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 47 毫秒
1.
The treatment of diluted solutions of the hydroxy diamides 6a and 6b in toluene with HCl gas at 100° gave the dimeric, 14‐membered cyclodepsipeptide 10 in up to 72% yield (Scheme 3). The same product was formed from the linear dimer of 6b , the depsipeptide 11 , under the same conditions (cf. Scheme 4). All attempts to prepare the cyclic seven‐membered monomer 9 , starting with different precursors and using different lactonization methods failed, and 10 was the only product which was isolated (cf. Scheme 6). For example, the reaction of the ester 20 with NaH in toluene at 80° led exclusively to the cyclodimer 10 . On the other hand, the base‐catalyzed cyclization of the hydroxy diester 22 , which is the ‘O‐analogue' of 20 , yielded neither the seven‐membered dilactone, nor the 14‐membered tetralactone, but only the known trimer 23 and tetramer 24 of 2,2‐dimethylpropano‐3‐lactone (cf. Scheme 7).  相似文献   

2.
The 1,2‐dithiolosultam derivative 14 was obtained from the (α‐bromoalkylidene)propenesultam derivative 9 (Scheme 1). Regioselective cleavage of the two ester groups (→ 1b or 2b ) allowed the preparation of derivatives with different substituents at C(3) in the dithiole ring (see 27 and 28 ) as well as at C(6) in the isothiazole ring (see 17 – 21 ; Scheme 2). Curtius rearrangement of the 6‐carbonyl azide 21 in Ac2O afforded the 6‐acetamide 22 , and saponification and decarboxylation of the latter yielded ‘sulfothiolutin’ ( 30 ). Hydride reductions of two of the bicyclic sultams resulted in ring opening of the sultam ring and loss of the sulfonyl group. Thus the reduction of the dithiolosultam derivative 14 yielded the alkylidenethiotetronic acid derivative 33 (tetronic acid=furan‐2,4(3H,4H)‐dione), and the lactam‐sultam derivative 10 gave the alkylidenetetramic acid derivative 35 (tetramic acid=1,5‐dihydro‐4‐hydroxy‐2H‐pyrrol‐2‐one) (Scheme 3). Some of the new compounds ( 14, 22, 26 , and 30 ) exhibited antimycobacterial activity. The oxidative addition of 1 equiv. of [Pt(η2‐C2H4)L2] ( 36a , L=PPh3; 36b , L=1/2 dppf; 36c , L=1/2 (R,R)‐diop) into the S? S bond of 14 led to the cis‐(dithiolato)platinum(II) complexes 37a – c . (dppf=1,1′‐bis(diphenylphosphino)ferrocene; (R,R)‐diop={[(4R,5R)‐2,2‐demithyl‐1,3‐dioxolane‐4,5‐diyl]bis(methylene)}bis[diphenylphosphine]).  相似文献   

3.
Cob(I) alamin ( 1(I) )-catalyzed reduction of the aldehyde 2 led to the two crystalline cyclopropanols 3 and 4 (see Scheme 2). The protolytic ring-opening starting from 3 produced the saturated aldehydes 6 and 7;8 was formed in traces only (see Scheme 3). The protolysis starting from 3 led, therefore, mainly to retention of configuration at the spiro C-atom ( 7 ); ring-opening with inversion was observed in traces only ( 8 ). Starting from 4 , the protolysis produced 9 and 7 ; the absence of 8 showed this protolysis to proceed 9 and 7 ; the absence of 8 showed this protolysis to proceed exclusively with inversion of configuration at the spiro center. Of the p-bromobenzoate 5 (cf. Scheme 2) the structure has been determined by X-ray analysis.  相似文献   

