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.
汪顺义  纪顺俊 《有机化学》2008,28(2):181-186
铜络合物催化的有机金属试剂与α,β-不饱和羰基化合物的1,4-共轭加成反应, 是形成新的C—C键的重要方法之一. 介绍了铜络合物催化的格氏试剂与α,β-不饱和羰基化合物的立体选择性1,4-共轭加成反应的研究进展.  相似文献   

2.
Based on the dependences v(CO) =a + b* for IR spectra of carbonyi complexes of transition metals, the inductive constants of the organometallic fragments M(CO) m Cp n and HgM(CO) m Cp n (M = Co, Mo, Mn, Fe, Re) have been determined. The acceptor properties of the organometallic fragments have been shown to change according to the order of the nucleophilicity of the anions: Fe(CO)2Cp > Re(CO)5 > Mn(CO)5 > Mo(CO)3Cp > Co(CO)4.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 8, pp. 1482–1484, August, 1994.  相似文献   

3.
王福冬  高坤  王长生 《化学学报》2007,65(20):2211-2216
使用B3LYP方法研究了有机铼化合物R3R5(NHR4)Re(=CHR1)(=NR2)中分子内α-氢转移反应, 探讨了不同取代基对α-氢转移反应势垒的影响. 研究发现, 可以通过改变取代基来影响过渡金属Re有机化合物中的α-氢转移反应. R1位置的取代基为Me或CMe3时, 可以较大程度降低α-氢转移反应的势垒. R2为H时, α-氢转移反应势垒最低. R3和R5位置为SiH3时的反应势垒最低. R4为CMe3时, α-氢转移反应势垒最低. 研究结果还表明, 取代基对于反应势垒的影响有加和性.  相似文献   

4.
Simple reactions of the most used functional groups allowing two molecular fragments to link under mild, sustainable conditions are among the crucial tools of molecular chemistry with multiple applications in materials science, nanomedicine, and organic synthesis as already exemplified by peptide synthesis and “click” chemistry. We are concerned with redox organometallic compounds that can potentially be used as biosensors and redox catalysts and report an uncatalyzed reaction between primary and secondary amines with organometallic electrophilic alkynes that is free of side products and fully “green”. A strategy is first proposed to synthesize alkynyl organometallic precursors upon addition of electrophilic aromatic ligands of cationic complexes followed by endo hydride abstraction. Electrophilic alkynylated cyclopentadienyl or arene ligands of Fe, Ru, and Co complexes subsequently react with amines to yield trans‐enamines that are conjugated with the organometallic group. The difference in reactivities of the various complexes is rationalized from the two‐step reaction mechanism that was elucidated through DFT calculations. Applications are illustrated by the facile reaction of ethynylcobalticenium hexafluorophosphate with aminated silica nanoparticles. Spectroscopic, nonlinear‐optical and electrochemical data, as well as DFT and TDDFT calculations, indicate a strong push–pull conjugation in these cobalticenium– and Fe– and Ru–arene–enamine complexes due to planarity or near‐planarity between the organometallic and trans‐enamine groups involving fulvalene iminium and cyclohexadienylidene iminium mesomeric forms.  相似文献   

5.
β‐Peptides and, to a certain extent, also mixed α,β‐peptides, are resistant to degradation by a variety of proteolytic enzymes that rapidly degrade natural α‐peptides. This is one of many characteristics that make β‐peptides an attractive class of compounds for drug‐discovery studies. On the other hand, modern organometallic reactions such as the Suzuki–Miyaura cross‐coupling have become standard tools in industry laboratories to derivatize side chains of α‐peptidic compounds to build up libraries of unnatural peptides. Combining both features, we prepared (4‐bromo)‐β3‐homophenylalanine derivatives 3 – 5 and 12 as precursors for Suzuki–Miyaura couplings. From these bromo compounds, we synthesized biaryl‐substituted β‐homoamino acids 6 , and analogs 13 and 15 of the anti‐AIDS drug Saquinavir.  相似文献   

