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 31 毫秒
1.
A series of heteroleptic [Ti 1 2X]? complexes have been selectively constructed from a mixture of TiIV ions, a pyridyl catechol ligand (H2 1 ; H2 1 =4‐(3‐pyridyl)catechol), and various bidentate ligands (HX) in the presence of a weak base, in addition to a previously reported [Ti 1 2(acac)]? (acac=acetylacetonate) complex. Comparative studies of these TiIV complexes revealed that [Ti 1 2(trop)]? (trop=tropolonate) is much more stable than the [Ti 1 2(acac)]? complex, which allows the replacement of acac with trop on the [Ti 1 2(acac)]? complex. This TiIV‐centered site‐selective ligand exchange reaction also takes place on a heteronuclear PdII? TiIV ring complex with the preservation of the PdII‐centered coordination structures. Intra‐ and intermolecular linking between two TiIV centers with a flexible or a rigid bis‐tropolone bridging ligand provided a tetranuclear and an octanuclear PdII? TiIV complex, respectively. These higher‐order structures could be efficiently constructed only through a stepwise synthetic route.  相似文献   

2.
Electronic Structures of Organometallic Complexes of f Elements. 60 [1] Structural, Single Crystal Optical and Magnetooptical Investigations on Trialkylphosphate Adducts of the Tris(cyclopentadienyl)lanthanide(III) (Ln = La, Pr) Moiety as well as Results of Comparing Optical Studies of [Pr(Ind)3(OP(OEt)3)] (Ind = indenyl) [Ln(Cp)3(OP(OR)3)] (Cp = η5‐cyclopentadienyl; Ln = La, R = Et ( 1 ); Ln = Pr, R = Me ( 2 ); Ln = Pr, R = Et ( 3 )) and [Pr(Ind)3(OP(OEt)3)] ( 4 ) have been synthesized and spectroscopically as well as partly structurally (only compounds 1 and 2 ) characterized. On the basis of variable temperature measurements of α absorption spectra of an oriented single crystal, the magnetic circular dichroism spectra of dissolved, and the luminescence spectra of powdered material, a nearly complete crystal field (CF) splitting pattern could be derived for 3 , and simulated by fitting the free parameters of a phenomenological Hamiltonian. The parameters used in the fit allowed the calculation of the global CF strength experienced by the Pr3+ central ion, the estimation of the nephelauxetic and relativistic nephelauxetic parameters, as well as the setup of experimentally based non‐relativistic and relativistic molecular orbital schemes in the f range. The optical spectra of compound 4 suggest that two different species exist at low temperatures, thus preventing a successful CF analysis.  相似文献   

3.
Treatment of Na[Re(CO)5] with RC?CCO2Et (R=phenyl, naphthalen‐1‐yl, phenanthren‐9‐yl and pyren‐1‐yl) followed by reaction with acetyl chloride and ethanol afforded the rhenacyclobutadienes Re{‐C(R)?C(CO2Et)C(OEt)?}(CO)4. Reactions of these rhenacyclobutadienes with HC?COEt produced rhenabenzenes Re{‐C(R)?C(CO2Et)C(OEt)?CHC(OEt)?}(CO)4. Except for R=Ph, new rhenacyclobutadienes with pendant alkenyl substituents Re{‐C(R)?C(C(OEt)?CH(CO2Et))C(OEt)?}(CO)4 were also isolated from these reactions. The NMR spectroscopic and X‐ray structural data, as well as the aromatic stabilization energy (ASE) values suggest that the rhenabenzenes are aromatic, with extensive delocalized π character.  相似文献   

4.
Oxygenation reactions of dialkylzinc solutions have been investigated. Me2Zn reacts with oxygen in the absence of water to give the bis(heterocubane) [(MeZn)6Zn(OMe)8], whereas Et2Zn and iPr2Zn afford the mono(heterocubanes) [(RZn)4(OR)4]. In the presence of small amounts of water (added during or after the oxygenation) the product types are reversed for Me2Zn and Et2Zn giving [(MeZn)4(OMe)4] and [(EtZn)6Zn(OEt)8], while being retained for iPr2Zn (giving [(iPrZn)4(OiPr)4]). Full characterization of all products by NMR spectroscopy, mass spectrometry, and elemental analyses is provided, and crystal structures of [(EtZn)6Zn(OEt)8] and [(iPrZn)4(OiPr)4] are reported. A rationalization of the different reactivities is attempted on the basis of DFT-calculated energies of some key reactants.  相似文献   

