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.
Five substituted cyclopentadienyl titanium trimethoxide complexes, RCpTi(OMe)3 (R=Me (2b), iPr (2c), Me3Si (2d), allyl (2e), PhCH2 (2f)), were prepared. By reacting RCpTi(OMe)3 with BF3OMe2, six RCpTiF2(OMe) (R=H (3a), Me (3b), iPr (3c), Me3Si (3d), allyl (3e), PhCH2 (3f)) were obtained. When activated with methylaluminoxane (MAO), the activities of RCpTiF2(OMe) system were less than those of RCpTi(OMe)3 system in solution polymerization of styrene, but the polymers made by RCpTiF2(OMe) exhibited higher Mw and melting point than those by RCpTi(OMe)3. Both systems produced polymers with similar syndiotacticities in the range 92.4-97.6%. Introduction of a substituent group into the Cp-ligand enhanced the melting points of the polymers, and meanwhile decreased the catalytic activities of RCpTi(OMe)3/MAO and RCpTiF2(OMe)/MAO systems, where the order of activity was RCp=Cp > MeCp > iPrCp > Me3SiCp > CH2CHCH2Cp > PhCH2Cp. Complexes 2a (CpTi(OMe)3) and 3a showed the highest activities respectively for both systems, and are three to four times more active than CpTiCl3. In bulk polymerization, the difference of activities between RCpTi(OMe)3/MAO and RCpTiF2(OMe)/MAO systems became small, where complexes 2e and 3e exhibited remarkably higher activities compared with their solution polymerization activities. The maximum polymerization activities were found at the polymerization temperature of 50 °C for most of the complexes. The influence of the polymerization time (tP), polymerization temperature (TP) and Al/Ti ratio on the activities of complexes 2b and 3b were investigated. It was observed that the initial rate of propagation of complex 2b was higher than that of complex 3b and the highest activities of both catalysts were reached at the relatively low Al/Ti ratio of 150 and decrease for larger ratios.  相似文献   

2.
New half-titanocenes, CpTiCl[(OCR2CH2)NMe(CH2CR2O)] [R,R′ = H (1), R,R′ = Me, H, (2), R,R′ = Me (3)], were prepared from CpTiCl3 (4) with the corresponding alcohols in the presence of triethylamine. X-ray analysis shows that 1 has slightly distorted trigonal bipyramidal geometry around Ti. These complexes exhibited moderate catalytic activities for syndiospecific styrene polymerization in the presence of MAO and the activity increased in the order: 2 > 1 > 4 > 3 (at 50 °C), 1 > 2 > 4 > 3 (at 70 °C and 90 °C).  相似文献   

3.
Reactions of [Ti(OPri)4] with various oximes, in anhydrous refluxing benzene yielded complexes of the type [Ti{OPri}4−n{L}n], where, n = 1-4 and LH = (CH3)2CNOH (1-4), C9H16CNOH (5-8) and C9H18CNOH (9-12). The compounds were characterized by elemental analyses, molecular weight measurements, FAB-mass, FT-IR and NMR (1H, 13C{1H}) spectral studies. The FAB-mass spectra of mono- (1), and di- (2), (6), (10) substituted products indicate their dimeric nature and that of tri- (3) and tetra- (4), (8) substituted derivatives suggest their monomeric nature. Crystal and molecular structure of [Ti{ONC10H16}4·2CH2Cl2] (8A) suggests that the oximato ligands bind the metal in a dihapto η2-(N, O) manner, leading to the formation of an eight coordinated species. Thermogravimetric curves of (3), (6) and (10) exhibit multi-step decomposition with the formation of TiO2 as the final product in each case, at 900 °C. Low temperature (∼600 °C) sol-gel transformations of (2), (3), (4), (6), (7) and (8) yielded nano-sized titania (a), (b), (c), (d), (e) and (f), respectively. Formation of anatase phase in all the titania samples was confirmed by powder XRD patterns, FT-IR and Raman spectroscopy. SEM images of (a), (b), (c), (d), (e) and (f) exhibit formation of nano-grains with agglomer like surface morphologies. Compositions of all the titania samples were investigated by EDX analyses. The absorption spectra of the two representative samples, (a) and (f) indicate an energy band gap of 3.17 eV and 3.75 eV, respectively.  相似文献   

