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 thermal 4 : 1 equilibrium mixture of 1,3,5,6- and 1,3,5,10-tetramethylheptalene ( 13a and 13b , resp.) has been prepared, starting from the thermal equilibrium mixture of dimethyl 6,8,10-trimethylheptalene-1,2- and -4,5-dicarboxylate ( 6a and 6b , resp.; cf. Scheme 5). These heptalenes undergo double-bond shifts (DBS) even at ambient temperature. Treatment of the mixture 13a / 13b 4 : 1 with [Cr(CO)3(NH3)3] in boiling 1,2-dimethoxyethane resulted in the formation of all four possible mononuclear Cr(CO)3 complexes 19a – 19d of 13a and 13b , as well as two binuclear Cr(CO)3 complexes 20a and 20b , respectively, in a total yield of 87% (cf. Scheme 7). The mixture of complexes was separated by column chromatography, followed by preparative HPLC (cf. Fig. 2). The structures of all complexes were established by X-ray crystal-structure analyses (complex 19b and 20b ; cf. Figs. 6 – 8) and extensive 1H-NMR measurements (cf. Table 3). In 20b , the two Cr(CO)3 groups are linked in a `syn'-mode to the highly twisted heptalene π-skeleton. The correspondence of the 1H-NMR data of 20a with that of 20b indicates that the two Cr(CO)3 groups in 20a also have a `syn'-arrangement. The thermal behavior of the mononuclear complexes 19a – 19d has been studied at 85° in hexafluorobenzene (HFB). At this temperature, all four complexes undergo rearrangement to the same thermal equilibrium mixture (cf. Table 8). The rates for the thermal equilibration of each complex have been determined by 1H-NMR measurements (cf. Figs. 912) and analyzed by seven different kinetic schemes (Chapt. 2). The equilibration rates are in agreement with two different haptotropic rearrangements that take place, namely intra- and inter-ring shifts of the Cr(CO)3 group, whereby both rearrangements are accompanied by DBS of the heptalene π-skeleton (cf. Scheme 9). All individual kinetic steps possess similar ΔG values in the range of 29 – 31 kcal⋅mol−1 (cf. Table 8). The occurrence of inter-ring haptotropic migrations of Cr(CO)3 groups has already been established for anellated aromatic systems (cf. Scheme 10); however, it is the first time that these rearrangements have been unequivocally demonstrated for Cr(CO)3 complexes of non-planar bicyclic [4n]annulenes, such as heptalenes. The mechanism of migration may be similar to that proposed for aromatic systems (cf. Schemes 10 and 11).  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.
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   

4.
A number of azulenes 1 , in particular those with π‐substituents at C(6) such as phenyl, 3,5‐dimethylphenyl, and 4‐biphenyl, have been reacted with 3 mol‐equiv. of dimethyl acetylenedicarboxylate (ADM) in MeCN at 110° (cf. Scheme 1). Main products had been, in all cases, the corresponding heptalene‐4,5‐dicarboxylates 2 . However, a whole number of side products, mainly rearranged (1+2)‐adducts with two molecules of ADM, in amounts of 0.2–9% were also isolated and characterized (cf. Scheme 2). The 2a,8a‐dihydro‐3,4‐ethenoazulene‐1,2‐dicarboxylates 14 , formed by energetically favorable ring closure from the solvent‐stabilized zwitterions 15 , resulting from bond heterolysis in the primary cycloadducts 12 (cf. Scheme 3), have been mechanistically identified as the pivotal intermediates responsible for the formation of all side product (cf. Schemes 5, 9, 12, and 13). Deuterium‐labeling experiments were in agreement with the proposed mechanisms, indicating that sigmatropic [1,5s]‐H shifts in 14 (cf. Scheme 6) as well as isoconjugate [1,4s]‐H shifts in resonance‐stabilized zwitterions of type 21 (cf. Scheme 9) are the crucial steps for side‐product formation. It is postulated that a concluding antarafacial 8e‐dyotropic rearrangement is responsible for the appearance of the 2,4a‐dihydrophenanthrene‐tetracarboxylates of type trans‐ 6 (cf. Scheme 9) in the reaction mixtures, which further rearrange thermally by a not fully understood mechanism into the isomeric tetracarboxylates 7 (cf. Schemes 10 and 11). Most surprising is the presence of a small amount (0.3–1%) of the azulene‐4,5,7,8‐tetracarboxylate 9 in the reaction mixture of azulene 1a and ADM. It is proposed that the formation of 9 is the result of a [1,5s]‐C shift in the spiro‐linked intermediates 24 , which, after prototropic shift and take‐up of a third molecule of ADM, disintegrate by a retro‐Diels‐Alder reaction into 9 and the phthalic diesters 30 (cf. Scheme 12). The UV/VIS spectra of the π‐substituted heptalene‐4,5‐dicarboxylates 2d – 2f and their double‐bond shifted (DBS) forms 2d – 2f (cf. Table 4 and Figs. 912) exhibit in comparison with the heptalene‐dicarboxylates 2a and 2′a , carrying a t‐Bu group at C(8), only marginal differences, which are mainly found in the relative intensity and position of heptalene bands II and III .  相似文献   

