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 328 毫秒
1.
Toluene solutions of M2(NMe2)6 (M = Mo, W) react with mesitylene selenol (Ar′SeH) to give M2(SeAr′) 6 complexes. MO2(OR)6 (R = tBu, CH2tBu) react with excess> 6 fold) Ar′SeH to give Mo2 (SeAr′)6, whilst W2(OR)6(py)2 (R = iPr, CH2tBu) react with excess (> 6 fold) Ar′SeH to give W2(OR)2(SeAr′)4. Reaction of MO2(OPri)6 with Ar′SeH produces Mo2(OPri)2 (SeAr′)4 which crystallizes in two different space groups. These areneselenato complexes are air-stable and insoluble in common organic solvents. X-ray crystallographic studies revealed that the Mo2(SeAr′)6 and W2(SeAr′)6 compounds are isostructural in the solid state and adopt ethane-like staggered configurations with the following important structural parameters, M---M (W---W/Mo---Mo) 2.3000(11)/2.2175(13) Å, M---Se 2.430 (av.)/2.440 (av.) Å, M---M---SE 97.0° (av.)°. In the solid state W2(OiPr)2(SeAr′)4 adopts the anti-configuration with crystallographically imposed Ci symmetry and W---W 2.3077(7) Å, W---Se 2.435 (av.) Å, W---O 1.858(6) Å; W---W---SE 100.27(3)°, 93.8(3)° and W---W---O 108.41(17)°. Mo2(OPri)2(SeAr′) 4 crystallizes in both P and A2/a space groups in which the molecules are isostructural with each other and the tungsten analogue. Important bond lengths and angles are Mo---Mo 2.180(24) Å, Mo---Se 2.432(av.) Å, Mo---O 1.872(9) Å, Mo---Mo---Se 99.39(9)°, 94.71(8)°, Mo---Mo---O 107.55(28)°.  相似文献   

2.
1,3-[2′,6′-Pyridinebis(methyleneoxy)]-1,3-bis(diphenyl)cyclodisiloxane (9) and 2,6-pyridinebis(1,1-diphenylethoxy)diphenylsilane (11) were obtained from 2,6-pyridinediol derivatives with dichlorodiphenylsilane. An N→Si interaction is present in 2,6-pyridinebis(1,1-diphenylethoxy)diphenylsilane, which also shows fluxional behavior. The activation energy of 13.2 kcal mol−1 for 11 was obtained for the intramolecular exchange between the phenyl groups from a variable-temperature 1H-NMR study. The compounds were characterized by 1H-, 13C- and 29Si-NMR and their structures were established by X-ray crystallographic studies.  相似文献   

3.
合成了一系列带有不同取代基的β-二亚胺配体及其Ni(Ⅱ)的配合物.利用核磁共振谱、元素分析和单晶X射线衍射等手段对配体及配合物进行了表征.元素分析和单晶结构分析表明,在相同的实验条件下苯基取代的β-二亚胺配体锂盐与NiCl2反应只能得到双配体化合物1;而2,6-二甲基苯基及2,6-二异丙基苯基取代的配体锂盐与NiCl2反应得二聚的单氯化物2和3,2个Ni原子通过双氯桥连接在一起.配合物2和3经烷基铝活化后催化乙烯聚合可得到高分子量聚乙烯,活性可达到2.0×105gPE/(molcat·h),分子量最高可达到100万以上.  相似文献   

4.
o-Bromo- and o-chloroaryloxyphosphines 1 may react with sodium in two competing ways: (i) metal halogen exchange followed by rapid intramolecular 1,3-rearrangement to give sodium o-hydroxylato-arylphosphines 2, later converted to their OSiMe3 derivatives 3, and (ii) reductive cleavage of the P---O bond to give diphosphines 4 or phosphides. The o-metallation is preferred with the more reactive bromides and bulky phosphino substituents or screened P---O bonds by substituents at 6-position. The reduction is favoured in the case of the less reactive aryl chlorides, small alkyl and flat phenyl substituents at phosphorus. Mixtures of meso- and rac-diphosphines are formed from asymmetric derivatives ArOPRR′. The meso-isomer of 1,2-di(tert-butyl)-1,2-diphenyldiphosphine is preferred.  相似文献   

