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 rhodium-catalysed hydroacylation of alkene is one of the most useful C–H bond activation processes. The C–C bond-forming reactions via C–H bond activation have extensively been the focus of study in the fields of organic and organometallic chemistry. In this work, density functional theory has been used to study Rh(I)-catalysed hydroacylation and hydrogenation of ethene with formic acid. All the intermediates and the transition states were optimised completely at the B3LYP/6-311++G(d,p) level (LANL2DZ(d) for Rh, P). Calculation results confirm that Rh(I)-catalysed hydroacylation of ethene is exothermic and the released Gibbs free energy is ? 60.39 kJ/mol. Rh(I)-catalysed hydrogenation of ethene is also exothermic and the released Gibbs free energy is ? 150.97 kJ/mol. Rh(I)-catalysed hydroacylation of ethene is the dominant reaction mode for Rh(I)-catalysed hydroacylation and hydrogenation of ethene with formic acid. In Rh(I)-catalysed hydroacylation of ethene, the H-transfer reaction is prior to the C–C bond-forming reaction. Therefore, the reaction mode ‘a’ (i.e. ca → M1 → TS1 → M2 → TS2a → M3a → TS3a → M4 → P1) is the dominant reaction pathway for Rh(I)-catalysed hydroacylation and hydrogenation of ethene. The theoretically predicted dominant product is propane acid.  相似文献   

2.
Density functional theory (DFT) was used to study the cobalt(I)-catalyzed enantioselective intramolecular hydroacylation of ketones and alkenes. All intermediates and transition states were fully optimized at the M06/6-31G(d,p) level (LANL2DZ(f) for Co). The results demonstrated that the ketone and alkene present different reactivities in the enantioselective hydroacylation. In ketone hydroacylation catalyzed by the cobalt(I)–(R,R)-Ph-BPE complex, reaction channel “a” to (R)-phthalide was more favorable than channel “b” to (S)-phthalide. Hydrogen migration was both the rate-determining and chirality-limiting step, and this step was endothermic. In alkene hydroacylation catalyzed by the cobalt(I)–(R,R)-BDPP complex, reaction channel “c” leading to the formation of (S)-indanone was the most favorable, both thermodynamically and kinetically. Reductive elimination was the rate-determining step, but the chirality-limiting step was hydrogen migration, which occurred easily. The results also indicated that the alkene hydroacylation leading to (S)-indanone formation was more energetically favorable than the ketone hydroacylation that gave (R)-phthalide, both thermodynamically and kinetically.
Graphical abstract A DFT study demonstrated that the ketone and alkene in the cobalt(I)-catalyzed enantioselective intramolecular hydroacylation showed different reactivities
  相似文献   

3.
Density functional theory (DFT) was used to investigate the Rh(I)-catalyzed intermolecular hydroacylation of vinylsilane with benzaldehyde. All intermediates and transition states were optimized completely at the B3LYP/6-31G(d,p) level (LANL2DZ(f) for Rh). Calculations indicated that Rh(I)-catalyzed intermolecular hydroacylation is exergonic, and the total free energy released is −110 kJ mol−1. Rh(I)-catalyzed intermolecular hydroacylation mainly involves the active catalyst CA2, rhodium–alkene–benzaldehyde complex M1, rhodium–alkene–hydrogen–acyl complex M2, rhodium–alkyl–acyl complex M3, rhodium–alkyl–carbonyl–phenyl complex M4, rhodium–acyl–phenyl complex M5, and rhodium–ketone complex M6. The reaction pathway CA2 + R2M1bT1bM2bT2b1M3b1T4bM4bT5bM5bT6bM6bP2 is the most favorable among all reaction channels of Rh(I)-catalyzed intermolecular hydroacylation. The reductive elimination reaction is the rate-determining step for this pathway, and the dominant product predicted theoretically is the linear ketone, which is consistent with Brookhart’s experiments. Solvation has a significant effect, and it greatly decreases the free energies of all species. The use of the ligand Cp′ (Cp′ = C5Me4CF3) decreased the free energies in general, and in this case the rate-determining step was again the reductive elimination reaction.  相似文献   

4.
In order to obtain a better understanding of the “periodic-acid-Schiff” reaction, also known as the “periodic-acid fuchsin-sulfurous-acid” reaction, three types of investigations were carried out

1) The Schiff reagent was replaced by other aldehyde reagents: benzidine or o-dianisidine. There was no significant change in the histological distribution and intensity of the reactions occurring after periodic acid oxidation.

