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.
The 1,5,6,8,10‐pentamethylheptalene‐4‐carboxaldehyde ( 4b ) (together with its double‐bond‐shifted (DBS) isomer 4a ) and methyl 4‐formyl‐1,6,8,10‐tetramethylheptalene‐5‐carboxylate ( 15b ) were synthesized (Schemes 3 and 7, resp.). Aminoethenylation of 4a / 4b with N,N‐dimethylformamide dimethyl acetal (=1,1‐dimethoxy‐N,N‐dimethylmethanamine=DMFDMA) led in DMF to 1‐[(1E)‐2‐(dimethylamino)ethenyl]‐5,6,8,10‐tetramethylheptalene‐2‐carboxaldehyde ( 18a ; Scheme 9), whereas the stronger aminoethenylation agent N,N,N′,N′,N″,N″‐hexamethylmethanetriamine (=tris(dimethylamino)methane=TDMAM) gave an almost 1 : 1 mixture of 18a and 1‐[(1E)‐2‐(dimethylamino)ethenyl]‐5,6,8,10‐tetramethylheptalene‐4‐carboxaldehyde ( 20b ; Scheme 11). Carboxylate 15b delivered with DMFDMA on heating in DMF the expected aminoethenylation product 19b (Scheme 10). The aminoethenylated heptalenecarboxaldehydes were treated with malononitrile in CH2Cl2 in the presence of TiCl4/pyridine to yield the corresponding malononitrile derivatives 23b, 24b , and 26a (Schemes 13 and 14). The photochemically induced DBS process of the heptalenecarboxaldehydes as ‘soft’ merocyanines and their malononitrile derivatives as ‘strong’ merocyanines of almost zwitterionic nature were studied in detail (Figs. 1029) with the result that 1,4‐donor/acceptor substituted heptalenes are cleaner switchable than 1,2‐donor/acceptor‐substituted heptalenes.  相似文献   

2.
Heterocycles Starting from Bis(alkoxycarbonyl)ketene Ethylene Acetals ( = Dialkyl 2-(1,3-Dioxolan-2-ylidene)propane-1,3-dioate). Synthesis and Properties of a New Class of Pyrazolium Betaines The readily available bis(alkoxycarbonyl)ketene ethylene acetals 1 react with bifunctional nucleophiles to give heterocycles 2–5 (Scheme 1). Their reactions with N,N-dialkylhydrazines lead to the pyrazolium betaines 7a–f (Scheme 4). Cyclic N,N-dialkylhydrazines give spiro compounds 7d–f . The reaction of thioketene acetal 12 and of the derivative 15 of methanetricarboxylic acid with N,N-dimethylhydrazine results in the formation of 3-(methylthio)- and 3-methoxypyrazolium betaine 7g and 7h , respectively (Scheme 4). The chemical reactivity of the synthesized pyrazolium betaines 7 was tested. The structure of the 3-(methylthio) derivative 7g was determined by X-ray analysis.  相似文献   

3.
We report the synthesis of N‐benzyl‐N‐[(E)‐buta‐1,3‐dienyl]propanamide ( 6 ) and its corresponding O‐silyl‐substituted ketene N,O‐acetal 7 and their Diels‐Alder reaction. Propanamide 6 reacted smoothly, whereas the yield obtained from 7 was low, probably due to polymerization of the dienophile induced by electron transfer. The ketene N,O‐acetals 27a – g were synthesized starting from the corresponding benzamides 25a – e (Scheme 9). The ketene N,O‐acetals 27a – g showed increased stabilities and underwent amino‐Claisen rearrangements under thermal conditions. Using catalysts, interesting side reactions leading either to the annulated systems rac‐ 35 – 37 or to a β‐lactam rac‐ 34 were observed.  相似文献   