5.
Hydrido complexes [MnH(CO)3L1–3] [L1 = 1,2‐bis‐(diphenylphosphanoxy)‐ethane ( 1 ); L2 = 1,2‐bis‐(diisopropylphosphanoxy)ethane ( 2 ); L3 = 1,3‐bis‐(diphenylphosphanoxy)‐propane ( 3 )] were prepared by treating [MnH(CO)5] with the appropriate bidentate ligand by heating to reflux. Photoirradiation of a toluene solution of complexes 1 and 2 in the presence of PPhn(OR)3–n (n = 0, 1; R = Me, Et) leads to the replacement of a CO ligand by the corresponding monodentate phosphite or phosphonite ligand to give new hydrido compounds of formula [MnH(CO)2(L1–2)(L)] [L = P(OMe)3 ( 1a – 2a ); P(OEt)3 ( 1b – 2b ); PPh(OMe)2 ( 1c – 2c ); PPh(OEt)2 ( 1d – 2d )]. All complexes were characterized by IR, 1H, 13C and 31P NMR spectroscopy. In case of compounds 2 and 3 , suitable crystals for X‐ray diffraction studies were isolated.  相似文献   

6.
The structural chemistry of crystalline adducts M2(OR)8(amine)2 (M = Ti: OR = OiPr, amine = NH2CH2Ph, NH2Bu; M = Ti: OR = OEt, amine = piperidine; M = Zr: OR = OiPr, amine = NH2CH2Ph, NH2C6H11) is discussed. The adducts were obtained by reaction of Ti(OEt)4 or M(OiPr)4 with primary or secondary amines. The monoamine adducts are centrosymmetric dimers in which the amines are coordinated axially to the M2μ-OR)2 ring and hydrogen-bonded to neighboring alkoxo ligands. The adducts are sufficiently stable if the hydrogen bond is strong. 15N NMR studies revealed that the amines are also coordinated in solution. Coordination polymers were obtained when diamines with two terminal NH2 groups are reacted with Ti(OiPr)4. The general structure is the same as that of Ti2(OiPr)8(RNH2)2. However, the diamines bridge the Ti2(OiPr)8 units. When Zr(OiPr)4 was reacted with NH2CH2CH2NHMe, the adduct Zr2(OiPr)8[NH2CH2CH2NHMe]2 was obtained where only the NH2 group is coordinated.  相似文献   

7.
Reactions of cis-dialkoxy-bis(acetylacetonato)titanium(IV), [(acac)2Ti(OR)2] (R = Et, Pr i ) with alkoxyalkanols (ROCH2CH2OH) (R = Me, Et, n-Bu) in 1:1 and 1:2 molar ratios in refluxing benzene under anhydrous conditions yield [(acac)2Ti(OR)2–n (OCH2CH2OR) n ] (n = 1 or 2) complexes, which were purified by distillation under reduced pressure. On the basis of i.r. and n.m.r. (1H- and 13C-) spectral studies, a cis-octahedral environment around TiIV is proposed. On keeping the distilled dark brown-red viscous liquid [(acac)2Ti(OEt)(OCH2CH2OBu)] for 2 weeks, orange yellow crystals of [(acac)2TiO]2 were obtained. A single crystal X-ray diffraction study suggests the product is a new modification of [(acac)2TiO]2.  相似文献   

8.
Compounds Ga(OR)3 (R = Me, Et, Pri, Bun, C2H4OMe) were synthesized by exchange reactions between gallium chloride and alkali metal alkoxides, the reetherefication of Ga(OPri)3 and Ga(OC2H4OMe)3 by other ROH (R = Me, Et), and anodic dissolution of metallic gallium in the presence of a electroconductive additive (LiCl, Bu4NBr). When solid GaCl3 is introduced into an alcoholic solution of NaOEt, stable soluble gallium oxoalkoxyhalides are formed. The same reaction with a GaCl3 solution in toluene or electrochemical synthesis produces nonvolatile Ga(OEt)3 samples, which have the polymer zigzag configuration [Ga(OR)4/2(OR)]. Mass spectrometry shows that only Ga(OPri)3 and freshly prepared X-ray amorphous Ga(OEt)3 samples (produced by reetherefication) are transferred to the gas phase. The spectra of the latter contain ions generated by penta-and hexanuclear oxoalkoxide molecules, along with fragments of orthospecies [Ga(OEt)3]2?4. IR spectra are described for all compounds synthesized.  相似文献   