4.
A convergent total synthesis of 13‐hydroxyisocyclocelabenzine was developed. (3S)‐Methyl 3‐amino‐3‐phenylpropanoate ( 4 ) was used as the chiral building block. The 3,4‐dihydro‐4‐hydroxyisoquinolin‐1(2H)‐one derivative ( 5 ), the key fragment for the total synthesis, was prepared by a novel base‐catalyzed lactone‐lactam ring enlargement (Scheme 3). The resulting target C(13) epimers 3a / 3b from macrocyclization (Scheme 4) were separated by repeated flash chromatography. The absolute configuration of the synthetic alkaloid was determined by an X‐ray crystal‐structure analysis, which enabled us to determine the absolute configuration (9S,13R) for natural 3a with positive [α]D.  相似文献   

5.
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   

6.
The synthesis of 4,6,8-trimethyl-1-[(E)-4-R-styryl]azulenes 5 (R=H, MeO, Cl) has been performed by Wittig reaction of 4,6,8-trimethylazulene-1-carbaldehyde ( 1 ) and the corresponding 4-(R-benzyl)(triphenyl)phosphonium chlorides 4 in the presence of EtONa/EtOH in boiling toluene (see Table 1). In the same way, guaiazulene-3-carbaldehyde ( 2 ) as well as dihydrolactaroviolin ( 3 ) yielded with 4a the corresponding styrylazulenes 6 and 7 , respectively (see Table 1). It has been found that 1 and 4b yield, in competition to the Wittig reaction, alkylation products, namely 8 and 9 , respectively (cf. Scheme 1). The reaction of 4,6,8-trimethylazulene ( 10 ) with 4b in toluene showed that azulenes can, indeed, be easily alkylated with the phosphonium salt 4b . 4,6,8-Trimethylazulene-2-carbaldehyde ( 12 ) has been synthesized from the corresponding carboxylate 15 by a reduction (LiAlH4) and dehydrogenation (MnO2) sequence (see Scheme 2). The Swern oxidation of the intermediate 2-(hydroxymethyl)azulene 16 yielded only 1,3-dichloroazulene derivatives (cf. Scheme 2). The Wittig reaction of 12 with 4a and 4b in the presence of EtONa/EtOH in toluene yielded the expected 2-styryl derivatives 19a and 19b , respectively (see Scheme 3). Again, the yield of 19b was reduced by a competing alkylation reaction of 19b with 4b which led to the formation of the 1-benzylated product 20 (see Scheme 3). The ‘anil synthesis’ of guaiazulene ( 21 ) and the 4-R-benzanils 22 (R=H, MeO, Cl, Me2N) proceeded smoothyl under standard conditions (powered KOH in DMF) to yield the corresponding 4-[(E)-styryl]azulene derivatives 23 (see Table 4). In minor amounts, bis(azulen-4-yl) compounds of type 24 and 25 were also formed (see Table 4). The ‘anil reaction’ of 21 and 4-NO2C6H4CH=NC6H5 ( 22e ) in DMF yielded no corresponding styrylazulene derivative 23e . Instead, (E)-1,2-bis(7-isopropyl-1-methylazulen-4-yl)ethene ( 27 ) was formed (see Scheme 4). The reaction of 4,6,8-trimethylazulene ( 10 ) and benzanil ( 22a ) in the presence of KOH in DMF yielded the benzanil adducts 28 to 31 (cf. Scheme 5). Their direct base-catalyzed transformation into the corresponding styryl-substituted azulenes could not be realized (cf. Scheme 6). However, the transformation succeeded smoothly with KOH in boiling EtOH after N-methylation (cf. Scheme 6).  相似文献   

7.
Heptaleno[1,2-c]furan-6-carbaldehydes such as 8 or their thiocarbaldehyde or iminomethyl derivatives easily undergo thermal cyclization, followed by a [1,5]- H shift, to give the corresponding heptalenodifurans, thienoheptalenofurans, as well as furoheptalenopyrroles (cf. Schemes 2 and 3). Generation of the 6-acetyl derivative of 8 from the corresponding secondary alcohol 15 with 1-hydroxy-1,2-benziodoxol-3(1H)-one 1-oxide (IBX) at 0° (cf. Scheme 4) leads directly to the formation of the cyclization product 16 which, upon standing at room temperature, undergoes the [1,5]-sigmatropic H-shift to the final difuran 17 . 1-Formylheptalene-4,5-dicarboxylates such as 9 can also be cyclized thermally, followed by the [1,5]-H shift, to the corresponding 8H-heptaleno[1,10-bc]furan-5,6-dicarboxylate 11 . On thiation with Lawesson′s reagent, 9 yields directly the corresponding heptalenothiophene 13 (cf. Scheme 3).  相似文献   

