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 effect of the nature of the anion on the performance of ionic rhodium catalysts has received little attention. Herein it is shown that the use of highly fluorous tetraphenylborate anions can enhance catalyst activity in both conventional and fluorous media. For hydrogenation catalysts of the type [Rh(COD)(dppb)][X] {COD=1,5‐cis,cis‐cyclooctadiene; dppb=1,4‐bis(diphenylphosphino)butane; X=BF4 ( 1a ), [BPh4] ( 1b ), [B{C6H4(SiMe3)‐4}4] ( 1c ), [B{C6H3(CF3)2‐3,5}4] ( 1d ), [B{C6H4(SiMe2CH2CH2C6F13)‐4}4] ( 1e ), [B{C6H4(C6F13)‐4}4] ( 1f ) and [B{C6H3(C6F13)2‐3,5}4] ( 1 g )} the activity towards the hydrogenation of 1‐octene in acetone increased in the order 1c < 1b < 1e < 1a < 1d ~ 1f < 1g with 1g being twice as active as the commonly applied 1a . Despite the fluorophilic character introduced by the substituted tetraarylborate anions, the presence of some perfluoroalkyl‐substituents in the cation was still required for achieving high partition coefficients. Therefore, [Rh(COD)(Ar2PCH2CH2PAr2)][X] {Ar=C6H4(SiMe2CH2CH2C6F13)‐4, X=[B{C6H3(C6F13)2‐3,5}4] ( 3f ); Ar=C6H4(SiMe(CH2CH2C6F13)2)‐4 and X=[B{C6H4(C6F13)‐4}4] ( 2g )} were prepared, which were active in the hydrogenation of 1‐octene, 2g even more so than 3f . Both these highly fluorous catalysts could be recycled with 99% efficiency through fluorous biphasic separation, whereas the corresponding BF4 complex of 2g ( 2a ) did not show any affinity for the fluorous phase.  相似文献   

2.
The reaction of [Ir(μ‐Cl)(COD)]2 with various fluorous derivatives of triphenylphosphane containing a para‐, meta‐, or ortho‐(1H,1H‐perfluoroalkoxy)‐substituted fluorous phosphane P(C6H4‐ORf)3 (Rf=CH2C7F15, CH2CH2CH2C8F17) and CO (1 atm) gives the corresponding trans‐[Ir(μ‐Cl)(CO){P(C6H4ORf)3}2]. The IR νCO values of these complexes give some information on the donor/acceptor properties of the phosphanes. These fluorous derivatives of triphenylphosphane, as well as a phosphane bearing two (1H,1H‐perfluoroalkyloxy) chains at the 3,5‐positions, were used in association with [Rh(μ‐Cl)(COD)]2 or [Rh(COD)2]PF6 in the reduction of methyl cinnamate, 2‐cyclohexen‐1‐one, cinnamaldehyde, and methyl α‐acetamidocinnamate in a two‐phase system D‐100/ethanol under 1 bar hydrogen at room temperature. Some differences in catalytic activity were observed in the reduction of methyl cinnamate, the most active catalyst being the rhodium complex containing the phosphane with the p‐fluorous ponytail. Recycling of the fluorous catalyst was possible, particularly using the p‐substituted phosphane, where no significant loss of catalyst or activity was observed, and generally with very low leaching of rhodium or phosphane in the organic phase.  相似文献   

3.
The atom transfer radical polymerization (ATRP) of methyl methacrylate (MMA) is often carried out under homogeneous conditions, so the residual metal catalyst in the polymer often influences the quality of the polymer and causes environmental pollution in the long run. Novel CuBr/4,4′‐bis(RfCH2OCH2)‐2,2′‐bpy complexes (Rf = n‐C9F19, n‐C10F21, or n‐C11F23; 2,2′‐bpy = 2,2′‐bipyridine) are insoluble in toluene at room temperature yet readily dissolve in toluene at elevated temperatures to form homogeneous phases for use as catalysts in the ATRP reaction, and the Cu complexes precipitate again upon cooling. The CuBr/4,4′‐bis(n‐C9F19CH2OCH2)‐2,2′‐bpy system produced the best results (e.g., polydispersity index by gel permeation chromatography = 1.26–1.41), in that the residual Cu content in the polymer was as low as 19.3 ppm when the ATRP of MMA was carried out in the thermomorphic mode. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  相似文献   

