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 46 毫秒
1.
The synthesis of polymer-bound thiol reagents, supported on macroporous 4% divinylbenzene co-polymer (Amberlite XE-305), via three synthetic approaches is described: (i) Alkylation or acylation of XE-305 with 3-nitro-4-halogen-substituted benzyl chloride or benzoyl halide yielding 3-nitro-4-halobenzene-bound species, followed by substitution of the activated polymeric halogen atom with sulfur (see Scheme 1). (ii) Formation of a thiol ether by a direct substitution of an active polymeric halogen by reaction with benzylthiol, followed by chlorination, thiolation, and reduction (see Scheme 2). (iii) Attachment of a prepared tailor-made disulfide to aminomethyl function of a polymeric support, followed by reduction (see Scheme 3). The polymers were tested for their free-thiol content by 5, 5′-dithiobis(2-nitrobenzoic acid) (Ellman's reagent12) in DMF. Their thiolytic activity was investigated in the removal of 2-nitrophenylsulphenyl (Nps) group from Nps-protected amino acid (Scheme 4). Site-site interaction between the polymer-bound thiol with its activated halide precursor to yield polymeric sulfide during displacement reaction, and the interconversion of the polymeric thiols into polymeric disulfides at equilibrium or during reaction with Nps-amino acids, observed, and is attributed to the flexibility of the polymeric matrices.  相似文献   

2.
The photoinduced substitution reactions of halogenated alkanes (1-haloadamantanes, 1-haloronorbornanes, menthyl chloride) with a homologous series of amines or alcohols (methylamine, 2-methyl-2-aminopropane, methanol, or 2-methyl-2-propanol) to form the corresponding alkane-substituted amines or ethers and HCl were investigated. The geometry of the bridgehead carbons made S(N)2 reactions impossible. Nonpolar reaction conditions were employed which made classical and nonclassical carbocation S(N)1 reaction pathways unlikely. The reaction rates were measured. Trapping experiments indicated that free radical reactions were uninvolved in the substitution product formation. A novel, photoinduced electron-transfer reaction mechanism involving a charge-transfer intermediate is proposed to explain the observed production of secondary amines and ethers. The excitation wavelength dependence (action spectrum) was measured and found to be comparable to the ultraviolet absorption spectra of the charge-transfer complexes. The stereochemical implications of the reaction mechanism were investigated. The formation of the methyl ether of (1R,2S,5R)-menthol was the only organic reaction product observed in the photoreaction between (1R,2S,5R)-menthyl chloride and methanol.  相似文献   

3.
Reactions of the nitrone CH3CH=N(CH3)O and the nitrile oxide CH3C[triple bond]NO with the nitrile complexes trans-[MCl2(N[triple bond]CCH3)2] (M = Pt, 1; Pd, 2) were investigated by theoretical methods at B3LYP and, for some processes, CCSD(T) levels of theory. The mechanisms of substitutions and cycloadditions were studied in detail. The former occur via a concerted asynchronous mechanism of dissociative type. The calculations of the metal-ligand bond energies in the starting complexes and substitution products and the analysis of structural features of the transition states indicate that the M-N bond dissociation (rather than M-O bond formation) is the step, which controls the reactivity of and in substitutions. The different chemical behaviours of the Pt and Pd complexes towards the 1,3-dipoles were investigated. The exclusive isolation of cycloaddition rather than substitution products in any solvents in the case of is both kinetically and thermodynamically controlled.The switch of the reaction mode from cycloaddition to substitution for 2 in CH2Cl2 solution is caused by the significantly lower Pd-N bond energy in comparison with the Pt-N bond energy, consistent with the higher lability of the Pd complexes. The different chemical behaviour of 2 in CH3CN and CH2Cl2 solvents is accounted for by the great excess of acetonitrile in the CH3CN solution rather than a different solvation character. The relative variation of Wiberg bond indices along the reaction path is proposed as a quantitative criterion for the classification of the reaction mechanism.  相似文献   

