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 343 毫秒
1.
采用三苯基膦羰基氢化铑作为催化剂,进行1-丁烯氢甲酰化合成戊醛反应,主要考察温度、铑浓度、配体浓度、丁烯浓度、合成气中H2和CO分压等因素对反应速率的影响.动力学研究表明温度、Rh浓度、丁烯浓度和H2分压的增加均可提高反应速度,CO分压和配体量的增加使反应速度降低.给出了RhH(CO)(PPh3)3催化1-丁烯氢甲酰化的反应动力学方程,并采用非线性最小二乘法对模型进行参数估值,计算值与实验值具有较好的一致性.  相似文献   

2.
新型双膦配体的合成及其在2-丁烯氢甲酰化反应中的应用   总被引:1,自引:0,他引:1  
合成了以联苯为骨架,以吲哚为取代基的双膦配体,并研究了该配体与Rh(acac)(CO)2原位生成的催化剂在2-丁烯氧甲酰化反应中的催化性能.考察了膦/铑比、反应温度、反应压力以及2-丁烯与Rh(acae)(CO)2摩尔比等因素对反应活性及区域选择性的影响.结果表明,在60℃反应时,醛的正异比高达28.5;当压力为2.0...  相似文献   

3.
 应用程序升温技术研究了氢甲酰化反应物CO,H2和C2H4在经PPh3修饰的Rh/SiO2(PPh3-Rh/SiO2)催化剂上的吸附-脱附行为. CO-TPD结果显示, Rh/SiO2催化剂在348, 398和525 K处有3个脱附峰, PPh3-Rh/SiO2催化剂仅在368 K处出现脱附峰,表明催化剂的吸附性能发生了明显变化. 采用原位红外光谱研究了PPh3-Rh/SiO2催化剂上CO的吸附态. 结果表明, 2040 cm-1处吸收峰归属于PPh3修饰的Rh粒子上线式吸附的CO. 这种吸附态既不同于Rh/SiO2多相催化剂表面Rh粒子上的CO吸附态,也不同于相应均相催化剂中的羰基配位态. TPD和FT-IR结果表明,在低压下PPh3-Rh/SiO2催化剂对氢甲酰化反应已具有相当的催化性能.  相似文献   

4.
以4,4'-二羟基二苯丙烷和2,4-二叔丁基苯酚为原料合成了一种新型双膦亚磷酸酯配体,并用此配体和Rh(acac)(CO)2原位形成的催化体系催化1-己烯的氢甲酰化反应.系统考察了反应温度、压力、P/Rh和溶剂四种反应参数对催化体系的催化性能影响.选择了最佳的反应条件,在铑浓度为0.75×10-3mol/L、P/Rh比为10、温度100℃、压力(H2/CO=1)2.0MPa的条件下反应1.0h,在溶剂甲苯中1-己烯的转化率可达到100%,醛选择性为98.7%,TOF为3498.6h-1.在相同的条件下与以三苯基膦和单膦亚磷酸三(2,4-二叔丁基苯基)酯为配体的铑催化剂相比较,以新型双膦亚磷酸酯为配体的铑催化剂的催化活性是PPh3的1.6倍,而与亚磷酸三(2,4-二叔丁基苯基)酯的催化活性相当.  相似文献   

5.
铑膦络合物是最常用的烯烃氢甲酰化反应的催化剂,目前改进这类催化剂的活性和选择性大多从选用新型膦配体着手,而很少研究改变铑络合物本身的配位环境.Rh(acac)(CO)_2等一价铑络合物是人们熟知的氢甲酰化催化剂,而类似结构的Rh(ON)(CO)_2型络合物(ON为氮氧阴离子配体)催化剂则尚未见报道.本文研究了Rh(ON)(CO)_2与膦配体组成的体系在常压下对烯烃氢甲酰化反应的催化作用.所用三种Rh(ON)(CO)_2型络合物的结构如下:  相似文献   

