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 203 毫秒
1.
刘龑  王正  丁奎岭 《化学学报》2012,(13):1464-1470
本文研究了一类结构可调手性单齿亚磷酰胺配体DpenPhos在Rh(I)催化的E和Z型β-脱氢氨基酸酯的不对称催化氢化反应中的应用.经过系统的反应条件和配体结构优化,发现N原子上含有H的亚磷酰胺配体与Rh(I)形成的催化剂通常比N原子上不含H的配体表现出更高的反应活性.在E型β-脱氢氨基酸酯的不对称氢化反应中,催化剂(R,R)-3k/Rh(I)表现突出,可以实现底物的常压催化氢化,取得了92%~96%的对映选择性,催化剂用量可降低至0.2 mol%;对于Z型β-脱氢氨基酸酯的不对称氢化反应,则(R,R)-3l/Rh(I)为最优催化剂,可以获得92%~98%ee值的氢化产物,特别是对于β-芳基取代衍生物的氢化反应,相应氢化产物的ee值可以达到96%~98%.该类催化剂为天然或非天然光学活性β氨基酸的合成提供了一个简便、高效的方法.  相似文献   

2.
采用共沉淀法制备了水溶性聚合物修饰的苯选择加氢制环己烯Ru-Zn催化剂, 并用X射线衍射、 透射电镜、 X射线能量色散谱、 X射线光电子能谱和氮气物理吸附等对加氢后催化剂进行了表征. 结果表明, 水溶性聚合物的种类和聚乙二醇-20000(PEG-20000)的用量对Ru-Zn催化剂微晶尺寸有显著影响. 在ZnSO4存在下, 随着Ru-Zn催化剂Ru微晶尺寸增加, 苯转化率降低, 环己烯最高收率则呈火山型变化趋势. 用0.4 g PEG-20000修饰的Ru-Zn催化剂[m(PEG-20000)∶m(Ru)=0.2]Ru的微晶尺寸为4.8 nm, 环己烯最高收率为62.2%. Ru微晶尺寸影响催化剂表面的Zn/Ru原子比, 进而影响Ru-Zn催化剂性能.  相似文献   

3.
手性环状氨基酸具有独特的结构和特殊的性能,是广泛存在于手性药物和手性催化剂中的一种结构单元.过渡金属催化的不对称氢化具有高效、绿色和经济的优点而成为合成该类化合物的首选方法.对该领域进行了综述,介绍了各种类型环状α-和β-脱氢氨基酸及其衍生物的不对称催化氢化研究现状,分析了各种类型钌(Ru)、铑(Rh)和铱(Ir)催化剂的优势和不足,并提出了该领域今后的研究重点和发展方向.  相似文献   

4.
碳纳米管担载纳米Ir催化生物质基乙酰丙酸合成γ-戊内酯   总被引:1,自引:0,他引:1  
以碳纳米管(CNTs)担载Ir纳米粒子为催化剂进行生物质基平台化合物乙酰丙酸(LA)选择加氢制备γ-戊内酯(GVL)的研究,并利用X射线衍射、X射线光电子能谱和透射电镜表征了使用前后的Ir/CNT催化剂,探讨了影响LA催化加氢制GVL反应性能的因素和该反应的可能路径.结果表明,与Ru,Rh和Pd等传统铂族金属相比,Ir/CNT催化剂不但可在温和条件下(50℃,2.0 MPa,H2)实现LA至GVL的完全转化,且可对多类直接源于生物质水解的含等量LA/甲酸的“真实”体系实现GVL的高效选择合成.  相似文献   

5.
Ru, Sn和Co促进的Pt/C催化剂电催化氧化甲醇的性能   总被引:4,自引:0,他引:4  
 采用溶胶法制备了用于阴离子膜直接甲醇燃料电池的Pt-M/C(M=Ru, Sn和Co)阳极电催化剂,并用X射线衍射和X射线能谱技术对催化剂进行了表征. 结果表明,制备的Pt-M合金颗粒分布均匀,粒径为2~6 nm, 其组成与前驱体中相应金属的原子比基本吻合. 用循环伏安法测定了催化剂在不同碱性条件下的活性. 结果显示,随着碱性的增加,甲醇的起始氧化电位降低,峰电流和催化剂的活性增大; 相同碱强度下催化剂活性顺序为Pt50Ru50/C>Pt50Sn50/C>Pt75Co25/C, 添加Ru可明显提高Pt/C催化剂的活性. Pt50Ru50/C在1.0 mol/L NaOH+1.0 mol/L CH3OH溶液中的峰电流密度可达634.7 mA/mg.  相似文献   