8.
Azulene-1-carbaldehydes which have Me substituents at C(3) and C(8) and no substituent at C(6) react with excess dimethyl acetylenedicarboxylate (ADM) in decalin at 200° to yield exclusively the Diels-Alder adduct at the seven-membered ring (cf. Scheme 3). The corresponding 1-carboxylates behave similarly (Scheme 4). Azulene-1-carbaldehydes which possess no Me substituent at C(8) (e.g. 11 , 12 in Scheme 2) gave no defined products when heated with ADM in decalin. On the other hand, Me substitutents at C(2) may also assist the thermal addition of ADM at the seven-membered ring of azulene-1-carbaldehydes (Scheme 6). However, in these cases the primary tricyclic adducts react with a second molecule of ADM to yield corresponding tetracyclic compounds. The new tricyclic aldehydes 16 and 17 which were obtained in up to 50% yield (Scheme 3) could quantitatively be decarbonylated with [RhCl(PPh3)3] in toluene at 140° to yield a thermally equilibrated mixture of four tricycles (Scheme 8). It was found that the thermal isomerization of these tricycles occur at temperatures as low as 0° and that at temperatures > 40° the thermal equilibrium between the four tricycles is rapidly established via [1,5]-C shifts. The establishment of the equilibrium makes the existence of two further tricycles necessary (cf. Scheme 8). However, in the temperature range of up to 85° these two further tricycles could not be detected by 1H-NMR. When heated in the presence of excess ADM in decalin at 180°, the ‘missing’ tricyclic forms could be evidenced by their tetracyclic trapping products ‘anti’- 45 and ‘anti’- 48 , respectively (Scheme 9).  相似文献   

9.
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   

10.
Palladium‐catalyzed tandem cyclization–cross‐coupling reaction of indolylborate 2 and vinyl bromide 5 was successfully applied in a short formal synthesis of olivacine. The reaction of 2 with 5 in the presence of Pd(OAc)2 readily afforded three kinds of products, triene derivative 6 and vinylindole derivative 7 , along with a small amount of the piperidine derivative 8 (Scheme 2). On the other hand, the reactions of 2 with bromide 10 or 15 were also examined (Schemes 4 and 5), and their outcome markedly depended on the relative ease of ring closure as a function of ring size. Irradiation of 6 with a high‐pressure mercury lamp (→ 9 ; Scheme 2), followed by removal of the N‐[(benzyloxy)carbonyl] group and subsequent oxidation afforded, after deprotection, pyridocarbazole 23 , and the conversion of 23 to olivacine is known (Scheme 6).  相似文献   

11.
Reaction Products from 3-Dimethylamino-2,2-dimethyl-2H-azirine and Phthalohydrazide or Maleohydrazide 3-Dimethylamino-2, 2-dimethyl-2H-azirine (1) reacts in dimethylformamide at room temperature with the six-membered cyclic hydrazides 2, 3-dihydrophthalazin-1, 4-dione (2) and 1, 2-dihydropyridazin-3, 6-dione (15) to give the zwitterionic compounds 3 and 16 , respectively (Scheme 1 and 7). The mechanism of these reactions is outlined in Scheme 1 for compound 3 (cf. also Scheme 8). The first steps are thought to be similar to the known reactions of 1 with the NH-acidic compounds saccharin and phthalimide (cf. [1]). Instead of ring expansion to the nine-membered heterocycle i (X=CONH, Scheme 8), a proton transfer followed by the loss of water gives 3 (Scheme 1). The structure of the zwitterionic compounds 3 and 16 is deduced from spectral data and the reactions of these compounds (see Schemes 2, 3, 4, 6 and 7). Methylation of 3 yields the iodide 4 , which is hydrolysed easily to the 2-imidazolin-5-one derivative 5 (Scheme 2). Hydrolysis of 3 under basic conditions leads to the amide 6 , which undergoes cyclization to 7 at 220–230° (Scheme 3). The analogous cyclization has been realized under acidic conditions in the case of 17 (Scheme 7). Catalytic reduction of 3 yields the tertiary amine 14 (Scheme 6), whereas the reduction with sodium borohydride leads to a mixture of 14 and the 2-imidazoline derivative 13 . The alcohol 11 , corresponding to the amine 14 , is obtained by sodium borohydride reduction of the 2-imidazolin-5-one 7 or of the amide 6 (Scheme 3). This remarkably easy reaction of 7 shows the unusual electrophilicity of the lactamcarbonyl group in this compound. The reduction of 6 to 11 is understandable only by neighbouring group participation of N (2′) in the dihydrophthalazine residue.  相似文献   