6.
田密  李海峰  王来来 《催化学报》2018,39(10):1646-1652
双环戊二烯(DCPD)是石脑油和燃料油裂解蒸汽的C5馏分中最重要的组分之一.DCPD经氢甲酰化反应可转化为具有广泛应用前景的三环癸烷不饱和单甲醛(TCDMA)和三环癸烷二甲醛(TCDDA),并可通过还原或胺化进一步转化为相应的醇和胺类化合物,用于农药、医药、润滑油和香料等的合成.但是,由于其分子结构中含有3,4-位和8,9-位两种不同活性的不饱和双键,因此DCPD氢甲酰化反应的产物通常非常复杂.过去数十年,研究者们为此相继开发了高转化率和高选择性的催化体系.但是反应条件相对都比较苛刻,尤其是对于双醛TCDDA的合成,通常需要较高的反应温度和反应压力以及大量的催化剂.本文以2'-联萘位置含有不同酯取代基(OCOMe,OCOPh,OCOAdamantyl和OCOPhCl)的三-H8-联萘单齿亚磷酸酯L1-L4为配体,以不同价态的金属铑前驱体为催化剂,开发了Rh催化DCPD氢甲酰化反应的新体系,并对亚磷酸酯配体、不同价态的金属铑催化剂前驱体、反应温度、反应时间、溶剂以及不同的底物和催化剂的S/C摩尔比对DCPD转化率和TCDDA选择性的影响进行了深入的研究.结果表明,当以金属铑前驱体Rh(acac)(CO)_2和配体L4-OCOPhCl为催化体系时,在DCPD氢甲酰化反应中表现出很高的活性,尤其是当S/C=4000时,TON值达到3286,并且该催化体系对于双醛TCDDA具有良好的选择性.值得注意的是,在相对温和的条件(6 MPa,120℃)下,Rh(Ⅰ)催化剂与氯苯酯基取代的三-H_8-联萘单齿亚磷酸酯所形成的配合物在催化DCPD的氢甲酰化反应中,DCPD的转化率达到99.9%,而双醛TCDDA的选择性达到98.7%.此外,我们采用L4-OCOPhCl作为模型单齿磷酸配体,在溶液中通过NMR对可能形成的Rh(Ⅰ)/亚磷酸酯催化物种进行了深入的考察.~(13)P NMR谱图表明,在DCPD的氢甲酰化反应中,催化物种[Rh(acac)(CO)(L4-OCOPhCl)]比[Rh(acac)(CO)(L2-OCOPh)]具有更好的稳定性,而且只有体积较大的配体L4-OCOPhCl才能与铑前驱体Rh(acac)(CO)_2进行很好的配位.  相似文献   

7.
 利用有机-无机杂化的概念,以三苯基膦直接修饰Rh/SiO2制备了PPh3-Rh/SiO2多相催化剂. 在浆态床烯烃氢甲酰化反应中,该催化剂在10 MPa,373 K温和条件下的活性和选择性远高于Rh/SiO2的活性和选择性,与相应的均相催化剂HRhCO(PPh3)3的性能相当,且具有易分离的优点. 31P MAS NMR和XPS技术表征结果表明,催化剂中的配体PPh3与高度分散的Rh之间存在配位作用,形成了兼具多相和均相催化性能的有机-无机杂化催化剂. 该催化剂对不同碳数烯烃的氢甲酰化反应都具有较好的催化性能.  相似文献   

8.
研究了由 Rh(acac) (CO) 2 和 1 ,4-双 (二苯膦基 )苯 (简称 DPPB)组成的体系对烯烃氢甲酰化反应的催化作用 .考察了反应温度、压力、P/Rh比、催化剂浓度等对 1 -己烯氢甲酰化反应的影响 ,得到了较适宜的反应条件 .在相同条件下比较了该体系催化 1 -己烯、 1 -辛烯、 1 -十二烯的氢甲酰化反应的性能 .结果表明 ,随着底物碳链的增长 ,反应活性呈提高趋势 .实验证明 ,DPPB-铑配合物对苯乙烯的氢甲酰化反应有较高的催化活性和选择性 .  相似文献   

9.
 以具有不同电子结构的有机膦直接修饰Rh/SiO2制备了有机-无机杂化L-Rh/SiO2催化剂,并考察了催化剂对1-己烯氢甲酰化反应的催化性能. 结果表明, P(OPh)3-Rh/SiO2催化剂上1-己烯氢甲酰化反应的TOF可高达4111 h-1,而PCy3-Rh/SiO2催化剂上生成醛的选择性为100%. 不同催化剂活性的顺序为P(OPh)3-Rh/SiO2>PPh3-Rh/SiO2>PCy3-Rh/SiO2,其选择性的顺序则与之相反. TG表征结果表明,有机膦与Rh之间相互作用强度的顺序为PCy3-Rh/SiO2>PPh3-Rh/SiO2>P(OPh)3-Rh/SiO2. 因此,有机膦给电子的能力越强,则与Rh之间相互作用的能力越强,催化剂体系越稳定,1-己烯氢甲酰化反应的活性越低,而庚醛的选择性越高.  相似文献   