6.
《化学学报》2012,70(13)
本文研究了一类结构可调手性单齿亚磷酰胺配体DpenPhos在Rh(I)催化的E和Z型β-脱氢氨基酸酯的不对称催化氢化反应中的应用.经过系统的反应条件和配体结构优化,发现N原子上含有H的亚磷酰胺配体与Rh(I)形成的催化剂通常比N原子上不含H的配体表现出更高的反应活性.在E型β-脱氢氨基酸酯的不对称氢化反应中,催化剂(R,R)-3k/Rh(I)表现突出,可以实现底物的常压催化氢化,取得了92%~96%的对映选择性,催化剂用量可降低至0.2 mol%;对于Z型β-脱氢氨基酸酯的不对称氢化反应,则(R,R)-3l/Rh(I)为最优催化剂,可以获得92%~98%ee值的氢化产物,特别是对于β-芳基取代衍生物的氢化反应,相应氢化产物的ee值可以达到96%~98%.该类催化剂为天然或非天然光学活性β氨基酸的合成提供了一个简便、高效的方法.  相似文献   

7.
采用水热技术制备了Zn掺杂的C/Nb_2O_5纳米催化剂.通过X射线衍射(XRD)、透射电子显微镜(TEM)和X射线光电子能谱(XPS)对催化剂进行了表征.结果表明,催化剂呈现较高的结晶度和较均匀的形貌,Zn元素以氧化锌的形式随机分布在C/Nb_2O_5纳米催化剂的表面,当锌/铌摩尔比为10%时,催化剂表面氧空穴(SOVs)的量会达到一个最优值.对染料罗丹明B(Rh B)和罗丹明6G(Rh6G)在可见光下的降解实验结果表明,与商业P25,Nb_2O_5以及C/Nb_2O_5相比,适量Zn掺杂的C/Nb_2O_5纳米催化剂具有更好的光催化活性.  相似文献   

8.
负载型纳米贵金属催化剂催化吡啶及其衍生物的加氢反应   总被引:2,自引:0,他引:2  
薛芳  林棋  杨朝芬  李贤均  陈华 《催化学报》2006,27(10):921-926
 制备了负载型高分散的纳米贵金属催化剂和含Ru的双金属催化剂,并考察了催化剂对吡啶及其衍生物加氢反应的催化性能. 结果表明, 5%Ru/C催化剂对吡啶加氢反应的催化活性高于5%Pd/C, 5%Pt/C和5%Ir/C. 在100 ℃, 3.0 MPa, 1 h和Ru/吡啶摩尔比=2.5/1000 的条件下, 5%Ru/C催化吡啶加氢的转化率大于99.9%, 生成哌啶的选择性为100%. 催化剂重复使用5次后,活性和选择性无明显下降. 在Ru催化剂中加入少量的Pd和Ir后催化剂活性没有明显的变化. 采用X射线衍射、高分辨透射电镜和X射线光电子能谱对还原后的5%Ru/C催化剂进行表征,结果表明Ru以高分散金属态存在,其平均粒径小于5 nm. 不同底物的加氢反应活性为: 吡啶≈2-甲基吡啶>2,6-二甲基吡啶>3-甲基吡啶>4-甲基吡啶>3,5-二甲基吡啶>2-甲氧基吡啶.  相似文献   

9.
高分散 Ru/MMT 催化剂的制备及其催化喹啉加氢性能   总被引:2,自引:0,他引:2  
 通过简单的离子交换法制备出高分散的蒙脱土 (MMT) 负载 Ru 催化剂, 采用 X 射线衍射、X 射线光电子能谱、程序升温还原和高分辨透射电子显微镜等手段对催化剂进行了表征. 结果表明, 金属 Ru 在蒙脱土层间高度分散, Ru 的平均粒径约 2 nm. 在喹啉加氢反应中, 该催化剂显示出很高的反应活性和选择性. 在 2 MPa 和 60 °C 的温和条件下, 以水为溶剂时, Ru/MMT 催化喹啉加氢生成 1,2,3,4-四氢喹啉的选择性高于 96.4%, 喹啉转化率达 99.2%. 当温度升高到 140 °C、压力增加到 3 MPa 时, 不需要补加催化剂就可以将喹啉一步加氢生成十氢喹啉, 选择性高达 98.1%.  相似文献   