4.
A fluoroalkene-soluble tetrakis[heptadecafluorononyl]-substituted Pd(II)-phthalocyanine complex has been studied for olefin (styrene, 1-octene, trans-2-octene and cyclohexene) hydrogenation with molecular hydrogen in an organic/fluorous biphasic system [n-hexane/perfluoromethylcyclohexane (PFMCH)]. The palladium complex was found to be an active catalyst for styrene (100% conversion, TON = 634) and 1-octene (92%, TON = 596) at 80 °C and 15 bar of H2 after 6 h of reaction time. The catalyst was recycled in nine consecutive reactions for the hydrogenation of styrene without the loss of activity or metal contamination.  相似文献   

5.
A kinetic model is derived from experimental data for the Fischer-Tropsch reaction on a precipitated iron catalyst. In this model, the effect of nonideality of the reaction medium on the reaction rate is taken into account by introducing fugacity coefficients derived from a modified Redlich-Kwong-Soave equation of state. Coefficients of the Schulz-Flory distribution for saturated and unsaturated hydrocarbon products were calculated as functions of CO and H2 fugacities in the reaction mixture. The proposed kinetic model is applicable atT= 523–623 K andP = 6–100 atm. A method based on the calculated critical parameters of the reaction mixture is proposed for the selection of suitable supercritical solvent and for the optimization of its concentration. The reaction rate and the total yield of CnH2n (n ≥ 2) olefins (including the desired fraction C5–C11) under supercritical conditions were demonstrated to be essentially higher than those for the reference process carried out in the absence of solvent.  相似文献   

6.
In this paper we report some performance properties of [(2-alkyl-1,3-dioxolane-4-yl)methyl]trimethylammonium bromides (C n -D-TAB, where C n =n-C9H19, n-C11H23, n-C13H27), [(2,2-dialkyl-1,3-dioxolane-4-yl)methyl] dimethylammonium bromides (C n C m -D-TAB, where C n =n-C8H17, n-C10H21 and C m =CH3 or C n =C m =n-C6H13), in relation to classical alkyltrimethylammonium bromides (C n -TAB, where C n =n-C12H25, n-C14H29 or C16H33) and dialkyldimethylammonium bromides (C n C m -DAB, where C n =C m =n-C10H21 and C12H25). Antistatic properties were measured by means of surface resistance (R s), half charge decay time τ1/2, and maximal voltage induced (U ind) on polyethylene (PE) film and polypropylene (PP) nonwoven fabric. Wetting ability is reported as the concentration of surfactant necessary for immersion of a cottonfabric ring in 100 s, according to the Polish Standard PN-74/C-04800. All acetal-type surfactants showed the antistatic effect. The best results were given by C9-D-TAB [R s=8.5·107 (PE), 1.1·108 Ω (PP); τ1/2=0.25 (PE), 0.30 s (PP); U ind=205 (PE), 220 V (PP)] and C11-D-TAB [R s=8.7·107 (PE), 1.3·108 Ω (PP); τ1/2=0.25 (PE), 0.30 s (PP); U ind=190 (PE), 340 V (PP)] in relation to pure PE foils and PP fabric: R s=3.0·1014 (PE), 9.8·1014 Ω (PP); τ1/2=>600 (PE), >600 s (PP); U ind=985 (PE), 940 V (PP). Both C9-D-TAB and C11-D-TAB indicate similar degrees of antistatic ability compared to C10C10-DAB. All C n -TAB compounds do not show any antistatic effects. It was concluded that addition of the 1,3-dioxolane moiety to the hydrophobic tail of quaternary ammonium salts promotes both antistatic and wetting properties. Additionally, methyl or n-hexyl group incorporation at the C-2 carbon atom in the 1,3-dioxolane ring significantly decreases the surface conductivity.  相似文献   