9.
N-methylaminoalkoxides of titanium of the type Ti(OR)4?n(O · CHR′ · CH2 · NR″R?)n where R = Et and Pr1; n = 1–4; and R′ = R″ = H, R? = Me; R′ = H, R″ = R? = Me; R′ = R″ = R? = Me, synthesized by the reactions of titanium alkoxides with aminoalcohols, show interesting variations in their properties like physical state, volatility and molecular complexity. I.r. and p.m.r. spectra of these derivatives have been recorded. A few interchange reactions with methanol and tert-butanol have also been carried out. These aminoalkoxides get cleaved with acetyl chloride and undergo insertion reactions with phenylisocyanate, thus providing the first examples of insertion reactions in such derivatives.  相似文献   

10.
The structures of three newly synthesized phosphonate‐substituted polyoxotitanates are reported. The Ti/O core of [Ti4O(OEt)12(PhenylPO3)] ( 1 ) is the building block for two larger phosphonate‐substituted nanoclusters, [Ti25O26(OEt)36(PhenylPO3)6] ( 2 ) and [Ti26O26(OEt)39(PhenylPO3)6]Br ( 3 ). All compounds exhibit a not previously recognized triply bridging binding mode of the phosphonate anchor with short connecting Ti? O bonds, the average of which is 2.010(7) Å. Comparison with previously reported work suggests that the binding mode of the phosphonate anchor is strongly dependent on the structure of the underlying substrate.  相似文献   

11.
Physicochemical analyses (solubility method, conductometry, and IR spectroscopy) revealed no complex formation in M(OR)n-Si(OR)4-ROH systems (M = Na, Ba, Al; R = Et, Pri), unlike in the systems containing alkoxides of two metals. IR and NMR spectroscopy showed that Si(OR)4 became reactive only due to microhydrolysis, which is accompanied most likely by the formation of intermediates (asymmetric molecules [Si(OH)n(OR)4 ? n ]). The hydrolysis was studied for model systems M(OEt)2 ? si(OEt)4 (M = Ba, Ca), and conditions for the synthesis of silicates were optimized.  相似文献   

12.
The insertion reaction of CS2 with Mg(NR2)2 (R= Et, iPr), MgR′2 (R′= Et, Ph) and R″MgBr (R″= iPr, Ph) respectively lead solid products, Mg(S2CNR2)2(THF)n ( 1 : R= Et, n=2; 2 : R= iPr, n=1), Mg(S2C′R)2(THF)2 ( 3 : ′R= Et, 4 : ′R= Ph), BrMg(S2C″R) (THF)3 ( 5 : ″R= iPr, 6 : ″R= Ph) in which the inserted carbon disulfides act as terminal chelating ligands. These compounds were characterized with 1H, 13C NMR, IR spectroscopy, mass spectrometry, elemental analyses, and X‐ray crystallography.  相似文献   

13.
Reaction of [Fe2(CO)9] with a half molar amount of R2PYPR2 (Y = CH2, R = Ph, Me, OMe or OPri; Y = N(Et), R = OPh, OMe or OCH2; Y = N(Me), R = OPri or OEt) leads to the ready formation of a product which on irradiation with ultraviolet light rapidly decarbonylates to the heptacarbonyl derivative [Fe2(μ-CO)(CO)6{μ-R2PYPR2}]. Treatment of the latter with a slight excess of the appropriate ligand results, under photochemical conditions, in the formation of the dinuclear pentacarbonyl complex [Fe2(μ-CO)(C))4{μ-R2PYPR2}2] but under thermal conditions in the formation of the mononuclear species [Fe(CO)3{R2PYPR2}]. Reaction of [Ru3(CO)12] with an equimolar amount of (RO)2PN(R′)P(OR)2 (R′ = Me, R = Pri or Et; R′ = Et, R = Ph or Me) under either thermal or photochemical conditions produces [Ru3(CO)10{μ-(RO)2PN(OR)2}] which reacts further with excess (RO)2PN(R′)P(OR)2 on irradiation with ultraviolet light to afford the dinuclear compound [Ru2(μ-CO)(CO4{μ-(RO)2PN(R′)P(OR)2}2]. The molecular structure of [Ru2(μ-CO)(CO)4{μ-(MeO)2PN(Et)P(OMe)2}2], which has been determined by X-ray crystallography, is described.  相似文献   