12.
Lactones from Epoxycyclodecanone Derivatives by Ring Enlargement Involving Fragmentation Reactions A stereospecific ring-enlargement reaction of alkyl esters of 2,3-epoxy-1-(3-hydroxypropyl)-10-oxocyclo-decanecarboxylic-acid derivatives is described, involving Grob fragmentation of in situ formed hemiacetals. The assignment of the relative configuration of the starting materials was accomplished on the basis of 1H-NMR data. The rearrangement of the epoxides 9 and 10 (with cis-orientation of the ester group and the epoxide ring, Scheme 1) gives the lactone 15 as the single and as the major product, respectively, with (Z)-configuration of the newly formed C?C bond. A concerted reaction mechanism is assumed. The formation of a small amount of 12 from 10 is probably due to a competitive two-step carbanion pathway. The reaction of the diastereoisomers 7 and 8 leads to the lactones 11 and 12 , respectively, as the only ring-enlargement products (Scheme 1), with (E)-configuration of the newly formed C?C bond. On the basis of our results, we cannot distinguish in this case between a concerted and a two-step carbanion mechanism. This type of reaction takes place only in the presence of an ester group; no ring enlargement was detected in case of compound 20 (Scheme 3), which is the de(alkoxycarbonyl) derivative of 9 . The eliminative opening of the epoxide ring in the epoxylactone 17 affords 11 as the single product (Scheme 2). A carbanion mechanism was assumed for this reaction.  相似文献   

