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.
The regioselectivity of the hydroformylation reaction of 2-methyl-3-(3-acetylpyrrol-1-yl)prop-1-ene catalyzed by an unmodified Rh catalyst has been investigated at the B3LYP/6-31G* level with Rh described by effective core potentials in the LANL2DZ valence basis set. Considering the population of all the H-Rh(CO)3-olefin transition state complexes, a regioselectivity ratio (B:L) of 12:88 has been obtained, in satisfactory agreement with the experiment producing the chiral linear aldehyde as the only product. The aldehyde, after complete diastereoselective cyclization, yields a 1:1 mixture of 1-acetyl-6R(S)-methyl-8R(S)-hydroxy-5,6,7,8-tetrahydroindolizine (having the same configuration on both stereogenic carbon atoms) and 2-acetyl-6-methyl-5,6-dihydroindolizine [Lett Org Chem (2006) 3:10-12]. The reason for such a high degree of diastereoselectivity has been elucidated examining the B3LYP/6-31G* potential energy surface for the reactions leading to the RR and RS diastereomers on a model system (without the acetyl substituent) and the actual compound. In the absence of a catalyst, a very high barrier is found along the reaction pathway, whereas spontaneous annulation occurs to a protonated pentahydroindolizine in the presence of H+. When a counterion (F-) is added, the proton on the newly formed tetrahedral carbon is abstracted, obtaining a structure closer to the final product (tetrahydroindolizine). Replacing H+ with Rh+, an initial adduct along the RS path much more favorable than any of those computed along the RR one is located because of the presence of the acetyl group. Tentative approaching paths obtained using [Rh(CO)3]+, bound to the aldehyde O, feature a higher barrier along the RS one, and offer a convincing explanation for the observed diastereoselectivity.  相似文献   

2.
The reaction rate of the oxidative addition and the following CO insertion step of methyl iodide with [Rh(acac)(CO)(P(OCH2)3CCH3)] is determined. The key finding is that while [Rh(acac)(CO)(P(OCH2)3CCH3)] oxidatively adds methyl iodide ca 300 times faster than the Monsanto catalyst, the CO insertion step is much slower. However, the rate-determining step of the oxidative addition reaction of the phosphorus-containing acetylacetonato-rhodium(I) complex, the carbonyl insertion step, is still in the same order or faster than the rate-determining oxidative addition step of iodomethane to [Rh(CO)2I2].  相似文献   

3.
The interaction between Rh(CO)(2)(acac=acetylacetonate ion) and human serum albumin (HSA) in a hydrosoluble nanostructured biocatalyst active in homogeneous hydroformylation was characterised by means of matrix assisted laser desorption ionization time-of-flight mass spectrometry (MALDI-TOFMS), high performance liquid chromatography mass spectrometry (HPLC-MS) and scanning electron microscopy (SEM). MALDI-TOFMS substantiates that the biocatalyst consists of a tetrameric structure of HSA that could bind up to 89 Rh(CO)(2)(+) units. A comparison between samples of pure HSA and the biocatalyst, both tryptic digested, showed a significant change in the tertiary structure of the protein in the HSA/Rh adduct, probably ascribable to the interaction of Rh(I) ions with sulphur atoms in the HSA moiety. SEM observations confirmed an evident denaturation of the protein and an outstanding correspondence between the surface distribution of Rh and S atoms; this is indirect evidence that the metal ion interacts strongly mainly with the sulphur atoms. Furthermore, an excellent agreement between calculated and measured (SEM) S/Rh elemental mean ratio was observed. Finally, an electrostatic interaction between Rh(I) and sulphur atoms was ruled out by sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE) findings.  相似文献   