14.
New Group 3 metal complexes of the type [LM(CH2SiMe3)2(THF)n] supported by tridentate phosphido‐diphosphine ligands [(o‐C6H4PR2) 2 PH; L1‐H : R = iPr; L2‐H : R = Ph] have been synthesized by reaction of L1‐H and L2‐H with [M(CH2SiMe3)3(THF)2)] (M = Y and Sc). All the new complexes [(o‐C6H4PR2) 2 PM(CH2SiMe3)2(THF)n] [M = Y, R = iPr (1), R = Ph (2); M = Sc, R = iPr (3), R = Ph (4)] were studied as initiators for the ring opening polymerization of lactide. The yttrium complexes ( 1 and 2 ) exhibited high activity and good polymerization control, shown by the linear fits in the plot of number‐averaged molecular weight (Mn) versus the percentage conversion and versus the monomer/initiator ratio and by the low polydispersity index values. Interestingly, very good molar‐mass control was observed even when L ‐Lactide was polymerized in the absence of solvent at 130 °C. A good molar‐mass control but lower activities were observed in the polymerization reaction of lactide promoted by the analogous scandium complexes 3 and 4 . © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 48: 1374–1382, 2010  相似文献   

15.
The dinuclear [NbCln(OR)(5‐n)]2 (n = 4, R = Et, 1 ; n = 4, R = CH2Ph, 2 ; n = 3, R = Et, 3 ; n = 2, R = Et, 4 ; n = 2, R = , 5 ), and [Nb(OEt)5]2, 6 , and the mononuclear niobium compounds NbCl42? OCH2CH(R′)OR] (R = Me, R′ = H, 7 ; R = Et, R′ = H, 8 ; R = CH2Cl, R′ = H, 9 ; R = CH2CH2OMe, R′ = H, 10 ; R = R′ = Me, 11 ), NbBr42? OCH2CH2OMe], 12 , and NbCl32? OCH2CH2OMe)(κ1? OCH2CH2OMe), 13 , were tested in ethylene polymerization. Optimized reaction conditions included the use of D‐MAO as co‐catalyst and chlorobenzene as solvent at 50 °C. Complex 7 , whose X‐Ray structure is described here for the first time, exhibited the highest activity ever reported for a niobium catalyst in alkene polymerization [151 kgpolymer × molNb?1 × h?1 × bar?1]. Compounds 1 , 3‐5 , 8 , 13 showed activities similar to that of 7 . Linear polyethylenes (characterized by FT‐IR, NMR, GPC, and DSC analyses) with a broad polydispersivity were obtained. © 2011 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2011  相似文献   

16.
A series of titanium(IV) complexes Ti(O‐i‐Pr)Cl3(THF)(PhCOR) (R = H ( 1 ), CH3 ( 2 ), or Ph ( 3 )) is prepared quantitatively from reactions of [Ti(O‐i‐Pr)Cl2(THF)(μ‐Cl)]2 with 2 molar equiv. PhCOR. Treatment of Ti(O‐i‐Pr)Cl3 with 2 molar equiv. of PhCOR affords the disubstituted complexes Ti(O‐i‐Pr)Cl3(PhCOR)2 (R = CH3 ( 4 ) or Ph ( 5 )). The 13C NMR study of these complexes shows that the relative bonding abilities are in the order of PhCOCH3 > PhCHO > PhCOPh. The molecular structure of 5 reveals that one of the benzophenone ligands is trans to the strongest 2‐propoxide ligand with a long Ti‐O(carbonyl) distance of 2.193(5) Å which is much longer than the other Ti‐O(carbonyl) distance of 2.097(4) Å by ?0.1 Å. All ligands cis to the alkoxide ligand are bending away from the alkoxide ligand with the RO‐Ti‐L angles ranging from 93.6(2) to 99.0(2)°.  相似文献   

