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 22 毫秒
1.
A new method for the smooth and highly efficient preparation of functionalized macrocyclic ketones has been developed. Pyrolysis of medium‐ and large‐ring 3‐vinylcycloalkanones by dynamic gas‐phase thermo‐isomerization (DGPTI) at 600–630° yielded, under insertion of a previously attached vinyl side chain by means of a 1,3‐C shift, the corresponding γ,δ‐unsaturated cycloalkanones. The yield of the two‐carbon ring‐expanded ketones greatly depended on the relative ring strains of substrate and product (5–87%, cf. Table 5). The formation of minor amounts of one‐carbon ring‐expanded cycloalkenes (<10%) can be ascribed to a subsequent decarbonylation step. A reaction mechanism involving initial cleavage of the weakest single bond in the molecule has been established (cf. Scheme 6). Recombination within the generated diradical intermediate in terminal vinylogous position led to the observed products, while reclosure gave recovered starting material. Substituents on the vinyl moiety were transferred locospecifically into the ring‐expanded products. An isopropenyl group did not significantly affect the isomerization process, whereas substrates bearing a prop‐1‐enyl group in β‐position enabled competing intramolecular H‐abstraction reactions, leading to acyclic dienones (cf. Schemes 911). DGPTI of the 13‐membered analogue directly yielded 4‐muscenone, which, upon hydrogenation, led to the valuable musk odorant (±)‐muscone. Increasing the steric hindrance on the vinyl moiety gave rise to diminishing amounts of the desired γ,δ‐unsaturated cycloalkanones. This novel two‐carbon ring‐expansion protocol was also successfully applied to 3‐ethynylcycloalkanones, giving rise to the corresponding ring‐expanded cyclic allenes (cf. Scheme 13).  相似文献   

2.
Acetylenic acetals of type I (Scheme 1) and acetylenic ketones of type III (Scheme 1), 37 and 38 (Scheme 7) are versatile synthetic precursors for the synthesis of various heterocycles by acid-catalyzed cyclization reactions. By this way, substituted 3-halofurans of type II and IV (Scheme 1) and flavones and styrylchromones (Scheme 7) can be synthesized in good-to-excellent yields. The high degree of regioselectivity in the synthesis of the 3-halofurans (Scheme 4) is the result of the regioselective β-addition of HX (X = Cl, Br, I) to the acetylenic aldehyde and acetylenic ketone moieties. A possible mechanism is depicted in Scheme 5. Since 3-halofurans can easily be metalated and substituted, this approach constitutes a new synthesis of highly substituted furans.  相似文献   

3.
On triplet excitation (E)- 2 isomerizes to (Z)- 2 and reacts by cleavage of the C(γ), O-bond to isomeric δ-ketoester compounds ( 3 and 4 ) and 2,5-dihydrofuran compounds ( 5 and 19 , s. Scheme 1). - On singulet excitation (E)- 2 gives mainly isomers formed by cleavage of the C(γ), C(δ)-bond ( 6–14 , s. Scheme 1). However, the products 3–5 of the triplet induced cleavage of the C(γ), O-bond are obtained in small amounts, too. The conversion of (E)- 2 to an intermediate ketonium-ylide b (s. Scheme 5) is proven by the isolation of its cyclization product 13 and of the acetals 16 and 17 , the products of solvent addition to b . - Excitation (λ = 254 nm) of the enol ether (E/Z)- 6 yields the isomeric α, β-unsaturated ε-ketoesters (E/Z)- 8 and 9 , which undergo photodeconjugation to give the isomeric γ, δ-unsaturated ε-ketoesters (E/Z)- 10 . - On treatment with BF3O(C2H5)2 (E)- 2 isomerizes by cleavage of the C(δ), O-bond to the γ-ketoester (E)- 20 (s. Scheme 2). Conversion of (Z)- 2 with FeCl3 gives the isomeric furan compound 21 exclusively.  相似文献   

