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 62 毫秒
1.
The hydrolysis and condensation processes of titanium iso-propoxide modified with catechol (C6H4(OH)2; H2cat) have been investigated by 1H, 13C and 17O nuclear magnetic resonance spectroscopy. The hydrolysis reactions of the modified titanium iso-propoxide in the system with Ti:tetrahydrofuran (THF):H2O = 1:20:x (x = 1, 2 and 5 in a molar ratio) are essentially completed in the initial stage (<1 h), and the condensation reactions also proceed significantly during this stage. Upon hydrolysis with H2O/Ti = 1, the iso-propoxy groups are selectively hydrolyzed and the catecholate groups remain bound to titanium. With H2O/Ti = 2 and 5, both the iso-propoxy and catecholate groups are hydrolyzed, and the hydrolysis of the iso-propoxy groups is relatively preferential. Approximately half the catecholate groups are stably bound to titanium, even after hydrolysis with H2O/Ti = 5.  相似文献   

2.
The reactions of bis(anilino)phosphine oxide (C 6 H 5 NH) 2 P(O) H with (C 5 H 5 )2TiCl2 or Me2SiCl2 in a 1:1 molar ratio in THF results in the isolation of new phosph(V)azane complexes (C5H5)2Ti[(N C6H5)2P(O)H] (1) or Cl 2 Si[(N C 6 H 5 )2P(O)H] (2), respectively. In these reactions, HCl or CH4 elimination occurs and N-Ti or N-Si bonds form directly between a bis(anilino)phosphine oxide ligand and organotitanium or organosilicon compounds. The products(1) and (2) have been fully characterized by elemental analysis as well as 1 H, 31 P, 29 Si NMR, and IR spectroscopy.  相似文献   

3.
Ethylene/styrene copolymerizations using Cp′TiCl2(O‐2,6‐iPr2C6H3) [Cp′ = Cp* (C5Me5, 1 ), 1,2,4‐Me3C5H2 ( 2 ), tert‐BuC5H4 ( 3 )]‐MAO catalyst systems were explored under various conditions. Complexes 2 and 3 exhibited both high catalytic activities (activity: 504–6810 kg‐polymer/mol‐Ti h) and efficient styrene incorporations at 25, 40°C (ethylene 6 atm), affording relatively high molecular weight poly (ethylene‐co‐styrene)s with unimodal molecular weight distributions as well as with uniform styrene distributions (Mw = 6.12–13.6 × 104, Mw/Mn = 1.50–1.71, styrene 31.7–51.9 mol %). By‐productions of syndiotactic polystyrene (SPS) were observed, when the copolymerizations by 1 – 3 ‐MAO catalyst systems were performed at 55, 70 °C (ethylene 6 atm, SPS 9.0–68.9 wt %); the ratios of the copolymer/SPS were affected by the polymerization temperature, the [styrene]/[ethylene] feed molar ratios in the reaction mixture, and by both the cyclopentadienyl fragment (Cp′) and anionic ancillary donor ligand (L) in Cp′TiCl2(L) (L = Cl, O‐2,6‐iPr2C6H3 or N=CtBu2) employed. Co‐presence of the catalytically‐active species for both the copolymerization and the homopolymerization was thus suggested even in the presence of ethylene; the ratios were influenced by various factors (catalyst precursors, temperature, styrene/ethylene feed molar ratio, etc.). © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 4162–4174, 2008  相似文献   

4.
Three proton-transfer salts of diphenylphosphinic acid (DPPA) with 2-amino-5-(X)-pyridine (AMPY, X = Cl, CN or CH3), namely, 2-amino-5-chloropyridinium diphenylphosphinate, C5H6ClN2+·C12H10O2P ( 1 , X = Cl), 2-amino-5-cyanopyridinium diphenylphosphinate, C6H6N3+·C12H10O2P ( 2 , X = CN), and 2-amino-5-methylpyridinium diphenylphosphinate, C6H9N2+·C12H10O2P ( 3 , X = CH3), have been synthesized and characterized by FT–IR and 1H NMR spectroscopy, and X-ray crystallography. The crystal structures of compounds 1 – 3 were determined in the space group P for 1 and 2 , and C2/c for 3 . All three compounds contain N—H…O hydrogen-bonding interactions due to proton transfer from the O=P—OH group of DPPA as donor to the pyridine N atom of AMPY as acceptor. The proton transfer of compounds 1 – 3 was also verified by 1H NMR and FT–IR spectroscopy. The stoichiometry of all three proton-transfer salts was determined to be 1:1 and the Benesi–Hildebrand equation was applied to determine the formation constant (KCT) and the molar extinction coefficient (ϵCT) in each case. Theoretical density functional theory (DFT) calculations were performed to investigate the optimized geometries, the molecular electrostatic potentials (MEP) and the highest occupied molecular orbitals (HOMO) and lowest unoccupied molecular orbitals (LUMO) of all three proton-transfer salts. The results showed good agreement between the experimental data and the DFT computational analysis.  相似文献   

