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 15 毫秒
1.
The reaction of [Cp*2RuBr]+Br with bromine in CH2Cl2 (CD2Cl2) in an inert atmosphere at room temperature produces the complexes [Cp*Ru(Br)C5Me4CH2Br]+Br3 (syn conformer), [Cp*Ru(Br)C5Me3(CH2Br)2]+ (syn and anti conformers), and [Ru(Br)(C5Me4CH2Br)2]+ (syn conformer). All complexes were characterized by 1H and 13C NMR spectroscopy; the former complex, by elemental analysis. These complexes were also prepared by the reaction of [Cp*RuC5Me4CH2]+BF4 with bromine in CH2Cl2. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 12, pp. 2712–2718, December, 2005.  相似文献   

2.
Ethylene homopolymerizations and copolymerizations were catalyzed by zirconocene catalysts entrapped inside functionalized montmorillonites that had been rendered organophilic via the ion exchange of the interlamellar cations of layered montmorillonite with hydrochlorides of L ‐amino acids (AAH+Cl?) or their methyl esters (MeAAH+Cl?), with or without the further addition of hexadecyltrimethylammonium bromide (C16H33N+Me3Br?; R4N+Br?). In contrast to the homogeneous Cp2ZrCl2/methylaluminoxane catalyst for ethylene homopolymerizations and copolymerizations with 1‐octene, the intercalated Cp2ZrCl2 activated by methylaluminoxane for ethylene homopolymerizations and copolymerizations with 1‐octene proved to be more effective in the synthesis of polyethylenes with controlled molecular weights, chemical compositions and structures, and properties, including the bulk density. The effects of the properties of the organic guests on the preparation and catalytic performance of the intercalated zirconocene catalysts were studied. © 2003 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 2187–2196, 2003  相似文献   

3.
Electrophilic bromination of monosubstituted aromatic compounds is effected in a pentaquadrupole mass spectrometer using BrCO+ and CH3NH2Br+ as mass-selected reagent ions. Reaction normally occurs at the ring and the brominated product can be mass selected in turn and caused to dissociate by Br˙ loss upon collision-induced dissociation. Linear free energy correlations with Brown substituent σ+ constants describe the extent of gas-phase bromine cation addition under the non-equilibrium, low-pressure and solvent-free conditions which pertain in quadruple collision cells. The electrophilic addition reaction proceeds via a σ-complex to the ring as suggested by MS3 spectra, except in the case of nitrobenzene, where substituent bromination is suggested by the occurrence of a competitive process in which the nitrosubstituent is displaced by bromine. The reactivity parameters ρ are ?0.23 and ?0.56 for the gaseous reagents, BrCO+ and CH3NH2Br+, respectively. Both values are much less negative than corresponding values for bromination in solution. The greater reactivity of BrCO+ is evident by the fact that it reacts even with the strongly deactivated substrates and this is consistent with a weak Br? CO bond. Competitive protonation occurs in the case of CH3NH2Br+ and, unlike bromination, the rate of this reaction does not correlate with σ+ values. This is suggested to be a consequence of protonation at the ring in some cases and at the substituent in others, including acetophenone and benzonitrile. Evidence for this is that, in contrast to its lack of correlation with substituent constants, the rate of protonation correlates linearly with proton affinity.  相似文献   

4.
This study focuses on a series of cationic complexes of iridium that contain aminopyridinate (Ap) ligands bound to an (η5‐C5Me5)IrIII fragment. The new complexes have the chemical composition [Ir(Ap)(η5‐C5Me5)]+, exist in the form of two isomers ( 1+ and 2+ ) and were isolated as salts of the BArF? anion (BArF=B[3,5‐(CF3)2C6H3]4). Four Ap ligands that differ in the nature of their bulky aryl substituents at the amido nitrogen atom and pyridinic ring were employed. In the presence of H2, the electrophilicity of the IrIII centre of these complexes allows for a reversible prototropic rearrangement that changes the nature and coordination mode of the aminopyridinate ligand between the well‐known κ2‐N,N′‐bidentate binding in 1+ and the unprecedented κ‐N3‐pseudo‐allyl‐coordination mode in isomers 2+ through activation of a benzylic C?H bond and formal proton transfer to the amido nitrogen atom. Experimental and computational studies evidence that the overall rearrangement, which entails reversible formation and cleavage of H?H, C?H and N?H bonds, is catalysed by dihydrogen under homogeneous conditions.  相似文献   

