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 78 毫秒
1.
1,3,4,6,8-Pentamethylazulene ( 9 ), when heated at 100° in supercritical CO2 at 150 bar in the presence of 4 equiv. of dimethyl acetylenedicarboxylate (ADM), led to the formation of 16% of a 1:1 mixture of dimethyl 3,5,6,8,10-pentamethylheptalene-1,2-dicarboxylate 12a ) and its double-bond-shifted isomer 12b as well as 4% of the corresponding azulene-1,2-dicarboxylate 13 (Scheme 4). The formation of the [1 + 2] adduct 11 (cf. Scheme 2) was not observed. Similarly, benz[a]azulene ( 25 ) yielded in supercritical CO2 (150°/170 bar) in the presence of 4 equiv. of ADM dimethyl benzo[d]heptalene-6,7-dicarboxylate ( 29 ; 30%) and dimethyl benzo[a]cyclopent[cd]azulene-1,2-dicarboxylate ( 28 ; 22%; Scheme 5). The reaction of 5,9-diphenylbenz[a]azulene ( 26 ) and ADM in supercritical CO2 (100°/150 bar) gave the corresponding benzo[d]heptalene-6,7-dicarboxylate 31 (22%) and dimethyl 5,9-diphenyl-4b,10-etheno-10H-benz[a]azulene-11,12-dicarboxylate( 30 ; 25%; Scheme 5).  相似文献   

2.
It is shown that heptaleno[1,2‐c]furans 1 , which are available in two steps from heptalene‐4,5‐dicarboxylates by reduction and oxidative dehydrogenation of the corresponding vicinal dimethanols 2 with MnO2 or IBX (Scheme 4), react thermally in a Diels–Alder‐type [4+2] cycloaddition at the furan ring with a number of electron‐deficient dipolarophiles to yield the corresponding 1,4‐epoxybenzo[d]heptalenes (cf. Schemes 6, 15, 17, and 19). The thermal reaction between dimethyl acetylenedicarboxylate (ADM) and 1 leads, kinetically controlled, via a sterically less‐congested transition state (Fig. 4) to the formation of the (M*)‐configured 1,4‐dihydro‐1,4‐epoxybenzo[a]heptalenes, which undergo a cyclic double‐bond shift to the energetically more‐relaxed benzo[d]heptalenes 4 (Schemes 6 and 7). Most of the latter ones exhibit under thermal conditions epimerization at the axis of chirality, so that the (M*)‐ and (P*)‐stereoisomers are found in reaction mixtures. The (P*)‐configured forms of 4 are favored in thermal equilibration experiments, in agreement with AM1 calculations (Table 1). The relative (P*,1S*,4R*)‐ and (M*,1S*,4R*)‐configuration of the crystalline main stereoisomers of the benzo[d]heptalene‐2,3‐dicarboxylates 4a and 4f , respectively, was unequivocally established by an X‐ray crystal‐structure determination (Figs. 1 and 2). Acid‐induced rearrangement of 4 led to the formation of the corresponding 4‐hydroxybenzo[a]heptalene‐2,3‐dicarboxylates 5 in moderate‐to‐good yields (Schemes 8, 13, and 14). When the aromatization reaction is performed in the presence of trifluoroacetic acid (TFA), trifluoroacetates of type 6 and 13 (Schemes 8, 12, and 13) are also formed via deprotonation of the intermediate tropylium ions of type 7 (Scheme 11). Thermal reaction of 1 with dimethyl maleate gave the 2,3‐exo‐ and 2,3‐endo‐configured dicarboxylates 14 as mixtures of their (P*)‐ and (M*)‐epimers (Scheme 15). Treatment of these forms with lithium di(isopropyl)amide (LDA) at ?70° gave the expected benzo[a]heptalene‐2,3‐dicarboxylates 15 in good yields (Scheme 16). Fumaronitrile reacted thermally also with 1 to the corresponding 2‐exo,3‐endo‐ and 2‐endo,3‐exo‐configured adducts 17 , again as mixtures of their (P*)‐ and (M*)‐epimers (Scheme 17), which smoothly rearranged on heating in dimethoxyethane (DME) in the presence of Cs2CO3 to the benzo[a]heptalene‐2,3‐dicarbonitriles 18 (Scheme 18). Some cursory experiments demonstrated that hex‐3‐yne‐2,5‐dione and (E)/(Z)‐hexa‐3‐ene‐2,5‐dione undergo also the Diels–Alder‐type cycloaddition reaction with 1 (Scheme 19). The mixtures of the stereoisomers of the 2,3‐diacetyl‐1,4‐epoxytetrahydrobenzo[d]heptalenes 22 gave, on treatment with Cs2CO3 in DME at 80°, only mixtures of the regioisomeric inner aldol products 24 and 25 of the intermediately formed benzo[a]heptalenes 23 (Scheme 20).  相似文献   