6.
Thiomaltol, a potential S,O‐coordinating molecule, has been utilized for the complexation of four different organometallic fragments, yielding the desired RuII, OsII, RhIII, and IrIII complexes having a “piano‐stool” configuration. In addition to the synthesis of these compounds with a chlorido leaving group, the analogous 1‐methylimidazole derivatives have been prepared, giving rise to thiomaltol‐based organometallics with enhanced stability under physiological conditions. The organometallic compounds have been characterized by NMR spectroscopy, elemental analysis, and X‐ray diffraction analysis. Their behavior in aqueous solution and their interactions with certain amino acids have been studied by ESI mass spectrometry. Their pH‐dependent stability has been investigated by 1H NMR in aqueous solution, and their cytotoxicity against three different cancer cell lines has been investigated. Furthermore, their capacity as topoisomerase IIα inhibitors as well as their effect on the cell cycle distribution and reactive oxygen species (ROS) generation have been elucidated.  相似文献   

7.
The interaction of estrogens and androgens with their corresponding receptors is known to play an important role in cancers of the breast and prostate. This paper reports the synthesis, characterization, and biochemical properties of a novel organometallic complex derived from 17α‐ethynyltestosterone, namely hexacarbonyl{μ‐[(20,21‐η:20,21‐η)‐(17α)‐17‐hydroxypregn‐4‐en‐20‐yn‐3‐one}dicobalt ([Co2(CO)6(17α‐ethynyltestosterone)]). The crystal and molecular structure of this compound was determined by single‐crystal X‐ray diffraction: it crystallizes in the monoclinic space group with a=24.6600(18) Å, b=12.9188(10) Å, c=26.3573(19) Å, β=108.651(2)°, and Z=12. A biochemical study showed that the compound is still recognized by the androgen receptor even when the relative binding affinity (RBA) is quite low (0.5%). This finding can be explained by the recently published 3D structure of the androgen receptor that shows that its binding site cannot accommodate a bulky substituent at the 17α position of the steroid.  相似文献   

8.
Coα‐(1H‐Imidazol‐1‐yl)‐Coβ‐methylcob(III)amide ( 4 ) was synthesized by methylation with methyl iodide of (1H‐imidazol‐1‐yl)cob(I)amide, obtained by electrochemical reduction of Coα‐(1H‐imidazol‐1‐yl)‐Coβ‐cyanocob(III)amide ( 5 ). The spectroscopic data and a single‐crystal X‐ray structure analysis indicated 4 to exhibit a base‐on constitution in solution and in the crystal. The crucial lengths of the axial Co−N and Co−CH3 bonds also emerged from the crystallographic data and were found to be smaller by 0.1 and 0.02 Å, respectively, than those in methylcob(III)alamin ( 2 ). The data of 4 support the view, that the `long' axial Co−N bonds as determined by X‐ray crystallography for the B12‐dependent methionine synthase, for methylmalonyl‐CoA mutase, and for glutamate mutase represent stretched Co−N bonds. The thermodynamic effect (the `trans influence') of the 1H‐imidazole base in 4 on the organometallic reactivity of this model for protein‐bound organometallic B12 cofactors was examined by studying Me‐group‐transfer equilibria in aqueous solution and using (5′,6′‐dimethyl‐1H‐benzimidazol‐1‐yl)cobamides (cobalamins) as reaction partners (Schemes 2 – 5, Table). In comparison with methylcob(III)alamin ( 2 ), 4 was found to be destabilized for an abstraction of the Co‐bound Me group by a CoIII electrophile. In contrast, the abstraction of the Co‐bound Me group by a radical(oid) CoII species was not significantly influenced thermodynamically by the exchange of the nucleotide base. Likewise, exploratory Me‐group‐transfer experiments with Me−CoIII and nucleophilic CoI corrinoids at pH 6.8 provided an apparent equilibrium constant near unity. However, this finding also was consistent with partial protonation of the imidazolylcob(I)amide at pH 6.8, suggesting an interesting pH dependence of the Megroup‐transfer equilibrium near neutral pH. Therefore, the replacement of the 5′,6′‐dimethyl‐1H‐benzimidazole base by an 1H‐imidazole moiety, as observed in methyl transferases and in C‐skeleton mutases, does not by itself strongly alter the inherent reactivity of the B12 cofactors in the crucial homolytic and nucleophilic‐heterolytic reactions involving the organometallic bond, but may help to enhance the control of the organometallic reactivity by protonation/deprotonation of the axial base.  相似文献   

9.
Stable organometallic derivatives of glucose were prepared either by treatment of 2,3,4,6-tetra-O-acetyl-1-thio-β-D-glucopyranose with organometallic hydroxides R3 MOH and oxides R2 MO or by the reaction between 2,3,4,6-tetra-O-acetyl-α-D-glucopyranosyl bromide with R3SnSLi.  相似文献   