5.
The 29Si, 13C and 1H NMR spectra of 11 mixtures of Me3SiI and Me3SiOSO2CF3 with DMF in CD2Cl2 show signals that are consistent with the formation of Me3SiOC+H(NMe2)X? but not with penta- or hexa-coordinate silicon species. The spectra of a 11 mixture of Me3SiBr and DMF show a rapidly exchanging, equilibrium mixture of Me3SiOC+H(NMe2)Br? and starting materials. No strong evidence for salt formation between DMF and Me3SiCl was obtained. The spectra of Me3SiX (X = I, Br, Cl, OSO2CF3) in CD3CN indicate that neither adduct formation nor extra coordination at silicon is significant.  相似文献   

6.
The reactivity of N‐heterocyclic carbenes (NHCs) and cyclic alkyl amino carbenes (cAACs) with arylboronate esters is reported. The reaction with NHCs leads to the reversible formation of thermally stable Lewis acid/base adducts Ar‐B(OR)2⋅NHC ( Add1 – Add6 ). Addition of cAACMe to the catecholboronate esters 4‐R‐C6H4‐Bcat (R=Me, OMe) also afforded the adducts 4‐R‐C6H4Bcat⋅cAACMe ( Add7 , R=Me and Add8 , R=OMe), which react further at room temperature to give the cAACMe ring‐expanded products RER1 and RER2 . The boronate esters Ar‐B(OR)2 of pinacol, neopentylglycol, and ethyleneglycol react with cAAC at RT via reversible B−C oxidative addition to the carbene carbon atom to afford cAACMe(B{OR}2)(Ar) ( BCA1 – BCA6 ). NMR studies of cAACMe(Bneop)(4‐Me‐C6H4) ( BCA4 ) demonstrate the reversible nature of this oxidative addition process.  相似文献   

7.
Motivated by the need for chemical strategies designed to tune peptide fragmentation to selective cleavage reactions, benzyl ring substituent influence on the relative formation of carbocation elimination (CCE) products from peptides with benzylamine‐derivatized lysyl residues has been examined using collision‐induced dissociation (CID) tandem mass spectrometry. Unsubstituted benzylamine‐derivatized peptides yield a mixture of products derived from amide backbone cleavage and CCE. The latter involves side‐chain cleavage of the derivatized lysyl residue to form a benzylic carbocation [C7H7]+ and an intact peptide product ion [(MHn)n+ – (C7H7)+](n‐1)+. The CCE pathway is contingent upon protonation of the secondary ε‐amino group (Nε) of the derivatized lysyl residue. Using the Hammett methodology to evaluate the electronic contributions of benzyl ring substituents on chemical reactivity, a direct correlation was observed between changes in the CCE product ion intensity ratios (relative to backbone fragmentation) and the Hammett substituent constants, σ, of the corresponding substituents. There was no correlation between the substituent‐influenced gas‐phase proton affinity of Nε and the relative ratios of CCE product ions. However, a strong correlation was observed between the π orbital interaction energies (ΔEint) of the eliminated benzylic carbocation and the logarithm of the relative ratios, indicating the predominant factor in the CCE pathway is the substituent effect on the level of hyperconjugation and resonance stability of the eliminated benzylic carbocation. This work effectively demonstrates the applicability of σ (and ΔEint) as substituent selection parameters for the design of benzyl‐based peptide‐reactive reagents which tune CCE product formation as desired for specific applications. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

8.
Treatment of Baylis–Hillman adducts 1 with bromo(dimethyl)sulfonium bromide, Br(Me2)S+Br?, in MeCN was found to stereoselectively afford (Z)‐ and (E)‐allyl bromides 2 . The reaction is rapid at room temperature, high‐yielding, and highly stereoselective.  相似文献   