4.
Four enol lactones, bearing phenyl or 1-naphthyl substituents on the alpha or beta positions [3-phenyl-6-methylenetetrahydro-2-pyranone (alpha Ph6H, IIc), 3-(1-naphthyl)-6-methylenetetrahydro-2-pyranone (alpha Np6H, IId), 4-phenyl-6-methylenetetrahydro-2-pyranone (beta Ph6H, IIIc), and 4-(1-naphthyl)-6-methylenetetrahydro-2-pyranone (beta Np6H, IIId)], available as pure R and S enantiomers, have been studied as alternate substrate inhibitors of chymotrypsin. Kinetic constants for substrate binding (Ks) and acylation (ka) were determined by a competitive substrate assay, using succinyl-L-Ala-L-Ala-L-Pro-L-Phe p-nitroanilide; the deacylation rate constant (kd) was determined by the proflavin displacement assay. All lactones undergo rapid acylation (ka varies from 17 to 170 min-1) that shows little enantioselectivity; there is, however, pronounced enantioselectivity in substrate binding for three of the lactones [Ks(R/S) = 40-110]. In each case it is the enantiomer with the S configuration that has the higher affinity. In all cases, deacylation rates are slow, and in two cases, acyl enzymes with half-lives of 4.0 and 12.5 h at pH 7.2, 25 degrees C, are obtained (for beta Ph6H and alpha Np6H, respectively). In these cases, high deacylation enantioselectivity is observed [kd(S/R) = 60-70], and the lactone more weakly bound as a substrate (R enantiomer) gives the more stable acyl enzyme. Two hypotheses, involving hindrance of the attack of water or an exchange of the ester and ketone carbonyl groups in the acyl enzyme, are advanced as possible explanations for the high stability of these acyl enzymes.  相似文献   

5.
Active carbon (A.C.), after treatment at 1123 K with hydrogen for more than 24 h, was found to be very effective as a support for ruthenium catalysis in ammonia synthesis. The activity of barium promoted Ru catalysts supported on this treated A.C. is affected remarkably by the Ru precursor. Ru3(CO)12 was found previously to be the most effective Ru precursor for oxide supports such as Al2O3, MgO, and CeO2 in ammonia synthesis. However, in this study, the Ru---Bao/A.C. catalyst prepared from Ru(acac)3 was found to yield the highest activity, while the catalyst prepared from Ru3(CO)12 resulted in the lowest activity among several precursors. Transmission electron microscopy and hydrogen chemisorption showed that the particle size of Ru obtained from the decomposition of Ru(acac)3 on hydrogen-treated A.C. is smaller than the particle size of Ru obtained from the decomposition of Ru3(CO)12. Additionally, RuCl3 was found to be an effective precursor for Ru/A.C. catalyst. It has been suggested that chlorine ions can be removed easily from A.C. by hydrogen reduction at 773 K, and that this results in the high activity.  相似文献   

6.
The reaction of [Rh(diene)(acac)] (diene=cyclooctadiene or norbornadiene; acac=acetylacetonate) with bidentate ligands of the type Ph2P(CH2)nSR (n=1, 2 or 3; R=Me, Et, Ph, not all combinations) or cis-Ph2PCHCHPPh2 leads to [Rh(diene)(LL)]+ or [Rh(LL)2]+, depending on the stoichiometry of the reaction. The complexes were fully characterized by 1H and 31P NMR spectroscopy.  相似文献   

7.
《Inorganica chimica acta》2006,359(9):2756-2759
Asymmetric hydrogenation of (Z)-α-acetamidocinnamic acid and methyl-(Z)-α-acetamidocinnamate by [Rh(COD)(DIPAMP)][BF4] catalyst was studied in ionic liquid/isopropanol two-phase catalytic system. In this system 97–100% conversion was achieved and the enantioselectivity values were over 90%. Application of [bmim][BF4] ionic liquid made it possible to recycle the catalyst in consecutive cycles. After four cycles, neither significant conversion nor enantioselectivity decrease was observed.  相似文献   

8.
The aim of this study was to investigate biohydrogen production from biofuel-reforming processes using new multi-component bulk-type cobalt-based catalysts. The addition of different components to improve the catalytic performance was studied. Monometallic cobalt catalyst and catalysts containing Ru (ca. 1%) and/or Na (ca. 0.5%) were characterized and tested in the 623-673 K temperature range in ethanol steam reforming (ESR) with a steam/carbon ratio (S/C) of 3. The catalysts showed a high performance for hydrogen production and, except for H(2) and CO(2), only small amounts of by-products were obtained, depending on the temperature and the catalyst used. The catalyst containing both Ru and Na (Co-Ru(Na)) showed the best catalytic behavior in ESR. It operated stably for at least 12 days under cycles of oxidative steam reforming of glycerol/ethanol mixtures (S/C=2) and activation under O(2).  相似文献   

