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.
N-aroyl beta-lactams are imides with exo- and endocyclic acyl centres which react with amines in aqueous solution to give the ring opened beta-lactam aminolysis product. Unlike the strongly base catalysed aminolysis of beta-lactam antiobiotics, such as penicillins and cephaloridines, the rate law for the aminolysis of N-aroyl beta-lactams is dominated by a term with a first-order dependence on amine concentration in its free base form, indicative of an uncatalysed aminolysis reaction. The second-order rate constants for this uncatalysed aminolysis of N-p-methoxybenzoyl beta-lactam with a series of substituted amines generates a Br?nsted betanuc value of +0.90. This is indicative of a large development of positive effective charge on the amine nucleophile in the transition state. Similarly, the rate constants for the reaction of 2-cyanoethylamine with substituted N-aroyl beta-lactams gives a Br?nsted betalg value of -1.03 for different amide leaving groups and is indicative of considerable change in effective charge on the leaving group in the transition state. These observations are compatible with either a late transition state for the formation of the tetrahedral intermediate of a stepwise mechanism or a concerted mechanism with simultaneous bond formation and fission in which the amide leaving group is expelled as an anion. Amide anion expulsion is also indicated by an insignificant solvent kinetic isotope effect, kH2ORNH2/kD2ORNH2, of 1.01 for the aminolysis of N-benzoyl beta-lactam with 2-methoxyethylamine. The Br?nsted betalg value decreases from -1.03 to -0.71 as the amine nucleophile is changed from 2-cyanoethylamine to propylamine. The Br?nsted betanuc value is more invariant although it changes from +0.90 to +0.85 on changing the amide leaving group from p-methoxy to p-chloro substituted. The sensitivity of the Br?nsted betanuc and betalg values to the nucleofugality of the amide leaving group and the nucleophilicity of the amine nucleophiles, respectively, indicate coupled bond formation and bond fission processes.  相似文献   

2.
The rate of aminolysis of benzylpenicillin and cephaloridine by hydroxylamine, unlike other amines, shows only a first order dependence on amine concentration. The rate enhancement compared with that predicted from a Bronsted plot for other primary amines with benzylpenicillin is greater than 10(6). This is much more than an alpha-effect and is compatible with rate-limiting formation of the tetrahedral intermediate due to a rapid intramolecular general acid catalysed breakdown of the intermediate. For cephaloridine, the rate enhancement is greater than 10(4) which demonstrates that beta-lactam C-N bond fission and expulsion of the leaving group at C3' are not concerted.  相似文献   

3.
Activated sulfonyl derivatives, similar to acyl ones, usually undergo aminolysis with amines in water as nucleophilic attack by the amine is preferred to hydrolysis. However, despite being active sulfonyl derivatives, four-membered heterocyclic sulfonamides, beta-sultams, do not undergo aminolysis in aqueous solution but preferentially react to give hydrolysis products only. The rate of the reaction of beta-sultams in buffered solutions of simple primary amines shows a first-order dependence on amine concentrations attributed to general base-catalyzed hydrolysis by the amine. Even N-benzyl-4,4-dimethyl-3-oxo-beta-sultam, which is both a beta-sultam and a beta-lactam, undergoes hydrolysis at the sulfonyl center rather than aminolysis at either the sulfonyl or acyl center. The solvent kinetic isotope effects (SKIE, k(H(2)O)/k(D(2)O)) for the amine-catalyzed hydrolyses are 1.4 and 1.9 for the hydrolysis of N-benzoyl-beta-sultam and N-benzyl-4,4-dimethyl-3-oxo-beta-sultam, respectively, compatible with a general base-catalyzed mechanism. The amine-catalyzed hydrolysis gives a Bronsted beta value of +0.9 for both N-benzoyl beta-sultam and N-benzyl-4,4-dimethyl-3-oxo-beta-sultam, indicating that the general base amine is almost fully protonated in the transition state. A general base-catalyzed mechanism for hydrolysis rather than nucleophilic attack was also deduced for the reaction of N-benzyl-4,4-dimethyl-3-oxo-beta-sultam with carboxylate anions based on a SKIE of 1.7-1.9 and rate constants which fit the Bronsted plot for amines. In contrast to acyl transfer reactions, those for sulfonyl transfer appear to show an inverse reactivity-selectivity relationshipthe most active compounds being the most selective. The lack of reactivity of beta-sultams toward amine nucleophiles appears to be related to the mechanism of ring opening of beta-sultams with a decreased reactivity toward amines relative to hydroxide ion, probably related to the expulsion of the relatively poor leaving group amide anion.  相似文献   

