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 87 毫秒
1.
Treatment of the biphenyl derivative [S=C{(NCH2But)2C6H3‐3,4}]2 or [Cl2Si{(NCH2But)2C6H3‐3,4}]2 with C8K afforded the new bis(carbene) 1 or the first bis(silylene) 2 , respectively. The X‐ray structure of 2 is presented.  相似文献   

2.
In an effort to develop new tripodal N-heterocyclic carbene (NHC) ligands for small molecule activation, two new classes of tripodal NHC ligands TIMER and TIMENR have been synthesized. The carbon-anchored tris(carbene) ligand system TIMER (R = Me, t-Bu) forms bi- or polynuclear metal complexes. While the methyl derivative exclusively forms trinuclear 3:2 complexes [(TIMEMe)2M3]3+ with group 11 metal ions, the tert-butyl derivative yields a dinuclear 2:2 complex [(TIMEt-Bu)2Cu2]2+ with copper(I). The latter complex shows both “normal” and “abnormal” carbene binding modes and accordingly, is best formulated as a bis(carbene)alkenyl complex. The nitrogen-anchored tris(carbene) ligands TIMENR (R = alkyl, aryl) bind to a variety of first-row transition metal ions in 1:1 stoichiometry, affording monomeric complexes with a protected reactivity cavity at the coordinated metal center. Complexes of TIMENR with Cu(I)/(II), Ni(0)/(I), and Co(I)/(II)/(III) have been synthesized. The cobalt(I) complexes with the aryl-substituted TIMENR (R = mesityl, xylyl) ligands show great potential for small molecule activation. These complexes activate for instance dioxygen to form cobalt(III) peroxo complexes that, upon reaction with electrophilic organic substrates, transfer an oxygen atom. The cobalt(I) complexes are also precursors for terminal cobalt(III) imido complexes. These imido complexes were found to undergo unprecedented intra-molecular imido insertion reactions to form cobalt(II) imine species. The molecular and electronic structures of some representative metal NHC complexes as well as the nature of the metal–carbene bond of these metal NHC complexes was elucidated by X-ray and DFT computational methods and are discussed briefly. In contrast to the common assumption that NHCs are pure σ-donors, our studies revealed non-negligible and even significant π-backbonding in electron-rich metal NHC complexes.  相似文献   

3.
In this review, the synthesis, reactivity and properties of linear and cyclic oligophosphanides are described. Specifically the structures and versatile reactivity of the anionic ligands (P4R4)2? (R = But, Ph, Mes), (P4HR4)? (R = Ph and Mes) and cyclo-(P5But4)? towards main group and transition metal complexes is elucidated. In addition, potential application of metal oligophosphanides as precursors for the preparation of metal phosphides is also briefly discussed.  相似文献   

4.
The 16-electron Co, Rh and Ir half-sandwich complexes of Cp#M[E2C2(B10H10)] and Cp#M(E2S2C6H4) (M = Co, Rh, Ir, Ru; E = S, Se) containing chelating 1,2-dicarba-closo-dodecaborane-1,2-dichalcogenolato ligands and benzenedithiolato ligands are promising precursors to build multimetallic clusters by reactions with low oxidation state late transition metal reagents. Such reactions lead to successful constructions of M–M bonds between iridium, rhodium, cobalt, ruthenium, and other late transition metals. Most of these complexes have been characterized by X-ray single crystal determinations and some have been studied by computational methods. Such theoretical studies reveal the covalent bonding nature of those multinuclear complexes. Some of these clusters have been found to have interesting nonlinear optical properties.  相似文献   

5.
The reaction of organoaluminum compounds containing O,C,O or N,C,N chelating (so called pincer) ligands [2,6-(YCH2)2C6H3]AliBu2 (Y = MeO 1, tBuO 2, Me2N 3) with R3SnOH (R = Ph or Me) gives tetraorganotin complexes [2,6-(YCH2)2C6H3]SnR3 (Y = MeO, R = Ph 4, Y = MeO, R = Me 5; Y = tBuO, R = Ph 6, Y = tBuO, R = Me 7; Y = Me2N, R = Ph 8, Y = Me2N, R = Me 9) as the result of migration of O,C,O or N,C,N pincer ligands from aluminum to tin atom. Reaction of 1 and 2 with (nBu3Sn)2O proceeded in similar fashion resulting in 10 and 11 ([2,6-(YCH2)2C6H3]SnnBu3, Y = MeO 10; Y = tBuO 11) in mixture with nBu3SniBu. The reaction 1 and 3 with 2 equiv. of Ph3SiOH followed another reaction path and ([2,6-(YCH2)2C6H3]Al(OSiPh3)2, Y = MeO 12, Me2N 13) were observed as the products of alkane elimination. The organotin derivatives 411 were characterized by the help of elemental analysis, ESI-MS technique, 1H, 13C, 119Sn NMR spectroscopy and in the case 6 and 8 by single crystal X-ray diffraction (XRD). Compounds 12 and 13 were identified using elemental analysis,1H, 13C, 29Si NMR and IR spectroscopy.  相似文献   