5.
Anilines with alkyl substituents on the phenyl ring (ArNH2 = 2,4,6-trimethylaniline; 2,3-, 2,4-, 2,6-, and 3,4-dimethylaniline; and 2,6-diisopropylaniline) react with MoO(X)2(dtc)2 (X = Cl or Br; dtc = diethyldithiocarbamate) in methanol in the presence of 2 equiv of triethylamine to form ionic imido complexes of the type [MoNAr(dtc)3]2[Mo6O19] or MoNAr(dtc)3]4[Mo8O26]. The same reaction in THF with butyllithium as base yields imido complexes of the type MoNAr(X)2(dtc)2. The structures of three ionic, five chloro, and two bromo complexes have been determined by X-ray crystallography. In all complexes, the molybenum center is a distorted pentagonal bipyramid. While the structures are similar, the angles of the imido linkages differ. The effect of the substituents on the phenyl ring of the imido ligand on the 95Mo NMR chemical shifts was determined. The Mo nucleus becomes more deshielded with the substituents in the following order: 3,4-Me2 < 2,3-Me2 < 2,4-Me2 < 2,6-Me2 < 2,4,6-Me3 < 2,6 isopropyl. Complexes with more deshielded 95Mo centers tend to have angles of the imido linkage that are closer to 180 degrees.  相似文献   

6.
The reactions of RNHSi(Me)2Cl (1, R=t-Bu; 2, R=2,6-(Me2CH)2C6H3) with the carborane ligands, nido-1-Na(C4H8O)-2,3-(SiMe3)2-2,3-C2B4H5 (3) and Li[closo-1-R′-1,2-C2B10H10] (4), produced two kinds of neutral ligand precursors, nido-5-[Si(Me)2N(H)R]-2,3-(SiMe3)2-2,3-C2B4H5, (5, R=t-Bu) and closo-1-R′-2-[Si(Me)2N(H)R]-1,2-C2B10H10 (6, R=t-Bu, R′=Ph; 7, R=2,6-(Me2CH)2C6H3, R′=H), in 85, 92, and 95% yields, respectively. Treatment of closo-2-[Si(Me)2NH(2,6-(Me2CH)2C6H3)]-1,2-C2B10H11 (7) with three equivalents of freshly cut sodium metal in the presence of naphthalene produced the corresponding cage-opened sodium salt of the “carbons apart” carborane trianion, [nido-3-{Si(Me)2N(2,6-(Me2CH)2C6H3)}-1,3-C2B10H11]3− (8) in almost quantitative yield. The reaction of the trianion, 8, with anhydrous MCl4 (M=Ti and Zr) in 1:1 molar ratio in dry tetrahydrofuran (THF) at −78 °C, resulted in the formation of the corresponding half-sandwich neutral d0-metallacarborane, closo-1-M[(Cl)(THF)n]-2-[1′-η1σ-N(2,6-(Me2CH)2C6H3)(Me)2Si]-2,4-η6-C2B10H11 (M=Ti (9), n=0; M=Zr (10), n=1) in 47 and 36% yields, respectively. All compounds were characterized by elemental analysis, 1H-, 11B-, and 13C-NMR spectra and IR spectra. The carborane ligand, 7, was also characterized by single crystal X-ray diffraction. Compound 7 crystallizes in the monoclinic space group P21/c with a=8.2357(19) Å, b=28.686(7) Å, c=9.921(2) Å; β=93.482(4)°; V=2339.5(9) Å3, and Z=4. The final refinements of 7 converged at R=0.0736; wR=0.1494; GOF=1.372 for observed reflections.  相似文献   