4.
The reaction of N,N′‐diarylselenoureas 16 with phenacyl bromide in EtOH under reflux, followed by treatment with NH3, gave N,3‐diaryl‐4‐phenyl‐1,3‐selenazol‐2(3H)‐imines 13 in high yields (Scheme 2). A reaction mechanism via formation of the corresponding Se‐(benzoylmethyl)isoselenoureas 18 and subsequent cyclocondensation is proposed (Scheme 3). The N,N′‐diarylselenoureas 16 were conveniently prepared by the reaction of aryl isoselenocyanates 15 with 4‐substituted anilines. The structures of 13a and 13c were established by X‐ray crystallography.  相似文献   

5.
A brief overview is presented of the field of organocatalysis using chiral H‐bond donors, chiral Brønsted acids, and chiral counter‐anions (Fig. 1). The role of TADDOLs (=α,α,α′,α′‐tetraaryl‐1,3‐dioxolane‐4,5‐dimethanols) as H‐bond donors and the importance of an intramolecular H‐bond for acidity enhancement are discussed. Crystal structures of TADDOLs and of their N‐, S‐, and P‐analogs (Figs. 2 and 3) point the way to proposals of mechanistic models for the action of TADDOLs as organocatalysts (Scheme 1). Simple experimental two‐step procedures for the preparation of the hitherto strongest known TADDOL‐derived acids, the bicyclic phosphoric acids ( 2 in Scheme 2) and of a phosphoric‐trifluorosulfonic imide ( 9 in Scheme 4), are disclosed. The mechanism of sulfinamide formation in reactions of TADDAMIN with trifluoro‐sulfonylating reagents is discussed (Scheme 3). pKa Measurements of TADDOLs and analogs in DMSO (reported in the literature; Fig. 5) and in MeO(CH2)2OH/H2O (described herein; Fig. 6) provide information about further possible applications of this type of compounds as strong chiral Brønsted acids in organocatalysis.  相似文献   

6.
The Ru—N bond distances in the title complex, [Ru(NO2)(C11H9N3)(C15H11N3)]BF4 or [Ru(NO2)(tpy)(azpy)]BF4, [tpy is 2,2′:6′,2′′‐ter­pyridine and azpy is 2‐(phenyl­azo)­pyridine], are Ru—Npy 2.063 (4), Ru—Nazo 2.036 (4), Ru—Nnitro 2.066 (3) Å, and Ru—Ntpy 2.082 (4), 1.982 (3) and 2.074 (4) Å. The azo N atom is trans to the nitro group. The azo N=N bond length is 1.265 (5) Å, which is the shortest found in such complexes to date. This indicates a multiple bond between Ru and the N atom of the nitro group, and π‐­backbonding [dπ(Ru) π*(azo)] is decreased.  相似文献   

7.
The photochemical reactions of 2‐substituted N‐(2‐halogenoalkanoyl) derivatives 1 of anilines and 5 of cyclic amines are described. Under irradiation, 2‐bromo‐2‐methylpropananilides 1a – e undergo exclusively dehydrobromination to give N‐aryl‐2‐methylprop‐2‐enamides (=methacrylanilides) 3a – e (Scheme 1 and Table 1). On irradiation of N‐alkyl‐ and N‐phenyl‐substituted 2‐bromo‐2‐methylpropananilides 1f – m , cyclization products, i.e. 1,3‐dihydro‐2H‐indol‐2‐ones (=oxindoles) 2f – m and 3,4‐dihydroquinolin‐2(1H)‐ones (=dihydrocarbostyrils) 4f – m , are obtained, besides 3f – m . On the other hand, irradiation of N‐methyl‐substituted 2‐chloro‐2‐phenylacetanilides 1o – q and 2‐chloroacetanilide 1r gives oxindoles 2o – r as the sole product, but in low yields (Scheme 3 and Table 2). The photocyclization of the corresponding N‐phenyl derivatives 1s – v to oxindoles 2s – v proceeds smoothly. A plausible mechanism for the formation of the photoproducts is proposed (Scheme 4). Irradiation of N‐(2‐halogenoalkanoyl) derivatives of cyclic amines 5a – c yields the cyclization products, i.e. five‐membered lactams 6a , b , and/or dehydrohalogenation products 7a , c and their cyclization products 8a , c , depending on the ring size of the amines (Scheme 5 and Table 3).  相似文献   