2) Periodic acid was replaced by another oxidizing agent: lead tetraacetate (dissolved in acetic acid). There was no significant change in the histological distribution of the reactions with the Schiff reagent, but some change in their intensity. It was concluded that 1,2-glycols and a-amino alcohols play the main role in the reactions with both oxidants. The presence of α-hydroxy acids in some types of mucous cells is suggested by the results with lead tetraacetate.

Incidently, glycogen and starch are not sufficiently oxidized by lead tetraacetate (in acetic acid) at room temperature to give positive reactions with the Schiff reagent, while cellulose and other periodic-acid-Schiff reactive substances are.

3) The staining of films of presumed reactive substances with the periodic-acid-Schiff technic C O the intense reactivity of many polysaccharides, mucopolysaccharides and mucoproteins, but not of ordinary proteins. (Hyaluronic and chondroitin sulfuric acid are, however, not reactive in vitro).

In conclusion, the periodic-acid-Schiff technic consists of an oxidation of 1,2-glycols and a-amino alcohols to produce aldehyde groups, which are then stained by the Schiff reagent. The “film test” reveals that these radicals are present in certain polysaccharides, mucopolysaccharides and mucoproteins.  相似文献   

5.
The formation and reactivity of pyridine diimine rhodium(I) alkyl complexes without β-hydrogens (Me, Bz, CH2SiMe3) is described. In contrast to the corresponding cobalt complexes, the rhodium complexes could not be activated to polymerise ethene. Rh ethyl complexes could not be prepared. Examples of hydrogen transfer to and from the ligand were observed, illustrating the active role the pyridine diimine ligand can play in the reactions of its complexes. Decomposition via loss of free ligand was observed in many cases, indicating that the pyridine diimine ligand is not a very suitable one for RhI.  相似文献   

6.
The dialkyl-μ-ethylidene-μ-methylene-bis (pentamethylcyclopentadienyl)-dirhodium complexes [{(C5Me5)Rh}2(μ-CH2)(μ-CHMe) (R)2] (4, P=Me; 5, Et; 6, n-Bu; 7, CH=CH2; and 8, Z-CH=CHMe) have been prepared from RMgBr and [{(C5Me5)Rh}2(μ-CH2)(μ-CHMe)(X)2] (2, X=Cl; 3, X=Br). Structures deduced from the NMR spectra show that the dialkyl complexes can exist in one trans and two cis forms. The decomposition of the dimethyl complex 4 is compared with that of the related di-μ-methylene complex; it reacts readily (30°C, MeCN solution) in the presence of one-electron oxidisers to give propene and methane and a little ethene and some butenes. Mass-spectrometric analysis of the 13C labelling in the organics originating from [{(C5Me5)Rh}2(μ-CH2)(μ-CHMe) (13CH3)2] shows that methane derives from the Rh---Me, ethene half from the ethylidene and half from coupling of Rh-methyl and a bridging methylene, while the propene arises almost entirely from the ethylidene and a rhodium methyl. The butenes come from coupling of ethylidene, methylene and a Rh-methyl, but only quite small amounts are formed; thus C+C coupling is the major decomposition path for the μ-ethylidenes, in contrast to the di-μ-methylene complexes where C+C+C coupling predominates. The divinyl complex [{(C5Me5)Rh}2(μ-CH2)(μ-CHMe) (CH=CH2)2] also underwent internal C+C coupling on reaction with AgBF4 in MeCN to give a mixture of the allyl and methylallyl cations [(C5Me5)Rh(η3-CH2CHCHR)(MeCN)]+(10, R=H; 11, R=Me).  相似文献   

7.
Fibronectin binds specifically to fibrin and is covalently cross-linked to the fibrin α chain by activated factor XIII (XIIIa). This reaction is important for wound healing. Here we investigate XIIIa-catalyzed cross-linking of fibronectin and some of its fragments to a recombinant fragment representing the COOH-terminal 30kDa of the fibrin α chain (αC30K:His 368–Val 610). Only fibronectin and those fragments containing an intact NH2-terminus were able to form cross-linked complexes. As many as 10 of the 17 lysines in αC30K can serve as amine donors in this reaction. Analysis of the rate of XIIIa-catalyzed cross-linking of fibronectin NH2-terminal peptides and fragments with αC30K revealed that the presence of the first type I “finger” module accelerates the cross-linking reaction; addition of fingers 2–5 had no further effect.  相似文献   