5.
Dimethyl heptalene‐4,5‐dicarboxylates
  • 1 The locants of heptalene itself are maintained throughout the whole work. See footnote 4 in [1] for reasoning.
  • undergo preferentially a Michael addition reaction at C(3) with α‐lithiated alkyl phenyl sulfones at temperatures below ?50°, leading to corresponding cis‐configured 3,4‐dihydroheptalene‐4,5‐dicarboxylates (cf. Table 1, Schemes 3 and 4). The corresponding heptalenofuran‐1‐one‐type pseudoesters of dimethyl heptalene‐4,5‐dicarboxylates (Scheme 5) react with [(phenylsulfonyl)methyl]lithium almost exclusively at C(1) of the furanone group (Scheme 6). In contrast to this expected behavior, the uptake of 1‐[phenylsulfonyl)ethyl]lithium occurs at C(5) of the heptalenofuran‐1‐ones as long as they carry a Me group at C(11) (Schemes 6 and 7). The 1,4‐ as well as the 1,6‐addition products eliminate, on treatment with MeONa/MeOH in THF, benzenesulfinate, thus leading to 3‐ and 4‐alkylated dimethyl heptalene‐4,5‐dicarboxylates, respectively (Schemes 813). The configuration of the addition reaction of the nucleophiles to the inherently chiral heptalenes is discussed in detail (cf. Schemes 1419) on the basis of a number of X‐ray crystal‐structure determinations as well as by studies of the temperature‐dependence of the 1H‐NMR spectra of the addition products.  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.
    Methyl heptalenecarboxylates of type A and B with π(1) and π(2) substituents in 1,4‐relation (Scheme 1) were synthetized starting with dimethyl 1‐methylheptalene‐4,5‐dicarboxylates 5b and 6b derived from 7‐isopropyl‐1,4‐dimethylazulene (=guaiazulene) and 1,4,6,8‐tetramethylazulene by thermal reaction with dimethyl acetylenedicarboxylate. The further general way of proceeding for the introduction of the π(1) and π(2) substituents is displayed in Scheme 3, and the thus obtained methyl heptalene‐5‐carboxylates of type A and B are listed in Table 1. The C?C bonds of the 2‐arylethenyl and 4‐arylbuta‐1,3‐dien‐1‐yl groups of π(1) and π(2) were in all cases (E)‐configured and showed s‐trans conformation at the C? C bonds (X‐ray and 1H‐NOE evidence) in the B ‐type as well as in the A ‐type heptalenes (cf. Figs. 5–12). All B ‐type heptalenes showed a strongly enhanced heptalene band I in the wavelength region 440–490 nm in hexane/CH2Cl2 9 : 1 (cf. Table 4 and Figs. 13–20). The A ‐type heptalenes showed in this region only weak absorption, recognizable as shoulders or simply tailing of the dominating heptalene bands II/III (Table 5). Absorption band I of the B ‐type heptalenes appeared almost at the same wavelength as the longest wavelength absorption band of comparable open‐chain α,ω‐diarylpolyenes (cf. Fig. 21). The cyclic double bond shift (DBS) of the A ‐ and B ‐type heptalenes could be photochemically steered in one or the other direction by selective irradiation (cf. Fig. 22).  相似文献   