10.
黄唯平  李君 《分子催化》2016,30(6):505-514
为了调控含碱性基团功能烯烃的催化氢甲酰化反应,设计合成了不同表面酸性的钛氧纳米管和Zr-掺杂钛氧纳米管负载Rh催化剂(Rh/TNTs,Rh/Zr-TNTs)。用XRD、XPS、FT-IR、TEM和低温N2吸脱附等对所合成的催化剂进行了结构及组成表征。催化剂没有显现出与Rh和Zr相关的XRD衍射峰,Rh在纳米管中高度分散;Zr-掺杂钛氧纳米管的比表面积比纯钛氧纳米管的要高,催化剂的比表面积随着载Rh量的增加而减小,其中Rh0百分率也会降低,催化剂中Rh能对CO进行化学吸附。所合成的催化剂能有效地催化氰基烯烃氢甲酰化,催化活性随载Rh量而变化,最佳值为0.13 w;提高催化温度虽能增加2-甲基-3-丁烯腈的转化率,但也有利于它的异构化;Rh/Zr-TNTs表面更强的B酸性有利于催化反应生成直链产物醛,负载催化剂能表现出良好的协同作用。  相似文献   

11.
The rhodium-phosphine complex catalyst Rh(CO)(acac)(PPh3)(Ⅰ) for 1-hexene hydroformylation was studied under the following reaction conditions: CO/H2=1(mole rate), pressure 1.0 MPa, temperature 25-120℃, by using the pressurized in-situ 1H NMR technique. Experimental results indicated that the formation of a rhodium hydride complex from (Ⅰ) began at room temperature and its amount increased with increasing of reaction temperature. This intermediate complex began to decompose at 100℃ and disapeared completely at 120℃. The intensity change of the proton signal was parallel to catalytical activity in hydroformylation of olefins. Under pure CO pressure the proton signal of Ph-H bond was not observed. There was a 0.2 ppm difference in proton chemical shifts of Rh-H bond under pure H2 pressure and under H2+CO pressure. The results showed that the rhodium-hydride carbonyl complex is the active intermediate in the industrial hydroformylation process.  相似文献   

12.
IntroductionHeterogenization of HRh( CO) ( PPh3 ) 3 ( PPh3 :triphenylphosphine) for hydroformylation andselective hydrogenation has received considerableattention in the past several decades from bothacademic and industrial interests[1] . Besides themilestone preparation of water- soluble TPPTS andthe corresponding rhodium- phosphine complexessuch as HRh( CO) ( TPPTS) 3 ( 1 ) [2 ,3 ] ,supportedliquid- phase catalysts ( SLPC) [4— 6] and supportedaqueous- phase catalysts ( SAPC) [7]…  相似文献   

13.
The coordination abilities of the novel N,N'-diphosphino-silanediamine ligand of formula SiMe(2)(NtolPPh(2))(2) (SiNP, 1) have been investigated toward rhodium, and the derivatives [RhCl(SiNP)](2) (2), [Rh(SiNP)(COD)][BF(4)] (3), and Rh(acac)(SiNP) (4) have been synthesized. The stability of the dinuclear frame of [RhCl(SiNP)](2) (2) toward incoming nucleophiles has been shown to be dependent on their π-acceptor ability. Indeed, the mononuclear complexes RhCl(SiNP)(L) (L = CO, 5; CN(t)Bu, 6) have been isolated purely and quantitatively upon reaction of 2 with CO and CN(t)Bu, respectively. Otherwise, PPh(3) and RhCl(SiNP) equilibrate with Rh(Cl)(SiNP)(PPh(3)) (7). Carbon electrophiles such as MeI and 3-chloro-1-proprene afforded the oxidation of rhodium(I) to rhodium(III) and the formation of RhCl(2)(η(3)-C(3)H(5))(SiNP) (8) and Rh(Me)(I)(SiNP)(acac) (10), respectively. The methyl derivative 10 is thermally stable and does not react either with CO or with CN(t)Bu even in excess. Otherwise, RhCl(2)(η(3)-C(3)H(5))(SiNP) (8) is thermally stable but reacts with CO, affording 3-chloro-1-proprene and RhCl(SiNP)(CO) (5). Finally, upon reaction of Rh(acac)(SiNP) (4) and 3-chloro-1-proprene, RhCl(acac)(η(1)-C(3)H(5))(SiNP) (9a) and [Rh(acac)(η(3)-C(3)H(5))(SiNP)]Cl (9b) could be detected at 233 K. At higher temperatures, 9a and 9b smoothly decompose, affording the dinuclear derivative [RhCl(SiNP)](2) (2) and the CC coupling product 3-allylpentane-2,4-dione.  相似文献   