8.
Esters of N-acylated amino acids and the sterically demanding phosphine 2-(di-ortho-tolylphosphino)phenol react within 1 h at room temperature with the Rh(I) centers of [Cl(μ-Cl)Rh(cyclooctene)2]2 to give products of oxidative addition of the ester carbonyl-O bond. The N-acyl carbonyl oxygen is bound to the Rh in these initial adducts, but is displaced upon addition of PMe3, PhPMe2, NH2NMe2, or the thioether function of a methionine derivative. Remarkably, both initial products from achiral amino acids and their ligand adducts are formed as single five-coordinate diastereomers in essentially quantitative yields. However, asymmetric induction by chiral amino acid derivatives of proline and phenylalanine on the stereochemistry at Rh was modest. Finally, the identities of infrared absorptions of acyl and amide groups in the complexes were established unequivocally by synthesis and spectroscopy of N-acetylglycine esters with a 13C label at either the ester or amide carbonyl group.  相似文献   

9.
Density functional theory has been used to study copper(I)-catalysed aziridination of diazoacetate with imines. All the intermediates and the transition states were optimised completely at B3LYP/6-31G(d) level. Calculation results confirm that copper(I)-catalysed aziridination of diazoacetate with imines is exothermic, and the total released Gibbs free energy is about ? 170 kJ/mol. Copper(I)-catalysed aziridination has two reaction modes: I and II, and thus the reaction mode I is dominant. The formation of the copper(I)–carbene–imine complex M3 (i.e. the attack of imines on copper–carbon(carbene) of copper–carbene intermediate M2) is the rate-determining step and the chirality-limiting step for copper-catalysed asymmetric aziridination. The reaction channel CA2 → M1a → TS1a → M2 → TS2a2 → M3a2 → TS3a2 → M4a2 → P1 is the most favourable one. The dominant products predicted theoretically are of (R)-chirality.  相似文献   

10.
N R Kallenbach  S D Drost 《Biopolymers》1972,11(8):1613-1620
Apparent second-order rate constants for complex formation between poly (I) and poly (C) and copolymers of C containing non-complementary I or U residues have been determined spectrophotometrically. The rate constants decrease as the concentration of either I or U in the C strands increases–the effect seems insensitive to the species of residue involved, when differences in the thermal stabilities of the poly (I) poly (C,I) and poly (I). poly (C,U) complexes are taken into account. These results suggest that low concentrations of relatively stable defects can alter the apparent kinetic “complexity” of polynucleotides as determined by hybridization methods (C0t analysis).  相似文献   

11.
This review deals with the application of Lewis super acids such as Al(III), In(III), and Sn(IV) triflates and triflimidates as catalysts in the synthesis of fragrance materials. Novel catalytic reactions involving C-C and C-heteroatom bond-forming reactions, as well as cycloisomerization processes are presented. In particular, Sn(IV) and Al(III) triflates were employed as catalysts in the selective cyclization of unsaturated alcohols to cyclic ethers, as well as in the cyclization of unsaturated carboxylic acids to lactones. The addition of thiols and thioacids to non-activated olefins, both in intra- and intermolecular versions, was efficiently catalyzed by In(III) derivatives. Sn(IV) Triflimidates catalyzed the cycloisomerization of highly substituted 1,6-dienes to gem-dimethyl-substituted cyclohexanes bearing an isopropylidene substituent. The hydroformylation of these unsaturated substrates, catalyzed by a Rh(I) complex with a bulky phosphite ligand, selectively afforded the corresponding linear aldehydes. The olfactory evaluation of selected heterocycles, carbocycles, and aldehydes synthesized is also discussed.  相似文献   

12.
《Inorganica chimica acta》1986,112(2):103-106
The A1 symmetry vCO of the carbonyl complexes [Mo(chel)(CO)4], [M(chel)(CO)2] [PF6] (M=Rh, Ir; chel=bipy, phen and substituted derivatives) are used for determining the electron donor-acceptor properties of the title ligands. The steric hindrance of the methyl groups in positions 2 and 9 of the phenanthroline favours the formation of Rh(I) and Ir(I) pentacoordinated derivatives.  相似文献   