4.
溶液中甲醇和二氯亚砜的化学反应   总被引:2,自引:0,他引:2  
用B3LYP方法和SCIPCM模型(模拟溶剂效应)研究了甲醇和二氯亚砜在两种非极性(ε<15)和两种极性(ε>15)溶剂中的反应(最终产物为氯代甲烷和二氧化硫). 反应过程由反应(1)和反应(2)组成, CH3OS(O)Cl是反应(1)的主要产物和反应(2)的反应物. 反应(2)有“前面取代”(经过渡态TS3f)和“背后取代”(先经CH3OS(O)Cl的电离, 再经过渡态TS3b)两种机理. 计算表明, 在气相和四种溶剂中反应(1)和(2)都是放热反应, 反应(1)具有相同的反应途径(经过渡态→中间体→过渡态), 溶剂的极性对反应(2)有很大的影响. 在气相和非极性溶剂中, TS3f的能量比(CH3OSO++Cl-)离子对(中间体IM2)的能量低, 反应(2)应为前面取代机理; 在极性溶剂中, IM2和TS3b的能量都比TS3f低, 反应(2)应为背后取代机理.  相似文献   

5.
Contributions to the Chemistry of Silicon and Germanium. XXXV. Halogenation of Higher Silanes with Tin(IV) Chloride or Mercury(II) Chloride Trisilane, n-tetrasilane, iso-tetrasilane, and n-pentasilane were chlorinated by SnCl4 or HgCl2 in 2,3-dimethylbutane at O°C. In each case, mainly the corresponding monochlorosilanes were formed by substitution at the primary silicon atom. With the dichloro- and trichlorosilanes formed, a double substitution at the same Si atom could not be observed. The most suitable reaction conditions for the preparation of the mono- and dichlorosilanes were ascertained. The composition of the reaction product concerning the different isomeric chlorosilanes was determined by gaschromatography.  相似文献   

6.
Poly-2-(2-azido-5-nitrobenzoyloxy)ethylmethacrylate (P-II-A) and poly-2-(4-azido-3-nitrobenzoyloxy)ethylmethacrylate (P-II-B) were synthesized from the reactions of 2-hydroxyethylmethacrylate with 2-chloro-5-nitrobenzoic acid and 4-chloro-3-nitrobenzoic acid, respectively, by substitution reactions of sodium azide with the corresponding chloronitrobenzoyl group. In addition, the degradation reaction of the 2-azido-5-nitrobenzoyl group in P-II-A and the transformation of the 4-azido-3-nitrobenzoyl group to 5-carboxylbenzofurazane-1-oxide ring in P-II-B by irradiation with ultraviolet (UV) light or by heating were investigated in detail. In the photochemical reaction the reaction of the azide group in P-II-A was affected by the presence of a spacer in the polymer chain. Moreover, in the thermochemical reaction the rates of the reactions of azide groups in P-II-A and P-II-B were controlled by the facility of molecular motion and the conformation of polymer chains.  相似文献   

7.
The gas phase reactions between HF and the protonated alkyl fluorides MeFH+, EtFH+, Pr(i)FH+, and Bu(t)FH+ have been studied using ab initio methods. The potential energy profiles for both nucleophilic substitution (S(N)2) and elimination (E2) pathways have been investigated. Both backside Walden inversion and frontside nucleophilic substitution reaction profiles have been generated. Backside substitution is very favourable, but shows relatively little variation with the alkyl group. Frontside substitution reaction barriers are only slightly higher than the barrier for backside substitution for HF + MeFH+, and the difference in barrier heights for frontside and backside displacement seems negligible for the larger alkyl groups. Reaction barrier trends have been analysed and compared with the results of similar studies of the H2O/ROH2+ and NH3/RNH3+ systems (R = Me, Et, Pr(i), and Bu(t)). Compared to the two other classes, protonated fluorides have extreme structures which, with the exception of the Me substrate, are weakly bound complexes between an alkyl cation and HF. The results nourish the idea that nucleophilic substitution reactions are better understood in view of competition between frontside and backside substitution than from the traditional S(N)1/S(N)2 perspective.  相似文献   

8.
在缚酸剂Na2CO3存在下,使溶胀的氯甲基化交联聚苯乙烯(CMCPS)微球表面的氯甲基与对羟基苯甲醛(HBA)发生亲核取代反应,制得了醛基(AL)化改性的交联聚苯乙烯(ALCPS)微球;利用所制得的ALCPS微球与甘氨酸(GL)发生缩合反应,制备了同时含有席夫碱配基与羧基的席夫碱型螯合树脂AGCPS微球。采用红外光谱法表征了微球功能基团的结构变化,重点研究了CMCPS微球醛基化改性反应,考察了影响亲核取代反应的主要因素,推测和探讨了反应的机理。研究表明,CMCPS微球表面的苄氯基团与HBA缩合反应的速率与HBA的浓度无关,属于典型的SN1反应;使用极性较强的溶剂DMF,在较高的反应温度(90℃)下,有利于亲核取代反应的进行。所得席夫碱型螯合树脂对铜离子具有较强的螯合吸附能力。  相似文献   