7.
The photophysical properties, which vary as R is varied, of a series of [Pt(N2O2)] complexes bearing bis(phenoxy)bipyridine auxiliaries with different substituents R=H (Pt-H) (1), 4,4′-2NH2 (Pt-NH2) (2), 4,4′-2tBu (Pt-tBu) (3), 4,4′-2CN (Pt-CN) (4), and 4,4′-2NO2 (Pt-NO2) (5) are investigated using density functional theory (DFT) and time-dependent density functional theory (TDDFT). The solvent effects are discussed in CH2Cl2, CH3CN and CH3OH solutions, respectively, by polarizable continuum model (PCM). It is anticipated that compared with σ-donor substituents, π-acceptors have more dramatic effects on the electronic and optical properties in this series of complexes. Introduction of π-electron withdrawing substituents on bipyridine ligand will benefit the LLCT (or MLCT) and prohibit the non-radiative pathways via d–d transitions by increasing the energy gap between the HOMO–LUMO and d–d transitions. The results also reveal that the lowest-energy excitations of all complexes show blue-shifts in the polarized solution and when the polarity of the solvent increases from CH2Cl2, CH3CN and CH3OH, the low-energy broad absorption band exhibit blue-shifts. The lowest-energy excitations and photoluminescence of all complexes are dominated by π(phenoxy)→π*(bpy/NO2) (LLCT) excited state mixed with some energetically dπ (Pt)→π*(bpy/NO2) (MLCT) transition.  相似文献   

8.
The oxidation of 2,6-di-tert-butylphenol with tert-butylhydroperoxide (ButO2H) has been studied using polymer (XAD4) anchored salicylaldoxime, 1,3-propylene-bis-salicylaldimine and o-phenylene-bis-salicylaldimine complexes of molybdenum and vanadium in acetonitrile. The predominant products formed in the oxidation reactions were 2,6-di-tert-butylbenzoquinone (BQ) and 3,3′-5,5′-tetra-tert-butyldiphenoquinone (dPQ), whereas with some only 2,6-di-tert-butylbenzoquinone was formed. This is the first reported use of polymer anchored molybdenyl and vanadyl complexes in selective oxidation of 2,6-di-tert-butylphenol. Solvent plays an important role in this reaction. The effects of varying the ligand, metal and the support on the catalytic activity in the oxidation of 2,6-di-tert-butylphenol have been studied. With polymer anchored MoO2(salpen), 81% of 2,6-di-tert-butylbenzoquinone was formed from 2,6-di-tert-butylphenol.  相似文献   

9.
Diaryl methane molecules (Ar–CH2–Ar) represent double rotor conformational problems. The simplest diaryl methane, diphenyl methane (Ph–CH2–Ph), governs certain symmetric conformational potential energy surface (PES) topology. With the replacement of one of the phenyl groups by a heterocyclic moiety, the PES topology may change dramatically. The induction of point-chirality, in the prochiral CH2 group, by axis-chirality or plane-chirality is explored within the framework of ‘dynamic chirality’.  相似文献   

10.
The syntheses of the 1,3,5-trimethyl- and tri-tert-butyl-1,3,5-triazacyclohexane-supported imido complexes [M(NR)(R′3tach)Cl2] (M = Ti or Zr (NMR only); R = But or 2,6-C6H3Pri2; R′ = Me or But) are reported, along with that of the thermally robust dibenzyl derivative [Ti(NBut)(Me3tach)(CH2Ph)2]. The tert-butylimido ligand in [Ti(NBut)(Me3tach)Cl2] undergoes exchange with ArNH2 (Ar = 4-C6H4Me or 2,6-C6H4Me or 2,6-C6H3Pri2) to form the corresponding arylimides [Ti(NAr)(Me3tach)Cl2]. The Me3tach ring in [Ti(NR)(Me3tach)Cl2] undergoes slow exchange with But3tach or Me3tacn (1,4,7-trimethyl-1,4,7-triazacyclononane) to give the ring-exchanged products [Ti(NR)(But3tach)Cl2] and [Ti(NR)(Me3tacn)Cl2], respectively. The complexes [Ti(NR)(Me3tach)X2] (R = But or 2,6-C6H3Pri2; X = Cl or CH2Ph) exhibit room-temperature dynamic NMR behaviour via an unusual trigonal twist of the facially coordinated Me3tach ligand, and the activation parameters for these processes have been measured and are discussed. The X-ray structures of [Ti(NR)(But3tach)Cl2] (R = But or 2,6-C6H3Pri2) and [Ti(NBut)(Me3tach)(X)2] [X= Cl or CH2Ph) are reported. Me3tach and But3tach = 1,3,5-trimethyl- and tri-tert-butyl-1,3,5-triazacyclohexane, respectively.  相似文献   