13.
Irradiation of 4-Allylated 2,6-Dimethylanilines in Methanol 4-Allyl-, 4-(1′-methylallyl)-, 4-(2′-butenyl)-, and 4-(1′,1′-dimethylallyl)-2,6-dimethylaniline ( 14–17 ; cf. Scheme 3) were obtained by the acid catalysed, thermal rearrangement of the corresponding N-allylated anilines in good yields. Aniline 14 , when irradiated with a high pressure mercury lamp through quartz in methanol, yielded as main product 4-(2′-methoxypropyl)-2,6-dimethylaniline ( 22 ; cf. Scheme 4) and, in addition, 2,6-dimethyl-4-propylaniline ( 18 ) and 4-cyclopropyl-2,6-dimethylaniline ( 23 ). The analogous products, namely erythro- and threo-4-(2′-methoxy-1′-methylpropyl)-2,6-dimethylaniline (erythro- and threo- 24 ), 2,6-dimethyl-4-(1′-methylpropyl)aniline ( 19 ), trans- and cis-2,6-dimethyl-4-(2′-methylcyclopropyl)aniline (trans- and cis- 25 ), as well as small amounts of 4-ethyl-2,6-dimethylaniline ( 26 ), were formed by irradiation of 15 in methanol (cf. Scheme 5). When this photoreaction was carried out in O-deuteriomethanol, erythro- and threo- 24 showed an up-take of one deuterium atom in the side chain. The mass spectra of erythro- and threo- 24 revealed that in 50% of the molecules the deuterium was located at the methyl group at C(1′) and in the other 50% at the methyl group at C(2′) (cf. Scheme 6). This is a good indication that the methanol addition products arise from methanolysis of intermediate spiro[2.5]octa-4,7-dien-6-imines (cf. Scheme 7). This assumption is further supported by the photoreaction of 17 in methanol (cf. Scheme 8) which led to the formation of 4-(2′-methoxy-1′,2′-dimethylpropyl)-2,6-dimethylaniline ( 28 ) as main product. The occurrence of a rearranged side chain in 28 can again be explained by the intervention of a spirodienimine 31 (cf. Scheme 9). In comparison with 14, 15 and 17 , the 2′-butenylaniline 16 reacted only sluggishly on irradiation in methanol (cf. Scheme 10). It is suggested that all photoproducts - except for the cyclopropyl derivatives which are formed presumably via a triplet di-π-methane rearrangement - arise from an intramolecular singlet electron-donor-acceptor complex between the aniline and ethylene chromophor of the side chain. Protonation of this complex at C(3′) or C(2′) will lead to diradicals (e.g. 33 and 34 , respectively, in Scheme 11). The diradicals of type 33 undergo ring closure to the corresponding spirodienimine intermediates (e.g. 31 ) whereas the diradicals of type 34 take up two hydrogen atoms to yield the photo-hydrogenated compounds (e.g. 21 ) or undergo to a minor extent fragmentation to side chain degraded products (e.g. 30 ; see also footnote 7).–Irradiation of 4-ally-2,6-dimethylaniline ( 14 ) in benzene or cyclohexane yielded the corresponding azo compound 38 (cf. Scheme 12), whereas its N,N-dimethyl derivative 41 was transformed into the cyclopropyl derivative 42 . The allyl moiety in 14 is not necessary for the formation of azo compounds since 2,4,6-trimethylaniline ( 39 ) exhibited the same type of photoreaction in benzene solution.  相似文献   

14.
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   

15.
The novel 8,14‐secoursatriene derivative 6 was synthesized starting from ursolic acid ( 1 ) via methyl esterification of the 17‐carboxylic acid group and benzoylation of the 3‐hydroxy group (→ 2 ; Scheme 1), ozone oxidation of the C(12)?C(13) bond (→ 3 ), dehydrogenation with Br2/HBr (→ 4 ), enol acetylation of the resulting carbonyl group (→ 5 ; Scheme 2), and ring‐C opening with the aid of UV light (→ 6 ). Ring‐C‐opened dienone derivative 7 of ursolic acid was also obtained via selective hydrolysis of 6 (Scheme 2). Both compounds 6 and 7 are key intermediates for the preparation of chiral decalin synthons from ursolic acid.  相似文献   

16.
The chemistry of two bicyclic endoperoxides, obtained by photooxygenation of 2,3‐benzotropone (=5H‐benzocyclohepten‐5‐one; 5 ) and of its ethyl carboxylate derivative 15 , was investigated with the aim of synthesizing the respective benzotropolone derivatives. The reaction of the endoperoxide 10 derived from 5 with thiourea gave the desired benzotropolone, i.e., 6‐hydroxy‐5H‐benzocyclohepten‐5‐one ( 11 ), in high yield (Scheme 1). On the other hand, the endoperoxide 16 derived from the ethyl carboxylate derivative 15 underwent an unprecedented transformation yielding mainly the ring‐contracted lactones 28 and 29 besides the expected substituted benzotropolone derivative 27 (Scheme 5). However, the thermolysis reaction of the same endoperoxide 16 resulted in the formation of four rearranged compounds with different skeletons (Scheme 3). The formation mechanism of all products is discussed (Schemes 4 and 6).  相似文献   