14.
Six calix[4]arenes each bearing two non-cyclic PR2 units attached at distal phenolic oxygen atoms, p-Bu t-calix[4]arene-25,27-(OPR2)2-26,28-(OR')2(R = OPh; R'= Prn, L1; R = OPh; R'= CH2CO2Et, L2; R= OPh; R'= CO2 cholesteryl, L3; R = Ph; R'= Prn, 4; R = Ph; R'= CH2CO2Et, L5; R = Ph; R'= CO2cholesteryl, L6) have been synthesized and their coordinative properties investigated. The diphosphites L1-L3, where the P centres are separated by 12 bonds, readily form chelate complexes provided the complexation reaction is achieved either by using a starting complex that possesses good leaving groups or by operating under high dilution in order to avoid oligomer formation. Thus, the cationic complexes [Rh(COD)L1]BF4 and [Rh(COD)L3]BF4 were both formed in high yield by reacting the appropriate diphosphite with either [Rh(COD)(THF)2]BF4 or [Rh(COD)2]BF4. At high dilution, reaction of L3 with the neutral complex [PdCl2(COD)] afforded the chelate complex [PdCl2L3] in 90% yield. The reaction of one equiv. of L1 with [Rh(acac)(CO)2] resulted in the formation of [Rh(acac)L1] without requiring high dilution conditions. When the latter reaction was carried out with 0.5 equiv. of L1, the bimetallic complex [{Rh(acac)(CO)}2(eta]1-P,eta1-P'-L1)] was formed instead. Reaction at high dilution of with the cyclometallated complex [Pd(o-C6H4CH2NMe2)(THF)2]BF4 gave the expected chelate complex [Pd(o-C6H4CH2NMe2)]BF4. The latter slowly converts in solution to an oligomer in which the ligand behaves as a (eta1-P,eta1-P') bridging ligand, thus leading to a less strained structure. All six ligands, when mixed with [Rh(acac)CO2], effectively catalyse the hydroformylation of octene and styrene. In the hydroformylation of octene, the linear aldehyde selectivities observed with L2 and L3 are significantly higher (linear : branched =ca. 10) than those obtained with the other 4 ligands of this study and also with respect to PPh3. Molecular modelling shows that the lower rim substituents of and form tighter pockets about the metal centre than do the other ligands and so sterically favour the formation of Rh(n-alkyl) intermediates over that of Rh(i-alkyl) ones. In styrene hydroformylation, all ligands result in the formation of unusually high amounts of the linear aldehyde, the b : l ratios being all close to 65 : 35. The highest activities were found when using an L/Rh ratio of 1/1.  相似文献   

15.
The two rhodium complexes [Rh(acac)(L(R))] (L(R)=(S,S)-5,11,17,23-tetra-tert-butyl-25,27-di(OR)-26,28-bis(1,1'-binaphthyl-2,2'-dioxyphosphanyloxy)calix[4]arene; 6: R=benzyl, 7: R=fluorenyl), each based on a hemispherical chelator forming a pocket about the metal centre upon chelation, are active in the hydroformylation of 1-octene and styrene. As expected for this family of diphosphanes, both complexes resulted in remarkably high selectivity towards the linear aldehyde in the hydroformylation of 1-octene (l/b≈15 for both complexes). Linear aldehyde selectivity was also observed when using styrene, but surprisingly only 6 displayed a marked preference for the linear product (l/b=12.4 (6) vs. 1.9 (7)). A detailed study of the structure of the complexes under CO or CO/H(2) in toluene was performed by high-pressure NMR (HP NMR) and FT-IR (HP-IR) spectroscopies. The spectroscopic data revealed that treatment of 6 with CO gave [Rh(acac)(CO)(η(1)-L(benzyl))] (8), in which the diphosphite behaves as a unidentate ligand. Subsequent addition of H(2) to the solution resulted in a well-defined chelate complex with the formula [RhH(CO)(2)(L(benzyl))] (9). Unlike 6, treatment of complex 7 with CO only led to ligand dissociation and concomitant formation of [Rh(acac)(CO)(2)], but upon addition of H(2) a chelate complex analogous to 9 was formed quantitatively. In both [RhH(CO)(2)(L(R))] complexes the diphosphite adopts the bis-equatorial coordination mode, a structural feature known to favour the formation of linear aldehydes. As revealed by variable-temperature NMR spectroscopy, these complexes show the typical fluxionality of trigonal bipyramidal [RhH(CO)(2)(diphosphane)] complexes. The lower linear selectivity of 7 versus 6 in the hydroformylation of styrene was assigned to steric effects, due to the pocket in which the catalysis takes place being less adapted to accommodate CO or larger olefins and, therefore, possibly leading to facile ligand decoordination. This interpretation was corroborated by an X-ray structure determination carried out for 7.  相似文献   