9.
Hydroformylation reactions of a series of alkenes and alkynes have been carried out using the heteronuclear Rh---W catalyst, (CO)4 hH(CO)(PPh3) (1). The results of these reactions have been compared with corresponding reactions using [Rh(OAc)2]2 as catalyst. Catalysis of a reaction of styrene using 1 gave a very high yield of the branched chain aldehyde containing only a trace of the straight chain isomer. Reactions of the phosphinoalkene, Ph2P(CH2)3CH=CH2 (7) and the corresponding alkyne, Ph2P(CH2)3CCH (11) gave similar products using either catalyst system with the alkryne reaction being significantly slower. Reaction of the alkenyl dithiane, H---CH2CH=CH2 (2), using the Rh---W catalyst (1) gave a higher ratio of linear to branched aldehydes (47 linear:53 branched) than the corresponding reaction using [Rh(OAc)2]2 (25 linear:75 branched). Reactions of vinyl acetate using 1 as catalyst gave a significant amount of linear aldehyde in contrast to reactions using [Rh(OAc)2]2 but reactions of allyl acetate gave similar products for both catalyst systems.  相似文献   

10.
Fox ME  McCague R 《Chirality》2005,17(4):177-185
The performance of a catalyst used in asymmetric synthesis is likely to be dependent upon its stereoisomeric purity. An impurity was detectable by (31)P NMR in early development batches of the asymmetric hydrogenation catalyst [(S,S)-Me-BPE-Rh(COD)](+)BF(4) (-) made from the ligand bis((2S,5S)-2,5-dimethylphospholano)ethane [(S,S)-Me-BPE]. Its identity as a stereoisomer with one chiral and one meso-phospholane ring was deduced by comparison of the (31)P NMR spectra and GC traces of the ligand with a deliberately synthesized mixture of isomers. Interestingly, the impurity corresponded to a trans-meso isomer formed by thermal (200 degrees C) pyramidal inversion at phosphorus of the initially synthesized cis-meso-phospholane when the ligand was purified by distillation. Low levels of this trans-meso/chiral impurity do not significantly impair the enantioselectivity of the rhodium complex as an asymmetric hydrogenation catalyst, but high levels of stereochemical impurities resulted in a loss of both enantioselectivity and activity. Therefore it is indeed important to establish that a catalyst used in asymmetric catalysis is sufficiently stereoisomerically pure. Owing to strict control of the stereochemical purity of the key hexane-2,5-diol intermediate, the impurity is not detected in production batches.  相似文献   

11.
The enantioselective polymerization of methyl-substituted epsilon-caprolactones using Novozym 435 as the catalyst was investigated. All substituted monomers could be polymerized except 6-methyl-epsilon-caprolactone (6-MeCL), which failed to propagate after ring opening. Interestingly, an odd-even effect in the enantiopreference of differently substituted monomers was observed. The combination of 4-methyl-epsilon-caprolactone with Novozym 435 showed good enantioselectivity also in bulk polymerization and resulted in enantiomerically enriched P((S)-4-MeCL) (eep up to 0.88). Subsequently, a novel initiator combining a primary alcohol to initiate the ring opening polymerization and a tertiary bromide to initiate atom transfer controlled radical polymerization (ATRP) was synthesized, and showed high initiator efficiencies (> 90%) in the ring opening polymerization of 4-methyl-epsilon-caprolactone in bulk. In addition, the enantioselectivity was retained (E = 11). By using Ni(PPh3)2Br2 as the ATRP catalyst, Novozym 435 could be effectively inhibited at the desired conversion of 4-methyl-epsilon-caprolactone, thus ensuring a high enantiomeric excess in the polymer backbone. At the same time, Ni(PPh3)2Br2 catalyzed the ATRP of methyl methacrylate resulting in the formation of P((S)-4-MeCL-b-MMA) block copolymers. By this combination of two inherently different polymerization reactions, chiral P((S)-4-MeCL-b-MMA) block copolymers can be conveniently obtained in one pot without intermediate workup.  相似文献   