4.
The hydrosulfido complexes CpRu(L)(L′)SH react with one equivalent of O-alkyl oxalyl chlorides (ROCOCOCl) to form the corresponding O-alkylthiooxalate complexes CpRu(L)(L′)SCOCO2R (L = L′ = PPh3 (1), (2); L = PPh3, L′ = CO (3); R = Me (a), Et (b)). The reactions of the hydrosulfido complexes with half equivalent of oxalyl chloride produce the bimetallic complexes [CpRu(L)(L′)SCO]2 (L = L′ = PPh3 (4), (5); L = PPh3, L′ = CO (6)). The crystal structures of CpRu(PPh3)2SCOCO2Me (1a) and CpRu(dppe)SCOCO2Et (2b) are reported.  相似文献   

5.
Optically active ligands of type Ph2PNHR (R = (R)-CHCH3Ph, (a); (R)-CHCH3Cy, (b); (R)-CHCH3Naph, (c)) and PhP(NHR)2 (R = (R)-CHCH3Ph, (d); (R)-CHCH3Cy, (e)) with a stereogenic carbon atom in the R substituent were synthesized. Reaction with [PdCl2(COD)2] produced [PdCl2P2] (1) (P = PhP(NHCHCH3Ph)2), whose molecular structure determined by X-ray diffraction showed cis disposition for the ligands. All nitrogen atoms of amino groups adopted S configuration. The new ligands reacted with allylic dimeric palladium compound [Pd(η3-2-methylallyl)Cl]2 to gave neutral aminophosphine complexes [Pd(η3-2-methylallyl)ClP] (2a-2e) or cationic aminophosphine complexes [Pd(η3-2-methylallyl)P2]BF4 (3a-3e) in the presence of the stoichiometric amount of AgBF4. Cationic complexes [Pd(η43-2-methylallyl)(NCCH3)P]BF4 (4a-4e) were prepared in solution to be used as precursors in the catalytic hydrovinylation of styrene. 31P NMR spectroscopy showed the existence of an equilibrium between the expected cationic mixed complexes 4, the symmetrical cationic complexes [Pd(η3-2-methylallyl)P2]BF4 (3) and [Pd(η3-2-methylallyl)(NCCH3)2]BF4 (5) coming from the symmetrization reaction. The extension of the process was studied with the aminophosphines (a-e) as well as with nonchiral monodentate phosphines (PCy3 (f), PBn3 (g), PPh3 (h), PMe2Ph (i)) showing a good match between the extension of the symmetrization and the size of the phosphine ligand. We studied the influence of such equilibria in the hydrovinylation of styrene because the behaviour of catalytic precursors can be modified substantially when prepared ‘in situ’. While compounds 3 and bisacetonitrile complex 5 were not active as catalysts, the [Pd(η3-2-methylallyl)(η2-styrene)2]+ species formed in the absence of acetonitrile showed some activity in the formation of codimers and dimers. Hydrovinylation reaction between styrene and ethylene was tested using catalytic precursors solutions of [Pd(η3-2-methylallyl)LP]BF4 ionic species (L = CH3CN or styrene) showing moderate activity and good selectivity. Better activities but lower selectivities were found when L = styrene. Only in the case of the precursor containing Ph2PNHCHCH3Ph (a) ligand was some enantiodiscrimination (10%) found.  相似文献   

6.
Reactions between [Fe(η-C5H5)(MeCO)(CO)(L)], L = PPh3 (1), PMe3 (2), PPhMe2 (3), PCy3 (4), CO (5), and B(C6F5)3 give new complexes [Fe(η-C5H5){MeCOB(C6F5)3}(CO)(L)] L = PPh3 (7), PMe3 (8), PPhMe2 (9), PCy3 (10), CO (11), where B(C6F5)3 coordinates selectively to the O-acyl groups. Hydrolysis of 7 gives [Fe(η-C5H5){HOB(C6F5)3}(CO)(PPh3)] (6). The X-ray structures of 6, 8 and 11 have been determined. Calculations, using density functional theory, demonstrate that the charge transfer to the acyl group on Lewis acid coordination is more significant in the σ than the π system. Both effects lead to a lengthening of the acyl C-O bond thus π populations cannot be inferred from the distance changes.  相似文献   