9.
Density functional theory computation indicates that bridge splitting of [PtIIR2(μ-SEt2)]2 proceeds by partial dissociation to form R2Pta(μ-SEt2)PtbR2(SEt2), followed by coordination of N-donor bromoarenes (L-Br) at Pta leading to release of PtbR2(SEt2), which reacts with a second molecule of L-Br, providing two molecules of PtR2(SEt2)(L-Br-N). For R=4-tolyl (Tol), L-Br=2,6-(pzCH2)2C6H3Br (pz=pyrazol-1-yl) and 2,6-(Me2NCH2)2C6H3Br, subsequent oxidative addition assisted by intramolecular N-donor coordination via PtIITol2(L-N,Br) and reductive elimination from PtIV intermediates gives mer-PtII(L-N,C,N)Br and Tol2. The strong σ-donor influence of Tol groups results in subtle differences in oxidative addition mechanisms when compared with related aryl halide oxidative addition to palladium(II) centres. For R=Me and L-Br=2,6-(pzCH2)2C6H3Br, a stable PtIV product, fac-PtIVMe2{2,6-(pzCH2)2C6H3-N,C,N)Br is predicted, as reported experimentally, acting as a model for undetected and unstable PtIVTol2{L-N,C,N}Br undergoing facile Tol2 reductive elimination. The mechanisms reported herein enable the synthesis of PtII pincer reagents with applications in materials and bio-organometallic chemistry.  相似文献   

10.
The relative importance of the rearrangement ions [M ? Br ? CO]+, [M ? Br2 ? CO]+ and [M ? HBr2 ? CO]+ in the mass spectra of the title compounds is compared with the amounts of α-methoxyketone formed on reduction of these compounds with a Zn/Cu couple in methanol. It is suggested that the quantitative correlation found reflects the electron releasing powers of the substituents on the α carbons.  相似文献   

11.
A new approach to main‐group H2 activation combining concepts of transition‐metal and frustrated Lewis pair chemistry is reported. Ambiphilic, metal‐like reactivity toward H2 can be conferred to 9,10‐dihydro‐9,10‐diboraanthracene (DBA) acceptors by the injection of two electrons. The resulting [DBA]2? ions cleave the H?H bond with the formation of hydridoborates under moderate conditions (T=50–100 °C; p<1 atm). Depending on the boron‐bonded substituents R, the addition is either reversible (R=C≡CtBu) or irreversible (R=H). The reaction rate is strongly influenced by the nature and the coordination behavior of the countercation (Li+ slower than K+). Quantum‐chemical calculations support the experimental observations and suggest a concerted, homolytic addition of H2 across both boron atoms. As proven by the successful conversion of Me3SiCl into Me3SiH, the system Li2[DBA]/H2 appears generally relevant for the hydrogenation of element–halide bonds.  相似文献   

12.
Syntheses and Structures of [ReNBr2(Me2PhP)3] and (Me2PhPH)[ fac ‐Re(NBBr3)Br3(Me2PhP)2] [ReNBr2(Me2PhP)3] ( 1 ) has been prepared by the reaction of [ReNCl2(Me2PhP)3] with Me3SiBr in dichloromethane. The bromo complex reacts with BBr3 under formation of [Re(NBBr3)Br2(Me2PhP)3] ( 2 ) or (Me2PhPH)[fac‐Re(NBBr3)Br3(Me2PhP)2] ( 3 ) depending on the experimental conditions. The formation of the nitrido bridge leads to a significant decrease of the structural trans influence of the nitrido ligand which is evident by the shortening of the Re‐(trans)Br bond from 2.795(1) Å in [ReNBr2(Me2PhP)3] to 2.620(1) Å in [fac‐Re(NBBr3)Br3(Me2PhP)2] and 2.598(1) Å in [Re(NBBr3)Br2(Me2PhP)3], respectively.  相似文献   