16.
The heteroscorpionate ligands [HB(taz)(2)(pz(R))](-) (pz(R) = pz, pz(Me2), pz(Ph)) and [HB(taz)(pz)(2)](-), synthesised from the appropriate potassium hydrotris(pyrazolyl)borate salt and 4-ethyl-3-methyl-5-thioxo-1,2,4-triazole (Htaz), react with [{Rh(cod)(μ-Cl)}(2)] to give [Rh(cod)Tx] {Tx = HB(taz)(2)(pz), HB(taz)(2)(pz(Me2)), HB(taz)(2)(pz(Ph)), HB(taz)(pz)(2)}; the heteroscorpionate rhodaboratrane [Rh{B(taz)(2)(pz(Me2))}{HB(taz)(2)(pz(Me2))}] is the only isolable product from the reaction of [{Rh(nbd)(μ-Cl)}(2)] with K[HB(taz)(2)(pz(Me2))]. Carbonylation of the cod complexes gave a mixture of [Rh(CO)(2)Tx] and [(RhTx)(2)(μ-CO)(3)] which reacts with PR(3) to give [Rh(CO)(PR(3))Tx] (R = Cy, NMe(2), Ph, OPh). In the solid state the complexes are square planar with the particular structure dependent on the steric and/or electronic properties of the scorpionate and ancillary ligands. The complex [Rh(cod){HB(taz)(pz)(2)}] has the heteroscorpionate κ(2)[N(2)]-coordinated to rhodium with the B-H bond directed away from the rhodium square plane while [Rh(cod){HB(taz)(2)(pz(Me2))}] is κ(2)[SN]-coordinated, with the B-H bond directed towards the metal. The complexes [Rh(CO)(PPh(3)){HB(taz)(2)(pz)}] and [Rh(CO)(PPh(3)){HB(taz)(2)(pz(Me2))}] are also κ(2)[SN]-coordinated but with the pyrazolyl ring cis to PPh(3); in the former the B-H bond is directed towards rhodium while in the latter the ring is pseudo-parallel to the rhodium square plane, as also found for [Rh(CO)(2){HB(taz)(2)(pz(Me2))}]. The analogues [Rh(CO)(PR(3)){HB(taz)(2)(pz(Me2))}] (R = Cy, NMe(2)) have the phosphines trans to the pyrazolyl ring. Uniquely, [Rh(CO)(PPh(3)){HB(taz)(2)(pz(Ph))}] is κ(2)[S(2)]-coordinated. A qualitative mechanism is given for the rapid ring-exchange, and hence isomerisation, observed in solution.  相似文献   

17.
Rhodium and iridium complexes bearing a tridentate [PEP] type ligand ([PEP] = {o-(Ph(2)P)C(6)H(4)}(2)E(Me); E = Ge or Sn) were synthesized through the phosphine exchange reaction accompanied by selective E-C bond cleavage. The ligand precursors {o-(Ph(2)P)C(6)H(4)}(2)EMe(2) (E = Ge or Sn) were readily obtained in excellent yields by treating {o-(Ph(2)P)C(6)H(4)}(2)Li with 0.5 equivalents of Me(2)ECl(2). Tris(triphenylphosphine)rhodium(i) carbonyl hydride M(H)(CO)(PPh(3))(3) (M = Rh, Ir) cleaved one of the E-Me bonds of {o-(Ph(2)P)C(6)H(4)}(2)EMe(2) exclusively to afford the trigonal bipyramidal (TBP) complexes, [PEP]M(CO)(PPh(3)). Square-planar rhodium complexes [PEP]Rh(PPh(3)) were also prepared from the reactions of tetrakis(triphenylphosphine)rhodium(i) hydride Rh(H)(PPh(3))(4) with {o-(Ph(2)P)C(6)H(4)}(2)EMe(2). Further, the trans influence of group 14 elements E (E = Si, Ge, Sn) in [PEP]Rh(PPh(3)) is discussed in terms of the (1)J(Rh-P) coupling constants, indicating that E exhibited a stronger trans labilizing effect in the order Sn < Ge < Si.  相似文献   