11.
Arylchlorogermylenes stabilized by two ortho side-chain donor ligands (aryl=2,6-(CH2NR2)2C6H3 with R=Et: 1 and R=i-Pr: 2) were synthesized; 1 was characterized by an X-ray analysis and intramolecular coordinations N→Ge were evidenced from the two side-chains. In this paper, all known stable and isolated halogermylenes are reviewed. Significant bond lengths of their X-ray structures are reported and compared. Surprisingly, attempts to obtain chlorogermylenes bearing a substituent with a strong electron-withdrawing effect such as 2,4,6-tris(trifluoromethyl)phenyl (Ar′) or 2,6-bis(trifluoromethyl)phenyl (Ar″) led us to new compounds, the diarylchlorogermanes Ar′2GeHCl or Ar″2GeHCl.  相似文献   

12.
The reactions of M(CO)4(R′-DAB) (M = Mo) or W; R′-DAB = R′-N=CHCH=NR′ (R′ = i-propyl, t-butyl, or cyclohexyl) with SnCl4 in dichloromethane solution result in the formation, in high yield, of the orange, diamagnetic, seven-coordinate oxidative-addition products M(CO)3(R′-DAB)(SnCl3)Cl. The reactions of Mo(CO)3(R′-DAB)(SnCl3)Cl (R′ = i-Pr or Cy) with an excess of alkyl isocyanide RNC (R = CHMe2, CMe3, or C6H11) in the presence of KPF6 lead to the formation of [Mo(CNR)4(R′-DAB)Cl]PF6 or [Mo(CNR)5(R′-DAB)](PF6)2 depending upon the reaction stoichiometry and reaction conditions. The monocationic chloro species are converted to [Mo(CNR)5(R′-DAB)](PF6)2 upon reflux with the stoichiometric amount of RNC. Under similar reactions conditions M(CO)3(t-Bu-DAB)(SnCl3)Cl (M = Mo or W) derivatives react with alkyl isocyanides with the reductive-elimination of the elements of SnCl4 and the formation of octahedral M(CO)3(CNR)(t-Bu-DAB). The dark red compounds [Mo(CNCMe3)5(R′-DAB)](PF6)2 (R′ = i-Pr or Cy) react readily with cyanide ions at ambient temperatures in methanol to yield [Mo(CNCMe3)4(R′-DAB)(CN)]PF6. Attempts to thermally dealkylate the parent complexes [Mo(CNCMe3)5(R′-DAB)](PF6)2 (R′ = i-Pr or Cy) to these same cyano species were unsuccessful.  相似文献   

13.
Structures of the following compounds have been obtained: N-(2-pyridyl)-N′-2-thiomethoxyphenylthiourea, PyTu2SMe, monoclinic, P21/c, a=11.905(3), b=4.7660(8), c=23,532(6) Å, β=95.993(8)°, V=1327.9(5) Å3 and Z=4; N-2-(3-picolyl)-N′-2-thiomethoxyphenyl-thiourea, 3PicTu2SeMe, monoclinic, C2/c, a=22.870(5), b=7.564(1), c=16.941(4) Å, β=98.300(6)°, V=2899.9(9) Å3 and Z=8; N-2-(4-picolyl)-N′-2-thiomethoxyphenylthiourea, 4PicTu2SMe, monoclinic P21/a, a=9.44(5), b=18.18(7), c=8.376(12) Å, β=91.62(5)°, V=1437(1) Å3 and Z=4; N-2-(5-picolyl)-N′-2-thiomethoxyphenylthiourea, 5PicTu2SMe, monoclinic, C2/c, a=21.807(2), b=7.5940(9), c=17.500(2) Å, β=93.267(6)°, V=2893.3(5) Å3 and Z=8; N-2-(6-picolyl)-N′-2-thiomethoxyphenylthiourea, 6PicTu2SMe, monoclinic, P21/c, a=8.499(4), b=7.819(2), c=22.291(8) Å, β=90.73(3)°, V=1481.2(9) Å3 and Z=4 and N-2-(4,6-lutidyl)-N′-2-thiomethoxyphenyl-thiourea, 4,6LutTu2SMe, monoclinic, P21/c, a=11.621(1), b=9.324(1), c=14.604(1) Å, β=96.378(4)°, V=1572.4(2) Å3 and Z=4. Comparisons with other N-2-pyridyl-N′-arylthioureas having substituents in the 2-position of the aryl ring are included.  相似文献   