17.
Starting from iodoalcohol 9 , the monoprotected dialdehyde 5 was synthesized (Scheme 2) and converted to 17 by reaction with oxo-phosphonate 15 (Scheme 3). The latter was prepared from 13 . Cyclisation of 17 to the target compound 18 failed. Also the attachment of thiol 22 to lactone 19 was unsatisfactory (Scheme 4). Therefore, the building blocks 28 and 29 were synthesized using diene 33 and diester 30 as starting material for 28 and 9 for 29 (Scheme 5 and 6). Hydroxy acid 28 was converted into formyl-ester 46 (Scheme 7). However, the condensation of its derivatives 48 and 49 with ‘Umpolung’ of the carbonyl reactivity was unsuccessful, probably due to steric hindrance.  相似文献   

18.
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   

19.
The aldol reaction of the C(3) carbanion of 7‐chloro‐1,3‐dihydro‐1‐methyl‐5‐phenyl‐2H‐1,4‐benzodiazepin‐2‐one ( 2 ) with a series of aromatic and aliphatic aldehydes at −78° afforded threo/erythro diastereoisomers 3 – 16 of 7‐chloro‐1,3‐dihydro‐3‐(hydroxymethyl)‐1‐methyl‐5‐phenyl‐2H‐1,4‐benzodiazepinones, substituted at the C(3) side chain, in a ratio from 55 : 45 to 94 : 6 (Scheme 1). Lewis acids exhibited limited effect on the syn/anti diastereoselectivity of this reaction, and kinetic control of the reaction was confirmed. 1H‐NMR Data suggested the assignment of the threo relative configuration to the first‐eluted diastereoisomers 3 , 5 , 7 , and 9 on reversed‐phase HPLC, and the erythro configuration to the second‐eluted counterparts 4 , 6 , 8 , and 10 , respectively. The structures and relative configurations threo and erythro of the diastereoisomers 5 and 6 , respectively, were established by single‐crystal X‐ray analysis, confirming the assignment based on the 1H‐NMR data. A tentative mechanistic explanation of the diastereoselectivity invokes the enolate anion of 1,3‐dihydro‐2H‐1,4‐benzodiazepin‐2‐one as the reactive species (Scheme 2). Acid‐catalyzed hydrolytic ring opening of 3 afforded threoβ‐hydroxy‐phenylalanine 17 , whereas from 4 , the N‐(benzyloxy)carbonyl derivative 18 of erythro‐β‐hydroxy‐phenylalanine was obtained (Scheme 3); in both cases, neither elimination of H2O from the C(3)−CHOH moiety nor epimerization at C(3) were observed. This result opens a new pathway to various configurationally uniform α‐amino‐β‐hydroxy carboxylic acids and their congeners of biological importance.  相似文献   

20.
The reactions of the enolizable thioketone (1R,4R)‐thiocamphor (=(1R,4R)‐1,7,7‐trimethylbicyclo[2.2.1]heptane‐2‐thione; 1 ) with (S)‐2‐methyloxirane ( 2 ) in the presence of a Lewis acid such as SnCl4 or SiO2 in anhydrous CH2Cl2 led to two diastereoisomeric spirocyclic 1,3‐oxathiolanes 3 and 4 with the Me group at C(5′), as well as the isomeric β‐hydroxy thioether 5 (Scheme 2). The analogous reactions of 1 with (RS)‐, (R)‐, and (S)‐2‐phenyloxirane ( 7 ) yielded two isomeric spirocyclic 1,3‐oxathiolanes 8 and 9 with Ph at C(4′), an additional isomer 13 bearing the Ph group at C(5′), and three isomeric β‐hydroxy thioethers 10, 11 , and 12 (Scheme 4). In the presence of HCl, the β‐hydroxy thioethers 5, 10, 11 , and 12 isomerized to the corresponding 1,3‐oxathiolanes 3 and 4 (Scheme 3), and 8, 9 , and 13 , respectively (Scheme 5). Under similar conditions, an epimerization of 3, 8 , and 9 occurred to yield the corresponding diastereoisomers 4, 14 , and 15 , respectively (Schemes 3 and 6). The structures of 9 and 15 were confirmed by X‐ray crystallography (Figs. 1 and 2). These results show that the Lewis acid‐catalyzed addition of oxiranes to enolizable thioketones proceeds with high regio‐ and stereoselectivity via an Sn 2‐type mechanism.  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