7.
The performance of jet fuel depends on the density (ρ), condensed phase heat of formation (▵fH°(c)), and specific impulse (ISP). Exo‐tricyclo[5.2.1.0(2,6)]decane (C10H16) or JP‐10 is now used as a suitable synthetic liquid jet fuel because it has the approximated values of ρ=1.1 g cm−3 and ▵fH°(c)=− 123 kJ mol−1 and a broad range between the melting and boiling points, i.e. TbpTmp=196.2 K. This work introduces a suitable pathway for calculation of the values of ρ, ▵fH°(c), and ISP of 13 well‐known isomers of JP‐10 and a series of saturated polycyclic hydrocarbons with general formula of CnHn (5≤n≤12) in order to specify high performance jet fuels. Although 13 compounds have larger values of ISP*ρ than JP‐10, only two compounds, tetraspiro[2.0.0.0.2.1.1.1]undecane and tetracyclo[3.2.0.0(2,7).0(4,6)]heptane, are suitable as jet fuels.  相似文献   

8.
The carbonylation of α-Bromo-p-xylene (BrCH2C6H4CH3, BX) with iron pentacarbonyl (Fe(CO)5) by phase-transfer catalysis was studied in an organic solvent/alkaline solution. The reaction mechanism was corrected and clarified. The concentrations of base, reactant and catalyst, volume ratio, the kind of catalysts, organometals and solvents were evaluated to find the optimum condition in this reaction. The technique of phase-transfer catalysis has a dramatic accelerating effect on the reaction. In examining eight kinds of phase-transfer catalysts, tetra-n-butylammonium cation and tetra-n-butylphosphonium cation were found to be the best for increasing the yield of bis(p-methylbenzyl) ketone ((p-CH3C6H4CH2)2CO, BMBK). The amount of phase-transfer catalyst, the concentration of NaOH, the molar ratio of BX to Fe(CO)5 and the volume ratio of aqueous to organic phase affected the product selectivity.  相似文献   

9.
One novel dysprosium coordination polymer [Dy(PIC)3(H2O)2]n (HPIC = pyridine-4-carboxylic acid) has been synthesized. X-ray analysis reveals that it forms the chain-like molecular structure through the bridged oxygen atoms of the carbonyl groups. The title coordination polymer crystallizes in the monoclinic system, space group C2/c, with lattice parameters: a=20.243(9) Å, b=11.576(5) Å, c=9.834(4) Å, β=110.601(2)°, V=2078(2) Å3, Dc=1.805 mg/m3, Z=4, F(000)=1100, GOF = 1.11, R1=0.0404. The photophysical property has been studied with ultraviolet absorption spectrum, excitation and emission spectrum. The luminescence spectra show the stronger blue emission than yellow emission.  相似文献   

10.
Two asymmetric alkylidene‐bridged dinuclear titanocenium complexes (CpTiCl2)25‐η5‐C9H6(CH2)nC5H4), 1 (n = 3) and 2 (n = 4) have been prepared by treating two equivalents of CpTiCl3 with the corresponding dilithium salts of the ligands C9H7(CH2)nC5H5 (n = 3, 4). Additionally, Ti(η55n‐BuC5H4C5H5)Cl2 (3) and Ti(η55n‐BuC9H6C5H5)Cl2 (4) were synthesized as corresponding mononuclear complexes. All complexes were characterized by 1H, 13C NMR, and IR spectroscopy. Homogenous ethylene polymerization catalyzation using those complexes has been conducted in the presence of methylaluminoxane (MAO). The influences of reaction parameters, such as [MAO]/[Cat] molar ratio, catalyst concentration, ethylene pressure, temperature, and time have been studied in detail. The results showed that the catalytic activities of both dinuclear titanocenes were higher than those of the corresponding mononuclear titanocenes. Although the two dinuclear complexes were different in only one [CH2] unit, the catalytic activity of 2 was about 50% higher than that of 1; however, the molecular weight of polyethylene (PE) obtained by 2 was lower than that obtained from 1. The molecular weight distribution of PE produced by these dinuclear complexes reached 6.9 and 7.3, respectively. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 101: 3317–3323, 2006  相似文献   