14.
The course of the intramolecular meta photocycloaddition of ring-substituted (E)- and (Z)- 6-phenylhex-2-enes depends on the position and nature of the substituent. In this paper, the effects of methyl and cyano groups and the fluorine atom are described. The results are in agreement with a reaction mechanism in which the excited phenyl ring becomes polarized when it is approached by the alkene. The dipolar character becomes particularly apparent in the case of fluorine which, depending on its position, can stabilize either the negative charge through its inductive effect or the positive charge through its mesomeric effect. The configuration of the terminal methyl group sterically influences the photoreaction. The Z alkenes readily undergo the 1′,3′-addition, but fail to add in the 2′,6′-mode, even if substituents which strongly promote this mode are present. The E alkenes seem to suffer from steric hindrance in the 2′,6′- and the 1′,3′-mode, but the electronic effects of activating substituents provide compensation and substituents may have a pronounced influence on the ratio of the two modes.  相似文献   

15.
The complexes (Hal)Nb(CO)3(PR3)3 (PR3 = PEt3, Hal = I; PR3 = PMe2Ph, Hal = Cl, Br, I) and (Hal)Nb(CO)4/2(dppe)1/2 (Hal = Br, I) have been prepared by oxidative halogenation of carbonylniobate with pyridinium halides (Hal = Cl, Br) or iodine (Hal = I). In the tricarbonyls, one CO and one PR3 are labile and can be displaced by a four-electron donating alkyne to give all-trans-[(Hal)Nb(CO)2(RCCR′)(PR3)2] (PR3 = PMe2Ph; Hal = Cl, Br, I: R, R′ = H, Et, Ph; R = H, R′ = Ph. PR3 = PEt3, Hal = I: R, R′ = Pr; R = H, R′ = Bu, Ph; R = Me, R′ = Et). In the case of acetylene, INb(CO)(HCCH)2(PEt3)2 is also formed. PR3 can be displaced by P(OMe) 3. In the tetracarbonyls, two CO ligands are replaced by two isonitriles to form INb(CO)2(CNR)2dppe (R = tBu, Cy), or by one alkyne to form (Hal)Nb(CO)2(PhCCPh)dppe (Hal = Br, I). In these complexes, the remaining CO ligands occupy cis positions. The structure of BrNb(CO)2(dppe)2·THF, INb(CO)2(dppe)2·hexane and INb(CO)2(PEt3)2(MeCCEt) have been determined by a single crystal X-ray diffraction study. The alkyne complexes are best regarded as octahedral with the centre of the alkyne ligand occupying the positions trans to the halide and the CC axis aligned with the OC---Nb---CO axis. The complexes (Hal)Nb(CO)2(dppe)2 adopt a trigonal prismatic structure with the halide capping the tetragonal face spanned by the four phosphorus functions. The crystal structure of a by-product, Br2Nb(CO)(H2CPhPCH2CH2PPh2)2·1/2THF has also been determined. The geometry is pentagonal bipyramidal, with one of the bromine atoms and the CO on the axis. Some 93 Nb NMR data for the NbI complexes are presented, and preliminary observations on the reactions between the π-alkyne complexes and H2 or H are reported.  相似文献   