3.
The synthesis of 5,9-diphenylbenz[a]azulene ( 1 ) from 1,3-diphenylcyclopent[a]indene-2,8-dione ( 4 ) and cyclopropene has been re-investigated. The reduction of the decarbonylated cycloadduct 5 with LiAlH4/AlCl3 in Et2O leads not only to the expected 7,10-dihydrobenz[a]azulene 6 , but also to small amounts of the cyclopropa[b]fluorenes exo- 7 and endo- 7 (cf. Scheme 2), the structures of which have been determined by X-ray crystal-structure analysis (cf. Fig. 1). The reaction of 1 with dialkyl acetylenedicarboxylates (ADR) in MeCN at 100° in the presence of 2 mol-% of catalysts such as [RuH2(PPh3)4] results mainly in the formation of the expected 8,12-diphenylbenzo[a]heptalene-6,7-dicarboxylates 3 . A thorough investigation of the reaction mixture of 1 and dimethyl acetylenedicarboxylate (ADM) revealed the presence of a number of intermediates and side products (Scheme 5). Most important was the isolation and identification of the cyclobutene intermediate 9a (cf. Fig. 4), which is formed by a zwitterionic rearrangement of the primary adduct 2a of 1 and ADM and represents the direct precursor of the heptalene-diester 3a . Compounds of type 9a have so far only been postulated as necessary intermediates in the thermal reaction of azulenes and ADR to give corresponding heptalenedicarboxylates. Compound 9a is photochemically unstable and undergoes rearrangement even under the influence of normal laboratory light into a mixture of trans- 10a and cis- 10a (Scheme 8). Both diastereoisomers are also found in the original reaction mixture of 1 and ADM, but not when the reaction is performed under exclusion of light. On heating in MeCN at 100°, or better in DMF at 150°, trans- 10a and cis- 10a undergo rearrangement to the fluoranthene-1,2-dicarboxylate 11a (Scheme 9), which is also present in the original reaction mixture of 1 and ADM. The catalysts do not accelerate the reaction of 1 and ADR, but they lead to better yields of the benzo[a]heptalene-6,7-dicarboxylates 3 , especially in the reaction of 1 with diisopropyl acetylenedicarboxylate (ADiP) (cf. Tables 1 and 2).  相似文献   

4.
The dehydrogenation reaction of the heptalene-4,5-dimethanols 4a and 4d , which do not undergo the double-bond-shift (DBS) process at ambient temperature, with basic MnO2 in CH2Cl2 at room temperature, leads to the formation of the corresponding heptaleno[1,2-c]furans 6a and 6d , respectively, as well as to the corresponding heptaleno[1,2-c]furan-3-ones 7a and 7d , respectively (cf. Scheme 2 and 8). The formation of both product types necessarily involves a DBS process (cf. Scheme 7). The dehydrogenation reaction of the DBS isomer of 4a , i.e., 5a , with MnO2 in CH2Cl2 at room temperature results, in addition to 6a and 7a , in the formation of the heptaleno[1,2-c]-furan-1-one 8a and, in small amounts, of the heptalene-4,5-dicarbaldehyde 9a (cf. Scheme 3). The benzo[a]heptalene-6,7-dimethanol 4c with a fixed position of the C?C bonds of the heptalene skeleton, on dehydrogenation with MnO2 in CH2Cl2, gives only the corresponding furanone 11b (Scheme 4). By [2H2]-labelling of the methanol function at C(7), it could be shown that the furanone formation takes place at the stage of the corresponding lactol [3-2H2]- 15b (cf. Scheme 6). Heptalene-1,2-dimethanols 4c and 4e , which are, at room temperature, in thermal equilibrium with their corresponding DBS forms 5c and 5e , respectively, are dehydrogenated by MnO2 in CH2Cl2 to give the corresponding heptaleno[1,2-c]furans 6c and 6e as well as the heptaleno[1,2-c]furan-3-ones 7c and 7e and, again, in small amounts, the heptaleno[1,2-c]furan-1-ones 8c and 8e , respectively (cf. Scheme 8). Therefore, it seems that the heptalene-1,2-dimethanols are responsible for the formation of the furan-1-ones (cf. Scheme 7). The methylenation of the furan-3-ones 7a and 7e with Tebbe's reagent leads to the formation of the 3-methyl-substituted heptaleno[1,2-c]furans 23a and 23e , respectively (cf. Scheme 9). The heptaleno[1,2-c]furans 6a, 6d , and 23a can be resolved into their antipodes on a Chiralcel OD column. The (P)-configuration is assigned to the heptaleno[1,2-c]furans showing a negative Cotton effect at ca. 320 nm in the CD spectrum in hexane (cf. Figs. 3–5 as well as Table 7). The (P)-configuration of (–)- 6a is correlated with the established (P)-configuration of the dimethanol (–)- 5a via dehydrogenation with MnO2. The degree of twisting of the heptalene skeleton of 6 and 23 is determined by the Me-substitution pattern (cf. Table 9). The larger the heptalene gauche torsion angles are, the more hypsochromically shifted is the heptalene absorption band above 300 nm (cf. Table 7 and 8, as well as Figs. 6–9).  相似文献   