5.
Highly porous titanium oxophenylphosphate (TPP-1) has been synthesized hydrothermally at 448 K using phenylphosphonic acid (PPA) as the organophosphorus source without the aid of any template or structure-directing agent (SDA). Powder XRD, TEM, SEM-EDS, N2 sorption, ICP-AES chemical analysis, 13C and 31P MAS NMR, UV-Vis, XPS spectroscopic tools, TG-DTA and NH3-TPD were used to characterize this material. XRD, N2 sorption and TEM image analysis suggested the disordered layered framework structure and the presence of large-size micropores along with mesopores in this material. Spectroscopic data suggested the presence of phenyl group, O, Ti and P in this open-framework material, where Ti centers can adopt both tetrahedral and octahedral geometry. TPP-1 showed fairly good H2 adsorption capacity under normal pressure to high pressure at 77 K. Physisorption data on hydrogen has suggested the potential application of this porous material in H2 storage.  相似文献   

6.
Linear low‐density polyethylene (LLDPE) can be prepared by addition of ethylene to a mixture of two catalysts. In this “tandem catalysis” scheme one catalyst dimerizes or oligomerizes ethylene to α‐olefins while the second site incorporates these α‐olefins into a growing polyethylene chain. A variety of classical catalyst combinations are available for this purpose. Better control over the polymerization process, and therefore product properties, is attained by the use of homogenous “single site” catalysts. The best‐behaved tandem processes take advantage of well‐defined catalysts that require stoichiometric quantities of activators. One such system employs [(C6H5)2PC6H4C(OB(C6F5)3)O‐κ2P,O]Ni(η3‐CH2CMeCH2) and {[(η5‐C5Me4)SiMe2(η1‐NCMe3)]TiMe}{MeB(C6F5)3}. The nickel sites are responsible for converting ethylene to 1‐butene or mixtures of 1‐butene with 1‐hexene. These olefins are copolymerized with ethylene at the titanium sites. It is possible to obtain a linear correlation between the branching content in the polymer product and the Ni/Ti ratio. The effect of ligand substitution at nickel has also been investigated. When the benzyl derivative [(C6H5)2PC6H4C(O‐B(C6F5)3)O‐κ2P,O]Ni(η3‐CH2C6H5) is used instead of the methallyl counterpart [(C6H5)2PC6H4C(OB(C6F5)3)O‐κ2P,O]Ni(η3‐CH2CMeCH2), one obtains, at a constant Ni/Ti ratio, considerably more branching in the final polymer structure. These results are rationalized in terms of a more efficient initiation when the more labile benzyl ligand is used.  相似文献   

7.
Hybrid TiO2/ormosil waveguiding films have been prepared by the sol-gel method at low thermal treatment temperature of 150C. The influence of processing parameters including the molar ratios of titanium butoxide (Ti(OBu)4)/3-glycidoxypropyltrimethoxysilane (GLYMO) and H2O/Ti(OBu)4 (expressed as R), especially aging of sol on the optical properties was investigated. The optical properties of films were measured with scanning electron microscope (SEM), UV/VIS/NIR spectrophotometer (UV-Vis), m-line and the scattering-detection method. The results indicate that the film thickness increases with the increase of sol aging time, but the variation of refractive index as a function of sol aging time depends on the relative ratios of GLYMO to Ti(OBu)4. Higher transmittance and lower attenuation of the planar waveguide can be obtained in the sol with lower Ti(OBu)4 contents and shorter aging time.  相似文献   

8.
5-Phenyl-2-pentene (5Ph2P) was found to undergo monomer-isomerization polymerization with TiCl3–R3Al (R = C2H5 or i-C4H9, Al/Ti > 2) catalysts to give a polymer consisting of exclusively 5-phenyl-1-pentene (5Ph1P) unit. The geometric and positional isomerizations of 5Ph2P to its terminal and other internal isomers were observed to occur during polymerization. The catalyst activity of alkylaluminum examined to TiCl3 was in the following order: (C2H5)3Al > (i-C4H9)3Al > (C2H5)2AlCl. The rate of monomer-isomerization polymerization of 5Ph2P with TiCl3–(C2H5)3Al catalyst was influenced by both the Al/Ti molar ratio and the addition of nickel acetylacetonate [Ni(acac)2], and the maximum rate was observed at Al/Ti = 2.0 and Ni/Ti = 0.4 in molar ratios.  相似文献   