10.
To investigate how remotely induced changes in ligand folding might affect catalysis by organometallic complexes, dynamic α-amino-iso-butyric acid (Aib) peptide foldamers bearing rhodium(I) N-heterocyclic carbene (NHC) complexes have been synthesized and studied. X-ray crystallography of a foldamer with an N-terminal azide and a C-terminal Rh(NHC)(Cl)(diene) complex showed a racemate with a chiral axis in the Rh(NHC) complex and a distorted 310 helical body. Replacing the azide with either one or two chiral L-α-methylvaline (L-αMeVal) residues gave diastereoisomeric foldamers that each possessed point, helical and axial chirality. NMR spectroscopy revealed an unequal ratio of diastereoisomers for some foldamers, indicating that the chiral conformational preference of the N-terminal residue(s) was relayed down the 1 nm helical body to the axially chiral Rh(NHC) complex. Although the remote chiral residue(s) did not affect the stereoselectivity of hydrosilylation reactions catalysed by these foldamers, these studies suggest a potential pathway towards remote conformational control of organometallic catalysts.  相似文献   

11.
We prepared the organometallic complex 17α‐(ferrocenylethynyl)estradiol (=[(3,17β‐dihydroxyestra‐1,3,5(10)‐trien‐17α‐yl)ethynyl]ferrocene; FcEE; 1 ) by a novel synthetic method. This metallocene possesses sufficient stability in aqueous media to permit the study of its biological properties. Thus, we were able to show that, despite the addition of a bulky substituent at the 17α position of the steroid, the metallocene is still well‐ recognized by an antibody specific to estradiol (CR=40%) and by both subtypes (ERα, ERβ) of the estrogen receptor (at 0°, RBA=28 and 37%, resp.). A DCI‐MS study of the stability of the carbocation [FcEE−OH]+ showed moderate stabilization of the carbocation, in agreement with the pK value of −0.72 found for the metallocene by means of Deno's method. The presence of the ferrocene allows the electrochemical detection of FcEE ( 1 ) by HPLC‐ED, with a detection limit of ca. 1 nM , suitable for quantitative pharmacological analysis.  相似文献   

12.
C. Lion  J.E. Dubois 《Tetrahedron》1973,29(21):3417-3423
α,α-Disubstituted ketones result from the reaction of organolithium compounds with substituted 4,4-dimethyl-2-oxazolines, 4, after hydrolysis of the intermediate enamine. Alkylation of the latter also permits the synthesis of α,α,α-trisubstituted ketones. α,α-Disubstituted ketones have been prepared by addition of organometallic reagents to 4,4-dimethyl-2-oxazolinium salts, 6. The efficiency of this method has been compared to that of other currently used methods.  相似文献   

13.
ABSTRACT

Stereocontrolled, stepwise synthesis of methyl α-L-rhamnopyranosyl-(1→2)-[α-D-glucopyranosyl-(1→3)]-α-L-rhamnopyranoside (A(E)B, 1) and methyl 2-acetamido-2-deoxy-β-D-glucopyranosyl-(1→2)-α-L-rhamnopyranosyl-(1→2)-[α-D-glucopyranosyl-(1→3)]-α-L-rhamnopyranoside (DA(E)B, 2) is described; these constitute the methyl glycosides of fragments of the O-specific polysaccharide of Shigella flexneri serotype 5a. Two routes to trisaccharide 1 were considered. Route 1 involved the coupling of a precursor to residue A and a disaccharide EB, whereas route 2 was based on the condensation of a precursor to residue E and a disaccharide AB. Rather surprisingly, the latter afforded the β-anomer of 1, namely methyl α-L-rhamnopyranosyl-(1→2)-[β-D-glucopyranosyl-(1→3)]-α-L-rhamnopyranoside as the major product. Route 1 was preferred. Overall, several observations made during this study suggested that, for the construction of higher fragments, a suitable precursor to rhamnose A would require protecting groups of low bulkiness at position 3 and 4. Therefore, the 2-O-acetyl-3,4-di-O-allyl-α-L-rhamnopyranosyl trichloroacetimidate (35) was the precursor of choice to residue A in the synthesis of the tetrasaccharide 2. The condensation product of 35 and methyl 2,3,4,6-tetra-O-benzyl-α-D-glucopyranosyl-4-O-benzyl-α-L-rhamnopyranoside was selectively deacylated and condensed to 2-trichloroacetamido-3,4,6-tri-O-acetyl-2-deoxy-α-D-glucopyranosyl trichloroacetimidate to afford the corresponding fully protected tetrasaccharide 45. Controlled stepwise deprotection of the latter proceeded smoothly to afford the target 2. It should be emphasised that the preparation of 45 was not straightforward, several donors and coupling conditions that were tested resulted only in the complete recovery of the acceptor. Distortion of several signals in the 13C NMR spectra of the fully or partially protected tetrasaccharide intermediates suggested that steric hindrance, added to the known low reactivity of HO-2 of rhamnosyl acceptors, probably played a major role in the outcome of the glycosidation attempts.  相似文献   