5.
6,10-Diphenylbenz[a]azulene ( 3 ) was reacted with dimethyl acetylenedicarboxylate (ADM) in the presence of 2 mol-% of [RuH2(PPh3)4] in MeCN at 100° to yield a 7:1 mixture of dimethyl 2,6-diphenyl-9,10-benzotricyclo[6.2.2.01,7]dodeca-2,4,6,9,11-pentaene-11,12-dicarboxylate ( 4 ) and dimethyl 8,12-diphenylbenzo[d]heptalene-6,7-dicarboxylate ( 5 ; Scheme 2). The tricycle 4 , when heated in DMF at 150° for 1 h led to the formation of 81.5% of the heptalene-6,7-dicarboxylate 5 and 15% of the starting azulene 3 . No rearrangement of tricycle 4 was observed, when it was heated at temperatures up to 180° in pseudocumene. The heptalene-6,7-dicarboxylate 5 was easily separated into its antipodes (PM)-and (MP)- 5 on a Chiracel column (cf. Fig. 2). On heating at 150° for 1 h, (MP)- 5 showed no racemization at all. The Ru-catalyzed reaction of benz[a]azulene ( 6 ) with ADM led to the formation of dimethyl 9,10-benzotricyclo[6.2.2.01,7]dodeca-2,4,6,9,11-pentaene-11,12-dicarboxylate ( 7 ; Scheme 3). However, the formation of the corresponding heptalene-6,7-dicarboxylate could not be observed.  相似文献   

6.
The benzo[a]heptalene formation from 4‐[(R‐sulfonyl)acetyl]heptalene‐5‐carboxylates 15 and 5‐[(R‐sulfonyl)acetyl]heptalene‐4‐carboxylates 16 (R=Ph or morpholino) in the presence of R′SO2CH2Li and BuLi has been investigated (Scheme 6). Only the sulfonyl moiety linked to the C?O group at C(4) of the heptalene skeleton is found at C(3) of the formed benzo[a]heptalene‐2,4‐diols 3 in accordance with the general mechanism of their formation (Scheme 3). Intermediates that might rearrange to corresponding 2‐sulfonylbenzo[a]heptalene‐1,3‐diols lose HO? under the reaction conditions to yield the corresponding cyclopenta[d]heptalenones of type 11 (Schemes 6 and 7). However, the presence of an additional Me group at C(α) of the lithioalkyl sulfones suppresses the loss of HO?, and 4‐methyl‐2‐sulfonylbenzo[a]heptalene‐1,3‐diols of type 4c have been isolated and characterized for the first time (Schemes 8 and 10). A number of X‐ray crystal‐structure analyses of starting materials and of the new benzo[a]heptalenes have been performed. Finally, benzo[a]heptalene 4c has been transformed into its 1,2,3‐trimethoxy derivative 23 , a benzo[a]heptalene with the colchicinoid substitution pattern at ring A (Scheme 11).  相似文献   

7.
3‐(Phenylsulfonyl)benzo[a]heptalene‐2,4‐diols 1 can be desulfonylated with an excess of LiAlH4/MeLi?LiBr in boiling THF in good yields (Scheme 6). When the reaction is run with LiAlH4/MeLi, mainly the 3,3′‐disulfides 6 of the corresponding 2,4‐dihydroxybenzo[a]heptalene‐3‐thiols are formed after workup (Scheme 7). However, the best yields of desulfonylated products are obtained when the 2,4‐dimethoxy‐substituted benzo[a]heptalenes 2 are reduced with an excess of LiAlH4/TiCl4 at ?78→20° in THF (Scheme 10). Attempts to substitute the PhSO2 group of 2 with freshly prepared MeONa in boiling THF led to a highly selective ether cleavage of the 4‐MeO group, rather than to desulfonylation (Scheme 13).  相似文献   