18.
长链烷基二苯基膦—铑配合物催化烯烃氢甲酰化反应研究   总被引:1,自引:0,他引:1  
陈骏如  陈华等 《分子催化》2001,15(6):413-415
研究了烷基二苯基膦-铑配合物RhCl(CO0(n-C8H17PPh2)2(1)和RhCl(Co)(n-C12H25PPh2)2(2)对1-辛烯氢甲酰化反应的催化性能。结果表明,配合物1比2具有更高的催化活性,而配合物2对生成正构醛的选择性更好;当催化剂浓度或膦/铑比增加时,配合物2催化成正构醛的选择性呈下降趋势,显示出与以PPh3为配体时的不同的性能。  相似文献   

19.
The treatment of [{Rh(μ-SH){P(OPh)(3)}(2)}(2)] with [{M(μ-Cl)(diolef)}(2)] (diolef=diolefin) in the presence of NEt(3) affords the hydrido-sulfido clusters [Rh(3)(μ-H)(μ(3)-S)(2)(diolef){P(OPh)(3)}(4)] (diolef=1,5-cyclooctadiene (cod) for 1, 2,5-norbornadiene (nbd) for 2, and tetrafluorobenzo[5,6]bicyclo[2.2.2]octa-2,5,7-triene (tfb) for 3) and [Rh(2)Ir(μ-H)(μ(3)-S)(2)(cod){P(OPh)(3)}(4)] (4). Cluster 1 can be also obtained by treating [{Rh(μ-SH){P(OPh)(3)}(2)}(2)] with [{Rh(μ-OMe)(cod)}(2)], although the main product of the reaction with [{Ir(μ-OMe)(cod)}(2)] was [RhIr(2)(μ-H)(μ(3)-S)(2)(cod)(2){P(OPh)(3)}(2)] (5). The molecular structures of clusters 1 and 4 have been determined by X-ray diffraction methods. The deprotonation of a hydrosulfido ligand in [{Rh(μ-SH)(CO)(PPh(3))}(2)] by [M(acac)(diolef)] (acac=acetylacetonate) results in the formation of hydrido-sulfido clusters [Rh(3)(μ-H)(μ(3)-S)(2)(CO)(2) (diolef)(PPh(3))(2)] (diolef=cod for 6, nbd for 7) and [Rh(2)Ir(μ-H)(μ(3)-S)(2)(CO)(2)(cod)(PPh(3))(2)] (8). Clusters 1-3 and 5 exist in solution as two interconverting isomers with the bridging hydride ligand at different edges. Cluster 8 exists as three isomers that arise from the disposition of the PPh(3) ligands in the cluster (cis and trans) and the location of the hydride ligand. The dynamic behaviour of clusters with bulky triphenylphosphite ligands, which involves hydrogen migration from rhodium to sulfur with a switch from hydride to proton character, is significant to understand hydrogen diffusion on the surface of metal sulfide hydrotreating catalysts.  相似文献   

20.
In immobilizing the rhodium complexes [Rh(acac)(CO)(P)] (1) and [Rh(acac)(P)2] (2) (P = Ph2PCH2CH2Si(OMe)3) onto SiO2, acetylacetone is found to be released through protonation of the acac ligand by the acidic silica-OH groups. The resulting complexes [Rh(O-{SiO2}(HO-{SiO2})(CO)(P-{SiO2})] (1a) and [Rh(O-{SiO2})(HO-{SiO2})(P-{SiO2})2] (2a) were successfully tested with respect to their catalytic action on 1-hexene hydroformylation as well as benzene and toluene hydrogenation. The reaction outcome, viz. the formation of aldehydes versus isomerization, depends strongly on the presence and concentration of a phosphine co-catalyst. Thus, while 1a gave only a 17% yield of aldehyde in the absence of phosphines, the yield is increased to 54% in the presence of phosphinated silica P-{SiO2} or even 94% if PPh3 is added to the solution. Without extra added phosphine, both 1a and 2a effect mainly the isomerization of 1-hexene to 2-hexene. Pre-catalyst 1a catalyzes also the hydrogenation of benzene at 10.5 atm H2 and 90 °C to give cyclohexane with a TOF of 608 h−1.  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