7.
The reaction of a precatalyst, [Cp∗Rh(bpy)(H2O)](OTf)2 (1), with sodium formate provided the hydride complex, [Cp∗Rh(bpy)(H)]+ (2), in situ, at pH 7.0, which was then evaluated in an aqueous, catalytic hydride transfer process with water soluble substrates that encompass 2-pentanone (3), cyclohexanone (4), acetophenone (5), propionaldehyde (6), benzaldehyde (7), and p-methoxybenzaldehyde (8). The initial rates, ri, of appearance of the reduction product alcohols at 23 °C provided a relative rate scale: 8 > 7 ≈ 6 > 5 > 4 > 3, while the effect of concentration of substrate, precatalyst, and sodium formate on ri, using 7 as an example, implicates [Cp∗Rh(bpy)(H)]+ formation as the rate-limiting step. The experimental kinetic rate expression was found to be: d[alcohol]/dt = kcat[1][HCO2Na]; substrate being pseudo zero order in water. The steric effects were also analyzed and appeared to be of less importance intra both the ketone and aldehyde series, but an inter series comparison appeared to show that the aldehydes had less of a steric effect on the initial rate, i.e., 7 > 4 by a factor of 3.6, while the aldehyde series appeared to have some moderate electronic influence on rates, presumably via electron donation to increase binding to the Cp∗Rh metal ion center, in accordance with these proposed concerted binding/hydride transfer reactions. A proposed catalytic cycle will also be presented.  相似文献   

8.
A series of mononuclear ruthenium complexes [RuCl(CO)(PMe3)3(CHCH-C6H4-R-p)] (R = H (2a), CH3 (2b), OCH3 (2c), NO2 (2d), NH2 (2e), NMe2 (2f)) has been prepared. The respective products have been characterized by elemental analyses, NMR spectrometry, and UV-Vis spectrophotometry. The structures of complexes 2c and 2d have been established by X-ray crystallography. Electrochemical studies have revealed that electron-releasing substituents facilitate monometallic ruthenium complex oxidation, and the substituent parameter values (σ) show a strong linear correlation with the anodic half-wave or oxidation peak potentials of the complexes.  相似文献   

9.
Ligand effects on the catalytic activity [and norbornene (NBE) incorporation] for both ethylene polymerization and ethylene/NBE copolymerization using half-titanocenes (titanium half-sandwich complexes) containing ketimide ligand of type Cp′TiCl2[NC(R1)R2] [Cp′ = Cp (1), C5Me5 (Cp, 2); R1,R2 = tBu,tBu (a), tBu,Ph (b), Ph,Ph (c)]-methylaluminoxane (MAO) catalyst systems have been investigated. CpTiCl2[NC(tBu)Ph] (1b) CpTiCl2(NCPh2) (1c), and CpTiCl2(NCPh2) (2c) were prepared and identified; the structure of CpTiCl2(NCPh2) (2c) was determined by X-ray crystallography. The catalytic activity for ethylene polymerization increased in the order: 1a > 1b > 1c, suggesting that an electronic nature of the ketimide ligand affects the activity. However, molecular weight distributions for resultant (co)polymers prepared by 1b,c and by 2c-MAO catalyst systems were bi- or multi-modal, suggesting that the ketimide substituent plays a key role in order for these (co)polymerizations to proceed with single catalytically-active species. CpTiCl2(NCtBu2) (1a) exhibited both remarkable catalytic activity and efficient NBE incorporation for ethylene/NBE copolymerization.  相似文献   