13.
[Fc2B2(Br)(μ‐NPEt3)2]+Br – a Ferrocenyl‐substituted Phosphoraneiminato Complex of Boron [Fc2B2(Br)(μ‐NPEt3)2]+Br has been prepared from ferrocenylboron dibromide, [Fe(η5‐C5H5)(η5‐C5H4BBr2)], and the silylated phosphoraneimine Me3SiNPEt3 in dichloromethane solution to give orange‐red single crystals which were characterized by IR, NMR and 57Fe Mössbauer spectra, as well as by a crystal structure determination. [Fc2B2(Br)(μ‐NPEt3)2]+Br · 3 CH2Cl2 ( 1 · 3 CH2Cl2): Space group P21/n, Z = 4, lattice dimensions at –50 °C: a = 1370.6(3), b = 2320.9(5), c = 1454.4(2), β = 95.38(1)°, R1 = 0.061. In the cation of 1 the ferrocenyl‐substituted boron atoms are connected by the nitrogen atoms of the [NPEt3] groups to form a planar B2N2 four‐membered ring. One of the boron atoms having planar, the other tetrahedral coordination.  相似文献   

14.
Summary The liquid phase oxidation of gold in donor-acceptor organic and aqueous-organic media has been studied. The compounds [AuCl(Me2S)], [AuBr(Me2S)], [AuBr3(Me2S)], [Me3S][AuBr4], [Me3S][AuBr4(Me2S)]·H2O, [Me3SO]-[AuBr4]·H2O, [Me3S][Au2Br7(Me2S)2]·3H2O, [Me3S]2-[Au2Br8]·2DMSO·H2O, [Me2(Bu)SO][AuBr4]·H2O and [Me3S]Br were isolated by dissolution of Au0 in DMSO-RX mixtures (R = H or Bu; X = Cl or Br). The products were characterized by elemental analysis and i.r. spectroscopy. The nature of the Au0-DMSO-RX systems and the oxidant species are discussed in terms of a newly-developed concept of donor-acceptor electron transport (DAET) systems.  相似文献   

15.
For a series of benzaldehydes only with a leaving group or with both a leaving group and a single methoxy substituent 18F-fluorination via nucleophilic aromatic substitution (SNAr) was studied in DMF and Me2SO. In general, the radiochemical yields were clearly higher in DMF than in Me2SO. In the fluorodehalogenation reaction (leaving group: halogen = Br, Cl), extremely low radiochemical yields were observed in Me2SO (<1%). By monitoring labeling reactions using HPLC, oxidation of the aldehyde function of the precursor was detected. Especially, 2-bromobenzaldehyde was oxidized fastest in Me2SO (within 3 min reaction time, 90% of the precursor was consumed; radiochemical yield = 1.0 ± 0.5%); however, in DMF oxidation was always kept at a low level during the entire reaction (<5% of the precursor was oxidized; radiochemical yield = 73.0 ± 0.2%). In DMF, nitrobenzaldehydes with a methoxy substituent (methoxy group in meta-position to the nitro group) were labeled with good radiochemical yields (4-methoxy-2-nitrobenzaldehyde: 87 ± 3%; 2-methoxy-4-nitrobenzaldehyde: 83 ± 3%; 2-methoxy-6-nitrobenzaldehyde: 79 ± 4%) comparable to the non-substituted nitrobenzaldehydes (2-nitrobenzaldehyde: 84 ± 3%; 4-nitrobenzaldehyde: 81 ± 5%). Moreover, for structurally similar compounds, radiochemical yields showed a good correlation with 13C-NMR ppm values of the aromatic carbon atom bearing the leaving group.  相似文献   

16.
《Tetrahedron》1988,44(14):4631-4636
The reaction between aromatic amines and phosphate radical has been investigated. The reaction is accelerated by electron-releasing substituents and retarded by electron-withdrawing substituents pointing to an electrophilic attack by the PO42xxx radical. σ+para values correlate the effect of substituents well. The rho value for PO42xxx is more positive than that for CO3xxx indicating higher reactivity of phosphate radical towards aromatic amines than the carbonate radical.  相似文献   