8.
A two-step synthesis of 4-methylcolchicine ( 13 ), starting from colchicine ( 2 ), has been developed (Scheme 5). In three steps, 4-ethylcolchicine ( 28 ) is also accessible from 2 (Scheme 8). Colchicine ( 2 ) and its derivatives 13 and 28 have been transformed into the benzo[a]heptalene derivatives 9 , 18 , and 34 , respectively, by Hofmann degradation of the corresponding deacetylcolchiceine 3, 19 , and 29 , respectively, followed by methylation of the two O-functions first with diazomethane and then with trimethoxonium tetrafluoroborate (Scheme 2 and 6). The thus formed tropylium salts gave, on deprotonation with Me3N in CHCl3, the expected pentamethoxybenzo[a]heptalenes 9, 18 , and 34 , respectively. X-Ray crystal-structure analysis of 9 (Fig.3) and 18 (Fig. 7), determination of the vicinal coupling constants of the H-atoms at the heptalene skeleton as well as the measurement of the racemization rate of the new benzo[a]heptalenes revealed a marked influence of the substituent at C(4) on the degree of twisting of the heptalene skeleton. The absolute configuration of the resolved heptalenes was deduced from their long-wavelength CD maxima around 350 nm. The heptalenes with a negative maximum in this range possess (7aP)-configuration.  相似文献   

9.
Treatment of 6,7‐diethoxy‐3,4‐dihydroisoquinoline ( 8 ) and its 1‐methyl derivative 12 with hydrazonoyl halides 10 in the presence of Et3N in THF under reflux afforded the corresponding 5,6‐dihydro‐1,2,4‐triazolo[3,4‐a]isoquinolines 11 and 13 , respectively, in high yield (Schemes 2 and 3). The products are formed via regioselective 1,3‐dipolar cycloaddition of the intermediate nitrilimines 9 with the isoquinoline C=N bond. Reaction of 6,7‐diethoxy‐3,4‐dihydroisoquinoline‐1‐acetonitrile ( 4a ) with ethyl α‐cyanocinnamates 15 in the presence of piperidine in refluxing MeCN yielded benzo[a]quinolizin‐4‐ones 16 (Scheme 4). Under the same conditions, 12 and arylidene malononitriles 19 reacted to give benzo[a]quinolizin‐4‐imines 20 (Scheme 5). Instead of 15 and 19 , mixtures of an aromatic aldehyde, and ethyl cyanoacetate or malononitrile, respectively, can be used in a one‐pot reaction.  相似文献   