4.
Reactions of 3-Dimethylamino-2,2-dimethyl-2H-azirine with NH-Acidic Heterocycles; Synthesis of 4H-Imidazoles In this paper, reactions of 3-dimethylamino-2,2-dimethyl-2H-azirine ( 1 ) with heterocyclic compounds containing the structure unit CO? NH? CO? NH are described. 5,5-Diethylbarbituric acid ( 5 ) reacts with 1 in refluxing 2-propanol to give the 4H-imidazole derivative 6 (Scheme 2) in 80% yield. The structure of 6 has been established by X-ray crystallography. Under similar conditions 1 and isopropyl uracil-6-carboxylate ( 7 ) yield the 4H-imidazole 8 (Scheme 3), the structure of which is deduced from spectral data and the degradation reactions shown in Scheme 3. Hydrolysis of 8 with 3N HCl at room temperature leads to the α-ketoester derivative 9 , which in refluxing methanol gives dimethyl oxalate and 5-dimethyl-amino-2,4,4-trimethyl-4H-imidazole ( 10 ). On hydrolysis the latter is converted to the known 2,4,4-trimethyl-2-imidazolin-5-one ( 11 ) [6]. Quinazolin-2,4 (1H, 3H)-dione ( 12 ) and imidazolidinetrione (parabanic acid, 14 ) undergo with 1 a similar reaction to give the 4H-imidazoles 13 and 15 , respectively (Schemes 4 and 5). In Scheme 6 two possible mechanisms for the formation of 4H-imidazoles from 1 and heterocycles of type 16 are formulated. The zwitterionic intermediate f corresponds to b in Scheme 1. Instead of dehydration as in the case of the reaction of 1 with phthalohydrazide [3], or ring expansion as with saccharin and cyclic imides [1] [2], f , undergoes ring opening (way A or B). Decarboxylation then leads to the 4H-imidazoles 17 .  相似文献   

5.
The intramolecular addition of unsaturated alkoxycarbenes leads in high yields and diastereoselectively to fused cyclopropanes (Scheme 1). Reaction of the halodiazirines 2 , 10 , 11 , and 20 with the unsaturated phenolates 1 , 8 , and 9 yielded intermediate alkoxydiazirines, and hence the homobenzofurans 5 , 12 – 16 , 22 , and 26 (Scheme 2). The intermediate alkoxydiazirine 25 was isolated at low temperature (Scheme 3). An equilibrium between the cyclopropane derivatives 12 and 27 , and 14 and 28 was established at 120°. At 200°, 12 rearranged to the chromene 29 , by disrotarory opening of the cyclopropane ring, followed by electrocyclization. Hydrogenation of 29 gave the (all-cis)-chroman 32 (Scheme 4). The homoindole 35 was obtained in good yields, presumably by an SRN1 reaction from 34 and 10 (Scheme 5).  相似文献   

6.
The protected disaccharide 44 , a precursor for the synthesis of allosamidin, was prepared from the glycosyl acceptor 8 and the donors 26–28 , best yields being obtained with the trichloroacetimidate 28 (Scheme 6). Glycosidation of 8 or of 32 by the triacetylated, less reactive donors 38–40 gave the disaccharides 46 and 45 , respectively, in lower yields (Scheme 7). Regioselective glycosidation of the diol 35 by the donors 38–40 gave 42 , the axial, intramolecularly H-bonded OH? C(3) group reacting exclusively (Scheme 5). The glycosyl acceptor 8 was prepared from 9 by reductive opening of the dioxolane ring (Scheme 3). The donors 26–28 were prepared from the same precursor 9 via the hemiacetal 25 . To obtain 9 , the known 10 was de-N-acetylated (→ 18 ), treated with phthalic anhydride (→ 19 ), and benzylated, leading to 9 and 23 (Schemes 2 and 3). Saponification of 23 , followed by acetylation also gave 9 . Depending upon the conditions, acetylation of 19 yielded a mixture of 20 and 21 or exclusively 20 . Deacetylation of 20 led to the hydroxyphthalamide 22 . De-N-acetylation of the 3-O-benzylated β-D -glycosides 11 and 15 , which were both obtained from 10 , was very sluggish and accompanied by partial reduction of the O-allyl to an O-propyl group (Scheme 2). The β-D -glycoside 30 behaved very similarly to 11 and 15 . Reductive ring opening of 31 , derived from 29 , yielded the 3-O-acetylated acceptor 32 , while the analogous reaction of the β-D -anomer 20 was accompanied by a rapid 3-O→4-O acyl migration (→ 34 ; Scheme 4). Reductive ring opening of 21 gave the diol 35 . The triacetylated donors 38–40 were obtained from 20 by debenzylidenation, acetylation (→ 36 ), and deallylation (→ 37 ), followed by either acetylation (→ 38 ), treatment with Me3SiSEt (→ 39 ), or Cl3CCN (→ 40 ).  相似文献   