13.
The goal of this study was to evaluate a new approach that can be applied for labeling biomolecules with (211)At. Many astatine compounds that have been synthesized are unstable in vivo, providing motivation for seeking different (211)At labeling strategies. The approach evaluated in this study was to attach astatide anions to soft metal cations, which are also complexed by a bifunctional ligand. Ultimately, this complex could in principle be subsequently conjugated to a biomolecule with the proper selection of ligand functionality. We report here the attachment of (211)At(-) and *I(-) (*I = (131)I or (125)I) anions to the soft metal cations Rh(III) and Ir(III), which are complexed by the 1,5,9,13-tetrathiacyclohexadecane-3,11-diol (16aneS4-diol) ligand. Radioactive *I(-) anions were used for preliminary studies directed at the optimization of reaction conditions and to provide a baseline for comparison of results with (211)At. Four complexes Rh[16aneS4-diol]*I/(211)At and Ir[16aneS4-diol]*I/(211)At were synthesized in high yield in a one-step procedure, and the products were characterized mainly by paper electrophoresis and reversed-phase HPLC. The influences of time and temperature of heating and concentrations of metal cations and sulfur ligand 16aneS4-diol, as well as pH on the reaction yields were determined. Yields of about 80% were obtained when the quantities of Rh(III) or Ir(III) cations and 16aneS4-diol ligand in the solutions were 62.5 nmol and 250 nmol, respectively, and the pH ranged 3.0-4.0. Syntheses required heating for 1-1.5 h at 75-80 degrees C. The influence of microwave heating on the time and completeness of the complexation reaction was evaluated and compared with the conventional method of heating in an oil bath. Microwave synthesis accelerates reactions significantly. With microwave heating, yields of about 75% for Rh[16aneS4-diol](131)I and Ir[16aneS4-diol](131)I complexes were obtained after only 20 min exposure of the reaction mixtures to microwave radiation. In conclusion, this study has shown that it is possible to attach an astatide anion to soft metal cations in a simple and fast one-step procedure, with high yields. These complexes will be evaluated as reagents for labeling biomolecules.  相似文献   

14.
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   

15.
During the 1860’s, Hugo Schiff studied many reactions between amines and aldehydes, some of which have been used in histochemistry, at times without credit to Schiff. Much controversy has surrounded the chemical structures and reaction mechanisms of the compounds involved, but modern analytical techniques have clarified the picture. I review these reactions here. I used molecular modeling software to investigate dyes that contain primary amines representing eight chemical families. All dyes were known to perform satisfactorily for detecting aldehydes in tissue sections. The models verified the correct chemical structures at various points in their reactions and also determined how decolorization occurred in those with “leuco” forms. Decolorization in the presence of sulfurous acid can occur by either adduction or reduction depending on the dye. The final condensation product with aldehyde was determined to be either a C-sulfonic acid adduct on the carbonyl carbon atom or an aminal at the same atom. Based on the various outcomes, I have placed the dyes and their reactions into five categories. Because Hugo Schiff studied the reactions between aldehydes and amines with and without various acids or alcohol, it is only proper to call each of them Schiff reactions that used various types of Schiff reagents.  相似文献   

16.
The mechanism of all elementary steps involved in the catalytic cycle of benzoylformate decarboxylase (BFD, E.C. 4.1.1.7) to generate the acyloin linkage is investigated by extensive molecular dynamics simulations. Models involving different charge states of amino acids and/or mutants of critical residues were constructed to understand the involvement of the catalytically active residues and the reactivity differences between different substrates in this reaction. Our calculations confirm that H70, S26, and H281 are catalytically active amino acids. H281 functions as a base to accept Hdonor in the first nucleophilic attack and as an acid in the second, to donate the proton back to Oacceptor. S26 assists H281 in deprotonation of the donor aldehyde and protonation of the acceptor aldehyde. In both the first and second nucleophilic attacks, H70 interacts with Oaldehyde and aligns it toward the nucleophilic center. H70 has been found to have an electrostatic effect on the approaching aldehyde whose absence would block the initiation of the reaction. The reactivity difference between benzaldehyde (BA) and acetaldehyde (AA) is mainly explained by the steric interactions of the acceptor aldehyde with the surrounding amino acids in the active center of the enzyme. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 32–46, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

17.
We report here two new, high purity, high yield, one-step syntheses of cis-[Rh(bpy)2X2][PF6] {X = Cl, Br, I, bpy = 2,2′-bipyridine} directly from the RhX3 · nH2O starting materials - one conventional and one microwave. The key to obtaining the pure complexes appears to be maintaining a 2:1 ratio of bipyridine to rhodium in solution; thus all reactants must be completely dissolved prior to the start of the reaction. Comparison of the two methods is also discussed. These complexes are pure enough for emission spectroscopy after minimal work-up. The complete characterization of all the halide complexes and the first cyclic voltammetry data on the cis-[Rh(bpy)2I2][PF6] complex are reported. The irreversible Rh(III)-Rh(I) redox potential becomes more positive from Cl to I. All three complexes show two reversible redox potentials corresponding to the bipyridine reductions. These data are consistent with the loss of the two halide ligands and formation of [Rh(bpy)2]+ upon reduction of Rh(III) to Rh(I).  相似文献   