9.
Summary Dichlorobis(methylsalicylato)titanium(IV) reacts with potassium or amine salts of dialkyl or diaryl dithiocarbamates in 11 and 12 molar ratios in anhydrous benzene (room temperature) or in boiling CH2Cl2 to yield mixed ligand complexes: (AcOC6H4O)2 Ti(S2CNR2)Cl (1) and (AcOC6H4O)2 Ti(S2CNR2)2 (2), R=Et, n-Pr, n-Bu, cyclo-C4H8 and cyclo-C5H10. These compounds are moisture sensitive and highly soluble in polar solvents. Molecular weight measurement in conjunction with i.r.,1H and13C n.m.r. spectral studies suggest coordination number 7 and 8 around titanium(IV) in (1) and (2) respectively.  相似文献   

10.
The reaction products of Cu(II) 2-chlorobenzoate and the imidazole (1), and of Cu(II) 2,6-dichlorobenzoate and the imidazole (2) formulated as CuL’2⋅2imd⋅2H2O and CuL”2⋅2imd⋅2H2O (L’=C7H4ClO2 , L”=C7H3Cl2O2 , imd=imidazole), were prepared and characterized by means of spectroscopic measurements and thermochemical properties. The blue (1) and green (2) complexes were obtained as solids with a 1:2:2 molar ratio of metal to carboxylate ligand to imidazole. When heated at a heating rate of 10 K min−1 the hydrated complexes, (1) and (2), lose some of the crystallization water molecules and then decompose to gaseous products. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

11.
Blue single crystals of Cu[μ3‐O3P(CH2)2COOH] · 2H2O ( 1 ) and Cu[(RS)‐μ3‐O3PCH(C2H5)COOH] · 3H2O ( 2 ) were prepared in aqueous solutions (pH = 2.5–3.5). 1 crystallizes in space group Pbca (no. 61) with a = 812.5(2), b = 919.00(9), and c = 2102.3(2) pm. Cu2+ is fivefold coordinated by three oxygen atoms stemming from [O3P(CH2)2COOH]2– anions and two water molecules. The Cu–O bond lengths range from 194.0(3) to 231.8(4) pm. The connection between the [O3P(CH2)2COOH]2– anions and the Cu2+ cations yields a polymeric structure with layers parallel to (001). The layers are linked by hydrogen bonds. 2 crystallizes in space group Pbca (no. 61) with a = 1007.17(14), b = 961.2(3), c = 2180.9(4) pm. The copper cations are surrounded by five oxygen atoms in a square pyramidal fashion with Cu–O bonds between 193.6(4) and 236.9(4) pm. The coordination between [O3PCH(C2H5)COOH]2– and Cu2+ results in infinite puckered layers parallel to (001). The layers are not connected by any hydrogen bonds. Each layer contains both R and S isomers of the [O3PCH(C2H5)COOH]2– dianion. Water molecules not bound to Cu2+ are intercalated between the layers. UV/Vis spectra suggest three d–d transition bands at 743, 892, 1016 nm for 1 and four bands at 741, 838, 957, and 1151 nm for 2 , respectively. Magnetic measurements suggest a weak antiferromagnetic coupling between Cu2+ due to a super‐superexchange interaction. Thermoanalytical investigations in air show that the compounds are stable up to 95 °C ( 1 ) and 65 °C ( 2 ), respectively.  相似文献   

12.
Summary Titanocene complexes ([Ti(5-C5H4 R)2 X 2];R = H, SiMe3;X=Cl, Br) react with Na2H2 EDTA in aqueous methanol to give an identical product ([Ti(EDTA)(H2O)] by cleavage of the halogen and cyclopentadienyl ligands. The structure of [Ti(EDTA)(H2O)] has been determined by X-ray diffraction; crystal data: monoclinica=13.923(6),b=7.048(3),c=13.252(5) Å, =90.81(1)°, space groupP21/c,Z=4. In this complex, Ti has a sevenfold coordination with a hexadentateEDTA 4– ligand and a water molecule occupying an additional coordination site.
Reaktion von Titanocen-Dihalogeniden mit Na2H2 EDTA. Kristallstruktur von [Ti(EDTA)(H2O)]
Zusammenfassung Titanocenkomplexe ([Ti(5-C5H4 R)2 X 2];R=H, SiMe3;X=Cl, Br) reagieren in wäßrigem Methanol mit Na2H2 EDTA unter Verdrängung der Halogen- und Cyclopentadienylliganden zum selben Produkt ([Ti(EDTA)(H2O)]). Die Struktur von ([Ti(EDTA)(H2O)]) wurde röntgenographisch bestimmt. Kristalldaten: monoklin,a=13.923(6),b=7.048(3),c=13.252(5) Å, =90.81(1)°, RaumgruppeP21/c,Z=4. In diesem Komplex ist das Titanatom mit einem sechszähnigenEDTA-Liganden und einem Wassermolekül, das eine zusätzliche Koordinationsstelle besetzt, siebenfach koordiniert.
  相似文献   