10.
用沉淀法制备了单金属纳米Ru(0)催化剂,考察了ZnSO4和La2O3作共修饰剂对该催化剂催化苯选择加氢制环己烯性能的影响,并用X射线衍射(XRD)、X射线荧光(XRF)光谱、X射线光电子能谱(XPS)、俄歇电子能谱(AES)、透射电镜(TEM)和N2物理吸附等手段对加氢前后催化剂进行了表征.结果表明,在ZnSO4存在下,随着添加碱性La2O3量的增加,ZnSO4水解生成的(Zn(OH)2)3(ZnSO4)(H2O)x(x=1,3)盐量增加,催化剂活性单调降低,环己烯选择性单调升高.当La2O3/Ru物质的量比为0.075时,Ru催化剂上苯转化率为77.6%,环己烯选择性和收率分别为75.2%和58.4%.且该催化体系具有良好的重复使用性能.传质计算结果表明,苯、环己烯和氢气的液-固扩散限制和孔内扩散限制都可忽略.因此,高环己烯选择性和收率的获得不能简单归结为物理效应,而与催化剂的结构和催化体系密切相关.根据实验结果,我们推测在化学吸附有(Zn(OH)2)3(ZnSO4)(H2O)x(x=1,3)盐的Ru(0)催化剂有两种活化苯的活性位:Ru0和Zn2+.因为Zn2+将部分电子转移给了Ru,Zn2+活化苯的能力比Ru0弱.同时由于Ru和Zn2+的原子半径接近,Zn2+可以覆盖一部分Ru0活性位,导致解离H2的Ru0活性位减少.这导致了Zn2+上活化的苯只能加氢生成环己烯和Ru(0)催化剂活性的降低.本文利用双活性位模型来解释Ru基催化剂上的苯加氢反应,并用Hückel分子轨道理论说明了该模型的合理性.  相似文献   

11.
Poly(propylene imine) dendrimer (G2-PPI) terminated by nitrogen-containing triolefinic macrocycle on the periphery (G2-M) was synthesized by a nucleophilic substitution reaction. The structure and composition of G2-M were characterized by FT-IR, 1H-NMR and elemental analysis. The Rh3+ dendrimer-stabilized catalyst (G2-M(Rh3+)) was prepared by utilizing G2-M as stabilizer and analyzed by UV–Vis spectroscopy, 1H-NMR spectrometry, XPS and XRD. G2-M(Rh3+) demonstrates excellent catalytic activity and selectivity for the hydrogenation of nitrile-butadiene rubber (NBR) and styrene-butadiene rubber (SBR), and Rh residue contents for HNBR and HSBR are only 35 and 13 ppm, respectively, without any post treatment.  相似文献   

12.
Addressed herein is the 20+ year-old question of whether the true benzene and cyclohexene hydrogenation catalysts derived from the organometallic precursor [Rh(eta5-C5Me5)Cl2]2, 1, are homogeneous or heterogeneous. The methodology employed is that developed earlier (Lin, Y.; Finke, R. G. Inorg Chem. 1994, 33, 4891, "A More General Approach to Distinguishing Homogeneous from Heterogeneous Catalysis..."). The kinetic evidence especially, but also the metal product (nanoclusters plus bulk metal), Hg0 poisoning and other experiments, provide compelling evidence that Rh0 nanoclusters are the true benzene hydrogenation heterogeneous catalyst derived from [Rh(eta5-C5Me5)Cl2]2, 1, at the required more vigorous conditions of 50-100 degrees C and 50 atm H2. However, the same methods reveal that the cyclohexene hydrogenation catalyst derived from 1 at the milder conditions of 22 degrees C and 3.7 atm H2 is a nonnanocluster, homogeneous catalyst, most likely the previously identified complex, [Rh(eta5-C5Me5)(H)2(solvent)] (Gill, D. S.; White, C.; Maitlis, P. M J. C. S. Dalton Trans. 1978, 617). In short, the present results solve the two-decade-old problem of identifying the true benzene and cyclohexene hydrogenation catalysts derived from [Rh(eta5-C5Me5)Cl2]2. Perhaps most significant is the demonstration that the methodology employed has the ability to identify both heterogeneous and homogeneous catalysts from the same catalyst precursor.  相似文献   