4.
Differential pulse polarography (d.p.p.) is used to study the degradation of cephalexin. Hydrogen sulphide, evolved during the degradation of cephalexin solutions, was removed continuously in a stream of nitrogen and determined periodically. Other electroactive degradation products were observed by d.p.p. of the degraded sample solutions. The degradation mechanism is highly dependent on pH, the initial concentration of cephalexin, temperature, the particular buffer used, and the presence of dissolved oxygen. The formation and degradation of the diketopiperazine derivative formed by intramolecular aminolysis, particularly at neutral pH, can be followed by means of its polarographic peak at -0.9 V (pH 7.4). Approximately half the total sulphur originally present in cephalexin is liberated as hydrogen sulphide at pH 7.4 at 37°C. Increasing the degradation temperature to 80°C and sweeping out the hydrogen sulphide with nitrogen increases the yield of a major product which gives a peak at -1.26 V. At pH 8.5 (80°C. 100 μg cephalexin ml-1) the percentage of the sulphur evolved as hydrogen sulphide increases with time, and a peak appears at -0.96 V (probably 2-hydroxy-3-phenyl-6-methylpyrazine) which increases as the peak at -1.26 V becomes smaller. Other products formed under different conditions (concentration, pH, temperature) are reported. At pH 3 (80°C) only 8% conversion via intramolecular aminolysis and 5% evolution of total sulphur is indicated after four hours.  相似文献   

5.
The Zn(2+)-tris (hydroxymethyl)aminomethane (Tris) system has a great catalytic effect on the hydrolysis and aminolysis of some beta-lactam antibiotics. In order to ascertain the mechanism of this catalysis we have analysed the effects of the beta-lactam antibiotic structure. First we studied the kinetics of the decomposition of imipenem, SCH 29482, aztreonam and nocardicin A in aqueous solution of Tris at 35.0 degrees C, 0.5 mol.dm-3 ionic strength and in the presence of metal ions (Zn2+, Cd2+, Co2+, Cu2+, Ni2+ and Mn2+). From these studies, we conclude that Tris and metal ions (in separate solutions) exert a great catalytic effect on the hydrolysis of imipenem and SCH 29482. We suggest that in metal ion solutions a 1:1 complex is formed between the metal ion and beta-lactam antibiotic, which is attacked by hydroxide ions. Studies of the degradation of the antibiotics studied in solutions of Tris and metal ions together indicate that the systems Cd(2+)-Tris and Zn(2+)-Tris have a great catalytic effect on the hydrolysis and aminolysis of imipenem and SCH 29482. We suggest that this catalysis takes place via a ternary complex in which the metal ion plays a double role by (a) placing the antibiotic and the Tris in the right position for the reaction and (b) lowering the pKa of the hydroxide group of Tris, which is coordinated with the metal ion, generating a strong nucleophile.  相似文献   

6.
Plots of log k(0) vs pH for the cyclization of trifluoroethyl and phenyl 2-aminomethylbenzoate to phthalimidine at 30 degrees C in H(2)O are linear with slopes of 1.0 at pH >3. The values of the second-order rate constants k(OH) for apparent OH(-) catalysis in the cyclization reactions are 1.7 x 10(5) and 5.7 x 10(7) M(-)(1) s(-)(1), respectively. These rate constants are 10(5)- and 10(7)-fold greater than for alkaline hydrolysis of trifluoroethyl and phenyl benzoate. The k(OH) for cyclization of the methyl ester is 7.2 x 10(3) M(-)(1) s(-)(1). Bimolecular general base catalysis occurs in the intramolecular nucleophilic reactions of the neutral species. The value of the Bronsted coefficient beta for the trifluoroethyl ester is 0.7. The rate-limiting step in the general base catalyzed reaction involves proton transfer in concert with leaving group departure. The mechanism involving rate-determining proton transfer exemplified by the methyl ester in this series (beta = 1.0) can then be considered a limiting case of the concerted mechanism. General acid catalysis of the neutral species reaction or a kinetic equivalent also occurs when the leaving group is good (pK(a) 相似文献   