12.
The ginsenoside Rh(2) and its aglycone 20(S)-protopanaxadiol are known to inhibit the binding of [(3)H]batrachotoxinin 20alpha-benzoate to site 2 on voltage-gated sodium channels and electrophysiological investigations conducted by others have shown that ginsenosides cause voltage-dependent inhibition of reconstituted forms of the sodium channel. Here we describe the actions of Rh(2) and 20(S)-protopanaxadiol on sodium channel function and release of neurotransmitters resulting from activation of native sodium channels in synaptic preparations isolated from whole mouse brain. Rh(2) and 20(S)-protopanaxadiol inhibited veratridine-dependent (tetrodotoxin-suppressible) depolarization of synaptoneurosomes as determined using the rhodamine 6G method although 20(S)-protopanaxadiol was more potent as an inhibitor than Rh(2). Veratridine- (sodium channel-) dependent release of the neurotransmitters L-glutamate and GABA was almost fully inhibited by 20(S)-protopanaxadiol, however, less complete inhibition was observed with Rh(2). At its maximum inhibitory concentration, Rh(2) also produced release of l-glutamate and GABA from synaptosomes, in contrast to 20(S)-protopanaxadiol. We conclude that low to moderate micromolar concentrations of Rh(2) and 20(S)-protopanaxadiol inhibit sodium channel function and sodium channel-activated release of neurotransmitters. Apparently the ginsenoside Rh(2) cannot achieve complete inhibition of sodium channel-activated transmitter release because at high concentrations it also stimulates release.  相似文献   

13.
A new convenient high-yield synthesis of the tris-cyclometalated complexes fac-[Rh(ppy)3] (4; ppy = 2-phenylpyridinato) was developed. Complex 4 was prepared in a kind of one-pot synthesis starting from in situ prepared [Rh(acac)(coe)2] (2) which was heated in refluxing 2-phenylpyridine for a short time. After purification by filtration over alumina, compound 4 was obtained in yields of 65%. Also [Rh(acac)(ppy)2] (3) was prepared in a similar manner by oxidative addition of Hppy in refluxing toluene in high yields. In contrast to previous findings with the analogous iridium compounds, there was not any hint at the formation of the isomer mer-[Rh(ppy)3] using similar reaction conditions as applied for iridium. Furthermore the compound [{Rh(μ-Cl)(ppy)2}2] (5) was prepared from [{Rh(μ-Cl)(coe)2}2] (1) and Hppy in refluxing toluene in nearly quantitative yield.  相似文献   

14.
The mechanisms by which G-protein-coupled receptors (GPCRs) activate G-proteins are not well understood due to the lack of atomic structures of GPCRs in an active form or in GPCR/G-protein complexes. For study of GPCR/G-protein interactions, we have generated a series of chimeras by replacing the third cytoplasmic loop of a scaffold protein bacteriorhodopsin (bR) with various lengths of cytoplasmic loop 3 of bovine rhodopsin (Rh), and one such chimera containing loop 3 of the human beta2-adrenergic receptor. The chimeras expressed in the archaeon Halobacterium salinarum formed purple membrane lattices thus facilitating robust protein purification. Retinal was correctly incorporated into the chimeras, as determined by spectrophotometry. A 2D crystal (lattice) was evidenced by circular dichroism analysis, and proper organization of homotrimers formed by the bR/Rh loop 3 chimera Rh3C was clearly illustrated by atomic force microscopy. Most interestingly, Rh3C (and Rh3G to a lesser extent) was functional in activation of GTPgamma35S/GDP exchange of the transducin alpha subunit (Galphat) at a level 3.5-fold higher than the basal exchange. This activation was inhibited by GDP and by a high-affinity peptide analog of the Galphat C terminus, indicating specificity in the exchange reaction. Furthermore, a specific physical interaction between the chimera Rh3C loop 3 and the Galphat C terminus was demonstrated by cocentrifugation of transducin with Rh3C. This Galphat-activating bR/Rh chimera is highly likely to be a useful tool for studying GPCR/G-protein interactions.  相似文献   