10.
Mono-demethylation of Cp2Ti(CH3)2 in dichloromethane with 1 M equivalent of [η5-(C5H4COOH)]Cr(CO)2NO (5), [η5-(C5H4COOH)]Cr(NO)2X] (X = Cl 6, X = I 7) and [η5-(C5H4COOH)]W(CO)3CH3 (8) gives Cp2Ti(CH3){[OC(O)C5H4]Cr(CO)2NO} (9), Cp2Ti(CH3){[OC(O)C5H4]Cr(NO)2Cl} (10), Cp2Ti(CH3){[OC(O)C5H4]Cr(NO)2I} (11) and Cp2Ti(CH3){[OC(O)C5H4]W(CO)3CH3} (12), respectively. The structure of 10 has been solved by X-ray diffraction studies. One of the nitrosyl groups is located at the site away from the exocyclic carbonyl carbon of the Cp(Cr) ring with twist angle of 178.1°. All the data reveals that Cp2Ti(CH3)- is a strong electron-donating group. The opposite correlation was observed on the chemical shift assignments of C(2)-C(5) in compounds 5-12, using HetCOR NMR spectroscopy, as compared with the NMR data of their ferrocene analogues. The electron density distribution in the cyclopentadienyl ring is discussed on the basis of 13C NMR data and those of 10 are compared with the calculations via density functional B3LYP correlation- exchange method.  相似文献   

11.
Syntheses, characterizations, electrochemistry and catalytic properties for styrene epoxidation of three manganese(III) compounds [MnIIIL1(H2O)(MeOH)](ClO4) (1) [MnIIIL1(N3)(H2O)]·dmf (2) [MnIIIL1(Cl)(H2O)] (3) derived from the Schiff base compartmental ligand N,N′-o-phenylenebis(3-ethoxysalicylaldimine) (H2L1) are reported. The three compounds are characterized by elemental analyses, IR, mass and UV–Vis spectra and conductance values. Single crystal X-ray structures of 1 and 2 have been determined. The structures of 1 and 2 show that these are mononuclear compounds having a salen type structure. In both structures, a dinuclear species is formed by bifurcated hydrogen bonding involving coordinated water molecule. The coordination of chloride in 3 is shown by conductance measurements. The compounds have also been characterized by UV–Vis and mass spectroscopic studies. Cyclic voltammetric and square wave voltammetric studies of the three compounds reveal that these undergo Mn(III)/Mn(II) reduction reversibly with the order of the ease of reduction as 3 > 2 > 1. This order has been explained proposing the composition of active species in solution. Catalytic properties for epoxidation of styrene by all the three complexes using PhIO and NaOCl as oxidant have been studied. The order of both the styrene conversion and styrene epoxidation using the three title compounds is 3 > 1 > 2. Again, it has been observed that more efficient conversion and epoxidation take place when PhIO is used as oxidant.  相似文献   

12.
The controlled/living cationic polymerization of styrene using R-OH/BF3OEt2 (R-OH = 1-phenylethanol (1), 2-phenyl-2-propanol (2) and 1-(4-methoxyphenyl)ethanol (3)) at 0 °C in CH2Cl2 and in the presence of water was investigated. With 1/BF3OEt2, the poor control over molecular weight and molecular weight distribution was ascribed to a competitive protonic initiation induced by water. The molecular weight of the polymers obtained with 2/BF3OEt2 and 3/BF3OEt2 at low water content ([H2O] ? 0.11 M) increased in direct proportion to the monomer conversion in agreement with the calculated values, assuming that one initiator molecule generates one polymer chain, but the molecular weight distribution was found relatively broad (Mw/Mn ∼ 1.8). 1H NMR analyses confirmed that polymerization proceeds via reversible activation of C-OH terminus, but some loss of hydroxyl functionality was revealed. Some trials using high water contents in the recipe ([H2O] ? 1.6 M) produced only traces of polymer due to catalyst decomposition.  相似文献   

13.
A series of novel octahedral nickel(II) dithiocarbamate complexes involving bidentate nitrogen-donor ligands (phen = 1,10-phenanthroline, bpy = 2,2′-bipyridine) or a tetradentate ligand (cyclam = 1,4,8,11-tetraazacycloteradecane) of the composition [Ni(BzMetdtc)(phen)2]ClO4 (1), [Ni(Pe2dtc)(phen)2]ClO4 (2), [Ni(Bzppzdtc)(phen)2]ClO4 · CHCl3 (3), [Ni(Bzppzdtc)(phen)2](SCN) (4), [Ni(BzMetdtc)(bpy)2]ClO4 · 2H2O (5), [Ni(Pe2dtc)(cyclam)]ClO4 (6), [Ni(BzMetdtc)2(cyclam)] (7), [Ni(Bz2dtc)2(cyclam)] (8) and [Ni(Bz2dtc)2(phen)] (9) (BzMetdtc = N,N-benzyl-methyldithiocarbamate(1-) anion, Pe2dtc = N,N-dipentyldithiocarbamate(1-) anion, Bz2dtc = N,N-dibenzyldithiocarbamate(1-) anion, Bzppzdtc = 4-benzylpiperazinedithiocarbamate(1-) anion), have been synthesized. Spectroscopic (electronic and infrared), magnetic moment and molar conductivity data, and thermal behaviour of the complexes are discussed. Single crystal X-ray analysis of 3 and 8 confirmed a distorted octahedral arrangement in the vicinity of the nickel atom with a N4S2 donor set. They represent the first X-ray structures of such type complexes. The catalytic influence of complexes 2, 3, 6, and 7 on graphite oxidation was studied and discussed.  相似文献   