7.
A computational and experimental study is presented, which provides the first evidence that amine has an opportunity to engage in bonding with pentathiepin to promote its decomposition. The study provides mechanistic insight into the process that gives rise to pentathiepin biological activity. Primary or secondary amine will allow for an intramolecular addition to the pentathiepin ring at the nearest sulfur (S1). In contrast, tertiary amine adds reversibly to S1, because nitrogen cannot lose its positive charge by deprotonation. This precludes the amine promotion step. An energetically low-lying process is characterized, corresponding to S3-loss triggered by nucleophilic activation with a primary or secondary amine. Pentathiepin desulfurization via S3-unit transfer is supported by a trapping study with norbornene. That the amine may confer an enhanced reactivity in the natural products varacin, 1, and lissoclinotoxin A, 2, adds to the understanding of the pathway for pentathiepin activation and may provide new design concepts that have potential applications for this class of biocides.  相似文献   

8.
Homocysteine thiolactone (tHcy) is deemed a risk factor for cardiovascular diseases and strokes, presumably because it acylates the side chain of protein lysine residues (“N‐homocysteinylation”), thereby causing protein damage and autoimmune responses. We analysed the kinetics of hydrolysis and aminolysis of tHcy and two related thiolactones (γ‐thiobutyrolactone and N‐trimethyl‐tHcy), and we have thereby described the first detailed mechanism of thiolactone aminolysis. As opposed to the previously studied (thio and oxo)esters and (oxo)lactones, aminolysis of thiolactones was found to be first order with respect to amine concentration. Anchimeric assistance by the α‐amino group of tHcy (through general acid/base catalysis) could not be detected, and the Brønsted plot (nucleophilicity versus pKa) for aminolysis yielded a slope (βnuc) value of 0.66. These data support a mechanism of aminolysis where the rate‐determining step is the formation of a zwitterionic tetrahedral intermediate. The βnuc value and steric factors dictate a regime whereby, at physiological pH values (pH 7.4), maximal reactivity of tHcy is exhibited with primary amine groups with a pKa value of 7.7; this allows the reactivity of various protein amino groups towards N‐homocysteinylation to be predicted.  相似文献   

9.
4-PEG接枝苯乙烯-马来酸酐交替共聚物的合成及功能化   总被引:2,自引:0,他引:2  
采用普通自由基聚合和可逆加成一断裂链转移(RAFT)自由基聚合方法合成了对位PEG取代苯乙烯(PEG-g-St)和马来酸酐的交替共聚物(P((PEG—g—St)-alt-MA)),”CNMR分析表明PEG-g-St和马来酸酐单元采取交替的序列结构.利用反应性基团-马来酸酐单元的水解以及胺解可以制备功能性的PEG聚合物.以月桂胺为模型小分子研究了该聚合物的胺解,得到4-PEG-苯乙烯与羧酸基团以及疏水烷烃的交替序列聚合物,该双亲聚合物在水溶液中形成组装体.  相似文献   

10.
Cyclic oligocarbonates are synthesized by two-phase interfacial reactions of aryl bischloroformates in the presence of an amine phase transfer catalyst and an alkali: nClOCOArOCOCl + 4nNaOH → (OArOCO)n + 2nNaCl + nNa2CO3 + 2nH2O. The organocarbonate product may be a macrocyclic oligomer or a linear chain polymer. A qualitative mechanism for this behavior has been proposed by Brunelle, Boden, and co-workers. Four steps are identifiable: activation of the aryl bischloroformate by the amine catalyst, hydrolysis of a portion of this intermediate at the aqueous/organic phase interface, oligomerization between activated and hydrolyzed moieties also at the interface, and chain terminating carbamate formation that leads to polymer. An important modification is made within the framework of the Brunelle and Boden mechanism. While the intramolecular cyclization reaction is formally second-order overall, it behaves as a first-order process. The kinetic constants for both this pseudo-first-order cyclization step and the corresponding second-order linearization step are simply related. It is speculated that the above relationship can be generalized for a whole class of pseudo-first- and second-order rate constants for similar macrocyclic reactions.  相似文献   

11.
A practical one-pot synthetic strategy for the efficient synthesis of a range of structurally interesting and bioactive quinoline-based tetracycles has been developed. A key step in the synthesis is a tandem three-component reaction of heteroaromatic amine, methyl 2-formylbenzoate and (t)butyl isonitrile, followed by TFA-mediated lactamization via intramolecular aminolysis of an adjacent ester. Results related to a kinase-panel screening for several selected compounds are also discussed in this article.  相似文献   