15.
The hydrosilation of prochiral ketones using catalysts prepared by alkylation of [1,2-bis(tetrahydroindenyl)ethane]titaniumIV(1,1′-binaphth-2,2′-diolate) with MeLi and n-BuLi, and (EtO)3SiH, Me(EtO)2SiH, [MeSi(H)O]4, Me3SiO[MeSi(H)O]nSiMe3 and MeSiH3 as the hydrosilane is described. The rates obtained with the MeLi based catalyst are one to two orders of magnitude faster than previously observed with MeLi based catalysts in the presence of MePhSiH2 and Ph2SiH2 and about the same as those observed with n-BuLi based catalysts. Me(EtO)2SiH, [MeSi(H)O]4 and Me3SiO[MeSi(H)O]nSiMe3 all undergo rapid redistribution in the presence of the catalyst to give MeSiH3, the actual hydrosilating agent in all three cases. Likewise, (EtO)3SiH redistributes to SiH4. The ee's for the hydrosilation product of acetophenone are consistently much higher (99%) for the n-BuLi based catalyst than for the MeLi based catalyst (40–50%). The hydride complex [(BTHIE)TiH]2 gives essentially the same enantioselectivity as the MeLi based catalyst. The ee's for a test set of dialkylketones are relatively insensitive to either the catalyst or the hydrosilane. Some possible mechansims that are consistent with the experimental results are discussed.  相似文献   

16.
A new class of polymer-supported (R,S)-BINAPHOS 1e in which the parent BINAPHOS has two alkoxy-substituents at the 3-positions of the phenyls, has been synthesised. Using its Rh(I) complex, asymmetric hydroformylation of olefins proceeded with higher enantioselectivity in some cases compared to the conventional polymer-supported 1c.  相似文献   

17.
The Baeyer-Villiger monooxygenase (BVMO) BmoF1 from Pseudomonas fluorescens DSM 50106 was shown before to enantioselectively oxidize different 4-hydroxy-2-ketones to the corresponding hydroxyalkyl acetates, being the first example of a BVMO-catalyzed kinetic resolution of aliphatic acyclic ketones. However, the wild-type enzyme exhibited only moderate E values (E approximately 55). Thus, the enantioselectivity was enhanced by means of directed evolution and optimization of reaction conditions since it was found that higher E values (E approximately 70 for wild-type BmoF1) could already be obtained when performing biotransformations in shake flasks rather than small tubes. In a first step, random mutations were introduced by error-prone polymerase chain reaction, and BmoF1 mutants (>3,500 clones) were screened for improved activity and enantioselectivity using a microtiter-plate-based screening method. Mutations S136L and L252Q were found to increase conversion compared to wild type, while several mutations (H51L, F225Y, S305C, and E308V) were identified enhancing the enantioselectivity to a varying extent (E approximately 75-90). In a second step, beneficial mutations were recombined by consecutive cycles of QuikChange site-directed mutagenesis resulting in a double mutant (H51L/S136L) showing both improved conversion and enantioselectivity (E approximately 86).  相似文献   