    8.
    A novel approach towards the synthesis of functionalized benzo[a]heptalenes 9 and 10 via a 6π‐electrocyclic ring closure – aromatization sequence of corresponding bis[prop‐2‐enoates] 5 and 6 has been developed (Scheme 1). The starting bis[prop‐2‐enoates] have been prepared from the corresponding dialdehydes 3a and 4a in a Wittig‐Horner reaction, and their UV/VIS properties have also been investigated (Fig. 1 and Table 1). The dehydrogenations of the corresponding diols 1 and 2 to dialdehydes with a number of oxidizing reagents, including MnO2 in CH2Cl2, tetrapropylammonium perruthenate (TPAP), and activated DMSO, have been studied in detail.  相似文献   

    9.
    A new method for the smooth and highly efficient preparation of polyalkylated aryl propiolates has been developed. It is based on the formation of the corresponding aryl carbonochloridates (cf. Scheme 1 and Table 1) that react with sodium (or lithium) propiolate in THF at 25 – 65°, with intermediate generation of the mixed anhydrides of the arylcarbonic acids and prop‐2‐ynoic acid, which then decompose almost quantitatively into CO2 and the aryl propiolates (cf. Scheme 11). This procedure is superior to the transformation of propynoic acid into its difficult‐to‐handle acid chloride, which is then reacted with sodium (or lithium) arenolates. A number of the polyalkylated aryl propiolates were subjected to flash vacuum pyrolysis (FVP) at 600 – 650° and 10−2 Torr which led to the formation of the corresponding cyclohepta[b]furan‐2(2H)‐ones in average yields of 25 – 45% (cf. Scheme 14). It has further been found in pilot experiments that the polyalkylated cyclohepta[b]furan‐2(2H)‐ones react with 1‐(pyrrolidin‐1‐yl)cyclohexene in toluene at 120 – 130° to yield the corresponding 1,2,3,4‐tetrahydrobenz[a]azulenes, which become, with the growing number of Me groups at the seven‐membered ring, more and more sensitive to oxidative destruction by air (cf. Scheme 15).  相似文献   

    10.
    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   

    11.
    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   

    12.
    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   

    13.
    A number of aryl 3‐arylprop‐2‐ynoates 3 has been prepared (cf. Table 1 and Schemes 3 – 5). In contrast to aryl prop‐2‐ynoates and but‐2‐ynoates, 3‐arylprop‐2‐ynoates 3 (with the exception of 3b ) do not undergo, by flash vacuum pyrolysis (FVP), rearrangement to corresponding cyclohepta[b]furan‐2(2H)‐ones 2 (cf. Schemes 1 and 2). On melting, however, or in solution at temperatures >150°, the compounds 3 are converted stereospecifically to the dimers 3‐[(Z)‐diarylmethylidene]‐2,3‐dihydrofuran‐2‐ones (Z)‐ 11 and the cyclic anhydrides 12 of 1,4‐diarylnaphthalene‐2,3‐dicarboxylic acids, which also represent dimers of 3 , formed by loss of one molecule of the corresponding phenol from the aryloxy part (cf. Scheme 6). Small amounts of diaryl naphthalene‐2,3‐dicarboxylates 13 accompanied the product types (Z)‐ 11 and 12 , when the thermal transformation of 3 was performed in the molten state or at high concentration of 3 in solution (cf. Tables 2 and 4). The structure of the dihydrofuranone (Z)‐ 11c was established by an X‐ray crystal‐structure analysis (Fig. 1). The structures of the dihydrofuranones 11 and the cyclic anhydrides 12 indicate that the 3‐arylprop‐2‐ynoates 3 , on heating, must undergo an aryl O→C(3) migration leading to a reactive intermediate, which attacks a second molecule of 3 , finally under formation of (Z)‐ 11 or 12 . Formation of the diaryl dicarboxylates 13 , on the other hand, are the result of the well‐known thermal Diels‐Alder‐type dimerization of 3 without rearrangement (cf. Scheme 7). At low concentration of 3 in decalin, the decrease of 3 follows up to ca. 20% conversion first‐order kinetics (cf. Table 5), which is in agreement with a monomolecular rearrangement of 3 . Moreover, heating the highly reactive 2,4,6‐trimethylphenyl 3‐(4‐nitrophenyl)prop‐2‐ynonate ( 3f ) in the presence of a twofold molar amount of the much less reactive phenyl 3‐(4‐nitrophenyl)prop‐2‐ynonate ( 3g ) led, beside (Z)‐ 11f , to the cross products (Z)‐ 11fg , and, due to subsequent thermal isomerization, (E)‐ 11fg (cf. Scheme 10), the structures of which indicated that they were composed, as expected, of rearranged 3f and structurally unaltered 3g . Finally, thermal transposition of [17O]‐ 3i with the 17O‐label at the aryloxy group gave (Z)‐ and (E)‐[17O2]‐ 11i with the 17O‐label of rearranged [17O]‐ 3i specifically at the oxo group of the two isomeric dihydrofuranones (cf. Scheme 8), indicating a highly ordered cyclic transition state of the aryl O→C(3) migration (cf. Scheme 9).  相似文献   