7.
At room temperature or under reflux in MeCN, 3-amino-2H-azirines 2 and 3,4-dihydro-2H-1,2-benzothiazin-3-one 1,1-dioxide ( 4 ) give 1,2,5-benzothiadiazonin-6-one 1,1-dioxides 5 in fair-to-good yield (Scheme 2). The structure of this novel type of heterocyclic compounds has been established by X-ray crystallography of 5a (Fig.). A ring expansion via a zwitterionic intermediate of type A ' is proposed as the reaction mechanism of the formation of 5 .  相似文献   

8.
Highly enantioselective Diels–Alder (DA) and inverse‐electron‐demand hetero‐Diels–Alder (HDA) reactions of β,γ‐unsaturated α‐ketoesters with cyclopentadiene catalyzed by chiral N,N′‐dioxide–Cu(OTf)2 (Tf=triflate) complexes have been developed. Quantitative conversion of β,γ‐unsaturated α‐ketoesters and excellent diastereoselectivities (up to 99:1) and enantioselectivities (up to >99 % ee) were observed for a broad range of substrates. Both aromatic and aliphatic β,γ‐unsaturated α‐ketoesters were found to be suitable substrates for the reactions. Moreover, the chemoselectivity of the DA and HDA adducts were improved by regulating the reaction temperature. Good to high chemoselectivity (up to 94 %) of the DA adducts were obtained at room temperature, and moderate chemoselectivity (up to 65 %) of the HDA adducts were achieved at low temperature. The reaction also featured mild reaction conditions, a simple procedure, and remarkably low catalyst loading (0.1–1.5 mol %). A strong positive nonlinear effect was observed.  相似文献   

9.
Boron-Trifluoride-Catalyzed Reactions of 3-Amino-2H-azirines with Amino-acid Esters and Amines After activation by protonation or complexation with BF3, 3-amino-2H-azirines 1 react with the amino group of α-amino-acid esters 3 to give 3,6-dihydro-5-aminopyrazin-2(1H)-ones 4 by ring enlargement (Scheme 2, Table 1). The configuration of 3 is retained in the products 4 . With unsymmetrically substituted 1 (R1 ≠ R2), two diastereoisomers of 4 (cis and trans) are formed in a ratio of 1:1 to 2:1. With β-amino-acid esters 5 and 7 , only openchain α-amino-imidamides 6 and 8 , respectively, are formed, but none of the seven-membered heterocycle (Scheme 3). Primary amines also react with BF3-complexed 1 to yield α-amino-imidamides of type 9 (Scheme 4, Table 2). Compound 9b is characterized chemically by its transformation into crystalline derivatives 10 and 12 with 4-nitrobenzoyl chloride and phenyl isothiocyanate, respectively (Scheme 5). The structure of 12 is established by X-ray crystallography. Mechanisms for the reaction of activated 1 with amino groups are proposed in Schemes 6 and 7.  相似文献   

10.
Two approaches for the conversion of d‐ glucose to (?) ‐neplanocin A ( 2 ), both based on the zirconocene‐promoted ring contraction of a vinyl‐substituted pyranoside, are herein evaluated (Scheme 1). In the first pathway (Scheme 2), the substrate possesses the α‐d‐ allo configuration (see 6 ) such that ultimate introduction of the nucleobase would require only an inversion of configuration. However, this precursor proved unresponsive to Cp2Zr (=[ZrCl2(Cp)2]), an end result believed to be a consequence of substantive nonbonded steric effects operating in a key intermediate (Scheme 5). In contrast, the C(2) epimer (see 7 ) experienced the desired metal‐promoted conversion to an enantiomerically pure polyfunctional cyclopentane (see 5 in Scheme 3). The substituents in this product are arrayed in a manner such that conversion to the target nucleoside can be conveniently achieved by a double‐inversion sequence (Scheme 4). Recourse to palladium(0)‐catalyzed allylic alkylation did not provide an alternate means of generating 2 .  相似文献   