14.
In neocoenzyme B12 (=(5′-deoxy-5′-adenosyl)-13-epicob(III)alamin; 5 ), an epimer of coenzyme B12 ( 1 ), the organometallic group and a propanamide side chain of the vitamin-B12 ligand compete for the same region in space. Interesting consequences for structure and organometallic reactivity of this isomer of 1 are to be expected. Neocoenzyme B12 ( 5 ; 89% yield) and methyl-13-epicobalamin ( 6 ; 88% yield) were prepared from neovitamin B12 ( 4 ) by electrochemical means (Fig. 3). The solution structure of the organometallic neovitamin-B12 derivative 5 was analyzed by homonuclear and heteronuclear NMR spectroscopy. Comparison of the structures of 1 and 5 informed on the structural consequences of the epimerization at C(13) and revealed a remarkable flexibility of the organometallic group in 5 . In 5 , both sterically interacting functionalities (organometallic group and propanamide side chain at C(13)) adapt their conformations dynamically to avoid significant mutual clashes. As one consequence of this structural adaptation, the major conformations of 5 feature counterclockwise and clockwise reorientations of the organometallic ligand with respect to its crystallographically determined position in coenzyme B12 ( 1 ). One of the dominant conformers of 5 exhibits an orientation of the organometallic functionality similar to that found in the crystal structure of the coenzyme-B12-dependent methylmalonyl CoA mutase. The present NMR study also revealed the significant population of syn-conformers of the organometallic adenosine group, another remarkable feature of the solution structure of 5 .  相似文献   

15.
The thermal decomposition of normal alkanes and linear poly(α-olefins) was found to occur predominately by random scissions, followed by volatilization into an open system. To show these processes the fragment distribution from high molecular weight linear polyethylene (fragments to C55), n-alkanes, polypropylene, and polyisobutylene (fragments to C60) are reported and discussed in depth. Based on the results, a mechanism of random scissions followed by volatilization is postulated for polyethylene and n-alkanes as well as for polyisobutylene and polypropylene. Polystyrene is shown not to follow the same mechanism, for little evidence for fragments above trimers was found. In addition, temperature rise–time measurements at the pyrolyzer probe are reported and explained.  相似文献   

16.
ABSTRACT

The stepwise synthesis of methyl α-D-glucopyranosyl-(1→3)-α-L-rhamnopyranosyl-(1→3)-α-L-rhamnopyranoside (EBC-OMe, 1), methyl α-L-rhamnopyranosyl-(1→2)-[α-D-glucopyranosyl-(1→3)]-α-L-rhamnopyranosyl-(1→3)-α-L-rhamnopyranoside (A(E)BC-OMe, 2), and methyl 2-acetamido-2-deoxy-β-D-glucopyranosyl-(1→2)-α-L-rhamnopyranosyl-(1→2)-[α-D-glucopyranosyl-(1→3)]-α-L-rhamnopyranosyl-(1→3)-α-L-rhamnopyranoside (DA(E)BC-OMe, 3) is described. Compounds 1, 2 and 3 constitute the methyl glycosides of fragments of the O-specific polysaccharide of Shigella flexneri serotype 5a. Methyl 2,4-di-O-benzoyl-α-L-rhamnopyranosyl-(1→3)-2,4-di-O-benzoyl-α-L-rhamnopyranoside was an appropriate BC precursor for the synthesis of 1. For the synthesis of the branched targets 2 and 3, a benzyl group was best suited at position 2 of rhamnose C. Thus, methyl 4-O-benzyl-α-L-rhamnopyranosyl-(1→3)-2,4-di-O-benzyl-α-L-rhamnopyranoside was the key intermediate to the BC portion. In all cases, 2,3,4,6-tetra-O-benzyl-α-D-glucopyranosyl fluoride was a convenient E precursor, when used in combination with titanium tetrafluoride. All along, attention was paid to steric hindrance as a factor of major impact on the condensation steps outcome. Therefore, based on previous experience, 2-O-acetyl-3,4-di-O-allyl-α-L-rhamnopyranosyl trichloroacetimidate and 3,4,6-tri-O-acetyl-2-deoxy-2-trichloroacetamido-α-D-glucopyranosyl trichloroacetimidate were used as donors. Both suited all requirements when used as key precursors for residues A and D in the synthesis of 3, respectively.  相似文献   