17.
Reaction of 2‐isopropyl‐(N,N‐diisopropyl)‐benzamide 5 with tBuLi in ether results in ortho deprotonation and the formation of a hemisolvate based on a tetranuclear dimer of ( 5 ‐Lio)2?Et2O. The solid‐state structure exhibits a dimer core in which the amide oxygen atoms fail to stabilize the metal ions but are instead available for interaction with two metalated monomers that reside peripheral to the core. Reaction of 5 with tBuLi in the presence of the tridentate Lewis base PMDTA (N,N,N′,N′′,N′′‐pentamethyldiethylenetriamine) takes a different course. In spite of the tertiary aliphatic group at the 2‐position in 5 , X‐ray crystallography revealed that a remarkable benzylic (lateral) deprotonation had occurred, giving the tertiary benzyllithium 5 ‐Lil?PMDTA. The solid‐state structure reveals that amide coordination and solvation by PMDTA combine to distance the Li+ ion from the deprotonated α‐C of the 2‐iPr group (3.859(4) Å), thus giving an essentially flat tertiary carbanion and a highly distorted aromatic system. DFT analysis suggests that the metal ion resides closer to the carbanion center in solution. In line with this, the same (benzylic) deprotonation is noted if the reaction is attempted in the presence of tridentate diglyme, with X‐ray crystallography revealing that the metal is now closer to the tertiary carbanion (2.497(4) Å). Electrophilic quenches of lithiated 5 have allowed, for the first time, the formation of quaternary benzylic substituents by lateral lithiation.  相似文献   

18.
The fast atom bombardment (FAB) mass spectra and low-energy collisional activation mass spectra of ions generated under FAB were investigated for twelve bisbenzylisoquinoline (BBI) alkaloids. The relative molecular mass of the free base and diquaternary BBI alkaloids can be obtained from FAB data. However, monoquaternary ammonium salts produce only an [M — X]+ ion and the relative molecular mass cannot be determined. For Type A (single ether linkage) BBI alkaloids, fragmentation occurs primarily from benzylic and ether cleavages. Thus, the total number of aromatic substituents (OH, OCH3 or OCH2O) can be determined for rings A-B, C-D, E and F. For Type B (two either linkages) BBI alkaloids, fragmentation occurs primarily from double benzylic cleavages. Hence only the total number of aromatic substituents can be determined for the upper half of the Type B BBI alkaloids, i.e. rings A-B and C-D. An unknown alkaloid was examined to illustrate the utility of the fragmentation schemes.  相似文献   

19.
Onium salts QZ (Z = Cl, Br) having a lipophilic (Q = R3NR', where R' = C16H33) or readily extractable (into organic phase) cation (Q = Ph4P) exhibit a high catalytic activity in phase-transfer alkaline hydrolysis of N-benzyloxycarbonylglycine 4-nitrophenyl ester in the two-phase system chloroform-borate buffer (pH 10). No catalytic effect is observed in the presence of hydrophilic ammonium salts [Et4NCl, Et3PhCH2NCl, Me2(NH2)+NCH2CH2+N(NH2)Me2·2Br-] and those insoluble in organic solvents [(Me)3+NNH(CH2)2COO-·2H2O, Me2(NH2)+NCH2CO-, Me2(NH2)+N(CH2)3SO3 -]. These data suggest extraction mechanism of the process. The activity of lipophilic cation Q is determined mainly by two factors: its extractibility, on the one hand, and the ability to form micelles, on the other.  相似文献   

20.
Comparative analysis of the oxidizing and complexing properties of the DMSO–HX (X = Cl, Br, I) and DMSO–HX–ketone (X = Br, I; the ketone is acetone, acetylacetone, or acetophenone) systems toward silver was performed. The reaction products are AgX (X = Cl, Br, I), [Me3S+]Ag n X m (n= 1, 2; m= 2, 3; X = Br, I) and [Me2S+CH2COR]AgX 2(R = Me, Ph; X = Br, I). The composition of the obtained complexes depends on both the DMSO : HX ratio and the nature of HX, as well as on the methods used to isolate solid products from the solution. It was noted that the formation of the [Me2S+CH2COMe]AgBr 2complex in the Ag0–DMSO–HBr–acetylacetone system occurs with cleavage of the acetylacetone C–C bond and follows a specific reaction course. The optimum conditions for production of the silver compounds in the title systems are determined.  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