14.
1-Ethynyl-2-phenyltetramethyldisilanes HCCSiMe2SiMe2C6H4X [X = NMe2 (1), H (2), CH3 (3), Br (4), CF3 (5)] are accessible from ClSiMe2SiMe2Cl, BrMgC6H4X and HCCMgBr in a two step Grignard reaction. The crystal structure of 1 as determined by single crystal X-ray crystallography exhibits a nearly planar PhNMe2 moiety and an unusual gauche array of the phenyl and the acetylene group with respect to rotation around the Si-Si bond. Full geometry optimization (B3LYP/6-31+G∗∗) of the gas phase structures of 1-5 affords minima for the gauche and the anti rotational isomers, both being very close in energy with a rotational barrier of only 3-5 kJ/mol. Experimental and calculated (time-dependent DFT B3LYP/TZVP) UV absorption data of 1-5 show pronounced electronic interactions of the HCC- and the C6H4X π-systems with the central Si-Si bond.  相似文献   

15.
This paper describes the synthesis of the first Ni(II) complexes with pyridoxal semicarbazone (PLSC), viz. Ni(PLSC)Cl2 · 3.5H2O (1), [Ni(PLSC)(H2O)3](NO3)2 (2), Ni(PLSC)(NCS)2 · 4H2O (3), [Ni(PLSC-2H)NH3] · 1.5H2O (4), as well as two new complexes with pyridoxal thiosemicarbazone (PLTSC), [Ni(PLTSC-H)py]NO3 (5) and [Ni(PLTSC-H)NCS] (6). Complexes 13 are paramagnetic and have most probably an octahedral structure, for complex 2 this was proved by X-ray diffraction analysis. In contrast, complexes 46 are diamagnetic and have a square-planar structure, and in the case of complex 5 this was also confirmed by X-ray structural analysis. In all cases the Schiff bases are coordinated as tridentate ligands with an ONX (X = O, PLSC; X = S, PLTSC) set of donor atoms. With the complexes involving the neutral form of PLSC and the monoanionic form of PLTSC, the PL moiety is in the form of a zwitterion. In addition to the above-mentioned techniques, all the complexes were characterized by measuring their molar conductivities, UV–Vis and partial IR spectra.  相似文献   

16.
TeX4 (X = Cl, Br) react in HCl/HBr with [Ph(CH3)2Te]X (X = Cl, Br) to give [PhTe(CH3)2]2[TeCl6] (1) and [PhTe(CH3)2]2[TeBr6] (2). The reaction of PhTeX3 (X = Cl, Br, I) in cooled methanol with [(Ph)3Te]X (X = Cl, Br, I) leads to [Ph3Te][PhTeCl4] (3), [Ph3Te][PhTeBr4] (4) and [Ph3Te][PhTeI4] (5). In the lattices of the telluronium tellurolate salts 1 and 2, octahedral TeCl6 and TeBr6 dianions are linked by telluronium cations through Te?Cl and Te?Br secondary bonds, attaining bidimensional (1) and three-dimensional (2) assemblies. The complexes 3, 4 and 5 show two kinds of Te?halogen secondary interactions: the anion-anion interactions, which form centrosymmetric dimers, and two identical sets of three telluronium-tellurolate interactions, which accomplish the centrosymmetric fundamental moiety of the supramolecular arrays of the three compounds, with the tellurium atoms attaining distorted octahedral geometries. Also phenyl C-H?halogen secondary interactions are structure forming forces in the crystalline structures of compounds 3, 4 and 5.  相似文献   