6.
The steric and electronic effects of bulky aryl and silyl groups on the Si–Si triple bonding in RSiSiR and the short Ga–Ga distance in Na2[RGaGaR] are investigated by density functional calculations. As typical bulky groups, Tbt = C6H2-2,4,6-{CH(SiMe3)2}3, Ar′ = C6H3-2,6-(C6H3-2,6-iPr2)2, Ar1 = C6H3-2,6-(C6H2-2,4,6-iPr3)2, SiMe(SitBu3)2, and SiiPrDis2 (Dis = CH(SiMe3)2) are investigated and characterized. The importance of large basis sets is emphasized for density functional calculations.  相似文献   

7.
The fluorine substituted thiourea 2,6-F2C6H3C(O)NHC(S)NEt2 was prepared in good yield from the reaction of 2,6-F2C6H3C(O)Cl with KSCN and Et2NH in acetone. Using this compound several heteroleptic, monocationic Pd(II), Pt(II) and Ru(II) complexes of the type cis-[M{κ2S,O-2,6-F2C6H3C(O)NC(S)NEt2}(L)]PF6 [M = Pt, Pd; L = (Ph3P)2, tBu2bipy, 1,10-phen] as well as [Ru(η6-p-cym){κ2S,O-2,6-F2C6H3C(O)NC(S)NEt2}(PPh3)]PF6 were prepared in high yields. The compounds were characterised by spectroscopic methods and, in one case, by single crystal X-ray diffraction.  相似文献   

8.
The reactions of [(η7-C7H7)Hf(η5-C5H5)] (1b) with the two-electron donor ligands tert-butyl isocyanide (tBuNC), 2,6-dimethylphenyl isocyanide (XyNC), 1,3,4,5-tetramethylimidazolin-2-ylidene (IMe) and trimethylphosphine (PMe3) are reported. The 1:1 complexes [(η7-C7H7)Hf(η5-C5H5)L] (2b, L = tBuNC; 3b, L = XyNC; 4b, L = IMe, 5b, L = PMe3) have been isolated in crystalline form, and their molecular structures have been determined by X-ray diffraction analyses. The stabilities of these hafnium complexes were probed via spectroscopic and theoretical methods, and the results were compared to those previously reported for the corresponding zirconium complexes derived from [(η7-C7H7)Zr(η5-C5H5)] (1a). The X-ray crystal structure of the PMe3 adduct [(η7-C7H7)Zr(η5-C5H5)(PMe3)] (5a) was also established.  相似文献   

9.
The reaction of the 1,2,4-triphosphaferrocene [Cp*Fe(η5-P3C2tBu2)] (1) with CuX (X = Cl, Br, I) in a 1:1 stoichiometric ratio leads to the formation of the oligomeric compounds [{Cu(μ-X)}66-X)Cu(MeCN)3{μ,η2-(Cp*Fe(η5-P3C2tBu2))}233-(Cp*Fe(η5-P3C2tBu2))}] (X = Cl (2), Br (3)) and [{Cu(μ-I)}3{Cu(μ3-I)}3Cu(μ6-I){μ,η2-(Cp*Fe(η5-P3C2tBu2))}31-(Cp*Fe(η5-P3C2tBu2))}] (4) revealing Cu(I) halide cages surrounded by 1,2,4-triphosphaferrocene moieties. The reaction of [Cp*Fe(η5-P3C2tBu2)] with CuI in a 1:4 stoichiometry leads to the formation of the two-dimensional polymer [{Cu(μ-I)}4{Cu(μ3-I)(MeCN)}233-(Cp*Fe(P3C2tBu2))}]n (5). The oligomeric compounds show dynamic behavior in solution monitored by 31P NMR spectroscopy. All compounds are additionally characterized by single crystal X-ray diffraction.  相似文献   