8.
Various substituted homotaurines (=3‐aminopropane‐1‐sulfonic acids) 6 were readily synthesized in satisfactory to good yields via the Michael addition of thioacetic acid to alk‐2‐enamides 3 (→ 4 ), followed by LiAlH4 reduction (→ 5 ) and performic acid oxidation (Scheme 1). The configuration of ‘anti’‐disubstituted homotaurine ‘anti’‐ 6h was deduced from the 3‐(acetylthio)alkanamide (=S‐(3‐amino‐1,2‐dimethyl‐3‐oxopropyl) ethanethioate)‘anti’‐ 4h formed in the Michael addition, which was identified via the Karplus equation analysis, and confirmed by X‐ray diffraction analysis. The current route is an efficient method to synthesize diverse substituted homotaurines, including 1‐, 2‐, and N‐monosubstituted, as well as 1,2‐, 1,N‐, 2,N‐, and N,N‐disubstituted homotaurines (Table).  相似文献   

9.
[(1,3-Dioxolan-2-ylidene)methyl]phosphonates and -phosphinates as [simple] Synthons in Heterocyclic Synthesis The readily available [(1,3-dioxolane-2-ylidene)methyl]phosphonates and -phosphinates 2a–f (Scheme 1) can be transformed with amines to aliphatic ketene N,O-and N,N-acetales (see Scheme 2, 2a → 3–7 ). Alkanediamines yield with 2a–f the imidazolidines 8a–f and the hexahydropyrimidines 9a–d (Scheme 3). the oxazolidine derivatives 10a–e and the thiazolidine 11 are accessible under special reaction conditions starting from 2a, b (Scheme 4). Hydrazines react with the CN-group-containing ketene O,O-acetals 2a–c to the pyrazoles 12a–g , whereof 12a, d, e can be cyclized to pyrazolo[1,5-a]pyrimidines 13a–d (Scheme 5). Amidines as starting materials transform 2a–c in an analogous way to the pyrimidine derivatives 14a–c (Scheme 6).  相似文献   

10.
D ‐Glyceraldehyde 3‐phosphate (=D ‐GAP; 2 ) was prepared by an improved chemical method (Scheme 2), and it was then employed to synthesize 1‐deoxy‐D ‐xylulose 5‐phosphate (=DXP; 3 ) which is enzymatically one of the key intermediates in the MEP ( 4 ) terpenoid biosynthetic pathway (Scheme 1). The recombinant DXP synthase of Rhodobacter capsulatus was used to catalyze the condensation of D ‐glyceraldehyde 3‐phosphate ( 2 ) and pyruvate (=2‐oxopropanoate; 1 ) to produce the sugar phosphate 3 (Scheme 2). The simple two‐step chemoenzymatic route described affords DXP ( 3 ) with more than 70% overall yield and higher than 95% purity. The procedure may also be used for the synthesis of isotope‐labeled DXP ( 3 ) by using isotope‐labeled pyruvate.  相似文献   