10.
It is shown in this ‘Part 2’ that heptaleno[1,2‐c]furans 1 react thermally in a Diels–Alder‐type [4+2] cycloaddition at the furan ring with vinylene carbonate (VC), phenylsulfonylallene (PSA), α‐(acetyloxy)acrylonitrile (AAN), and (1Z)‐1,2‐bis(phenylsulfonyl)ethene (ZSE) to yield the corresponding 1,4‐epoxybenzo[d]heptalenes (cf. Schemes 1, 5, 6, and 8). The thermal reaction of 1a and 1b with VC at 130° and 150°, respectively, leads mainly to the 2,3‐endo‐cyclocarbonates 2,3‐endo‐ 2a and ‐ 2b and in minor amounts to the 2,3‐exo‐cyclocarbonates 2,3‐exo‐ 2a and ‐ 2b . In some cases, the (P*)‐ and (M*)‐configured epimers were isolated and characterized (Scheme 1). Base‐catalyzed cleavage of 2,3‐endo‐ 2 gave the corresponding 2,3‐diols 3 , which were further transformed via reductive cleavage of their dimesylates 4 into the benzo[a]heptalenes 5a and 5b , respectively (Scheme 2). In another reaction sequence, the 2,3‐diols 3 were converted into their cyclic carbonothioates 6 , which on treatment with (EtO)3P gave the deoxygenated 1,4‐dihydro‐1,4‐epoxybenzo[d]heptalenes 7 . These were rearranged by acid catalysis into the benzo[a]heptalen‐4‐ols 8a and 8b , respectively (Scheme 2). Cyclocarbonate 2,3‐endo‐ 2b reacted with lithium diisopropylamide (LDA) at ?70° under regioselective ring opening to the 3‐hydroxy‐substituted benzo[d]heptalen‐2‐yl carbamate 2,3‐endo‐ 9b (Scheme 3). The latter was O‐methylated to 2,3‐endo‐(P*)‐ 10b . The further way, to get finally the benzo[a]heptalene 13b with MeO groups in 1,2,3‐position, could not be realized due to the fact that we found no way to cleave the carbamate group of 2,3‐endo‐(P*)‐ 10b without touching its 1,4‐epoxy bridge (Scheme 3). The reaction of 1a with PSA in toluene at 120° was successful, in a way that we found regioisomeric as well as epimeric cycloadducts (Scheme 5). Unfortunately, the attempts to rearrange the products under strong‐base catalysis as it had been shown successfully with other furan–PSA adducts were unsuccessful (Scheme 4). The thermal cycloaddition reaction of 1a and 1b with AAN yielded again regioisomeric and epimeric adducts, which could easily be transformed into the corresponding 2‐ and 3‐oxo products (Scheme 6). Only the latter ones could be rearranged with Ac2O/H2SO4 into the corresponding benzo[a]heptalene‐3,4‐diol diacetates 20a and 20b , respectively, or with trimethylsilyl trifluoromethanesulfonate (TfOSiMe3/Et3N), followed by treatment with NH4Cl/H2O, into the corresponding benzo[a]heptalen‐3,4‐diols 21a and 21b (Scheme 7). The thermal cycloaddition reaction of 1 with ZSE in toluene gave the cycloadducts 2,3‐exo‐ 22a and ‐ 22b as well as 2‐exo,3‐endo‐ 22c in high yields (Scheme 8). All three adducts eliminated, by treatment with base, benzenesulfinic acid and yielded the corresponding 3‐(phenylsulfonyl)‐1,4‐epoxybenzo[d]heptalenes 25 . The latter turned out to be excellent Michael acceptors for H2O2 in basic media (Scheme 9). The Michael adducts lost H2O on treatment with Ac2O in pyridine and gave the 3‐(phenylsulfonyl)benzo[d]heptalen‐2‐ones 28a and 3‐exo‐ 28b , respectively. Rearrangement of these compounds in the presence of Ac2O/AcONa lead to the formation of the corresponding 3‐(phenylsulfonyl)benzo[a]heptalene‐1,2‐diol diacetates 30a and 30b , which on treatment with MeONa/MeI gave the corresponding MeO‐substituted compounds 31a and 31b . The reductive elimination of the PhSO2 group led finally to the 1,2‐dimethoxybenzo[a]heptalenes 32a and 32b . Deprotonation experiments of 32a with t‐BuLi/N,N,N′,N′‐tetramethylethane‐1,2‐diamine (tmeda) and quenching with D2O showed that the most acid C? H bond is H? C(3) (Scheme 9). Some of the new structures were established by X‐ray crystal‐diffraction analyses (cf. Figs. 1, 3, 4, and 5). Moreover, nine of the new benzo[a]heptalenes were resolved on an anal. Chiralcel OD‐H column, and their CD spectra were measured (cf. Figs. 8 and 9). As a result, the 1,2‐dimethoxybenzo[a]heptalenes 32a and 32b showed unexpectedly new Cotton‐effect bands just below 300 nm, which were assigned to chiral exciton coupling between the heptalene and benzo part of the structurally highly twisted compounds. The PhSO2‐substituted benzo[a]heptalenes 30b and 31b showed, in addition, a further pair of Cotton‐effect bands in the range of 275–245 nm, due to chiral exciton coupling of the benzo[a]heptalene chromophore and the phenylsulfonyl chromophore (cf. Fig. 10).  相似文献   

11.
Syntheses and Investigations of [Oxazolo[2,3-a]isoindol-9b(2H)-yl]phosphonates and -phosphinates: a New Class of Heterocycles We attempted to synthesize diethyl (1-methyl-2-phthalimidoethyl)phosphonate ( 14a ) in a Michaelis-Becker reaction using diethyl sodiophosphonate ( 13 ) and the tosylate 12a of (2-hydroxypropyl)phthalimide as starting materials. Instead of TsO substitution in 12a by the nucleophile 13 , the carbonyl C-atom of the phthalimido moiety was attacked by 13 , followed by an intramolecular nucleophilic substitution at C(2) of the side chain leading to the (oxazolo[2,3-a]isoindolyl)phosphonate 15a (Scheme 1). Similarly, 12a and N-(2-bromoethyl)phthalimide ( 12b ) reacted with butyl (benzene)sodiophosphinate ( 18 ) to the (oxazolo[2,3-a]isoindolyl)(phenyl)phosphinates 20a and 20b , respectively (Scheme 2). The attempt to synthesize enantiomerically pure 2-substituted (2-phthalimidoethyl)phosphonates 27 starting from L -α-amino-acids failed, too (Scheme 3): the main products of the reaction of the N,N-phthaloyl-O1-tosyl-L -aminoalcohols 25a–d with 13 were the 3-substituted (oxazolo[2,3-a]isoindolyl)-phosphonates 26a–d , the desired 27b and 27c being observed as by-products in the 31P-NMR spectrum.  相似文献   