10.
The reaction of triorganotin(IV) compound Ph2LSnCl (1), (L = 2,6-(t-BuOCH2)2C6H3), with (Bu3Sn)2O resulted to the isolation of Ph2LSn(μ-OH)Bu3SnCl (2), in which a monomeric triorganotin(IV) hydroxide Ph2LSnOH intermolecularly coordinates Bu3SnCl moiety. Compound 2 was characterized by combination of 1H, 13C and 119Sn NMR spectroscopy, ESI/MS, elemental analysis and X-ray diffraction.  相似文献   

11.
Metal carbene complexes have been at the forefront of organic and organometallic synthesis and are instrumental in guiding future sustainable chemistry efforts. While classical Fischer and Schrock type carbenes have been intensely studied, compounds that do not fall within one of these categories have attracted attention only recently. In addition, applications of carbene complexes rarely take advantage of redox processes, which could open up a new dimension for their use in practical processes. Herein, we report an umpolung of a nucleophilic palladium carbene complex, [{PC(sp2)P}tBuPd(PMe3)] ({PC(sp2)P}tBu = bis[2-(di-iso-propylphosphino)-4-tert-butylphenyl]methylene), realized by successive one-electron oxidations that generated a cationic carbene complex, [{PC(sp2)P}tBuPdI]+, via a carbene radical, [{PC˙(sp2)P}tBuPdI]. An EPR spectroscopic study of [{PC˙(sp2)P}tBuPdI] indicated the presence of a ligand-centered radical, also supported by the results of reactions with 9,10-dihydroanthracene and PhSSPh. The cationic carbene complex shows electrophilic behavior toward nucleophiles such as NaH, pTolNHLi, PhONa, and PMe3, resulting from an inversion of the electronic character of the Pd–Ccarbene bond in [{PC(sp2)P}tBuPd(PMe3)]. The redox induced umpolung is reversible and unprecedented.  相似文献   

12.
In spite of the great importance of the (P, V, T) data of phosphonium–based ionic liquids, only limited information on these data seems to be available in the open literature. In this work, we present the results for the density measurements of the trihexyltetradecylphosphonium chloride, [(C6H13)3P(C14H29)][Cl] and trihexyltetradecylphosphonium dicyanamide, [(C6H13)3P(C14H29)][N(CN2)] with an estimated uncertainty of ±0.5 kg · m?3. The ranges of temperature and pressure are T = (273.15 to 318.15) K and p = (0.1 to 25) MPa for [(C6H13)3P(C14H29)][Cl] and T = (273.15 to 318.15) K and p = (0.1 to 35) MPa for [(C6H13)3P(C14H29)][N(CN2)]. The high consistency of our data for [(C6H13)3P(C14H29)][Cl] compared with those measured by other authors allowed all the experimental data for this IL to be combined and correlated using the Goharshadi–Morsali–Abbaspour equation of state over a wide range of temperature and pressure. From this equation, thermomechanical coefficients as the isothermal compressibility, thermal expansivity, thermal pressure, and internal pressure were calculated for the two ILs. The Sanchez–Lacombe equation of state was used also for (P, V, T) correlation and the estimation of the free volume in these phosphonium ionic liquids. Finally ionic volumes for trihexyltetradecylphosphonium cation and several anions available in the literature made possible the calculation of the free (hole) volume.  相似文献   