9.
The reaction of quinazoline-4(3H)-thiones 2a-d with amino acid ester hydrochlorides in boiling solvents, under the basic catalysis, afforded the corresponding substitution products (3-6)a-e in low yield. The reaction could be improved by carrying it without a solvent yielding imidazo[1,2-c]- and pyrimido[1,2-c]quinazolines (7-10)a-e . The antibacterial and antifungal activities of the prepared compounds were tested.  相似文献   

10.
To prepare 10-trifluoromethyl analogues of important antimalarials such as artemether and artesunate, the substitution reaction of the 10-trifluoromethyl hemiketal 6 and bromide 4 derived from artemisinin was investigated. While 6 appeared to be unreactive under various conditions, bromide 4 could easily undergo substitution with methanol under electrophilic assistance or noncatalyzed conditions. Optimization of the reaction revealed the role of CH(2)Cl(2) as solvent to avoid the competitive elimination process and the crucial influence of hexafluoro-2-propanol (HFIP) in increasing the rate and the stereoselectivity of the substitution reaction (de >98%). The efficiency of this reaction was exemplified with various alcohols and carboxylates (yield up to 89%).  相似文献   

11.
The mechanism of reaction of the β-aminocrotonate of butanediol (βACB) with 4-chloro-2-hexene (4C2H), a model compound for the allylic chlorine in polyvinyl chloride), was studied in THF or dichloroethane at 60°C by gas and liquid chromatography. The reaction, which needs ZnCl2 as a catalyst, leads to substitution products through the primary amine group and the hydrogen atom of the trisubstituted double bond. βACB reacts with HCl to give NH4Cl and a set of complex organic products. NH4Cl and the substitution products are able to complex ZnCl, inhibiting its catalytic activity. In combination with other stabilizers, βACB strongly induces the substitution reaction versus the dehydro-chlorination. In the polymer at 190°C, it increases very much the time of action of the stabilizers; it acts as an HC1 acceptor but also it may be substituted on the polymer even without catalysts. Synergistic effects are observed with epoxy compounds or indole derivatives.  相似文献   

12.
Polymer-supported 1,5,7-triazabicyclo[4.4.0]deca-5-ene (TBD) was prepared by the reaction of chloromethylated polystyrene resins, cross-linked with 2 mol% of divinylbenzene, with TBD. The reaction of benzoic acid with bromobutane was carried out in toluene or acetonitrile in the presence of polymer-supported TBD or polymer-supported 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU). The rate of esterification was dependent on the ring substitution of the supported bases and the solvent. The reduced ring substitution resulted in the increased swellability of the polymeric beads containing benzoate or bromide ions for toluene and thereby in the increased reaction rate. The rate with the high ring-substituted bases increased in acetonitrile, because of the high swellability of the immobilized salts and the high nucleophilicity of benzoate anion in the solvent. The supported bases were also effective for alkylation of an active methylene compound with bromoalkane.  相似文献   

13.
3,5-Dinitro-4-(phenylsulfonyl)pyrazole (5) obtained by oxidation of 3,5-dinitro-4-(phenylthio)pyrazole with 30% H2O2 in AcOH was involved into nucleophilic substitution reaction with thiophenol, which proceeded with substitution of the phenylsulfonyl group at position 4. N-Methyl-3,5-dinitro-4-(phenylsulfonyl)pyrazole obtained by methylation of 5 with dimethyl sulfate was involved into nucleophilic substitution reaction with thiophenol, p-bromophenol, and morpholine with the regioselective substitution of the nitro group at position 5 to form 5-R-3-nitro-4-(phenylsulfonyl)pyrazoles.  相似文献   