11.
An asymmetric 3‐oxa‐pentamethylene bridged dinuclear titanocenium complex (CpTiCl2)25‐η5‐C9H6(CH2CH2OCH2CH2)C5H4) ( 1 ) has been prepared by treating two equivalents of CpTiCl3 with the corresponding dilithium salts of the ligand C9H7(CH2CH2OCH2 CH2)C5H5. The complex 1 was characterized by 1H‐, 13C‐NMR, and elemental analysis. Homogenous ethylene polymerization catalyzed using complex 1 has been conducted in the presence of methylaluminoxane (MAO). The influences ofreaction parameters, such as [MAO]/[Cat] molar ratio, catalyst concentration, ethylene pressure, temperature, and time have been studied in detail. The results show that the catalytic activity and the molecular weight (MW) of polyethylene produced by 1 /MAO decrease gradually with increasing the catalyst concentration or polymerization temperature. The most important feature of this catalytic system is the molecular weight distribution (MWD) of polyethylene reaching 12.4, which is higher than using common mononuclear metallocenes, as well as asymmetric dinuclear titanocene complexes like [(CpTiCl2)25‐η5‐C9H6(CH2)nC5H4)] (n = 3, MWD = 7.31; n = 4, MWD = 6.91). The melting point of polyethylene is higher than 135°C, indicating highly linear and highly crystalline polymers. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  相似文献   

12.
The title compound {[Ca(CHZ)2(H2O)](NTO)2⋅3.5H2O}n was synthesized by using an aqueous solution of calcium 3‐nitro‐1,2,4‐triazol‐5‐onate and carbohydrazide (CHZ, NH2NHCONHNH2). Its molecular structure was determined by X‐ray diffraction and its crystals have monoclinic form, with space group C2/c, where a=2.4483(4) nm, b=1.2581(2) nm, c =1.6269(3) nm, β=121.168(12)°, V=4.2879(13) nm3, Z=8, dc=1.727 g⋅cm−3, μ (Mo Kα)=3.9 cm−1, M=557.47, F(000)=2312. The coordination polyhedron is a tricapped trigonal prism in a tetradecahedron with a coordination number of nine. The whole molecule has many long chains formed through the carbohydrazide bridges, and every long chain is unlimited along the c axis. The long chains are linked by hydrogen bonds to form the crystal structure.  相似文献   

13.
The unique properties of surface active agents which comprise a perfluoroalkyl group of about eight carbon atoms linked to a hydrophilic group have been recognised for some time. For example, such surfactants give very low surface tensions to aqueous systems, they are very effective at low concentrations, they have activity in organic systems, they are often stable in hostile environments and they can exhibit oil and water repellency when they are absorbed on substrates. Prior to the work described in this paper, the fluorochemical group for these surfactants had been made by two routes, electrochemical fluorination and telomerisation. Both of these suffer certain disadvantages so a programme of work was initiated to see whether another route to suitable perfluoroalkyl compounds could be discovered. Work in our laboratories showed that when tetrafluoroethylene is treated with fluoride ion in an aprotic solvent such as dimethylformamide, a mixture of oligomers ranging from trimers to heptamers is produced. These oligomers are highly branched unsaturated compounds. The pentamer is the most abundant and the manner in which its chemistry has been exploited to yield surfactants with a highly branched fluorochemical group (unlike those prepared by other methods) is outlined. For example, the pentamer will react with phenol or substituted phenols to give phenyl ethers, C10F19OC6H4X (X = H or a substituent). Such ethers can be converted into surface active compounds by either the introduction of hydrophilic groups onto the benzene ring (e. g. sulphonation) or by converting existing substituents into hydrophilic groups (e. g. if X = CH3, by oxidation to COOH). Hydrophilic groups may be attached directly to the fluorocarbon by for example reacting the pentamer with a hydroxyl-ended ethylene oxide condensate (C10F20+HO(CH2CH2O)nR→C10F19O(CH2CH2O)nR + HF). Reaction of TFE-pentamer with inorganic nucleophiles e. g. OH-, can cause a certain amount of degradation but the products of such degradation can be used to give other surface active compounds.  相似文献   