11.
The diastereoselectivity of the addition of NH3 and MeNH2 to glyconolactone oxime sulfonates and the structures of the resulting N‐unsubstituted and N‐methylated glycosylidene diaziridines were The 15N‐labelled glucono‐ and galactono‐1,5‐lactone oxime mesylates 1* and 9* add NH3 mostly axially (>3 : 1; Scheme 4), while the 15N‐labelled mannono‐1,5‐lactone oxime sulfonate 19* adds NH3 mostly equatorially (9 : 1; Scheme 7). The 15N‐labelled mannono‐1,4‐lactone oxime sulfonate 30* adds NH3 mostly from the exo side (>4 : 1; Scheme 9). The configuration of the N‐methylated pyranosylidene diaziridines 17, 18, 28 , and 29 suggests that MeNH2 adds to 1, 9, 19 , and 23 mostly to exclusively from the equatorial direction (>7 : 3; Schemes 5 and 8). The mannono‐1,4‐lactone oxime sulfonate 30 adds MeNH2 mostly from the exo side (85 : 15; Scheme 10), while the ribo analogue 37 adds MeNH2 mostly from the endo side (4 : 1; Scheme 10). Analysis of the preferred and of the reactive conformers of the tetrahedral intermediates suggests that the addition of the amine to lactone oxime sulfonates is kinetically controlled. The diastereoselectivity of the diaziridine formation is rationalized as the result of the competing influences of intramolecular H‐bonding during addition of the amines, steric interactions (addition of MeNH2), and the kinetic anomeric effect. The diaziridines obtained from 2,3,5‐tri‐O‐benzyl‐D ‐ribono‐ and ‐D ‐arabinono‐1,4‐lactone oxime methanesulfonate ( 42 and 48 ; Scheme 11) decomposed readily to mixtures of 1,4‐dihydro‐1,2,4,5‐tetrazines, pentono‐1,4‐lactones, and pentonamides. The N‐unsubstituted gluco‐ and galactopyranosylidene diaziridines 2, 4, 6, 8 , and 10 are mixtures of two trans‐substituted isomers ( S / R ca. 19 : 1, Scheme 2). The main, (S,S)‐configured isomers S are stabilised by a weak intramolecular H‐bond from the pseudoaxial NH to RO? C(2). The diaziridines 12 , derived from GlcNAc, cannot form such a H‐bond; the (R,R)‐isomer dominates ( R / S 85 : 15; Scheme 3). The 2,3‐di‐O‐benzyl‐D ‐mannopyranosylidene diaziridines 20 and 22 adopt a 4C1 conformation, which does not allow an intramolecular H‐bond; they are nearly 1 : 1 mixtures of R and S diastereoisomers, whereas the OH5 conformation of the 2,3:5,6‐di‐O‐isopropylidene‐D ‐mannopyranosylidene diaziridines 24 is compatible with a weak H‐bond from the equatorial NH to O? C(2); the (R,R)‐isomer is favoured ( R / S ≥7 : 3; Scheme 6). The mannofuranosylidene diaziridine 31 completely prefers the (R,R)‐configuration (Scheme 9).  相似文献   

12.
A detailed structural analysis has been performed for N,N′‐bis(4‐chlorophenyl)‐7,8,11,12‐tetraoxaspiro[5.6]dodecane‐9,10‐diamine, C20H22Cl2N2O4, (I), N,N′‐bis(2‐fluorophenyl)‐7,8,11,12‐tetraoxaspiro[5.6]dodecane‐9,10‐diamine, C20H22F2N2O4, (II), and N,N′‐bis(4‐fluorophenyl)‐7,8,11,12‐tetraoxaspiro[5.6]dodecane‐9,10‐diamine, C20H22F2N2O4, (III). The seven‐membered ring with two peroxide groups adopts a twist‐chair conformation in all three compounds. The lengths of the C—N and O—O bonds are slightly shorter than the average statistical values found in the literature for azepanes and 1,2,4,5‐tetraoxepanes. The geometry analysis of compounds (I)–(III), the topological analysis of the electron density at the (3, ?1) bond critical points within Bader's quantum theory of `Atoms in molecules' (QTAIM) and NBO (natural bond orbital) analysis at the B3LYP/6‐31G(d,2p) level of theory showed that there are nO→σ*(C—O), nN→σ*(C—O) and nO→σ*(C—N) stereoelectronic effects. The molecules of compounds (I) and (III) are packed in the crystals as zigzag chains due to strong N—H…O and C—H…O hydrogen‐bond interactions, whereas the molecules of compound (II) form chains in the crystals bound by N—H…O, C—H…π and C—H…O contacts. All these data show that halogen atoms and their positions have a minimal effect on the geometric parameters, stereoelectronic effects and crystal packing of compounds (I)–(III), so that the twist‐chair conformation of the tetraoxepane ring remains unchanged.  相似文献   