12.
Tetrahydrobenzo[a]pyrrolizidines (= octahydro-1H-pyrrolo[2,1-a]isoindoles) and tetrahydrobenzo[a]indo-lizidines, (= decahydropyrido[2,1-a]isoindoles) were prepared stereoselectively in four steps through an amineinduced ring-opening of 3-bromo-2,5-dimethylthiophene 1,1-dioxide ( 1 ) with L -prolinol ( 9 ), piperidine-2-methanol ( 10 ), and piperidine-2-ethanol ( 11 ), yielding the dienes (2S)-1-[(2E,4Z)-4-bromohexa-2,4-dienyl]pyrrolidine-2-methanol ( 12 ), 1-[(2E,4Z)-4-bromohexa-2,4-dienyl]piperidine-2-methanol ( 13 ), and 1-[(2E,4Z)-4-bromo-hexa-2,4-dienyl]piperidine-2-ethanol ( 14 ; Scheme2), which, after conversion into their α,β-unsaturated esters, cyclized in a TiCl4-catalyzed intramolecular Diets-Alder reaction (Scheme3). A discussion on the mechanism of the ring opening reaction including semiempirical and ab initio calculations is also presented.  相似文献   

13.
Alkylation reactions of 3‐(X‐sulfonyl)benzo[a]heptalene‐2,4‐diols (X=Ph, morpholin‐4‐yl) and their dimethyl ethers were studied. The diols form with K2CO3/MeI in aqueous media the 1‐methylated benzoheptalenes, but in yields not surpassing 20% (Table 1). On the other hand, 2,4‐dimethoxybenzo[a]heptalenes can easily be lithiated at C(3) with BuLi and then treated with alkyl iodides to give the 3‐alkylated forms in good yield (Table 2). Surprising is the reaction with two equiv. or more of t‐BuLi since the alkylation at C(4) is accompanied by the reductive elimination of the X‐sulfonyl group at C(3) (Table 3). Most exciting is also the course of 2,4‐dimethoxy‐3‐(phenylsulfonyl)benzo[a]heptalenes in the presence of an excess of MeLi. After the expected exchange of MeO against Me at C(4) (Scheme 6), rearrangement takes place under formation of 4‐benzyl‐2‐methoxybenzo[a]heptalenes and concomitant loss of the sulfonyl group at C(3) (Table 4). In the case of X=morpholin‐4‐yl, rearrangement cannot occur. However, the intermediate benzyl anions of Type E (Scheme 8) react easily with O2 of the air to build up corresponding benzo[a]heptalene‐4‐methanols (Table 6).  相似文献   

14.
It is shown that heptalene‐4,5‐dicarboxylates 2 , which react with lithiated methyl sulfones mainly in a Michael fashion at C(3) (cf. Scheme 2), so that the formation of 3‐sulfonylbenzo[a]heptalene‐2,4‐diols 5 is repressed or completely suppressed, can be transformed into corresponding pseudo‐esters 15 (Scheme 4). These pseudo‐esters, on treatment with lithiated methyl sulfones, followed by addition of BuLi, furnish the 3‐sulfonylbenzo[a]heptalene‐2,4‐diols 5 in excellent‐to‐moderate yields without formation of Michael adducts or their follow‐up products (cf. Scheme 5 and 6). The reaction of the pseudo‐ester 15a with Li[13C]H2SO2Ph, followed by treatment with non‐labeled LiCH2SO2Ph and then BuLi, led to the exclusive formation of 3‐(phenylsulfonyl)‐[1‐13C]benzo[a]heptalene‐2,4‐diol 5a* (Scheme 9). This experiment demonstrates that the (phenylsulfonyl)acetyl groups at C(4) and C(5) of the heptalene core retain their individual positions in the course of the benzo[a]heptalene‐2,4‐diol formation. These findings are only compatible with an intramolecular rearrangement mechanism as depicted in Scheme 10.  相似文献   

15.
Oxidation of 1,2-Thiazoles; A Convenient Approach to 1,2-Thiazol-3(2H)-one 1,1-Dioxides The 1,2-thiazoles obtained from 3-chloroalk-2-enals and ammonium thiocyanate ( 7 → 9 , Scheme 1) are easily transformed to 1,2-thiazol-3(2H)-one 1,1-dioxidcs 10 on treatment with H2O2 in AcOH at 80°. Hydrogenation of 10 in AcOH yields the corresponding saturated 1,2-thiazolidin-3-one 1,1-dioxides 16 (Scheme 3). Cycloalka[c]-1,2-thiazoles 18 are prepared from 2-[(thiocyanato)methyliden]cycloalkan-1-ones and ammonia (Scheme 4). Surprisingly, oxidation of 18a with H2O2 in AcOH yields the tricyclic oxaziridine 19.  相似文献   