    14.
    It has been found that dimethyl heptalene-4,5-dicarboxylates, when treated with 4 mol-equiv. of lithiated N,N-dialkylamino methyl sulfones or methyl phenyl sulfone, followed by 4 mol-equiv. of BuLi in THF in the temperature range of ?78 to 20°, give rise to the formation of 3-[(N,N-dialkylamino)sulfonyl]- or 3-(phenylsul-fonyl)benzo[a]heptalene-2,4-diols of. (cf. Scheme 4, and Tables 2 and 3). Accompanying products are 2,4-bis{[(N,N-dialkylamino)sulfonyl]methyl}- or 2,4-bis[(phenylsulfonyl)methyl]-4,10a-dihydro-3H-heptaleno[1,10-bc]furan-3-carboxylates as mixtures of diastereoisomers of. cf. Scheme 4, and (Tables 2 and 3) which are the result of a Michael addition reaction of the lithiated methyl sulfones at C(3) of the heptalene-4,5-dicarboxylates, followed by (sulfonyl)methylation of the methoxycarbonyl group at C(5) and cyclization of. (cf. Scheme 5). It is assumed that the benzo[a]heptalene formation is due to (sulfonyl)methylation of both methoxycarbonyl groups of the heptalene-4,5-dicarboxylates of. (cf. Schemes 6 and 8). The resulting bis-enolates 35 are deprotonated further. The thus formed tris-anions 36 can then cyclize to corresponding tris-anions 37 of cyclopenta[d]heptalenes which, after loss of N,N-dialkylamido sulfite or phenyl sulfinate, undergo a ring-enlargement reaction by 1,2-C migration finally leading to the observed benzo[a]heptalenes of. (cf. Schemes 8 and 9). The structures of the new product types have been finally established by X-ray crystal-structure analyses (cf. Figs. 1 and 2 as well as Exper. Part).  相似文献   

    15.
    Reaction of 2,2,4,4-tetramethylcyclobutane-1,3-dione ( 2 ) under phase-transfer-catalysis (PTC) conditions (CHCl3/aqueous NaOH) yielded a complex mixture of unexpected products (Scheme 2). From the organic phase, three ring-enlarged products 7 – 9 with a cyclopentane-1,3-dione (cf. 7 and 9 ) or a cyclopentenone skeleton (cf. 8 ) were isolated in low yield. After acidification of the aqueous phase, the oily residue was treated with CH2N2, and methyl 3-oxopentanoate 12 and dimethyl 2-hydroxybutanedioate 13 were obtained in almost equal amounts. The structures of 8 and 9 were established by X-ray crystal-structure analysis (Fig.). Mechanisms for the formation of the products, initiated by nucleophilic attack of trichloromethanide ion and opening of the cyclobutane ring, are proposed in Schemes 3 and 4.  相似文献   