13.
An effective route to functionalized 2H‐chromene (=2H‐1‐benzopyran) derivatives 4 is described (Scheme 1). This involves the reaction of a 1,1‐diactivated alkene, resulting from the reaction of dimedone (=5,5‐dimethylcyclohexane‐1,3‐dione; 1a ) with methyl chloroglyoxylate (ClC(O)COOMe), benzyl carbonochloridate (ClC(O)OCH2Ph) or 3,5‐dinitrobenzoyl chloride (3,5‐(NO2)2C6H3C(O)Cl), and a dialkyl acetylenedicarboxylate (=dialkyl but‐2‐ynedioate) in the presence of Ph3P which undergo intramolecular Wittig reaction to produce 2H‐chromene derivatives (Scheme 1).  相似文献   

14.
The protected poly‐Aib oligopeptides Z‐(Aib)n‐N(Me)Ph with n=2–6 were prepared according to the ‘azirine/oxazolone method’, i.e., by coupling amino or peptide acids with 2,2,N‐trimethyl‐N‐phenyl‐2H‐azirin‐3‐amine ( 1a ) as an Aib synthon (Scheme 2). Following the same concept, the segments Z‐(Aib)3‐OH ( 9 ) and H‐L ‐Pro‐(Aib)3‐N(Me)Ph ( 20 ) were synthesized, and their subsequent coupling with N,N′‐dicyclohexylcarbodiimide (DCC)/ZnCl2 led to the protected heptapeptide Z‐(Aib)3‐L ‐Pro‐(Aib)3‐N(Me)Ph ( 21 ; Scheme 3). The crystal structures of the poly‐Aib oligopeptide amides were established by X‐ray crystallography confirming the 310‐helical conformation of Aib peptides.  相似文献   

15.
The 4‐exo cyclizations of two types of carbamoyl radicals onto O‐alkyloxime acceptor groups were studied as potential routes to 3‐amino‐substituted azetidinones and hence to penicillins. A general synthetic route to ‘benzaldehyde oxime oxalate amides’ (= 2‐[(benzylideneamino)oxy]‐2‐oxoacetamides; see, e.g., 10c ) of 2‐{[(benzyloxy)imino]methyl}‐substituted thiazolidine‐4‐carboxylic acid methyl esters 9 was developed (Scheme 3). It was shown by EPR spectroscopy that these compounds underwent sensitized photodissociation to the corresponding carbamoyl radicals but that these did not ring close. An analogous open‐chain precursor, benzaldehyde O‐(benzylaminoacetaldehyde‐O‐benzyloxalyl)oxime, 15 , lacking the 5‐membered thiazolidine ring, was shown by EPR spectroscopy to release the corresponding carbamoyl radical (Scheme 4). The latter underwent 4‐exo cyclization onto its C?NOBn bond in non‐H‐atom donor solvents. The rate constant for this cyclization was determined by the steady‐state EPR method. Spectroscopic evidence indicated that the reverse ring‐opening process was slower than cyclization.  相似文献   

16.
An effective route to novel 4‐(alkylamino)‐1‐(arylsulfonyl)‐3‐benzoyl‐1,5‐dihydro‐5‐hydroxy‐5‐phenyl‐2H‐pyrrol‐2‐ones 10 is described (Scheme 2). This involves the reaction of an enamine, derived from the addition of a primary amine 5 to 1,4‐diphenylbut‐2‐yne‐1,4‐dione, with an arenesulfonyl isocyanate 7 . Some of these pyrrolones 10 exhibit a dynamic NMR behavior in solution because of restricted rotation around the C? N bond resulting from conjugation of the side‐chain N‐atom with the adjacent α,β‐unsaturated ketone group, and two rotamers are in equilibrium with each other in solution ( 10 ? 11 ; Scheme 3). The structures of the highly functionalized compounds 10 were corroborated spectroscopically (IR, 1H‐ and 13C‐NMR, and EI‐MS), by elemental analyses, and, in the case of 10a , by X‐ray crystallography. A plausible mechanism for the reaction is proposed (Scheme 4).  相似文献   