14.
1-(4-Isobutylphenyl)ethanol (IBPE) was carbonylated to 2-(4-isobutylphenyl)propionic acid (ibuprofen) in an aqueous/organic two phase system using the water-soluble Pd(tppts)3 catalyst [tppts = P(C6H4-m-SO3Na)3] in the presence of p-CH3C6H4SO3H at 363 K, 15 MPa CO pressure and a palladium concentration of 150 ppm without addition of organic solvents. Under these conditions the conversion of IBPE was 83% and the selectivity to ibuprofen 82% with no decomposition of the Pd(tppts)3 catalyst. Both the activity and selectivity were strongly influenced by the tppts/Pd molar ratio and the nature of the added Brønsted acid. Maximum efficiency was observed for P/Pd = 10. Acids of weakly or non-coordinating anions, such as p-CH3C6H4SO3H, CF3COOH or HPF6 afforded carbonylation. No catalytic activity was observed in the presence of acids of strongly coordinating anions, such as HI. The water-soluble Pd/dppps catalyst [dppps = Ar2-nPhnP-(CH2)3-PPhnAr2-n; Ar = C6H4-m-SO3Na; n = nń = 0: 86% and n = 0, nń = 1: 14%] exhibited low catalytic activity and the major product obtained was the linear isomer of ibuprofen, 3-(4-isobutylphenyl)propionic acid (3-IPPA) with selectivities up to 78%. Replacement of tppts by a ligand containing less −SO3Na groups such as monosulphonated triphenylphosphine (tppms) gives rise to a dramatic drop in the catalytic activity and selectivity to ibuprofen. No catalytic activity was observed using palladium catalysts modified with 2-pyridyldiphenylphosphine (PyPPh2) and tris(2-pyridyl)phosphine (PPy3) which are both water soluble in their protonated form. A catalytic cycle is proposed to explain the observed results. ©1997 SCI  相似文献   

15.
High-optical-quality ternary [(Y1-xGdx)0.99Dy0.01]2O3 (x = 0 and 0.4) ceramics were successfully fabricated by vacuum sintering with in-line transmittances of 71.4%-75.1% at 574 nm, the wavelength of Dy3+ emission (the 4F9/2 → 6H13/2 transition). Substitution of Gd3+ for Y3+ significantly affected the photoluminescent and scintillant properties of the ceramics. Gd3+ addition effectively increased lattice parameters and theoretical densities of the ceramic samples; this contributed to enhanced X-ray absorption coefficients. Both [(Y1-xGdx)0.99Dy0.01]2O3 (x = 0 and 0.4) ceramics displayed typical Dy3+ emissions from 4F9/2 → 6HJ (J = 15/2, 13/2, 11/2) transitions under UV and X-ray irradiations. By incorporating Gd3+ into the lattice, a stronger excitation peak of Gd3+ due to internal f-f transitions relative to Dy3+ was observed at 276 nm; subsequent ceramics obtained a sharper PL intensity and a warmer hue via effective energy transfer from Gd3+ to Dy3+. Using a Gd3+ dopant generally reduces the total photoluminescence/photoluminescence excitation intensities and light output; it also delays the lifetime and afterglow of the transparent ceramics.  相似文献   