18.
The interaction of Li+, a weak activator of pyruvate kinase, with substrate and inhibitor complexes of the enzyme has been investigated by magnetic resonance techniques. Proton relaxation rate (PRR) titrations indicate that the dissociation constant of Li+ from the ternary enzyme-Mn(II)-phosphoenolpyruvate (P-enolpyruvate) complex is 15 mm at 5 °C and 17 mm at 30 °C. The electron paramagnetic resonance spectrum of the enzyme-Mn(II)-Li(I)-P-enolpyruvate complex is the superposition of spectra for two distinct species (Reed, G. H., and Cohn, M. (1973) J. Biol. Chem.248, 6436–6442). Low temperatures favor the form giving rise to the more nearly isotropic spectrum, whereas high temperatures favor the species giving rise to the anisotropic “K+-like” spectrum. 7Li nuclear magnetic resonance data are consistent with a model in which the two forms observed by epr correspond to differing Mn(II) to Li(I) distances. The form giving rise to the anisotropic spectrum is characterized by a Mn(II) to Li(I) distance of 4.7 Å, and in the more isotropic form this distance is approximately 9 Å. The 4.7 Å separation of the Mn(II) and Li(I) in the anisotropic form of the complex compares favorably with the 4.9 Å separation of Mn(II) and T1(I) (Reuben, J., and Kayne, F. J. (1971) J. Biol. Chem.246, 6227–6234) in the P-enolpyruvate complex, although T1+ is a much better activator of the pyruvate kinase reaction. Thus, a change in the distance between the monovalent and divalent cations does not account quantitatively for the lower activation by Li+, inasmuch as more than 50% of the enzyme-Mn(II)-Li(I)-P-enolpyruvate complex has the “active” conformation with respect to the separation of the cations and the epr spectrum of the complex. As reported previously (Reed, G. H., and Morgan, S. D. (1974) Biochemistry13, 3537–3541), the dissociation constant of oxalate and the epr spectrum for the ternary complex of pyruvate kinase with Mn(II) and oxalate are not influenced by the species of monovalent cation present. The nuclear relaxation rates of Li+ are increased in the presence of the ternary oxalate complex, although the separation of the Mn(II) and Li(I) appears to be much greater than for the “anisotropic” form of the P-enolpyruvate complex.  相似文献   

19.
Many proteins require N-linked glycosylation for conformational maturation and interaction with their molecular chaperones. In Drosophila, rhodopsin (Rh1), the most abundant rhodopsin, is glycosylated in the endoplasmic reticulum (ER) and requires its molecular chaperone, NinaA, for exit from the ER and transport through the secretory pathway. Studies of vertebrate rhodopsins have generated several conflicting proposals regarding the role of glycosylation in rhodopsin maturation. We investigated the role of Rh1 glycosylation and Rh1/NinaA interactions under in vivo conditions by analyzing transgenic flies expressing Rh1 with isoleucine substitutions at each of the two consensus sites for N-linked glycosylation (N20I and N196I). We show that Asn(20) is the sole site for glycosylation. The Rh1(N20I) protein is retained within the secretory pathway, causing an accumulation of ER cisternae and dilation of the Golgi complex. NinaA associates with nonglycosylated Rh1(N20I); therefore, retention of nonglycosylated rhodopsin within the ER is not due to the lack of Rh1(N20I)/NinaA interaction. We further show that Rh1(N20I) interferes with wild type Rh1 maturation and triggers a dominant form of retinal degeneration. We conclude that during maturation Rh1 is present in protein complexes containing NinaA and that Rh1 glycosylation is required for transport of the complexes through the secretory pathway. Failure of this transport process leads to retinal degeneration.  相似文献   

20.
2-Methyl-4, 6-cyclohexadienaldehyde and n-butyraldehyde were treated with sodium in p-xylene to yield the aromatized “leaf alcohol reaction” product, 2-methyl-benzylalcohol, in a better yield than that with the cyclohexadienaldehyde alone. n-Butyric acid isolated from the reaction mixture unequivocally showed the operation of the “crossed Cannizzaro disproportionation” in this reaction, aliphatic aldehyde serving as the hydride donor. 2-Propyl-5-ethyl-4, 6-cyclohexadienaldehyde was obtained by the NaOH/H2O-EtOH Michael-Aldol condensation of leaf aldehyde, gave 2-propyl-5-ethyl-benzylalcohol along with caproic acid.

On the basis of “leaf alcohol-reaction” mechanism, it was obtained following benzyl-alkohols; 2-methyl-, 2-propyl-, 2-methyl-5-ethyl-, 2-propyl-5-ethyl-benzylalcohol, from leaf alcohol and crotylalcohol.  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