13.
采用多元醇还原法将2.4~5.4 nm范围内粒径均一、尺寸可控的Ru纳米粒子负载在ZrO2上,研究了Ru的粒径对Ru/ZrO2催化剂上苯部分加氢性能的影响.采用紫外-可见吸收光谱(UV-Vis)、N2物理吸附、H2化学吸附、H2-程序升温脱附(H2-TPD)、粉末X射线衍射(XRD)、透射电子显微镜(TEM)和X射线光电子能谱(XPS)等手段对催化剂进行了系统的表征.研究表明,用于还原的醇的种类及添加剂乙酸钠的浓度对Ru粒径有显著影响.在苯部分加氢反应中,Ru/ZrO2催化剂有明显的粒径效应.随着Ru粒径的增大,苯的转换频率(TOF)提高,环己烯初始选择性(S0)则呈火山型变化趋势,选择性最高时的Ru粒径为4.4 nm.1,2-丙二醇还原得到的Ru/ZrO2催化剂上S0及环己烯得率最高,分别可达82%和39%.结合催化剂的表征和加氢结果,讨论了Ru粒径影响苯部分加氢活性和选择性的原因.  相似文献   

14.
Highly enantioselective hydrogenation of beta-alkyl-substituted (E)-beta-(acylamino)-acrylates catalyzed by Ru((R)-Xyl-P-Phos)(C(6)H(6))Cl(2) complex (cat. 1c) was achieved in up to 99.7% ee. Moderate to good enantioselectivities in the hydrogenation of corresponding (Z)-isomers in the presence of [Rh((R)-Xyl-P-Phos)(COD)]BF(4) (cat. 2c) were also obtained. The results demonstrated that the electronic and steric properties of the dipyridylphosphine ligands as well as the different transition metal ions have significant influences on the catalytic properties in the hydrogenation of beta-(acylamino)acrylates.  相似文献   

15.
开发高效的催化剂用于催化还原CO2转化为甲酸和它的盐类已经成为研究的热点,是因为将CO2转化为C1产物不仅可以解决CO2的含量升高带来的环境问题,还可以解决化石能源燃烧日趋严重的问题。贵金属配合物催化CO2转化为甲酸和甲酸盐类是目前这类反应最有效的方式,尤其是Ru、Ir和Rh等贵金属。我们之前的研究结果表明Ir(Ⅲ), Ru(Ⅱ)类配合物催化还原CO2转化为甲酸盐的活性是由配合物Ru―H键的成键性质决定的。它们能高活性的催化CO2是由于它们都含有同一种特点的Ru―H键,是由Ru的sd2杂化轨道和H的1s轨道杂化而成的,而且这一特点可以被活性氢的对位配体显著影响。鉴于硼基配体具有强的对位效应,我们基于高活性的均相催化剂Ru(PNP)(CO)H2 (PNP = 2, 6-二(二叔丁基磷甲基)-吡啶)设计了Ru-PNP-HBcat和Ru-PNP-HBpin,并计算了二者催化还原CO2的活性。Bcat和Bpin配体是实验上常用的硼基配体。我们的计算结果表明Ru-PNP-HBcat和Ru-PNP-HBpin有比Ru-PNP-H2更长的Ru―H键、亲核性更强的活性氢,其Ru―H键中的Ru原子的d轨道杂化成分的贡献也比Ru-PNP-H2的更少。相应地Ru-PNP-HBcat和Ru-PNP-HBpin活化CO2的能垒比Ru-PNP-H2低。而且Ru-PNP-H2、Ru-PNP-HBcat和Ru-PNP-HBpin催化CO2转化为甲酸盐的能垒分别为76.2、67.8、54.4 kJ∙mol-1,表明Ru-PNP-HBpin具有最高的催化活性。因此,钌配合物催化还原CO2的活性可由硼基配体强的对位效应和Ru―H键的成键性质来调控。  相似文献   