14.
A number of 2(2-hydroxyphenyl)2H-benzotriazoles substituted in the 5-position with reactive hydroxyl and carboxyl groups have been synthesized. They include compounds with a tert butyl substitution in the 3-position or without substitution. The latter compounds were subjected to a reaction with N-hydroxy-methyl-(meth)acrylamide to form acrylic polymerizable 3(meth)acrylamidomethyl-2(2-hydroxyphenyl)2H-benzotriaz-oles. Hydroxyl reactive compounds were allowed to react with long chain acids, compounds with carboxyl groups with long chain hydrocarbon, fluorocarbon or silicon oligomer alcohols. The polymerizable 2(2-hydroxyphenyl)2H-benzotriazoles were then copolymerized with methyl methacrylate. Films made from such polymers showed by contact angle measurements, substantial migration of the fluorocarbon or silicon component to the surface of the film.  相似文献   

15.
The zirconium imido complex Cp2(THF)Zr=NSi(t-Bu)Me2 (1) reacts with allylic ethers, chlorides, and bromides to give exclusively the products of the SN2' reaction; i.e., attack at the allylic position remote from the leaving group with migration of the double bond. The primary amine products can be isolated in excellent yields, after in situ Cbz protection, in the presence of variety of functional groups. Good diastereoselectivity and complete stereoselectivity allowed the formation of enantioenriched allylic amines from enantioenriched allylic ethers. Regiospecific substitution with 1 has also been achieved with allylic fluorides, which are notoriously poor substrates in other substitution reactions. On the basis of rate and kinetic isotope effect studies, we propose a general mechanism for the allylic substitution reactions with 1 which involves dissociation of THF and binding of the substrate, followed by the substitution step. In a DFT study of the substitution reaction, we identified a six-membered closed transition state for the substitution step and other relevant stationary points along the reaction coordinate. This study shows that the substitution reaction can be described as a concerted asynchronous [3,3]-sigmatropic rearrangement. This detailed knowledge of the reaction mechanism provides a rationale for the origins of the observed regio-, diastereo-, and stereoselectivity and of the unusual reactivity profile observed in the reaction.  相似文献   

16.
The evolution of the ν C (SINGLE BOND) Cl bands of the infrared spectrum of a Bernoullian though slightly isotactic poly(vinyl chloride) (PVC), with both the degree of SN2 substitution reaction with sodium benzenethiolate, as studied earlier, and the increase of the nucleophile infrared bands, has been studied by FTIR spectroscopy. In a parallel way, the changes in the same bands, in particular those at 615 and 637 cm−1, presumably induced by SN2 substitution, have been estimated, theoretically, by comparing the sequential order and the number of the distinct conformationally sensitive vibration modes of C(SINGLE BOND)Cl bond, prior and after substitution, for a series of polymer sequences containing the reactive sites, namely the isotactic mmr tetrad and the heterotactic rmrr pentad, according to earlier work. The experimental behaviour of the νC(SINGLE BOND)Cl bands is found to be in close agreement with the theoretical expectations, thereby allowing two main conclusions to be drawn: (i) during the early stage going up to conversions of 10–12%, the reaction proceeds in a nearly exclusive manner, by the mmr and rmrr terminal of long isotactic and syndiotactic sequences, respectively; and (ii) any reaction event throughout the substitution process proves to be highly dependent upon the local environment in which each of the foregoing reactive structures finds itself. In summary, the local configurational nature of the mechanisms of analogous reactions of polymers is strongly suggested on the grounds of the results given herein. © 1996 John Wiley & Sons, Inc.  相似文献   

17.
DFT calculations at the B3LYP/6-31G(d,p) level have been performed to explore the substitution reactions of silylenoid H(2)SiLiF with XH(n) hydrides, where XH(n) = CH(4), NH(3), H(2)O, HF, SiH(4), PH(3), H(2)S, and HCl. We have identified a previously unreported reaction pathway on each reaction surface, H(2)SiLiF + XH(n) --> H(3)SiF + LiXH(n-1), which involved the initial formation of an association complex via a five-membered cyclic transition state to form an intermediate followed by the substituted product H(3)SiF with LiXH(n-1) dissociating. These theoretical calculations suggest that (i) there is a very clear trend toward lower activation barriers and more exothermic interactions on going from left to right along a given row in the periodic table, and (ii) for the second-row hydrides, the substitution reactions are more exothermic than for the first-row hydrides and the reaction barriers are lower. The solvent effects were considered by means of the polarized continuum model (PCM) using THF as a solvent. The presence of THF solvent disfavors slightly the substitution reaction. Compared to the previously reported insertions and H(2)-elimination reactions of H(2)SiLiF and XH(n), the substitution reactions should be most favorable.  相似文献   