12.
Beta-sultams are the sulfonyl analogues of beta-lactams, and N-acyl beta-sultams are novel inactivators of the class C beta-lactamase of Enterobacter cloacae P99. They sulfonylate the active site serine residue to form a sulfonate ester which subsequently undergoes C-O bond fission and formation of a dehydroalanine residue by elimination of the sulfonate anion as shown by electrospray ionization mass spectroscopy. The analogous N-acyl beta-lactams are substrates for beta-lactamase and undergo enzyme-catalyzed hydrolysis presumably by the normal acylation-deacylation process. The rates of acylation of the enzyme by the beta-lactams, measured by the second-order rate constant for hydrolysis, kcat/K(m), and those of sulfonylation by the beta-sultams, measured by the second-order rate constant for inactivation, k(i), both show a similar pH dependence to that exhibited by the beta-lactamase-catalyzed hydrolysis of beta-lactam antibiotics. Electron-withdrawing groups in the aryl residue of the leaving group of N-aroyl beta-lactams increase the rate of alkaline hydrolysis and give a Bronsted beta(lg) of -0.55, indicative of a late transition state for rate-limiting formation of the tetrahedral intermediate. Interestingly, the corresponding Bronsted beta(lg) for the beta-lactamase-catalyzed hydrolysis of the same substrates is -0.06, indicative of an earlier transition state for the enzyme-catalyzed reaction. By contrast, although the Bronsted beta(lg) for the alkaline hydrolysis of N-aroyl beta-sultams is -0.73, similar to that for the beta-lactams, that for the sulfonylation of beta-lactamase by these compounds is -1.46, compatible with significant amide anion expulsion/S-N fission in the transition state. In this case, the enzyme reaction displays a later transition state compared with hydroxide-ion-catalyzed hydrolysis of the beta-sultam.  相似文献   

13.
The aminolysis of poly(styrene-co-methyl acrylate) (SMA) by octadecylamine in solution and in the melt has been reported in Part I. We now have studied the aminolysis of poly(ethylene-co-methyl acrylate) (EMA) with the same amine in the melt and compared the two sets of data in this paper. With EMA, the data confirmed and precised the catalytic mechanism proposed in Part I. The best tautomeric catalysts are the ones which form an eight-atom ring structure with the ester and amine groups. With EMA aminolysis is faster than with SMA because of the steric hindrance of phenyl groups in SMA. But EMA aminolysis remains a rather slow reaction. In a corotating twin-screw extruder the conversion was only around 4% at 220°C with a mean residence time of 150 s. It was also shown that the EMA/octadecylamine/catalyst system, like the SMA system, is homogeneous in the molten state at temperatures around 200°C.  相似文献   

14.
Semiempirical calculations were used to conduct a comprehensive study of the thiolysis of the fundamental core of penicillins and cephalosporins. The significance of the intramolecular protonation of the β‐lactam nitrogen in the formation and cleavage of the tetrahedral intermediate ( T in Scheme 1 ) was examined in two thiols bearing substituents of different basicity in β with respect to the thiol group in the attacking nucleophile, namely 2‐mercaptoethanol ( 6 ) and 2‐mercaptoethylamine ( 7 ). Based on the results, the rate‐determining step in the reaction of penicillins is the cleavage of the tetrahedral intermediate, consistent with an intramolecular acid catalysis process in their thiolysis by 2‐mercaptoethylamine. On the other hand, the rate‐determining step in the reaction of cephalosporins, which possess an appropriate leaving group at position 3', is the formation of the tetrahedral intermediate, so the desolvation energy of the nucleophile is a major contributor to the overall energy of the process. This differential behavior between the two types of β‐lactam bicycles arises from the presence of the acetate group at 3' and the delocalization of π electrons over the N5–C4–C3 system in cephalosporins; this favors the formation of a thiolate with the 5‐ethoxymethylene‐1,3‐thiazine group in the cleavage of the tetrahedral intermediate, which is stabilized by an intramolecular hydrogen bond between N5 and the alcohol or amine group in β of the attacking thiol. The theoretical results are consistent with previous experimental data showing that, unlike penicillins, cephalosporins undergo no intramolecular acid catalysis in their thiolysis. © 2005 Wiley Periodicals, Inc. Int J Chem Kinet 37: 434–443, 2005  相似文献   