13.
The synthesis, complete characterization, and solid state conformation of a new series of p-tert-butylcalix[5]arene (ButC5) mono-, di-, tri- and pentaanions are reported. X-ray structures of the alkali metal salts illustrate the strong influence of the alkali metal ion on the structure of the calixanion. The strength of the alkali metal base and its reaction stoichiometry play an important role in the conformation and level of deprotonation of the resulting anion. Reaction of ButC5 in a 2:1 molar ratio with alkali metal bases M2CO3 (M = Rb, Cs), or in a 1:1 ratio with M2CO3 (M = Na, K), MOH (M = Na, K, Rb, Cs) or MH (M = Li, Na) produces ButC5 monoanions, but ButC5 reacts in a 1:1 molar ratio with M2CO3 (M = Rb, Cs) or a 1:2 molar ratio with MOC(CH3)3 (M = Na, K) to afford ButC5 dianions. Due to the steric bulkiness of the But group no polymeric structures are observed. Alkali metal salts of trianionic ButC5 were obtained in high yields from reactions of ButC5 with MOC(CH3)3 (M = Li, Na, K), BunLi, LiH and LiOH in a 1:3 molar ratio. Pentaanionic ButC5 salts were obtained by the reaction with MOC(CH3)3 (M = Li, Na, K) or BunLi in a 5:1 ratio. X-ray crystal structures of ButC5 · Na and ButC5 · Cs indicate that the size of the alkali metal influences the level of cation-π arene interactions and therefore the conformation of the ButC5 unit; for example, ButC5 · Na has a cone conformation while ButC5 · Cs shows a flattened cone conformation. Cation-π arene interactions are observed in most of the calixarene salts.  相似文献   

14.
The heterometallic cluster complexes {(p-Cymene)Ru[S2C2(B10H10)]}Mo(CO)2{(CO)3Ru[S2C2(B10H10)]} (2) and {(p-Cymene)Ru[Se2C2(B10H10)]}2Mo(CO)2 (3) (p-Cymene = η6-4-isopropyl-toluene) have been synthesized from the reactions of 16-electron half-sandwich ruthenium 1,2-dichalcogenolate carborane complexes (p-Cymene)Ru[E2C2(B10H10)] (E = S(1a), Se(1b)) with Mo(CO)3(Py)3 in the presence of BF3 · Et2O. The complexes of 2 and 3 were characterized by elemental analysis and IR, NMR spectra. The molecular structure of 2 has been characterized by single-crystal X-ray diffraction analysis. Complex 2 is unsymmetrical and the two Ru–Mo single bonds (2.7893(14), 2.8189(13) Å) are each supported by a symmetrically bridging o-carborane-1,2-dithiolato ligand.  相似文献   

15.
The first charge-transfer (CT) complexes containing the cationic ferrocenyl donor CpFeCpCH2N+(C2H5)3 and polyoxometalate (POM) acceptors of the Lindqvist structural type [M6O19]2? (M=Mo, W) with the ratio of ferrocenyl:POM of 1:1, (Bu4N)[CpFeCpCH2N(C2H5)3][Mo6O19] (1) and (Bu4N)[CpFeCpCH2N(C2H5)3][W6O19] (2), were synthesized in high yields (67–71%) by traditional solution synthetic method, and characterized by elemental analysis, IR spectroscopy, UV–vis diffuse reflectance spectrum, luminescent spectrum and single crystal X-ray diffraction. The X-ray structure of the two novel CTCs were both solved in the monoclinic space group P21/n and show the close interaction of the hydrogen atoms of the CpFeCpCH2N+(C2H5)3 with the oxygen atoms on the surface of the POM. The UV–vis diffuse reflectance spectrum in the solid state indicates the presence of a new CT band at λmax = 576 nm and 590 nm for 1 and 2, respectively, attributed to CT transitions between the ferrocenyl donors and the POM acceptors. The luminescent spectroscopy of 1 exhibits the weakened fluorescence signals compared to that of the corresponding POM and the cationic donor, however, 2 has an intense emission at about ca. 394 nm and may be excellent candidates for potential solid-state photofunctional materials.  相似文献   

16.
The optoelectronic and nonlinear optical (NLO) properties of a soluble 2,(3)-(tetra-tert-butylphthalocyaninato)titanium(IV) oxide (tBu4PcTiO) in solutions and in the solid states have been described. The nonlinear response demonstrated that tBu4PcTiO exhibited strong RSA at 532 nm for both solution and solid-state based experiments. The decrease in the effective intensity dependent nonlinear absorption coefficient with increasing input intensities possibly results from high order triple state transitions of the excited-state population. No evidence of film fatigue or degradation was observed in the PMMA/tBu4PcTiO film, after numerous scans at varying laser intensity. The doping of tBu4PcTiO into poly[2-methoxy-5-(2′-ethylhexyloxy)-p-phenylene-vinylene] (MEH-PPV) results in the apparent increases of the open circuit voltage (Voc) and the short circuit photocurrent density under illumination with 40 mW cm−2 white-light. The light absorption of tBu4PcTiO incorporated into polymer represents the dominant contribution to the enhancement of the photocurrent. The dependence of the short circuit photocurrent in an ITO/tBu4PcTiO-doped MEH-PPV/Al cell on the incident light intensity (Iin) between 30 and 200 mW cm−2 was also investigated.  相似文献   