18.
We examined the enzymatic function of recombinant CYP2C19 in enantiomeric hexobarbital (HB) 3'-hydroxylation, and searched the roles of amino acid residues, such as Phe-100, Phe-114, Asp-293, Glu-300, and Phe-476 of CYP2C19 in the stereoselective HB 3'-hydroxylation, using a yeast cell expression system and site-directed mutagenesis method. CYP2C19 wild-type exerted substrate enantioselectivity of (R)-HB>(S)-HB and metabolite diastereoselectivity of 3'(R)<3'(S) in 3'-hydroxylation of HB enantiomers. The substitution of Asp-293 by alanine failed to yield an observable peak at 450 nm in its reduced carbon monoxide-difference spectrum. CYP2C19-E300A and CYP2C19-E300V with alanine and valine, respectively, in place of Glu-300 exerted total HB 3'-hydroxylation activities of 45 and 108%, respectively, that of the wild-type. Interestingly, these two mutants showed substrate enantioselectivity of (R)-HB<(S)-HB, which is opposite to that of the wild-type, while metabolite diasteroselectivity remained unchanged. The replacement of Phe-476 by alanine increased total HB 3'-hydroxylation activity to approximately 3-fold that of the wild-type. Particularly, 3'(S)-OH-(S)-HB-forming activity elevated to 7-fold that of the wild-type, resulting in the reversal of the substrate enantioselectivity. In contrast, the substitution of phenylalanine at positions 100 and 114 by alanine did not produce a remarkable change in the total activity or the substrate enantioselectivity. These results indicate that Glu-300 and Phe-476 are important in stereoselective oxidation of HB enantiomers by CYP2C19.  相似文献   

19.
The effect of pH and temperature on structure, stability, activity and enantioselectivity of haloalkane dehalogenase DbjA from Bradyrhizobium japonicum USDA110 was investigated in this study. Conformational changes have been assessed by circular dichroism spectroscopy, functional changes by kinetic analysis, while quaternary structure was studied by gel filtration chromatography. Our study shows that the DbjA enzyme is highly tolerant to pH changes. Its secondary and tertiary structure was not affected by pH in the ranges 5.3-10.3 and 6.2-10.1, respectively. Oligomerization of DbjA was strongly pH-dependent: monomer, dimer, tetramer and a high molecular weight cluster of the enzyme were distinguished in solution at different pH conditions. Moreover, different oligomeric states of DbjA possessed different thermal stabilities. The highest melting temperature (T(m) = 49.1 ± 0.2 °C) was observed at pH 6.5, at which the enzyme occurs in dimeric form. Maximal activity was detected at 50 °C and in the pH interval 7.7-10.4. While pH did not have any effect on enantiodiscriminination of DbjA, temperature significantly altered DbjA enantioselectivity. A decrease in temperature results in significantly enhanced enantioselectivity. The temperature dependence of DbjA enantioselectivity was analysed with 2-bromobutane, 2-bromopentane, methyl 2-bromopropionate and ethyl 2-bromobutyrate, and differential activation parameters Δ(R-S)ΔH and Δ(R-S)ΔS were determined. The thermodynamic analysis revealed that the resolution of β-bromoalkanes was driven by both enthalpic and entropic terms, while the resolution of α-bromoesters was driven mainly by an enthalpic term. Unique catalytic activity and structural stability of DbjA in a broad pH range, combined with high enantioselectivity with particular substrates, make this enzyme a very versatile biocatalyst. Enzyme EC3.8.1.5 haloalkane dehalogenase.  相似文献   

20.
The structure of [V(eta5-C5H5)2(CH3C(O)CHC(O)CH3)](O3SCF3) (1) (=[VCp2(acac)](O3SCF3)), a dual-function anti-cancer agent with anti-angiogenic and anti-mitotic properties, was determined by single-crystal X-ray diffraction. The geometry is well described as a pseudo-tetrahedral like structure with the centroids of the cyclopentadienyl rings and the two oxygen atoms of the acetylacetonate ring in the ancillary positions of the central vanadium (IV) atom. The bisector of the V(acac) fragment deviates from the C2 axis of the ligand framework by only 4 degrees, compared to a deviation of 7 degrees for the V(acac) fragment in the tetramethylethano-bridged vanadocene acetyl acetonate complex. Crystal data for 1: space group, P2(1)/c; a=7.5544(9) A, b=14.936(2) A, c=16.193(2) A, beta=102.901(2) degrees, V= 1781.0(4) A3; Z=4; R=0.0506 for 2310 reflections with I> 2sigma(I). This report also details the electron paramagnetic resonance, UV/Vis spectroscopy, electrochemical properties and the biological activity profile of this potent anti-cancer agent.  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