17.
Syntheses of [Me3SbM(CO)5] [M = Cr (1), W (2)], [Me3BiM(CO)5] [M = Cr (3), W (4)], cis-[(Me3Sb)2Mo(CO)4] (5), [tBu3BiFe(CO)4] (6), crystal structures of 1-6 and DFT studies of 1-4 are reported.  相似文献   

18.
The hydridic reactivity of the complex W(CO)(H)(NO)(PMe3)3 (1) was investigated applying a variety of protic donors. Formation of organyloxide complexes W(CO)(NO)(PMe3)3(OR) (R = C6H5 (2), 3,4,5-Me3C6H2 (3), CF3CH2 (4), C6H5CH2 (5), Me (6) and iPr (7)) and H2 evolution was observed. The reactions of 1 accelerated with increasing acidity of the protic donor: Me2CHOH (pKa = 17) < MeOH (pKa = 15.5) < C6H5CH2OH (pKa = 15) < CF3CH2OH (pKa = 12.4) < C6H2Me3OH (pKa = 10.6) < C6H5OH (pKa = 10).Regioselective hydrogen bonding of 1 was probed with two of the protic donors furnishing equilibrium formation of the dihydrogen bonded complexes ROH···HW(CO)(NO)(PMe3)3 (R = 3,4,5-Me3C6H2,3a and iPr, 7a) and the ONO hydrogen bonded species ROH···ONW(CO)(H)(PMe3)3 (R = C6H2Me3,3b and iPr, 7b) which were studied in hexane and d8-toluene solutions using variable temperature IR and NMR spectroscopy. Quantitative IR experiments at low temperatures using 3,4,5-trimethylphenol (TMP) confirmed the two types of competitive equilibria: dihydrogen bonding to give 3aH1 = −5.8 ± 0.4 kcal/mol and ΔS1 = −15.3 ± 1.4 e.u.) and hydrogen bonding to give 3b (ΔH2 = −2.8 ± 0.1 kcal/mol and ΔS2 = −5.8 ± 0.3 e.u.). Additional data for the hydrogen bonded complexes 3a,b and 7a,b were determined via NMR titrations in d8-toluene from the equilibrium constants Kδ) and KR1) measuring either changes in the chemical shifts of HW(Δδ) or the excess relaxation rates of HWR1) (3a,b: ΔHδ) = −0.8 ± 0.1 kcal/mol; ΔSδ) = −1.4 ± 0.3 e.u. and ΔHR1) = −5.8 ± 0.4 kcal/mol; ΔSR1) = −22.9 ± 1.9 e.u) (7a,b: ΔHδ) = −2.3 ± 0.2 kcal/mol; ΔSδ) = −11.7 ± 0.9 e.u. and ΔHR1) = −2.9 ± 0.2 kcal/mol; ΔSR1) = −14.6 ± 1.0 e.u). Dihydrogen bonding distances of 1.9 Å and 2.1 Å were derived for 3a and 7a from the NMR excess relaxation rate measurements of HW in d8-toluene. An X-ray diffraction study was carried out on compound 2.  相似文献   

19.
Chiral “P-N-P” ligands, (C20H12O2)PN(R)PY2 [R = CHMe2, Y = C6H5 (1), OC6H5 (2), OC6H4-4-Me (3), OC6H4-4-OMe (4) or OC6H4-4-tBu (5)] bearing the axially chiral 1,1′-binaphthyl-2,2′-dioxy moiety have been synthesised. Palladium allyl chemistry of two of these chiral ligands (1 and 2) has been investigated. The structures of isomeric η3-allyl palladium complexes, (R′ = Me or Ph; Y = C6H5 or OC6H5) have been elucidated by high field two-dimensional NMR spectroscopy. The solid state structure of [Pd(η3-1,3-Ph2-C3H3){κ2-(racemic)-(C20H12O2)PN(CHMe2)PPh2}](PF6) has been determined by X-ray crystallography. Preliminary investigations show that the diphosphazanes, 1 and 2 function as efficient auxiliary ligands for catalytic allylic alkylation but give rise to only moderate levels of enantiomeric excess.  相似文献   

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