © 2004 by Sharutin, Egorova, Tsiplukhina, Gerasimenko, Pushilin.  相似文献   

11.
The heat of formation of benzophenone oxide, Ph2CO2, was measured using photoacoustic calorimetry. The enthalpy of the reaction Ph2CN2 + O2 → Ph2CO2 + N2 was found to be ?48.0 ±0.8 kcal mol?1 and ΔHf(Ph2CN2) was determined by measuring the reaction enthalpy for Ph2CN2 + EtOH → Ph2CHOEt + N2 (?53.6 ±1.0 kcal mol?1). Taking ΔHf(PhCHOEt) = ?10.6 kcal mol?1 led to ΔHf(Ph2CN2) = 99.2 ± 1.5 kcal mol?1 and hence to ΔHf(Ph2CO2) = 51.1 ± 2.0 kcal mol?1. The results imply that the self-reaction of benzophenone oxide i.e., 2Ph2CO2 → 2Ph2CO + O2 is exothermic by ?76.0 ±4.0 kcal mol?1.  相似文献   

12.
The reaction of niobium pentachloride with three equivalents of 2-t-butylphenol in carbon tetrachloride afforded [NbCl2(OC6H4C(CH3)3-2)3]. The identity of the complex has been established by elemental analyses, molar conductance, molecular weight determination, IR, 1H, and 13C-NMR and UV-Vis spectral studies. Based upon these studies, a square–pyramidal geometry around niobium has been proposed. Thermal behavior of the complex has been studied by TGA and DTA. Acceptor behavior of [NbCl2(OC6H4C(CH3)3-2)3] toward Ph3P, Ph3As, Ph3PO, Ph3AsO, and an uncommon ligand arsenictrithiophenoxide As(SPh)3 allows the isolation of 1 : 1 addition compounds as shown by physicochemical, IR, and 1H-NMR spectral studies. The formation of [NbCl2(OC6H4C(CH3)3-2)3] · As(SPh)3 appears to be the first adduct of its class and suggests the suitability of As(SPh)3 as a ligand.  相似文献   

13.
A stopped-flow investigation of the reversible addition of Ph3P to [(C8H11)Co(C5H5)]+ indicates the rate law, kobs = k1[Ph3P] + k?1. The low Δ2 of 21.0 ± 1.2 kJ mol?1 and the negative ΔS2 of ?114 ± 5 J K?1 mol?1 are consistent with rapid addition to the enyl ligand. The higher Δ2 of 86.2 ± 5.1 kJ mol?1 and the positive ΔS2 of +60 ± 17 J K?1 mol?1are as expected for the reverse dissociation. Preliminary studies show that the related complex [(C7H9)Co(C5H5)]+ is at least 65 times more electrophilic towards Ph3P.  相似文献   

14.
A series of novel titanium(IV) complexes bearing tetradentate [ONNO] salan type ligands: [Ti{2,2′‐(OC6H3‐5‐t‐Bu)2‐NHRNH}Cl2] (Lig1TiCl2: R = C2H4; Lig2TiCl2: R = C4H8; Lig3TiCl2: R = C6H12) and [Ti{2,2′‐(OC6H2‐3,5‐di‐t‐Bu)2‐NHC6H12NH}Cl2] (Lig4TiCl2) were synthesized and used in the (co)polymerization of olefins. Vanadium and zirconium complexes: [ M{2,2′‐(OC6H3‐3,5‐di‐t‐Bu)2‐NHC6H12NH}Cl2] (Lig4VCl2: M = V; Lig4ZrCl2: M = Zr) were also synthesized for comparative investigations. All the complexes turned out active in 1‐octene polymerization after activation by MAO and/or Al(i‐Bu)3/[Ph3C][B(C6F5)4]. The catalytic performance of titanium complexes was strictly dependent on their structures and it improves for the increasing length of the aliphatic linkage between nitrogen atoms (Lig1TiCl2 << Lig2TiCl2 < Lig3TiCl2) and declines after adding additional tert‐Bu group on the aromatic rings (Lig3TiCl2 < Lig4TiCl2). The activity of all titanium complexes in ethylene polymerization was moderate and the properties of polyethylene was dependent on the ligand structure, cocatalyst type, and reaction conditions. The Et2AlCl‐activated complexes gave polymers with lover molecular weights and bimodal distribution, whereas ultra‐high molecular weight PE (up to 3588 kg mol?1) and narrow MWD was formed for MAO as a cocatalyst. Vanadium complex yielded PE with the highest productivity (1925.3 kg molv?1), with high molecular weight (1986 kg mol?1) and with very narrow molecular weight distribution (1.5). Copolymerization tests showed that titanium complexes yielded ethylene/1‐octene copolymers, whereas vanadium catalysts produced product mixtures. © 2014 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2014 , 52, 2111–2123  相似文献   

15.
《Polyhedron》1986,5(4):959-965
119Sn NMR and Mössbauer spectroscopic data have been recorded for 12 compounds of formula Ph3SnO2CC6H4X(X = H, Me-2, NH2-2, NMe2-2, Cl-2, Cl-3, Cl-4, OH-2, OH-4, SMe-4 or OMe-2) and Ph3SnO2CC10H7-1. On the basis of these measurements all the compounds are assigned a coordination number of four at tin in solution, and, with the exception of the Cl-2 and OH-2 derivatives, retain this structure in the solid. Both of these latter compounds exhibit enhanced Mössbauer quadrupole splittings (3.71 and 2.97 mm s−1), which are attributed to carboxylate- and hydroxyl-bridged structures, respectively. The variable-temperature Mössbauer spectra of two compounds (X = OMe-2 or OH-2) are discussed.  相似文献   