16.
A series of dinuclear Rh(II) complexes, [Rh(2)(μ-OAc)(4)(H(2)O)(2)] (HOAc = acetic acid) (1), [Rh(2)(μ-gly)(4)(H(2)O)(2)] (Hgly = glycolic acid) (2), [Rh(2)(μ-CF(3)CO(2))(4)(acetone)(2)] (3), and [Rh(2)(bpy)(2)(μ-OAc)(2)(OAc)(2)] (4), were found to serve as H(2)-evolving catalysts in a three-component system consisting of tris(2,2'-bipyridine)ruthenium(II) (Ru(bpy)(3)(2+)), methylviologen (MV(2+)), and ethylenediaminetetraacetic acid disodium salt (EDTA). It was also confirmed that thermal reduction of water into H(2) by MV(+)˙, in situ generated by the bulk electrolysis of MV(2+), is effectively promoted by 1 as a H(2)-evolving catalyst. The absorption spectra of the photolysis solution during the photocatalysis were monitored up to 6 h to reveal that the formation of photochemical or thermal byproducts of MV(+)˙ is dramatically retarded in the presence of the Rh(II)(2) catalysts, for the H(2) formation rather than the decomposition of MV(+)˙ becomes predominant in the presence of the Rh(II)(2) catalysts. The stability of the Rh(II)(2) dimers was confirmed by absorption spectroscopy, (1)H NMR, and ESI-TOF mass spectroscopy. The results indicated that neither elimination nor replacement of the equatorial ligands take place during the photolysis, revealing that one of the axial sites of the Rh(2) core is responsible for the hydrogenic activation. The quenching of Ru*(bpy)(3)(2+) by 1 was also investigated by luminescence spectroscopy. The rate of H(2) evolution was found to decrease upon increasing the concentration of 1, indicating that the quenching of Ru*(bpy)(3)(2+) by the Rh(ii)(2) species rather than by MV(2+) becomes predominant at the higher concentrations of 1. The DFT calculations were carried out for several possible reaction paths proposed (e.g., [Rh(II)(2)(μ-OAc)(4)(H(2)O)] + H(+) and [Rh(II)(2)(μ-OAc)(4)(H(2)O)] + H(+) + e(-)). It is suggested that the initial step is a proton-coupled electron transfer (PCET) to the Rh(II)(2) dimer leading to the formation of a Rh(II)Rh(III)-H intermediate. The H(2) evolution step is suggested to proceed either via the transfer of another set of H(+) and e(-) to the Rh(II)Rh(III)-H intermediate or via the homolytic radical coupling through the interaction of two Rh(II)Rh(III)-H intermediates.  相似文献   

17.
We have investigated a series of enantiopure phosphine-phosphite ligands (P(1)-P(2) = ligands 1-4) in the rhodium-catalyzed asymmetric hydrogenation reaction. Intermediate [Rh(P(1)-P(2))(cod)]BF(4) and [Rh(P(1)-P(2))(5)]BF(4) complexes (cod = 1,5-cyclooctadiene; 5 = methyl acetamidoacrylate ester) were observed by (31)P[(1)H] NMR. The [Rh(P(1)-P(2))(cod)]BF(4) complexes were precursors to active catalysts of the asymmetric hydrogenation reaction of several prochiral dehydroamino acid derivatives under mild reaction conditions (1 bar of hydrogen and 20 degrees C). The enantiomeric excess reached up to 99%.  相似文献   