17.
New triphenylantimony(V) o‐amidophenolates (AP‐Me,Et)SbPh3 (1) and (AP‐Me,iPr)SbPh3 (2) with unsymmetrically substituted N‐aryl groups and (AP‐Et,Et)SbPh3 (3) with symmetrical N‐aryl group {AP‐R1,R2 is 4,6‐di‐tert‐butyl‐N‐[2‐alkyl(R1),6‐alkyl(R2)‐phenyl]‐o‐amidophenolate dianion} were synthesized and characterized in detail. Complexes were examined for dioxygen activity. The unsymmetrical complexes 1 and 2 were found to form different geometrical isomers (A and B) of spiroendoperoxides [L‐R1,R2(O2)]SbPh3 (4 and 5, respectively) with different dispositions of peroxide group and N‐aryl fragment (methyl and peroxide group are on the same side of the molecule in the less shielded isomer A, and on different sides in the more hindered isomer B). The isomer A prevails over isomer B, reflecting the possibility of steric control on the dioxygen‐binding reaction. Complex 3, where R1 = R2 = Et, formed the isomers 6A and 6B as 50:50. The ratio 4A:4B was 60:40 (for methyl‐ethyl containing complex 4) and it increased up to 80:20 for methyl‐isopropyl‐containing 5. The molecular structure of isomers 4A and 4B was confirmed by X‐ray analysis. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

18.
The first metal‐carbon bond β‐form paddlewheel complexes containing a Pd24+ core, [Pd(η2‐dithio)]2(μ‐dppa)( μ‐SCNMe2) (dithio = S2P(OEt)2, 2 ; S2COEt, 3 ; S2CNC4H8, 4 ), were prepared by the reactions of the α‐form paddlewheel‐type Pd2+4 dipalladium complex [Pd2 (μ‐Hdppa)2(μ‐SCNMe2)2][Cl]2, 1 with various dithio‐ligands, [NH4][S2P(OEt)2], [K][S2COEt] and [NH4][S2CNC4H8], in methanol at ambient temperature (Hdppa = bis(diphenylphosphino)amine). Electronic spectra and two X‐ray structures of the Pd2+4 species have been determined.  相似文献   

19.
Heterobimetallic alkoxides of Cu(II) of the types [Cu{η4-Ti2(OEt)9}Cl] (1) and [Cu{η3-Ti2(OR)9}2] [R = Pri (2), R = Et(3)] have been prepared for the first time by the reactions of CuCl2 · xROH with KTi2(OR)9 in 1:1 and 1:2 molar ratios, respectively, in benzene medium. The chloro(nonaalkoxo dititanato)copper(II) complexes undergo chloride replacement reactions by a variety of monodentate alkoxo (OPri, OEt) and chelating [Al(OPri) 4 , Al(OEt) 4 , Nb(OPri) 6 , Zr2(OPri) 9 , Sn2(OPri) 9 , and Sn2(OEt) 9 ] ligands to form interesting hetero(bi-and tri-)metallic complexes. Alcoholysis (with methyl alcohol and tert-butyl alcohol) and hydrolytic [with Ba(OH)2 · 8H2O powder] reactions of a few typical compounds have also been investigated. All of these have been characterized by elemental analyses, molecular weight determinations, spectral (i.r. and visible) and magnetic studies. On attempted volatilization under reduced pressure these complexes liberated titanium alkoxides as a volatile component leaving nonvolatile residues.  相似文献   

20.
The interaction between magnesium and titanium alkoxides is studied in order to chose the best precursors for synthesis of MgTiO3. No reaction between magnesium and titanium methoxides and isopropoxides occurs. The solubility diagrams for Mg(OR)2-Ti(OR)4-ROH, R = Et,-Bu at 20°C are studied. Magnesium ethoxotitanates of variable composition MgnTi4-n (OEt)16-2n2nEtOH (n=2.0-0) which are structural analogs of Ti4(OR)16 (R = Me, Et) are isolated. This is a quite unusual example of statistical distribution of heteroatoms in molecular structures of metal alkoxides. Among the systems of metal alkoxides with simple aliphatic radicals only Mg(OBu)2-Ti(OBu)4-BuOH gives a convenient precursor for the synthesis of MgTiO3. A simple scheme of preparation of magnesium titanate from the alkoxide solutions is suggested. The phase purity of MgTiO3 is to a considerable extent dependent on the hydrolysis conditions. The alkoxy-derived magnesium titanate is obtained in the form of a uniform fine powder, it can be sintered into dense ceramics in the temperature range of 1140–1220°C which is 150–200°C lower in comparison with the conventional powders.  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