16.
1-Mesityl allene ( 1 ), 1-mesityl-3-methyl allene ( 2 ) and 1-mesityl-3,3-dimethyl allene ( 3 ) rearrange thermally at 150–190° in decane via [1,5s]sigmatropic H-shifts to yield the o-quinodimethanes 4 , which cyclise to give the 1,2-dihydronaphthalenes 5 and 6 and/or undergo [1,7 a]sigmatropic H-shifts to give 1-mesityl-(Z)-buta-1, 3-dienes (Z)- 7 and (Z)- 8 , respectively (Schemes 1,3,4 and 5) in almost quantitative yields. The activation parameters of these isomerisations are given in Table 1. 1-Mesityl-1-methyl allene ( 9 ) isomerises at 190° to give 4,5,7-trimethyl-1,2-dihydronaphthalene ( 17 ) in 50% yield (Scheme 6). 2′-Isopropylphenyl allene ( 10 ) in decane rearranges at 170° to 1-(Z)-propenyl-2-isopropenyl-benzene ((Z)- 19 , Scheme 7). Deuterium labelling experiments show that the rate determining step is an aromatic [1,5s]sigmatropic hydrogen shift from an sp3- to an sp-hybridised carbon atom. The primary kinetic isotopic effect (kH/kD) is 3.45, while the secondary βisotopic effect is 1.20 (Scheme 7 and Table 2).  相似文献   

17.
The reaction of guaiazulene ( 4 ) and dimethyl acetylenedicarboxylate (ADM) in tetralin or toluene, catalyzed by 5 mol-% of trifluoroacetic acid (TFA) at ambient temperature, leads to the formation of the corresponding heptalene-4,5-dicarboxylate 6 and a guaiazulenyl-substituted 2,2a,4a,8b-tetrahydrocyclopent[cd]azulene derivative 7 beside the expected guaiazulenyl-substituted ethenedicarboxylates (E)- 5 and (Z)- 5 as main products (Scheme 2). The structure of 7 was unequivocally established by an X-ray crystal-structure analysis (Fig. 1). Precursor of 7 must be the 2a,4a-dihydrocyclopent[cd]azulene-3,4-dicarboxylate 9 which reacts, under TFA catalysis, with a second molecule of 4 (Scheme 3). No formation of products of type 7 has been observed in the TFA-catalyzed reaction of 4,6,8-trimethyl- and 1,4,6,8-tetramethylazulene ( 13 and 16 , respectively) and ADM (Scheme 4). On the other hand, the TFA-catalyzed reaction of azulene ( 18 ) itself and ADM at ambient temperature gives rise to a whole variety of new products (Scheme 5), the major part of which is derived from dimethyl 2a,4a-dihydrocyclopent[cd]azulene-3,4-dicarboxylate ( 25 ) as the main intermediate (Scheme 6). Nevertheless, for the formation of the 2a,4a,6,8b-tetrahydrocyclobut[a]azulene derivatives (E)- 24a and (E)- 24b , a corresponding 2a,8b-dihydro precursor 29 has to be postulated as crucial intermediate (Scheme 8).  相似文献   