11.
Synthesis of 3-(2-Carboxy-4-pyridyl)-and 3-(6-Carboxy-3-pyridyl)-DL-alanine As starting materials for potential photochemical approaches to betalaines C(R = COOH) and to muscaflavine F(R = COOH), β-(2-carboxy-4-pyridyl)- and β-(6(carboxy-3-pyridyl))-DL-alanine ( A and D with R = COOH or 4 and 11 ), respectively, were prepared (Scheme 1). The synthesis of 4 (= A, R = COOH) started with the 2-[(4-pyridyl)methyl]malonate 1 and proceeded via the N-oxide 2 , cyanation and hydrolysis (Scheme 2). Amino acid 11 was obtained from (3-pyridyl)methyl-bromide ( 6 ) via the malonate 7 by an analogous sequence of reactions (Scheme 3).  相似文献   

12.
The enantioselective tandem reaction of β,γ‐unsaturated α‐ketoesters with β‐alkynyl ketones was realized by a bimetallic catalytic system of achiral AuΙΙΙ salt and chiral N,N′‐dioxide‐MgΙΙ complex. The cycloisomerization of β‐alkynyl ketone and asymmetric intermolecular [4+2] cycloaddition with β,γ‐unsaturated α‐ketoesters subsequently occurred, providing an efficient and straightforward access to chiral multifunctional 6,6‐spiroketals in up to 97 % yield, 94 % ee and >19/1 d.r. Besides, a catalytic cycle was proposed based on the results of control experiments.  相似文献   

13.
We describe here a synthesis of the morphine partial structures 28 and 36 , and of their enantiomers, which uses 7-methoxy-benzofurancarboxylic acid as starting material. A key intermediate in this scheme is compound 15 , which is converted, via 1,2-ketone shift, into 22 . This latter is stereospecifically reduced to the alcohol 24 and converted to the amide 25 . The diastereomer of 25 is afforded by stereospecific introduction of a ethoxycarbonyl group in 15 to yield the β-ketoester 31 , followed by Curtius degradation of the acid 32 to the acylamine 34 . An efficient method for removal of the methoxy group in methoxy-dihydro-benzofurans is presented (Scheme 9), as is the functionalization of the N-atom in 27 with concurrent complex formation between the free hydroxy group and boric-acid. The aromatization of the furan ring (Scheme 10) with DDQ gave the expected benzofuran derivative 30 .  相似文献   

14.
A highly enantioselective formal conjugate allyl addition of allylboronic acids to β,γ‐unsaturated α‐ketoesters has been realized by employing a chiral NiII/N,N′‐dioxide complex as the catalyst. This transformation proceeds by an allylboration/oxy‐Cope rearrangement sequence, providing a facile and rapid route to γ‐allyl‐α‐ketoesters with moderate to good yields (65–92 %) and excellent ee values (90–99 % ee). The isolation of 1,2‐allylboration products provided insight into the mechanism of the subsequent oxy‐Cope rearrangement reaction: substrate‐induced chiral transfer and a chiral Lewis acid accelerated process. Based on the experimental investigations and DFT calculations, a rare boatlike transition‐state model is proposed as the origin of high chirality transfer during the oxy‐Cope rearrangement.  相似文献   

15.
The first diastereo‐ and enantioselective inverse electron demand hetero‐Diels–Alder reaction of β,γ‐unsaturated α‐ketoesters with allylsilanes is described. Chiral copper(II) catalysts successfully activate the β,γ‐unsaturated α‐ketoesters and promote the reaction with allylsilanes with excellent enantioselectivities. This process represents a new entry to chiral oxanes.  相似文献   

