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 93 毫秒
1.
In the reaction of ethyl isothiocyanatoacetate with diamines, followed by cyclization of the intermediate product, 3‐monosubstituted thiohydantoins have been obtained. It was found that the reaction course depends on the purity of the isothiocyanate used and also, in the case of dialkylaminoamines, the self‐cyclization occurs. Besides the dialkylamino derivatives of 3‐monosubstituted 2‐thiohydantoins also new monoalkylamino, amino and heterocyclic derivatives were synthesized. The aryldiazonium derivative of 3‐monosubstituted 2‐thiohydantoin yielded both respective phenol derivative after hydrolysis and the product of coupling with 2‐naphthol.  相似文献   

2.
The application of a straightforward biocatalytic technology for the reduction of racemic 2‐monosubstituted 3‐thiazolines, which are easily prepared via Asinger‐multicomponent reaction, is reported. The biocatalytic reduction yields racemic 2‐monosubstituted 3‐thiazolidines, which are difficult to be prepared by means of classic chemical routes, in moderate to high yields. Moreover, our study clarifies the stereochemical reaction course of the biocatalytic reduction. Furthermore, the efficiency of this biocatalytic technology is demonstrated in an experiment at an elevated substrate concentration of 60 mM leading to 96% conversion.  相似文献   

3.
α‐Amino nitriles tethered to alkenes through a urea linkage undergo intramolecular C‐alkenylation on treatment with base by attack of the lithionitrile derivatives on the N′‐alkenyl group. A geometry‐retentive alkene shift affords stereospecifically the E or Z isomer of the 5‐alkenyl‐4‐iminohydantoin products from the corresponding starting E ‐ or Z N ′‐alkenyl urea, each of which may be formed from the same N ‐allyl precursor by stereodivergent alkene isomerization. The reaction, formally a nucleophilic substitution at an sp2 carbon atom, allows the direct regioselective incorporation of mono‐, di‐, tri‐, and tetrasubstituted olefins at the α‐carbon of amino acid derivatives. The initially formed 5‐alkenyl iminohydantoins may be hydrolyzed and oxidatively deprotected to yield hydantoins and unsaturated α‐quaternary amino acids.  相似文献   

4.
The synthesis and structural elucidation of some novel 5,5′‐disubstituted spiro and nonspiro‐bis‐hydantoins are reported. The Bucherer Burge's method has been modified for the preparation of some 5,5′‐substituted bis(imidazolidine‐2,4‐dione) derivatives starting with diketones ( 1–5 ) and dialdehydes ( 6 , 7 ). In some cases, diastereoisomeric mixtures of compounds were obtained. The resulting bis‐hydantoins ( 8–11 , 13 , 14 ) have not to our knowledge been previously reported in the literature.  相似文献   

5.
The visible intramolecular charge transfer band of some carbanion monosubstituted 3‐(p‐halo‐phenyl)‐pyridazinium ylids in benzene solutions has been considered as an indicator of the ylid stability. The catalytic effect of light on 3 + 3 dipolar thermal dimerization reaction was revealed. The dimerization reaction was found to be of the second order in the ylid concentration. The rate of dimerization was found to be promoted by increasing the electronegativity of the substituents on the ylid carbanion. The associated intramolecular charge transfer band oscillator strength has been correlated with the rate constant. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 34: 613–619, 2002  相似文献   

6.
The reaction of an N‐monosubstituted amidine with a β‐ketoester to afford a pyrimidinone is sluggish at best under normal conditions. We now report that this reaction can be effected in moderate yield under high pressure. Thus, 2,6‐dichloro‐4‐pyridyl‐(N‐prop‐2‐ynyl)carboxamidine (4b) was reacted with three α‐substituted‐β‐ketoesters (2b‐d) at 10–16 kbar to afford herbicidal 2‐(2,6‐dichloro‐4‐pyridyl)‐3‐(prop‐2‐ynyl)‐4(3H)‐pyrimidinones 5b and 5c in 15 ‐ 43% yield. This result expands the scope of reactions promoted by application of high pressure.  相似文献   