17.
With the aim to develop a general approach to a total synthesis of aminoacylated t‐RNAs and analogues, we describe the synthesis of stabilized, aminoacylated RNA fragments, which, upon ligation, could lead to aminoacylated t‐RNA structures. Novel RNA phosphoramidites with fluoride‐labile 2′‐O‐[(triisopropylsilyl)oxy]methyl (=tom) sugar‐protecting and N‐{{2‐[(triisopropylsilyl)oxy]benzyl}oxy}carbonyl (=tboc) base‐protecting groups were prepared (Schemes 4 and 5), as well as a solid support containing an immobilized N6‐tboc‐protected adenosine with an orthogonal (photolabile) 2′‐O‐[(S)‐1‐(2‐nitrophenyl)ethoxy]methyl (=(S)‐npeom) group (Scheme 6). From these building blocks, a hexameric oligoribonucleotide was prepared by automated synthesis under standard conditions (Scheme 7). After the detachment from the solid support, the resulting fully protected sequence 34 was aminoacylated with L ‐phenylalanine derivatives carrying photolabile N‐protecting groups (→ 42 and 43 ; Scheme 9). Upon removal of the fluoride‐labile sugar‐ and nucleobase‐protecting groups, the still stabilized, partially with the photolabile group protected precursors 44 and 45 , respectively, of an aminoacylated RNA sequence were obtained (Scheme 9 and Fig. 3). Photolysis of 45 under mild conditions resulted in the efficient formation of the 3′(2′)‐O‐aminoacylated RNA sequence 46 (Fig. 4). Additionally, we carried out model investigations concerning the stability of ester bonds of aminoacylated ribonucleotide derivatives under acidic conditions (Table) and established conditions for the purification and handling of 3′(2′)‐O‐aminoacylated RNA sequences and their stabilized precursors.  相似文献   

18.
The reaction of the intermediate ketene N,Se‐hemiacetal 3 , prepared from cyanomethylene derivatives 1 by treatment with Et3N and aryl isoselenocyanates 2 , with bis‐electrophiles 6, 7, 9 , and 11 in DMF affords tetrahydro‐1H‐1,3‐selenazine (=1,3‐selenazinane) derivatives 8, 10 , and 12 in good yield (Scheme 2 and Tables 1–3). Chemical and spectroscopic evidence for the structures of the new compounds are described. The structures of 8d and 12e are established by X‐ray crystallography (Figs. 1 and 2).  相似文献   

19.
The reaction of N‐alkylbenzaldimines with carbon monoxide (CO) in the presence of cobalt (Co) catalysts resulted in the formation of N‐alkylphthalimidines (Table 1). Their formation is proposed to occur by C? H activation of the aryl ring, migratory insertion of the hydride species into the benzaldimine functionality, CO coordination, and insertion into the Co? C bond, followed by reductive elimination of the N‐alkylphthalimidine and regeneration of the starting Co species (Scheme 4). Deuterium (2H)‐labeling NMR studies are consistent with this mechanism (Scheme 5).  相似文献   

20.
TADDOL (=α,α,α′,α′‐Tetraaryl‐1,3‐dioxolane‐4,5‐dimethanol) and the corresponding dichloride are converted to TADDAMINs (=(4S,5S)‐2,2,N,N′‐tetramethyl‐α,α,α′,α′‐tetraphenyl‐1,3‐dioxolan‐4,5‐dimethanamines) (Scheme 2) and ureas, 12 – 15 , and to TADDOP derivatives with seven‐membered O? P? O ester rings (Schemes 3 and 4). Cl/P‐Replacement via the Michaelis? Arbuzov reaction (Scheme 7) on mono‐ and dichlorides, derived from TADDOL, are described. It was not possible to obtain phosphines with the P‐atom attached to the benzhydrylic C‐atom of the TADDOL skeleton (Schemes 6 and 7). The X‐ray crystal structures (Figs. 1 and 2) of ten of the more than 30 new TADDOL derivatives are discussed. Full experimental details are presented.  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