16.
The isocratic normal-phase high-performance liquid chromatography of a series of triphenylphosphine (PPh3)-substituted homo- and hetero-dinuclear metal carbonyl complexes [MM′ (CO)10−n(PPh3)n, where M,M′ = Mn, Re; N = 1,2] is reported. A column packed with silica bonded with phenyl groups was used after preliminary experiments showed that columns packed with conventional silica, and with silica bonded with amion-cyano groups were unsatisfactory for separation. The mobile phases used were hexane-toluene (8:2) and hexane-dichloromethane (90:10). The results suggest that besides the symmetry-imposed polarity of the complexes, the nature of the metal and substituent ligand also determine their retention characteristics.  相似文献   

17.
The radical copolymerization of (2,6-diphenyl) phenyl methacrylate (1) with methyl methacrylate in DMF with AIBN at 70°C has the reactivity ratios r1 = 0.071 and r2 = 1.42, from which Q1 = 1.45 and e1 = 1.20. The copolymers had Mns in the range of 10,000–40,000 and Tgs ranging from 406 to 480 K from which the hypothetical Tg for poly-1 was deduced as 500 K (227°C). Unlike 1, (2,6-diphenyl) phenyl acrylate could be polymerized to oligomers with Mn of the order of 2500.  相似文献   

18.
The new chloro(cyclopentadienyl)silanes Cp′SiHyCl3−y (Cp′=Me4EtC5, y=1: 1; Cp′=Me4C5H, y=1: 2; y=0: 3; Cp′=Me3C5H2, y=1: 4 and pentachloro(cyclopentadienyl)disilanes Cp′Si2Cl5 (Cp′=Me5C5 5, Me4EtC5 6, Me4C5H 7, Me3C5H2 8, Me3SiC5H4 9) are synthesized in good yields via metathesis reactions. Treatment of 1–9 with LiAlH4 leads under Cl–H exchange to the hydridosilyl compounds Cp′SiH3 (Cp′=Me4EtC5 10, Me4C5H 11, Me3C5H2 12) and to the hydridodisilanyl compounds Cp′Si2H5 (Cp′=Me5C5 13, Me4EtC5 14, Me4C5H 15, Me3C5H2 16, Me3SiC5H4 17). Complexes 1–17 are characterized by 1H, 13C, and 29Si-NMR spectroscopy, IR spectroscopy, mass spectrometry and CH-analysis. The structures of 6, 7 and 9 are determined by single-crystal X-ray diffraction analysis. Pyrolysis studies of the cyclopentadienylsilanes 10–12 and disilanes 13–17 show their suitability as precursors in the MOCVD process.  相似文献   

19.
A series of thiapyrylium pentamethine dyes (4 and 12-15) bearing 2,2'-di-tert-butyl-6,6'-diphenyl, 2,2'-di-tert-butyl-6,6'-bis(2,6-dimethylphenyl), 2,2'-di-tert-butyl-6,6'-bis(2-methylphenyl), 2,2',6,6'-tetrakis(2,6-dimethylphenyl), and 2,2',6,6'-tetrakis(2-methylphenyl) substituents, respectively, were prepared and their linear optical properties and electrochemical redox properties were measured and compared to thiapyrylium pentamethine dyes 3 and 5. The tert-butyl and 2,6-dimethylphenyl substituents give nearly identical chromophores with respect to values of lambda(max), molar extinction coefficients (epsilon), bandwidths at half-height (nu(1/2)), and lack of absorption in the visible spectrum. The 2-methylphenyl substituent imparts linear optical properties that are intermediate between those of the tert-butyl and phenyl substituents. The 2,6-dimethylphenyl and 2-methylphenyl substituents impart greater oxidative stability based on anodic shifts in oxidation potential.  相似文献   

20.
Substitution of alkyl groups on the ortho-position of 3-phenylsydnone causes a steric hindrance in coplanarity of the sydnone and phenyl rings. This was proved from the shift of the polarographic half-wave potentials (in excess of the polar effects), from the ultra-violet spectra, and from scale models. The behaviour of 3-o-tolylsydnone resembles more that of 3-benzylsydnone than that of 3-phenylsydnone. In 3′,4′-dihydroquinolino[1′,2′-c]-sydnone, the —CH2 CH2—bridge brings the sydnone and phenyl rings into a nearly coplanar position, shown on scale models, and its polarographic and spectrophotometric behaviour resembles that of 3-phenylsydnone.  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