16.
Complexes of the type [Cu(CH3CN)4]+[A] ([A] = [B(C6F5)4] (1), [B{C6H3(CF3)2}] (2), [(C6F5)3B–C3H4N2–B(C6F5)3] (3)) are synthesized and characterized. Their utilization as catalysts in cyclopropanation and aziridination reactions of olefins, forming three membered rings is explored. The compounds are found to catalyze both reactions in moderate to good yields being the best results obtained with compound 1. The more weakly the nitrile ligands are coordinated to the metal center, the better is the catalytic performance of the catalyst.  相似文献   

17.
亚甲基蓝光度法研究基于CaO2的Fenton反应条件   总被引:1,自引:0,他引:1       下载免费PDF全文
刘娇  孟范平  王震宇  刘启元 《化工学报》2011,62(9):2520-2526
CaO2作为原位Fenton 氧化修复中H2O2持续供源的作用逐渐受到关注。利用亚甲基蓝分光光度法评价了基于CaO2的Fenton反应中催化剂种类、初始pH值、CaO2用量、催化剂和CaO2比例、磷酸缓冲溶液浓度对羟基自由基(HO·)产率的影响。结果表明,采用Fe2+作为催化剂,在pH值为4、CaO2相似文献   

18.
《分离科学与技术》2012,47(5):676-682
Using the liquid-liquid partition method, the formation of Co(II) complexes with 1-alkyl-2-methylimidazoles (where alkyl = C4H9 trough C12H25) at 25°C and at a fixed ionic strength of the aqueous phase (I = 0.5, (HL)NO3, KNO3) were studied. Dichloromethane, trichloromethane, and 2-ethylhexanol were used as diluents. The tetrahedral and octahedral complexes were formed. Stability constants (βn) of the complexes as well partition ratios (Pn) of the extracted species were determined. It was shown that both βn and Pn increased with an increasing 1-alkyl chain length. Tetrahedral complexes are more readily extractable by organic solvent. Their Pn values are the highest.  相似文献   

19.
The yields of gaseous products (H2, CO, CO2, and C n H2n + 2 at n = 1−4) from brown coal and brown coal-KOH compounds were determined under conditions of nonisothermal heating (4°C/min) to 800°C followed by an isothermal exposure (1 h, 800°C). It was found that, in the presence of the alkali, the yields of H2, CO, C2H6, and C3H8 increased; the yields of CO2 and CH4 decreased; and the formation of isobutane was completely suppressed. Changes in the gas compositions were explained by the alkali degradation of C-C bonds in the organic matter of coal and by the thermally initiated dehydrogenation and dealkylation reactions of arene and alkane structural fragments, in which KOH molecules served as H-atom donors in the formation of H2 and alkanes.  相似文献   

20.
Poly(di(ω-alkylphenyl)stannane)s, [Sn(C n H2n Ph)2] m with n = 2–4, and a copolymer of di(3-propylphenyl)stannane and dibutylstannane of weight-average molar masses of 2–8 · 104 g/mol were synthesized by dehydropolymerization of stannanes of the composition H2SnR2 using Wilkinson’s catalyst [RhCl(PPh3)3]. At least two methylene groups were required as spacers between the phenyl group and the tin atom for polymerization to occur. The polystannanes were characterized by, among other techniques, 1H, 13C and 119Sn NMR spectroscopy, thermal analysis and X-ray diffraction. The polymers featured properties different from those of the corresponding poly(dialkylstannane)s. Specifically, the [Sn(C n H2n Ph)2] m family displayed glass transitions at remarkably low temperatures, down to ca. −50 °C, and a lower value for a copolymer (−68 °C). Polymers [Sn(CnH2nPh)2]m with n = 2 and 3 and a copolymer at room temperature were of a gel-like concistence, which enabled facile orientation with shear forces. Finally, the temperature-dependent electrical conductivity was determined for poly(di(3-propylphenyl)stannane), which followed the law of typical semiconductors, with an activation energy for conduction of 0.12 eV.  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