17.
A nickel‐catalyzed asymmetric diarylation reaction of vinylarenes enables the preparation of chiral α,α,β‐triarylated ethane scaffolds, which exist in a number of biologically active molecules. The use of reducing conditions with aryl bromides as coupling partners obviates the need for stoichiometric organometallic reagents and tolerates a broad range of functional groups. The application of an N‐oxyl radical as a ligand to a nickel catalyst represents a novel approach to facilitate nickel‐catalyzed cross‐coupling reactions.  相似文献   

18.
The binding property of hydrogen on organometallic compounds consisting of Co, and Ni transition metal atoms bound to C m H m rings (m = 4, 5) is studied through density functional theory calculation. CoC m H m and NiC m H m complexes can store up to 3.49 wt% hydrogen with an average binding energy of about 1.3 eV. The adsorption characteristics of hydrogen to organometallic compounds are investigated by analyzing vibrational spectra of CoC4H4(H2) n and NiC4H4(H2) n (n = 0, 1, 2). The kinetic stability of these hydrogen-covered organometallic complexes is assured by analyzing the energy gap between the highest occupied molecular orbitals and the lowest unoccupied molecular orbitals. It is also discussed the application of 18-electron rule in predicting maximum number of hydrogen molecules that could be adsorbed by these organometallic compounds.  相似文献   

19.
Matrix‐assisted laser desorption/ionization in‐source decay (MALDI‐ISD) is initiated by hydrogen transfer from matrix molecules to the carbonyl oxygen of peptide backbone with subsequent radical‐induced cleavage leading to c′/z? fragments pair. MALDI‐ISD is a very powerful method to obtain long sequence tags from proteins or to do de novo sequencing of peptides. Besides classical fragmentation, MALDI‐ISD also shows specific fragments for which the mechanism of formation enlightened the MALDI‐ISD process. In this study, the MALDI‐ISD mechanism is reviewed, and a specific mechanism is studied in details: the N‐terminal side of Cys residue (Xxx‐Cys) is described to promote the generation of c′ and w fragments in MALDI‐ISD. Our data suggest that for sequences containing Xxx‐Cys motifs, the N–Cα bond cleavage occurs following the hydrogen attachment to the thiol group of Cys side‐chain. The c?/w fragments pair is formed by side‐chain loss of the Cys residue with subsequent radical‐induced cleavage at the N–Cα bond located at the left side (N‐terminal direction) of the Cys residue. This fragmentation pathway preferentially occurs at free Cys residue and is suppressed when the cysteines are involved in disulfide bonds. Hydrogen attachment to alkylated Cys residues using iodoacetamide gives free Cys residue by the loss of ?CH2CONH2 radical. The presence of alkylated Cys residue also suppress the formation of c?/w fragments pair via the (Cβ)‐centered radical, whereas w fragment is still observed as intense signal. In this case, the z? fragment formed by hydrogen attachment of carbonyl oxygen followed side‐chain loss at alkylated Cys leads to a w fragment. Hydrogen attachment on peptide backbone and side‐chain of Cys residue occurs therefore competitively during MALDI‐ISD process. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

20.
The mass spectra of some α-substituted phenyl-α,α′-dimethoxyl ketones (compounds 1) and their 2,4-dinitrophenylhydrazones (compounds 2) and semicarbazones (compounds 3) have been studied. The characteristic fragments at m/z (M ? 73) from compounds 1, m/z (M ? 253) from compounds 2 and m/z (M ? 130) from compounds 3 are abundant and proposed to be [ArCROCH3]+. Fragmentations yielding [M+ ? 49] from compounds 2 are abnormal and probably involve the methoxyl and nitro groups. The intense peak at m/z 130 due to [CH3OCH2CNNHCONH2]+ from compounds 3 corresponds to α-cleavage of the molecular ion. Some other fragments from these new compounds are interpreted in this paper.  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