13.
Reacting stoichiometric amounts of 1‐(diphenylphosphino)ferrocene­carboxylic acid and [Ti(η5‐C5HMe4)22‐Me3SiC[triple‐bond]CSiMe3)] produced the title carboxyl­atotitanocene complex, [{μ‐1κ2O,O′:2(η5)‐C5H4CO2}{2(η5)‐C5H4P(C6H5)2}{1(η5)‐C5H(CH3)4}2FeIITiIII] or [FeTi(C9H13)2(C6H4O2)(C17H14P)]. The angle subtended by the Ti/O/O′ plane, where O and O′ are the donor atoms of the κ2‐carboxy­late group, and the plane of the carboxyl‐substituted ferrocene cyclo­penta­dienyl is 24.93 (6)°.  相似文献   

14.
Phosphorus-doped aluminium oxide thin films were deposited in a flow-type ALE reactor from AlCl3, H2O and from either P2O5 or trimethyl-phosphate. Structural information of the films was obtained from Fourier transform infrared (FTIR) spectra. Rutherford backscattering spectroscopy (RBS) was used to quantitatively determine the composition of the films. The P/Al intensity ratios calculated from X-ray fluorescence (XRF) results were in a linear relation with the P/Al concentration ratios calculated from RBS results. For comparison, the intensity ratios of the phosphorus peak (P=O) at about 1250 cm–1 and the aluminium peak (Al-O) at about 950 cm–1 were determined from the IR absorption spectra. The calibration of FTIR peak intensities was done by plotting the intensity ratios of phosphorus and aluminium peaks against the P/Al concentration ratios measured by RBS. FTIR gave also a linear calibration curve with RBS but the method is less suitable for routine analysis of P/Al ratio than XRF.  相似文献   

15.
Abstract

Nickel(II) complexes with a combination of trithiocyanuric acid and diamines or triamines of composition [Ni(aepa)(ttcH)(H2O)], [Ni(dien)(ttcH)(H2O)], [Ni(dpta)(ttcH)(H2O)] H2O, [Ni(phen)2(ttcH)]H2O, [Ni(phen)3](ttcH)-5H2O and [Ni(1,2-pn)3](ttcH)-H2O (aepa = N-(2-aminoethyl)-1,3-propanediamine,dien = diethylenetriamine,dpta = dipropylenetnamine, phen = 1,10-phenanthroline, 1,2-pn = 1,2-diaminopropane. ttcH3 = trithiocyanuric acid) have been prepared. The compounds have been characterized by means of elemental analysis, IR and electronic spectroscopies and magnetochemical measurements. Selected complexes were studied by thermal analysis. The compounds can be characterized as distorted octahedral Ni(II) complexes. It was found that the trithiocyanuric dianion can act either as a bidentate ligand or be situated out of the coordination sphere of nickel. The crystal and molecular structure of [Ni(dpta)(ttcH)(H2O)] H2O was determined. Crystals are monoclinic, space group P21/n, with a = 20.316(4), b = 7.967(2), c = 21.401(4) Å, β = 99.23(3)°, K=3419.1(13)Å3, Z = 4, T = 293 K. The nickel(II) atom is six-coordinated by three nitrogen atoms from dipropylene-triamine, nitrogen and sulphur from trithiocyanuric acid, and an oxygen atom from a water molecule in a distorted octahedral geometry.  相似文献   

16.
The adducts of bis(O,O′-dialkylmonoselenophosphato)cobalt(II) complexes, Co{O(Se)P(OR)2}2(L)4 (where R?=?n-Pr, i-Pr; L?=?C5H5N, NC5H4Me-2, NC5H4Me-3), were synthesized by in situ reactions of CoCl2?·?6H2O, Lewis base, and NaO(Se)P(OR)2. The single crystal structure of Co{O(Se)P(OiPr)2}2(C5H5N)4 shows distorted octahedral geometry around cobalt(II) and monoselenophosphates are trans. The CoN4 forms a square plane. These bis(O,O′-dialkylmonoselenophosphato)cobalt(II) adducts were characterized by elemental analyses, spectroscopic techniques (UV-Vis, infrared, 1H and 31P), and magnetic moment measurements.  相似文献   