18.
Sodium [1,3-13C2]cyclopentadienide in tetrahydrofuran (THF) has been prepared from the corresponding labelled [13C2]cyclopentadiene which was synthesized from 13CO2 and (chloromethyl)trimethylsilane (cf. Scheme 10) according to an established procedure. It could be shown that the acetate pyrolysis of cis-cyclopentane-1,2-diyl diacetate (cis- 22 ) at 550 ± 5° under reduced pressure (60 Torr) gives five times as much cyclopentadiene as trans- 22 . The reaction of sodium [1,3-13C2]cyclopentadienide with 2,4,6-trimethylpyrylium tetrafluoroborate in THF leads to the formation of the statistically expected 2:2:1 mixture of 4,6,8-trimethyl[1,3a-13C2], -[2,3a-13C2]-, and -[1,3-13C2]azulene ( 20 ; cf. Scheme 7 and Fig. 1). Formylation and reduction of the 2:2:1 mixture [13C2]- 20 results in the formation of a 1:1:1:1:1 mixture of 1,4,6,8-tetramethyl[1,3-13C2]-, -[1,3a-13C2]-, -[2,3a-13C2]-, -[2,8a-13C2]-, and -[3,8a-13C2]azulene ( 5 ; cf. Scheme 8 and Fig. 2). The measured 2J(13C, 13C) values of [13C2]- 20 and [13C2]- 5 are listed in Tables 1 and 2. Thermal reaction of the 1:1:1:1:1 mixture [13C2]- 5 with the four-fold amount of dimethyl acetylenedicarboxylate (ADM) at 200° in tetralin (cf. Scheme 2) gave 5,6,8,10-tetramethyl-[13C2]heptalene-1,2-dicarboxylate ([13C2]- 6a ; 22%), its double-bond-shifted (DBS) isomer [13C2]- 6b (19%), and the corresponding azulene-1,2-dicarboxylate 7 (18%). The isotopically isomeric mixture of [13C2]- 6a showed no 1J(13C,13C) at C(5) (cf. Fig. 3). This finding is in agreement with the fact that the expected primary tricyclic intermediate [7,11-13C2]- 8 exhibits at 200° in tetralin only cleavage of the C(1)? C(10) bond and formation of a C(7)? C(10) bond (cf. Schemes 6 and 9), but no cleavage of the C(1)? C(11) bond and formation of a C(7)? C(11) bond. The limits of detection of the applied method is ≥96% for the observed process, i.e., [1,3a-13C2]- 5 + ADM→ [7,11-13C2]- 8 →[1,6-13C2]- 9 →[5,10a-13C2]- 6a (cf. Scheme 6).  相似文献   

19.
It is shown that the 2-(hydroxymethyl)-1-methylazulenes 6 are being oxidized by activated MnO2 in CH2Cl2 at room temperature to the corresponding azulene-1,2-dicarbaldehydes 7 (Scheme 2). Extension of the MnO2 oxidation reaction to 1-methyl- and/or 3-methyl-substituted azulenes led to the formation of the corresponding azulene-1-carbaldehydes in excellent yields (Scheme 3). The reaction of unsymmetrically substituted 1,3-dimethyl-azulenes (cf. 15 in Scheme 4) with MnO2 shows only little chemoselectivity. However, the observed ratio of the formed constitutionally isometric azulene-1-carbaldehydes is in agreement with the size of the orbital coefficients in the HOMO of the azulenes. The reaction of guaiazulene ( 18 ) with MnO2 in dioxane/H2O at room temperature gave mainly the expected carbaldehyde 19 . However, it was accompanied by the azulene-diones 20 and 21 (Scheme 5). The precursor of the demethylated compound 20 is the carbaldehyde 19 . Similarly, the MnO2 reaction of 7-isopropyl-4-methyalazulene ( 22 ) as well as of 4,6,8-trimethylazulene ( 24 ) led to the formation of a mixture of the corresponding azulene-1,5-diones and azulene-1,7-diones 20 / 23 and 25 / 26 , respectively, in decent yields (Schemes 6 and 7). No MnO2 reaction was observed with 5,7-dimethylazulene.  相似文献   

20.
The thermal reaction of azulene-1-carbaldehydes 5 and 6 with excess dimethyl acetylenedicarboxylate (ADM) in decalin leads mainly to the formation of (1 + 1) and (1 + 2) adducts arising from the addition of ADM at the seven-membered ring of the azulenes (cf. Schemes 2 and 4). The (1 + 2) adducts are formed in a homo-Diels-Alder reaction of ADM and isomeric tricyclic carbaldehydes which are derived from the primary tricyclic carbaldehydes by reversible [1s5s]-C shifts (cf. Schemes 3 and 5). The thus formed pentacyclic carbaldehydes seem to undergo deep-seated skeletal rearrangements (cf. Scheme 7) which result finally in the formation of the formyl-tetrahydrocyclopenta[bc]acenaphthylene-tetraesters 12 and 19 , respectively. In other cases, e.g., azulene-1-carbaldehydes 7 and 8 (cf. Scheme 8), the thermal reaction with excess ADM furnishes only the already known tetracycfic (1 + 2) adducts of type anti- 26 to ‘anti’- 29 . The thermal reaction of 1,3,4,8-tetramethylazulene ( 9 ) with excess ADM in decalin resulted in the formation of two (1 + 2) and one (1 + 3) adduct in low yields (cf. Scheme 9). The latter turned out to be the 2,6-bridged barrelene derivative 32 . There are structural evidences that 32 is formed by similar pathways as the formyl-tetrahydrocyclopenta[bc]acenaphthylene-tetraesters (cf. Schemes 7 and 11). [2H3]Me-Labelling experiments are in agreement with the proposed mechanisms (cf. Scheme 13).  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