16.
α-Chamigren-3-one (+) -8 bearing an axial CI-atom at C(8) exists as a largely dominant conformer with Me—C(5) at the envelope-shaped enone ring pointing away from CIax?C(8) at the cyclohexane ring (= B) in the ‘normal’ chair conformation, as shown by 1H-NMR. In contrast, the α-chamigren-3-ols (+) -9 and (+) -10 , obtained from hydride reduction of (+) -8 , show a temperature-dependent equilibrium of conformers where the major conformers have ring B in the inverted-chair (and twist-boat for (+) -9 ) conformation to avoid repulsions between Me?C(5) and CIax–C(8) (Scheme 1). This is in agreement with the conformation of the epoxidation product (+) -12 of (+) -9 where Me–C(5) is pushed away from CIax–C(8) in a ring-B chair similar to that of (+) -8 (Scheme 2). Introduction of a pseudoequatorial Br-atom at C(2) of (+) -8 , as in enone (+) -15 (Scheme 3), does not affect the conformation; but a pseudoaxial Br? C(2) experiences repulsive interactions with Heq–C(7), as shown by the 1H-NMR data of the isomeric enone (+) -16 where the ‘normal’-chair conformer Cβ -16 is in an equilibrium with the inverted chair conformer ICβ -16 (Scheme 3). These results and the accompanying paper allow a unifying view on the conformational behavior of marine polyhalogenated α-chamigrenes. This view is supported by the acid-induced isomerization of α-chamigrene (+) -9 (inverted chair) to β-chamigrene (+) -17 (‘normal’ chair; Scheme 4), the driving force being the lesser space requirement of CH2?C(5) than of Me–C(5). This explains why β-chamigrenes are so common in nature.  相似文献   

17.
Oxa‐bridged oxazocines bearing three chiral carbon centers were synthesized efficiently through a bimetallic catalytic asymmetric tandem reaction of β,γ‐unsaturated α‐ketoesters with diazoimides. The process contained a rhodium‐promoted in situ generation of isomünchnone from diazoimide decomposition, and a [4+3]‐cycloaddition of β,γ‐unsaturated α‐ketoester catalyzed by a chiral N,N′‐dioxide‐ZnII complex. Ligand‐accelerated catalysis was found, and a possible transition‐state model was proposed to explain the origin of stereoselectivity.  相似文献   

18.
The 1,3‐dipolar cycloaddition of 2,2,4,4‐tetramethyl‐3‐thioxocyclobutanone S‐methylide ( 2a ), generated in situ by thermal extrusion of N2 from the corresponding 2,5‐dihydro‐1,3,4‐thiadiazole 1a , with electron‐deficient acetylenic compounds yields spirocyclic 2,5‐dihydrothiophene derivatives of type 4 (Scheme 2). Mixtures of diastereoisomers are obtained in the case of propiolates. The strained cyclooctyne also undergoes smooth cycloadditions with thioketone S‐methylides (Scheme 3). Under acidic conditions, the spirocyclic products of type 4 and 6a isomerize, via opening of the cyclobutanone ring and aromatization of the five‐membered ring, to thiophene derivatives of type 7 (Scheme 4).  相似文献   

19.
The Zip-reaction: A New Method for the Synthesis of Macrocyclic Polyaminolactams The 21- and 25-membered aminolactams 11 and 25 were synthesized from the 13-membered lactam 4 . To introduce the ring enlargement unit (a propylamino group) 4 was N-alkylated using acrylonitrile and the resulting product hydrogenated. Repetition of this reaction sequence gave 3 , which was converted in the presence of base in 90% yield to the ring-enlarged macrocyclic base 11 (Scheme 2). In a similar but stepwise synthesis consisting of two separate ring-enlargement reactions 4 was transformed to 11 via 13 (Scheme 4). Introducing three ringenlargement units into 4 the 25-membered aminolactam 25 was synthesized in 84% yield (Scheme 5). The mechanism of the ring-enlargement reaction is given in Scheme 3. In comparison to a zip-fastener or zipper this reaction is called “zipreaction”.  相似文献   

20.
The syntheses of 1, 2-annelated 1, 4-benzodiazepines (IV, Y = N) and 4, 1-benzoxazepines (IV, Y = 0) are described (Scheme 1). The key step is a nucleophilic aromatic substitution of 2-substituted piperazines (II, Z = N? CH3), piperidines (II, Z = CH2) or pyrrolidines (II, Z= (CH2)0) with activated aryl halides (I).  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