15.
The influences of buffer pH, buffer concentration and buffer electrolyte on the migration behavior and separation of 12 cephalosporin antibiotics in capillary zone electrophoresis using three different types of buffer electrolyte, including phosphate, citrate, and 2-(N-morpholino)ethanesulfonate (MES), were investigated. The results indicate that, although buffer pH is a crucial parameter, buffer concentration also plays an important role in the separation of cephalosporins, particularly when cefuroxime and cefazolin, cephalexin and cefaclor, or cefotaxime and cephapirin are present as analytes at the same time. The electrophoretic mobility of cephalosporins and electroosmotic mobility measured in citrate and MES buffers are remarkably different from those measured in phosphate buffer. With citrate buffer, optimum buffer concentration is confined to a small range (35-40 mM), whereas buffer concentrations up to 300 mM can be used with MES buffer. Complete separations of 12 cephalosporins could be satisfactorily achieved with these three buffers under various optimum conditions. However, the separability of 12 cephalosporins with citrate or MES buffer is better than that with phosphate buffer. As a consequence of a greater electrophoretic mobility of cephalosporins than the electroosmotic mobility with citrate buffer at pH below about 5, some cephalosporins are not detectable. The cloudiness of the peak identification and of the magnitudes of the electrophoretic mobility of cefotaxime and cefuroxime reported previously are clarified. In addition, the pKa values of cephradine, cephalexin, cefaclor, and cephapirin attributed to the deprotonation of either an amino group or a pyridinium group are reported, and the migration behavior of these cephalosporins in the pH range studied is quantitatively described.  相似文献   

16.
The pH-independent, acid-catalyzed and base-catalyzed hydrolyses of N-acyloxymethylazetidin-2-ones all occur at the ester function. The pH-independent hydrolysis involves rate-limiting alkyl C-O fission and formation of an exocyclic beta-lactam iminum ion. This iminium ion is then trapped by water at the exocyclic iminium carbon atom, rather than at the beta-lactam carbonyl carbon atom, to form the corresponding N-hydroxymethylazetidin-2-ones. Calculations carried out at the B3LYP/6-31+G(d) level of theory also support that nucleophilic attack by water takes place at the exocyclic carbon rather than at the beta-lactam carbonyl carbon of the iminium ion. The mechanism for the acid-catalyzed pathway involves a preequilibrium protonation, probably at the beta-lactam nitrogen, followed by rate-limiting alkyl C-O fission with formation of an exocyclic iminum ion. The base-catalyzed hydrolysis involves rate-limiting hydroxide attack at the ester carbonyl carbon. These results imply formation of a beta-lactam system containing a positively charged amide nitrogen atom that hydrolyzes via a pathway that preserves the beta-lactam structure in the product and provide further evidence that cleavage of the beta-lactam C-N bond is not as facile as is commonly imagined.  相似文献   

17.
The reactions of the title substrate (1) with a series of secondary alicyclic amines are subjected to a kinetic investigation in 44 wt% ethanol‐water, at 25.0°C, ionic strength 0.2 M (KCl). Under amine excess over the substrate, pseudo‐first‐order rate coefficients (kobs) are obtained. Plots of kobs against [NH], where NH is the free amine, are nonlinear upwards, except the reactions of piperidine, which show linear plots. According to the kinetic results and the analysis of products, a reaction scheme is proposed with two tetrahedral intermediates, one zwitterionic (T±) and another anionic (T), with a kinetically significant proton transfer from T± to an amine to yield T (k3 step). By nonlinear least‐squares fitting of an equation derived from the scheme to the experimental points, the rate microcoefficients involved in the reactions are determined. Comparison of the kinetics of the title reactions with the linear kobs vs. [NH] plots found in the same aminolysis of O‐ethyl 4‐nitrophenyl dithiocarbonate (2) in the same solvent shows that the rate coefficient for leaving group expulsion from T± (k2) is larger for 2 due to a stronger push by EtO than PhO. The k3 value is the same for both reactions since both proton transfers are diffusion controlled. Comparison of the title reactions with the same aminolysis of phenyl 4‐nitrophenyl thionocarbonate (3) in water indicates that (i) the k2 value is larger for the aminolysis of 1 due to the less basic nucleofuge involved and the small solvent effect on k2, (ii) the k3 value is smaller for the reactions of 1 due to the more viscous solvent, (iii) the rate coefficient for amine expulsion from T± (k−1) is larger for the aminolysis of 1 than that of 3 due to a solvent effect, and (iv) the value of the rate coefficient for amine attack (k1) is smaller for the aminolysis of 1 in aqueous ethanol, which can be explained by a predominant solvent effect relative to the electron‐withdrawing effect from the nucleofuge. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 839–845, 1999  相似文献   