18.
Biologically active 1,3-dihydroxy-2-(6-substituted-9-purinyl)cyclohexanes (VI-X) were prepared by a reaction between trans-2-amino-1,3-cyclohexanediol (I) and 4,6-dichloro-5-amino-pyrimidine (II), followed by cyclization and nucleophilic substitution. Compounds VI and VII showed a weak inhibitory effect against Xanthomonas oryzae, a pathogenic bacteria of leaf blight of rice plant.  相似文献   

19.
It has been shown that 1,1'-oxalyldiimidazole (ODI) is formed as an intermediate in the imidazole-catalyzed reaction of oxalate esters with hydrogen peroxide. Therefore, the kinetics of the chemiluminescence reaction of 1,1'-oxalyldiimidazole (ODI) with hydrogen peroxide in the presence of a fluorophore was investigated in order to further elucidate the mechanism of the peroxyoxalate chemiluminescence reaction. The effects of concentrations of ODI, hydrogen peroxide, imidazole (ImH), the general-base catalysts lutidine and collidine, and temperature on the chemiluminescence profile and relative quantum efficiency in the solvent acetonitrile were determined using the stopped-flow technique. Pseudo-first-order rate constant measurements were made for concentrations of either H2O2 or ODI in large excess. All of the reaction kinetics are consistent with a mechanism in which the reaction is initiated by a base-catalyzed substitution of hydrogen peroxide for imidazole in ODI to form an imidazoyl peracid (Im(CO)2OOH). In the presence of a large excess of H2O2, this intermediate rapidly decays with both a zero- and first-order dependence on the H2O2 concentration. It is proposed that the zero-order process reflects a cyclization of this intermediate to form a species capable of exciting a fluorophore via the "chemically initiated electron exchange mechanism" (CIEEL), while the first-order process results from the substitution of an additional molecule of hydrogen peroxide to the imidazoyl peracid to form dihydroperoxyoxalate, reducing the observed quantum yield. Under conditions of a large excess of ODI, the reaction is more than 1 order of magnitude more efficient at producing light, and the quantum yield increases linearly with increasing ODI concentration. Again, it is proposed that the slow initiating step of the reaction involves the substitution of H2O2 for imidazole to form the imidazoyl peracid. This intermediate may decay by either cyclization or by reaction with another ODI molecule to form a cyclic peroxide that is much more efficient at energy transfer with the fluorophore. The reaction kinetics clearly distinguishes two separate pathways for the chemiluminescent reaction.  相似文献   

20.
The kinetics of substitution, by thiourea, of ethylenediamine (en) or N,N'-dimethylethylenediamine (Me(2)en) coordinated to palladium(II) in the complexes [Pd(4,4'-R(2)bpy)(en)](PF(6))(2) (bpy = 2,2'-bipyridine; R = H or Me), [Pd(en)(2)](PF(6))(2) and [Pd(Me(2)en)(2)](PF(6))(2) have been studied at 25 degrees C, pH 7 and various ionic strength values, in the presence of calf thymus DNA. The rate of the reaction in water depends on ionic strength, pH, and nucleophile concentration; at fixed pH and ionic strength the k(obsd) values are correlated to the square of the thiourea concentration. This rate law is not altered by the presence of DNA, but the rate of reaction is influenced, depending on the nature of ancillary ligand, L-L, bound to palladium. DNA inhibits the substitution process when L-L is bpy or 4,4'-Me(2)bpy and catalyzes the same reaction when L-L is en or Me(2)en. These opposite kinetic effects can be related to the noncovalent interactions of the various complexes with the DNA double helix. Inhibition of the reactivity of the complexes [Pd(4,4'-R(2)bpy)(en)](2+) is due to protection of the reaction center from nucleophile attack by DNA. Acceleration of the reaction when L-L is en or Me(2)en is related to the dependence of the rate of reaction on pH. If, due to the higher activity of water under the electric field of phosphate groups, hydronium ion concentration on DNA surface is higher than in the bulk solution, the enzyme-like dependence of the rate of reaction on [DNA] is due to progressive accumulation of the complexes around the double helix. Regardless of the complexes' nature, the rate constant values obtained in DNA at pH 7 correspond to values determined in water at pH 5. This pH value on the DNA surface, lower by about two units with respect to the bulk solution, is in good agreement with theoretical predictions. Acceleration of ethylenediamine substitution has been observed for all of the complexes studied in the presence of sodium polyvinylsulfonate.  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