16.
Organotin compounds R3Sn(CH2)n+2OC6H4C6H4Y (R3=Ph3, Ph2Bu; Y=H, CN; n=1-3) and RX2Sn(CH2)n+2OC6H4C6H4Y (R=Ph, Bu; Y=H, CN; X=Br, I; n=1-3) have been synthesised and characterised by 1H-, 13C-, 119Sn-NMR and Mössbauer spectroscopies. X-ray crystallography reveals tetrahedral geometries for Ph3Sn(CH2)4OC6H4C6H5 and Ph3Sn(CH2)3OC6H4C6H4CN, a six-coordinated, bromine-bridged dimeric structure for PhBr2Sn(CH2)3OC6H4C6H5 containing a mer-Br3C2OSn coordination sphere about tin and a five-coordinated monomeric structure for PhBr2Sn(CH2)3OC6H4C6H4CN. In all cases there is strong alignment of mesogenic groups in the solid-state but only PhBr2Sn(CH2)3OC6H4C6H4CN shows any indication of liquid-crystal behaviour. Wurtz polymerisation of RBr2Sn(CH2)5OC6H4C6H5 (R=Ph, Bu), both of which contain non-chelating ether functions, generated polystannanes (RR′Sn)n with Mn 2.3×105; Mw 3.0×105; Mw/Mn 1.30 and Mn 1.3×105; Mw 2.5×105; Mw/Mn 1.96, respectively, while no polymer was obtained from chelated PhBr2Sn(CH2)3OC6H4C6H5  相似文献   

17.
The reaction of the bulky bis(imidazolin‐2‐iminato) ligand precursor (1,2‐(LMesNH)2‐C2H4)[OTs]2 ( 1 2+ 2[OTs]?; LMes=1,3‐dimesityl imidazolin‐2‐ylidene, OTs=p‐toluenesulfonate) with lithium borohydride yields the boronium dihydride cation (1,2‐(LMesN)2‐C2H4)BH2[OTs] ( 2 + [OTs]?). The boronium cation 2 + [OTs]? reacts with elemental sulfur to give the thioxoborane salt (1,2‐(LMesN)2‐C2H4)BS[OTs] ( 3 + [OTs]?). The hitherto unknown compounds 1 2+ 2[OTs]?, 2 + [OTs]?, and 3 + [OTs]? were fully characterized by spectroscopic methods and single‐crystal X‐ray diffraction. Moreover, DFT calculations were carried out to elucidate the bonding situation in 2 + and 3 +. The theoretical, as well as crystallographic studies reveal that 3 + is the first example for a stable cationic complex of three‐coordinate boron that bears a B?S double bond.  相似文献   

18.
The kinetics of the reaction of trans-bromo(tetracarbonyl)phenylcarbyne-tungsten (IIIb) with N = PPh3, AsPh3, SbPh3, P(p-C6H4CH3)3, P(p-C6H4Cl)3, P(p-C6H4F)3, P(p-C6H4NMe2)3, Ph2AsCH2AsPh2 and P(OPh)3 have been studied in octane, n-butyl bromide, 1,1,2-trichloroethane and various other solvents. The formation of the monosubstituted carbyne complex follows a first-order rate law. The rates of reaction depend neither on the nature of the substitting nucleophile nor on its concentration. The rates decrease with increasing polarity of the solvent. Under high CO-pressure the substitution is partially reversible. The activation parameters are ΔH 98–108 kJ mol?1 and ΔS 26–53 J K?1 mol?1. The results are discussed on the basis of a dissociative mechanism.  相似文献   

19.
A new catalytic reaction of C-phenylation of methyl acrylate with bismuth derivatives Ph3Bi(O2CR)2 (R = H, Me, Et, Bu, CF3, Ph) or Ph3BiCl2 in a 3 : 1 ratio was studied. The reaction was performed in the presence of 4 mol % palladium compounds PdCl2, Pd(OAc)2, Pd2(dba)3, Pd(Ph3P)2Cl2, and PdLCl2 (L = dppm, dppe, dppp, dppb, dppf, binap, xantphos) at 50°C in CH3CN, DMF, THF, or CH2Cl2 and gave methyl cinnamate (yield up to 85% based on the starting bismuth compound), diphenyl (up to 138%), and benzene (up to 59%).  相似文献   

20.
Interconversion of the molybdenum amido [(PhTpy)(PPh2Me)2Mo(NHtBuAr)][BArF24] (PhTpy=4′‐Ph‐2,2′,6′,2“‐terpyridine; tBuAr=4‐tert‐butyl‐C6H4; ArF24=(C6H3‐3,5‐(CF3)2)4) and imido [(PhTpy)(PPh2Me)2Mo(NtBuAr)][BArF24] complexes has been accomplished by proton‐coupled electron transfer. The 2,4,6‐tri‐tert‐butylphenoxyl radical was used as an oxidant and the non‐classical ammine complex [(PhTpy)(PPh2Me)2Mo(NH3)][BArF24] as the reductant. The N?H bond dissociation free energy (BDFE) of the amido N?H bond formed and cleaved in the sequence was experimentally bracketed between 45.8 and 52.3 kcal mol?1, in agreement with a DFT‐computed value of 48 kcal mol?1. The N?H BDFE in combination with electrochemical data eliminate proton transfer as the first step in the N?H bond‐forming sequence and favor initial electron transfer or concerted pathways.  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