18.
Five functional silanes--3-aminopropyltriethoxysilane (APTES), 3-aminopropyltrimethoxysilane (APTMS), N-(2-aminoethyl)-3-aminopropyltriethoxysilane (AEAPTES), N-(2-aminoethyl)-3-aminopropyltrimethoxysilane (AEAPTMS), and N-(6-aminohexyl)aminomethyltriethoxysilane (AHAMTES)--were assessed for the preparation of hydrolytically stable amine-functionalized silica substrates. These can be categorized into three groups (G1, G2, and G3) based on the intramolecular coordinating ability of the amine functionality to the silicon center. Silanizations were carried out in anhydrous toluene as well as in the vapor phase at elevated temperatures. Aminosilane-derived layers prepared in solution are multilayers in nature, and those produced in the vapor phase have monolayer characteristics. In general, vapor-phase reactions are much less sensitive to variations in humidity and reagent purity, are more practical than the solution-phase method, and generate more reproducible results. Intramolecular catalysis by the amine functionality is found to be important for both silanization and hydrolysis. The primary amine group in the G1 silanes (APTES and APTMS) can readily catalyze siloxane bond formation and hydrolysis to render their silane layers unstable toward hydrolysis. The amine functionality in the G3 silane (AHAMTES) is incapable of intramolecular catalysis of silanization so that stable siloxane bonds between the silane molecules and surface silanols do not form easily. The secondary amine group in the G2 silanes (AEAPTES and AEAPTMS), on the other hand, can catalyze siloxane bond formation, but the intramolecular catalysis of bond detachment is sterically hindered. The G2 silanes are the best candidates for preparing stable amine-functionalized surfaces. Between the two G2 aminosilanes, AEAPTES results in more reproducible silane layers than AEAPTMS in the vapor phase due to its lower sensitivity to water content in the reaction systems.  相似文献   

19.
Since bromides are well-known precursors of primary amines, diethylaluminum 12 bromo-1-dodecyl oxide has been prepared and used as an initiator for the ring-opening polymerization of ?-caprolactone and L-lactide. Uner strictly controlled conditions, the end-functionalization of the polyesters in quantitative and the bromo end-group is easily converted into an azide group whatever the polymeric backbone. The subsequent reduction of the azide into the expected primary amine has been investigated by catalytic transfer hydrogenation (CTH) in DMF and by hydrolysis in the presence of triphenylphosphine in THF, respectively.The hydrolysis reaction (PΦ3/H2O) is perturbed by a coupling reaction, which involves a protonate secondary amine and leads to a twofold increase in the polyester molecular weight. The CTH method gives rise to the expected ω-NH2 poly (?-caprolactone), in contrast to polylactide which seems to be unstable toward the nascent amine end group. Whatever the polarity of the medium (DMF or THF), aminolysis of polylactides is observed to occur and leads to the formation of an internal amide. © 1994 John Wiley & Sons, Inc.  相似文献   

20.
Beta-Amino alcohols N-2'-pyridylmethyl substituted 3 have been prepared in excellent yields under mild conditions by the first Lewis acid-catalyzed aminolysis of 1,2-epoxides 1 with the bihaptic amine 2-picolylamine (2) with use of 5 mol % of Al(OTf)(3) under solvent-free conditions. As a representative of a new class of ionic liquids, cis-5-[(4'-methylphenyl)sulfonyl]-1,2,3,4,4a,5,6,11a-octahydropyrido[1,2-a]quinoxalin-11-ium methanesulfonate (6) and its chloride derivative 7 have been synthesized under environmentally friendly conditions by the one-pot aminolysis of cyclohexene oxide (1a) with 2 and intramolecular cyclization of the resulting 2-[(pyridin-2'-yl)methylamino]cyclohexanol (3a).  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