    16.
    Regioselective Pd0‐catalyzed cross‐coupling of substrates, which bear bispropargylic leaving groups with silyl‐protected alkynes, has provided access to a variety of 1,3‐diethynylallenes, a new family of modules for three‐dimensional acetylenic scaffolding. In enantiomerically pure form, these C‐rich building blocks could provide access – by oxidative oligomerization – to a fascinating new class of helical oligomers and polymers with all‐carbon backbones (Fig. 2). In the first of two routes, a bispropargylic epoxide underwent ring opening during Sn 2′‐type cross‐coupling, and the resulting alkoxide was silyl‐protected, providing 1,3‐diethynylallenes (±)‐ 8 , (±)‐ 12 (Scheme 3), and (±)‐ 15 (Scheme 5). A more general approach involved bispropargylic carbonates or esters as substrates (Scheme 68), and this route was applied to the preparation of a series of 1,3‐diethynylallenes to investigate how their overall stability against undesirable [2+2] cycloaddition is affected by the nature of the substituents at the allene moiety. The investigation showed that the 1,3‐diethynylallene chromophore is stable against [2+2] cycloaddition only when protected by steric bulk and when additional π‐electron delocalization is avoided. The regioselectivity of the cross‐coupling to the bispropargylic substrates is entirely controlled by steric factors: attack occurs at the alkyne moiety bearing the smaller substituent (Schemes 9 and 10). Oxidative Hay coupling of the terminally mono‐deprotected 1,3‐diethynylallene (±)‐ 49 afforded the first dimer 50 , probably as a mixture of two diastereoisomers (Scheme 12). Attempts to prepare a silyl‐protected tetraethynylallene by the new methodology failed (Scheme 13). Control experiments (Schemes 1416) showed that the Pd0‐catalyzed cross‐coupling to butadiyne moieties in the synthesis of this still‐elusive chromophore requires forcing conditions under which rapid [2+2] cycloaddition of the initial product cannot be avoided.  相似文献   

    17.
    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   

    18.
    The C? C coupling of the two bicyclic, unsaturated dicarboximides 5 and 6 with aryl and heteroaryl halides gave, under reductive Heck conditions, the C‐aryl‐N‐phenyl‐substituted oxabicyclic imides 7a – c and 8a – c (Scheme 3). Domino‐Heck C? C coupling reactions of 5, 6 , and 1b with aryl or heteroaryl iodides and phenyl‐ or (trimethylsilyl)acetylene also proved feasible giving 8, 9 , and 10a – c , respectively (Scheme 4). Reduction of 1b with LiAlH4 (→ 11 ) followed by Heck arylation and reduction of 5 with NaBH4 (→ 13 ) followed by Heck arylation open a new access to the bridged perhydroisoindole derivatives 12a , b and 14a , b with prospective pharmaceutical activity (Schemes 5 and 6).  相似文献   

    19.
    A new concept for molecular switches, based on thermal or photochemical double-bond shifts (DBS) in [4n]annulenes such as heptalenes or cyclooctatetraenes, is introduced (cf. Scheme 2). Several heptalene-1,2- and -4,5-dicarboxylates (cf. Scheme 4) with (E)-styryl and Ph groups at C(5) and C(1), or C(4) and C(2), respectively, have been investigated. Several X-ray crystal-structure analyses (cf. Figs. 1–5) showed that the (E)-styryl group occupies in the crystals an almost perfect s-trans-conformation with respect to the C?C bond of the (E)-styryl moiety and the adjacent C?C bond of the heptalene core. Supplementary 1H-NOE measurements showed that the s-trans-conformations are also adopted in solution (cf. Schemes 6 and 9). Therefore, the DBS process in heptalenes (cf. Schemes 5 and 8) is always accompanied by a 180° torsion of the (E)-styryl group with respect to its adjacent C?C bond of the heptalene core. The UV/VIS spectra of the heptalene-1,2- and -4,5-dicarboxylates illustrated that it can indeed be differentiated between an ‘off-state’, which possesses no ‘through-conjugation’ of the π-donor substituent and the corresponding MeOCO group and an ‘on-state’ where this ‘through-conjugation’ is realized. The ‘through-conjugation’, i.e., conjugative interaction via the involved s-cis-butadiene substructure of the heptalene skeleton, is indicated by a strong enhancement of the intensities of the heptalene absorption bands I and II (cf. Tables 3–6). The most impressive examples are the heptalene-dicarboxylates 11a , representing the off-state, and 11b which stands for the on-state (cf. Fig.8).  相似文献   

    20.
    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   

    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