7.
The novel and one‐pot synthesis of 5‐hydroxyl hydantoins from Ugi five‐component condensation procedures is reported. The transformation was promoted via a microwave heating method and provided 5‐hydroxyl hydantoins in workable yields, representing a direct approach to this synthetically challenging and biologically interesting pharmacophore.  相似文献   

8.
The synthesis of 6β-hydroxy- and 6β,7β-dihydroxy-8-alkyl-8-azabicyclo[3.2.1]octane-3-spiro-5′-hydantoins was stereoselectively achieved by Bucherer-Bergs reaction of the corresponding ketones. An α configuration on C3 was proposed for all hydantoins on the basis of spectral data.  相似文献   

9.
A facile, efficient and metal‐free synthetic approach to 3‐monosubstituted unsymmetrical 1,2,4,5‐tetrazines is presented. Dichloromethane (DCM) is for the first time recognized as a novel reagent in the synthetic chemistry of tetrazines. Using this novel approach 11 3‐aryl/alkyl 1,2,4,5‐tetrazines were prepared in excellent yields (up to 75 %). The mechanism of this new reaction, including the role of DCM in the tetrazine ring formation, has been investigated by 13C labeling of DCM, and is also presented and discussed as well as the photophysical and electrochemical properties.  相似文献   

10.
An asymmetric conjugate addition of 3‐monosubstituted oxindoles to a range of (E)‐1,4‐diaryl‐2‐buten‐1,4‐diones, catalyzed by commercially available cinchonine, is described. This organocatalytic asymmetric reaction affords a broad range of 3,3′‐disubstituted oxindoles that contain a 1,4‐dicarbonyl moiety and vicinal quaternary and tertiary stereogenic centers in high‐to‐excellent yields (up to 98 %), with excellent diastereomeric and moderate‐to‐high enantiomeric ratios (up to 99:1 and 95:5, respectively). Subsequently, cyclization of the 1,4‐dicarbonyl moiety in the resultant Michael adducts under different Paal–Knorr conditions results in two new kinds of 3,3′‐disubstituted oxindoles—3‐furanyl‐ and 3‐pyrrolyl‐3‐alkyl‐oxindoles—in high yields and good enantioselectivities. Notably, the studies presented here sufficiently confirm that this two‐step strategy of sequential conjugate addition/Paal–Knorr cyclization is not only an attractive method for the indirect enantioselective heteroarylation of 3‐alkyloxindoles, but also opens up new avenues toward asymmetric synthesis of structurally diverse 3,3′‐disubstituted oxindole derivatives.  相似文献   

11.
This work documents the influence of the position of single carboxymethyl group on the β‐cyclodextrin skeleton on the enantioselectivity. These synthesized monosubstituted carboxymethyl cyclodextrin (CD) derivatives, native β‐cyclodextrin, and commercially available carboxymethyl‐β‐cyclodextrin with degree of substitution approximately 3 were used as additives into the BGE consisting of phosphate buffer at 20 mmol/L concentration, pH 2.5, and several biologically significant low‐molecular‐mass chiral compounds were enantioseparated by CE. The results indicate that different substituent location on β‐cyclodextrin skeleton has a significant influence on the enantioseparation of the investigated enantiomers. The enantioselectivity of 2IO‐regioisomer was better than with native β‐cyclodextrin. Comparable results to native β‐cyclodextrin were obtained for 6IO‐ regioisomer and the enantioselectivity of 3IO‐regioisomer was even worse than with native β‐cyclodextrin. Commercially available derivative of CD provides better resolutions than the monosubstituted carboxymethyl CD derivatives for most of the investigated analytes.  相似文献   