17.
Three lanthanide coordination polymers constructed from infinite rod‐shaped secondary building units (SBUs), [Nd2(H2O)2(cis‐chdc)2(trans‐chdc)]?2H2O ( 1 ), Nd2(H2O)4(trans‐chdc)3 ( 2 ), and [Sm2(H2O)2(cis‐chdc)(trans‐chdc)2]?4H2O ( 3 ) (chdcH2=1,4‐cyclohexanedicarboxylic acid), were hydrothermally synthesized and structurally characterized. The structures of 1 – 3 are modulated by different ratios of the cis and trans configurations of chdc2? ligands, which was achieved by temperature control in the hydrothermal reactions. Crystal‐structure analysis revealed that 1 is a four‐connected pcu‐type rod packing network built from cross‐linking of rod‐shaped neodymium–oxygen SBUs by cis‐ and trans‐chdc2? ligands in a 2:1 ratio, 2 displays a complicated six‐connected hex‐type rod packing structure built by connection of rod‐shaped neodymium–oxygen SBUs and trans‐chdc2? ligands, and 3 features an unprecedented five‐connected rod packing pattern constructed from rod‐shaped samarium–oxygen SBUs and cis‐ and trans‐chdc2? ligands in a 1:2 ratio.  相似文献   

18.
In order to investigate the influence of Cl/SO42− molar ratios and hydrolysis temperature on the hydrolysis process and TiO2 pigment, H2TiO3 was prepared with a low concentration of titanyl sulfuric–chloric acid solution by hydrothermal hydrolysis. Under the optimal hydrolysis conditions, 1.5–2.2 μm of H2TiO3 samples were achieved. After doping and calcination, anatase TiO2 pigments demonstrated excellent performance: the achromic ability of tinctorial strength (TCS) and blue phase index (SCX) were 1,429 and 4.07, respectively. As hydrolysis was a significant step in the process, the structure was simplified to a periodic structure of Ti[OH](H2O)3Cl(SO4) to simulate the cluster structures. Based on experimental results and density functional theory (DFT) calculation, the hydrolysis mechanism was presumed to be a process of anionic (OH, Cl and SO42−) competition reaction to explain the formation of anatase-type H2TiO3, and the crystal growth direction of H2TiO3 was also confirmed to be a (OA) and b (OB).  相似文献   

19.
The title compound, octa‐tert‐butoxybis­[μ3‐2,2′‐(N‐methyl­imino)­diethanolato]­di‐μ‐oxo‐tetratitanium(IV), [Ti2O{(OCH2CH2)2(NCH3)}{(CH3)3CO}4]2 or [Ti4(C5H11NO2)2(C4H9O)8O2], lies about an inversion centre, and displays the less usual zigzag configuration. One O atom of the N‐methyl­diethoxo­amine ligand bridges the symmetry‐related Ti atoms, while the other bridges the two independent Ti atoms, with the N atom binding to give a facial configuration. Four tBuO ligands and a bridging oxide complete the respective five‐ and sixfold coordination of the two Ti atoms. The Ti—O bond lengths range in a self‐consistent fashion from 1.7624 (17) to 2.0878 (18) Å, while the Ti—N bond length is 2.374 (2) Å.  相似文献   

20.
The influence of the molar ratio h = [H2O]/[Ti(OR)4] (R = Pr i ) on the kinetics of the titanium-oxo-alkoxy clusters (TOAC) nucleation was studied. Clusters were formed by the titanium tetraisopropoxide Ti(OPr i )4 chemical reaction with H2O in n-propanol solution, with the fixed concentration of Ti(OPr i )4 (c = 0.04 M), molar ratio h ∈ {11, 14, 17, 20} and temperature T ∈ {298, 308, 318} K. It was determined that the isothermal rate of clusters nucleation is a power law function of the molar ratio h. The kinetic parameter β value changes complexly as h and T change. The value of apparent activation energy of the nucleation process (E a) decreases with the increase of value h. It was found that nucleation is a reaction with complex kinetics whose elementary stages are hydrolysis Ti(OR)4 to Ti(OR)3OH and formation of titanium-oxo-alkoxy clusters [Ti n + βOβ](OR)4n + 2β through the alcoxolation reaction.  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