18.
A series of heptametallic cyanide cages are described; they represent soluble analogues of defect-containing cyanometalate solid-state polymers. Reaction of 0.75 equiv of [Cp*Ru(NCMe)3]PF6, Et(4)N[Cp*Rh(CN)3], and 0.25 equiv of CsOTf in MeCN solution produced (Cs subset [CpCo(CN)3]4[Cp*Ru]3)(Cs subset Rh4Ru3). 1H and 133Cs NMR measurements show that Cs subset Rh4Ru3 exists as a single Cs isomer. In contrast, (Cs subset [CpCo(CN)3]4[Cp*Ru]3) (Cs subset Co4Ru3), previously lacking crystallographic characterization, adopts both Cs isomers in solution. In situ ESI-MS studies on the synthesis of Cs subset Rh4Ru3 revealed two Cs-containing intermediates, Cs subset Rh2Ru2+ (1239 m/z) and Cs subset Rh3Ru3+ (1791 m/z), which underscore the participation of Cs+ in the mechanism of cage formation. 133Cs NMR shifts for the cages correlated with the number of CN groups bound to Cs+: Cs subset Co4Ru4+ (delta 1 vs delta 34 for CsOTf), Cs subset Rh4Ru3 where Cs+ is surrounded by ten CN ligands (delta 91), Cs subset Co4Ru3, which consists of isomers with 11 and 10 pi-bonded CNs (delta 42 and delta 89, respectively). Although (K subset [Cp*Rh(CN)3]4[Cp*Ru]3) could not be prepared, (NH4 subset [Cp*Rh(CN)3]4[Cp*Ru]3) (NH4 subset Rh4Ru3) forms readily by NH4+-template cage assembly. IR and NMR measurements indicate that NH4+ binding is weak and that the site symmetry is low. CsOTf quantitatively and rapidly converts NH4 subset Rh4Ru3 into Cs subset Rh4Ru3, demonstrating the kinetic advantages of the M7 cages as ion receptors. Crystallographic characterization of CsCo4Ru3 revealed that it crystallizes in the Cs-(exo)1(endo)2 isomer. In addition to the nine mu-CN ligands, two CN(t) ligands are pi-bonded to Cs+. M subset Rh4Ru3 (M = NH4, Cs) crystallizes as the second Cs isomer, that is, (exo)2(endo)1, wherein only one CN(t) ligand interacts with the included cation. The distorted framework of NH4 subset Rh4Ru3 reflects the smaller ionic radius of NH4+. The protons of NH4+ were located crystallographically, allowing precise determination of the novel NH4...CN interaction. A competition experiment between calix[4]arene-bis(benzocrown-6) and NH4 subset Rh4Ru3 reveals NH4 subset Rh4Ru3 has a higher affinity for cesium.  相似文献   

19.
Abstract

Two new ferrocenyl iminopyridyl ligands, L1 and L2, have been synthesized and characterized using spectroscopic and analytical techniques. Both ligands were used to prepare new Rh(I) and Ru(II) ferrocenyl complexes 14. The structures of the complexes were confirmed using 1H and 13C nuclear magnetic resonance spectroscopy, high resolution electrospray ionization mass spectrometry, and infrared spectroscopy. The complexes were tested as catalysts in the hydroformylation of 1-octene. Rh ferrocenyl complexes 1 and 4 produced aldehydes under mild conditions while the Ru-ferrocenyl complexes 2 and 3 required higher temperature and pressure for effective hydroformylation to occur. The catalysts display excellent aldehyde chemoselectivity with varying regeoselectivity depending on temperature and pressure conditions employed. At high temperatures, the Rh ferrocenyl precatalysts favor formation of branched aldehydes due to increased isomerization at high temperatures. The Ru ferrocenyl precatalysts displayed less hydroformylation activity; however, the complexes show good chemoselectivity for aldehydes with no hydrogenation products formed.  相似文献   

20.
Hydrogen transfer reduction processes are attracting increasing interest from synthetic chemists in view of their operational simplicity. Reaction of [Ph2PNHCH2‐C4H3S] with [Ru(η6‐benzene)(µ‐Cl)Cl]2, [Rh(µ‐Cl)(cod)]2 and [Ir(η5‐C5Me5)(µ‐Cl)Cl]2 gave a range of new monodendate complexes [Ru(Ph2PNHCH2‐C4H3S)(η6‐benzene)Cl2], 1, [Rh(Ph2PNHCH2‐C4H3S)(cod)Cl], 2, and [Ir(Ph2PNHCH2‐C4H3S)(η5‐C5Me5)Cl2], 3, respectively. All new complexes were fully characterized by analytical and spectroscopic methods. 1H? 31P NMR, 1H? 13C HETCOR or 1H? 1H COSY correlation experiments were used to confirm the spectral assignments. 1–3 are suitable catalyst precursors for the transfer hydrogenation of acetophenone derivatives. Notably [Ru(Ph2PNHCH2‐C4H3S)(η6‐benzene)Cl2], 1, acts as an excellent catalyst, giving the corresponding alcohols in 98–99% yields in 30 min at 82 °C (TOF ≤200 h?1) for the transfer hydrogenation reaction in comparison to analogous rhodium or iridium complexes. This transfer hydrogenation is characterized by low reversibility under these conditions. Copyright © 2011 John Wiley & Sons, Ltd.  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