12.
The structure of 4‐methoxy‐1‐naphthol, C11H10O2, (I), contains an intermolecular O—H...O hydrogen bond which links the molecules into a simple C(2) chain running parallel to the shortest crystallographic b axis. This chain is reinforced by intermolecular π–π stacking interactions. Comparisons are drawn between the crystal structure of (I) and those of several of its simple analogues, including 1‐naphthol and some monosubstituted derivatives, and that of its isomer 7‐methoxy‐2‐naphthol. This comparison shows a close similarity in the packing of the molecules of its simple analogues that form π‐stacks along the shortest crystallographic axes. A substantial spatial overlap is observed between adjacent molecules in such stacks. In this group of monosubstituted naphthols, the overlap depends mainly on the position of the substituents carried by the naphthalene moiety, and the extent of the overlap depends on the substituent type. By contrast with (I), in the crystal structure of the isomeric 7‐methoxy‐2‐naphthol there are no O—H...O hydrogen bonds or π–π stacking interactions, and sheets are formed by O—H...π and C—H...π interactions.  相似文献   

13.
Previously unknown azomethylene derivatives of 4‐chloro‐5H‐1,2,3‐dithiazole 5–7 were synthesized by the reaction of the Appel salt 1 with N‐monosubstituted hydrazones 2–4. It was shown that they could be transformed into heterocyclic compounds 8–10.  相似文献   

14.
Sulphamoyl chlorides and chlorosulphonyl isocyanate react with monosubstituted hydrazones and alkylhydrazonates to sulphamoyl hydrazones and sulphamoyl hydrazonates respectively. Reaction of benzil monoalkylhydrazones with chlorosulphonyl isocyanate results in formation of 2‐alkyl‐4,5‐aryl‐2H‐ [1λ6,2,3,6]‐thiatriazine‐1,1‐dioxides.  相似文献   

15.
The catalytic asymmetric synthesis of both α‐substituted and α,α‐disubstituted (quaternary) β‐tetralones through direct α‐functionalization of the corresponding β‐tetralone precursor remains elusive. A designed Brønsted base‐squaramide bifunctional catalyst promotes the conjugate addition of either unsubstituted or α‐monosubstituted β‐tetralones to nitroalkenes. Under these reaction conditions, not only enolization, and thus functionalization, occurs at the α‐carbon atom of the β‐tetralone exclusively, but adducts including all‐carbon quaternary centers are also formed in highly diastereo‐ and enantioselective manner.  相似文献   

16.
A series of 3,7‐diaryl‐6,7‐dihydro‐5H‐6‐substituted‐thiazolo[3,2‐a]pyrimidin‐5‐ones ( 3a , 3b , 3c , 3d , 3e , 3f , 3g , 3h , 3i , 3j ) were synthesized by the [4+2] cycloaddition reaction of 2‐arylideneamino‐4‐arylthiazoles (2a–j) with in situ generated monosubstituted ketenes under PTC conditions coupled with ultrasonication in good to excellent yields. All the synthesized compounds were characterized on the basis of their spectral (IR, PMR, and Mass) and analytical data. The synthesized compounds were also screened for their insecticidal activity against Helicoverpa armigera and some of them showed promising activity.  相似文献   