17.
Hypercoordinated diorganotin(IV) dichloride, [2-(Et2NCH2)C6H4]2SnCl2 (1), was prepared by reacting [2-(Et2NCH2)C6H4]Li with SnCl4. Halide-exchange reactions between 1 and the appropriate potassium halides gave [2-(Et2NCH2)C6H4]2SnX2 [X = F (2), Br (3), I (4)]. Reaction of 1 with excess of Na2S gave the cyclo-[{2-(Et2NCH2)C6H4}2SnS]2 (5). The solution behavior of the title compounds in solution was investigated by multinuclear (1H, 13C, 19F and 119Sn) NMR spectroscopy, including variable temperature studies. Single-crystal X-ray diffraction analysis revealed for all compounds intramolecular N  Sn coordination thus resulting in distorted octahedral (C,N)2SnX2 configurations. Planar chirality is observed as result of the non-planarity of the SnC3N rings; all compounds, however, crystallizing as racemates. The isomers are linked by extensive hydrogen bonds to give different supramolecular architectures in the crystals of compounds 1, 3 and 4.  相似文献   

18.
The synthesis and molecular structure of the zero-valent platinum-mono-carbene-bis-alkene complexes [Pt0(NHC)(dimethyl fumarate)2] (NHC = 1,3-dimesityl-imidazol-2-ylidene (1a); 1,3-dimesityl-dihydroimidazol-2-ylidene (2a); diphenyl-dihydroimidazol-2-ylidene (2b) are described. Two routes have been evaluated for the synthesis of 1a and 2a, involving reaction of a zero-valent platinum compound either with an isolated carbene ligand, or with an in situ generated carbene ligand. The in situ method proved to be easier and gave similar yields of about 50% after crystallization. Attempts have been made to synthesize similar compounds with N-phenyl and N-alkyl groups, of which the latter met with little success. However, (1,3-diphenyl-dihydroimidazol-2-ylidene)-bis(η2-dimethyl fumarate) platinum(0) (2b) could be obtained in 49% yield, after crystallization, from the appropriate Wanzlick dimer.Compound 1a reacts with H2 and D2 in sequences of oxidative addition, migration–insertion involving dimethyl fumarate, and reductive elimination to form neutral hydrido platinum (II) carbene complexes, probably containing a metallacyclic (R)–CO  Pt unit.  相似文献   

19.
Three palladium(II) complexes and four platinum(II) complexes having general formula CpFe{1,2-C5H3(PPh2)(CH2SR)}MCl2 (M = Pd, R = Ph, Et and tBu; M = Pt, R = Ph, Et, tBu and Cy) have been synthesized by reaction of the corresponding CpFe{1,2-C5H3(PPh2)(CH2SR)} ligands with PdCl2(CH3CN)2 or PtCl2(CH3CN)2. These complexes have been fully characterized in solution and in solid state. In all cases, monomeric square planar complexes were obtained as pure diastereoisomers.  相似文献   

20.
Catalytic cyclopropanation reactions of olefins with ethyl diazoacetate were carried out using copper(I) diphosphinoamine (PPh2)2N(R) (R = iPr, H, Ph and –CH2–C6H4–CHCH2) complexes at 40 °C in chloroform. High yields of the cyclopropanes were obtained in all cases. The rate of the reaction was influenced by the nuclearity of the complex and the binding mode of the ligand which was either bridging or chelating. Comparison of isostructural complexes shows that the rate follows the order R = iPr > H > Ph, where R is the substituent on the N. However, cyclopropane formation versus dimerization of the carbene, and trans to cis ratios of cyclopropane was similar in all cases. The nearly identical selectivity for different products formed was indicative of a common catalytic intermediate. A labile “copper–olefin” complex which does not involve the phosphine or the counterion is the most likely candidate. The differences in the reaction rates for different complexes are attributed to differences in the concentration of the catalytically active species which are in equilibrium with the catalytically inactive copper–phosphinoamine complex. To test the hypothesis a diphosphinoamine polymer complexed to copper(I) was used as a heterogeneous catalyst. Leaching of copper(I) and deactivation of the catalyst confirmed the proposed mechanism.  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