17.
The molecules of racemic 3‐benzoylmethyl‐3‐hydroxyindolin‐2‐one, C16H13NO3, (I), are linked by a combination of N—H...O and O—H...O hydrogen bonds into a chain of centrosymmetric edge‐fused R22(10) and R44(12) rings. Five monosubstituted analogues of (I), namely racemic 3‐hydroxy‐3‐[(4‐methylbenzoyl)methyl]indolin‐2‐one, C17H15NO3, (II), racemic 3‐[(4‐fluorobenzoyl)methyl]‐3‐hydroxyindolin‐2‐one, C16H12FNO3, (III), racemic 3‐[(4‐chlorobenzoyl)methyl]‐3‐hydroxyindolin‐2‐one, C16H12ClNO3, (IV), racemic 3‐[(4‐bromobenzoyl)methyl]‐3‐hydroxyindolin‐2‐one, C16H12BrNO3, (V), and racemic 3‐hydroxy‐3‐[(4‐nitrobenzoyl)methyl]indolin‐2‐one, C16H12N2O5, (VI), are isomorphous in space group P. In each of compounds (II)–(VI), a combination of N—H...O and O—H...O hydrogen bonds generates a chain of centrosymmetric edge‐fused R22(8) and R22(10) rings, and these chains are linked into sheets by an aromatic π–π stacking interaction. No two of the structures of (II)–(VI) exhibit the same combination of weak hydrogen bonds of C—H...O and C—H...π(arene) types. The molecules of racemic 3‐hydroxy‐3‐(2‐thienylcarbonylmethyl)indolin‐2‐one, C14H11NO3S, (VII), form hydrogen‐bonded chains very similar to those in (II)–(VI), but here the sheet formation depends upon a weak π–π stacking interaction between thienyl rings. Comparisons are drawn between the crystal structures of compounds (I)–(VII) and those of some recently reported analogues having no aromatic group in the side chain.  相似文献   

18.
During studies of aziridination of α,β‐unsaturated amides with diaziridine, we found that we could prepare both the cis‐ and trans‐aziridinecarboxamides by choosing an appropriately substituted diaziridine. While 3‐monosubstituted diaziridine 2 was suitable for the trans‐selective aziridination, employment of 3,3‐dialkyldiaziridine 1 resulted in the formation of cis‐aziridine carboxamides, irrespective of the geometry of the substrate (Scheme 1 and Tables 1 and 2). To elucidate the unique nonstereospecificity and to expand these aziridinations to asymmetric ones, several optically active diaziridines were newly prepared. Aziridination with an optically active 3‐monosubstituted diaziridine, 3‐cyclohexyl‐1‐[(1R)‐1‐phenylethyl]diaziridine 16 , proceeded smoothly with high trans‐selectivity as well as excellent enantioselectivity (up to 98% ee; see Table 3). On the other hand, highly enantioselective cis‐aziridination was achieved (>99% ee) with optically active 3,3‐dimethyl‐1‐[(1R)‐1‐phenylethyl]diaziridine 15 , though the yield was low (4%). This aziridination was considered to proceed stepwise by way of the enolate intermediate (Scheme 2). Careful inspection of the stereochemistry and its solvent‐dependence suggested that the diastereoselection of the reaction was kinetically controlled: the 1,4‐addition of N‐lithiated diaziridine was a crucial step for determination of the stereochemical course of the aziridination (Figs. 24).  相似文献   

19.
N-Methyl-2-methyl-3-(benzotriazol-l-yl)propanamide, on treatment with butyllithium forms a dianion which on treatment with alkyl and benzyl halides, aldehydes and ketones affords monosubstituted products; with ethyl p-toluate, a lactam is formed. The alkylated derivatives eliminate benzotriazole in the presence of base to afford trisubstituted α,β-unsaturated amides.  相似文献   

20.
Due to the mesomeric interaction of the nitrogen lone pair with the As=C double bond, the perfluoroarsapropene derivative F3CAs=C(F)NEt2 ( 1 ) is sufficiently stable to serve as a ligand in transition metal carbonyl complexes. 1 was coordinated to chromium by reaction with the photochemically generated labile complex Cr(CO)5(THF), yielding the monosubstituted pentacarbonyl derivative Cr(CO)5[F3CAs=C(F)NEt2] ( 2 ). Already at room temperature, this is slowly transformed into the binuclear complex [F3CAs=C(F)NEt2][Cr(CO)5]2 ( 3 ) by replacing 1 from a neighbouring molecule by the stronger donor 2 . In a closed system 3 obviously exists in an equilibrium with 1 and 2 . Both complexes are related to the previously studied derivatives of the 2‐dimethylamino‐perfluoro‐1‐phosphapropene ligand. The products were identified by spectroscopic (IR, NMR) investigations and comparison with the related phosphaalkene complexes.  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