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 296 毫秒
1.
In order to trap ‘thiocarbonyl-aminides’ A , formed as intermediates in the reaction of thiocarbonyl compounds with phenyl azide, a mixture of 2,2,4,4-tetramethyl-3-thioxocyclobutanone ( 1 ), phenyl azide, and fumarodinitrile ( 8 ) was heated to 80° until evolution of N2 ceased. Two interception products of the ‘thiocarbonylaminide’ A (Ar?Ph) were formed: the known 1,4,2-dithiazolidine 3 (cf. Scheme 1) and the new 1,2-thiazolidine 12 (Scheme 2). The structure of the latter was established by X-ray crystallography (Fig.1). The analogous ‘three-component reaction’ with dimethyl fumarate ( 9 ) yielded, instead of 8 , in addition to the known interception products 3 and 6 (Scheme 1), two unexpected products 15 and 16 (Scheme 3), of which the structures were elucidated by X-ray crystallography (Fig.2). Their formation is rationalized by a primary [2 + 3] cycloaddition of diazo compound 18 with 1 to give 19 , followed by a cascade of further reactions (Scheme 4).  相似文献   

2.
The racemic spirosesquiterpenes β-acorenol ( 1 ), β-acoradiene ( 2 ), acorenone-B ( 3 ) and acorenone ( 4 ) (Scheme 2) have been synthesized in a simple, flexible and highly stereoselective manner from the ester 5 . The key step (Schemes 3 and 4), an intramolecular thermal ene reaction of the 1,6-diene 6 , proceeded with 100% endo-selectivity to give the separable and interconvertible epimers 7a and 7b . Transformation of the ‘trans’-ester 7a to (±)- 1 and (±)- 2 via the enone 9 (Scheme 5) involved either a thermal retro-ene reaction 10 → 12 or, alternatively, an acid-catalysed elimination 11 → 13 + 14 followed by conversion to the 2-propanols 16 and 17 and their reduction with sodium in ammonia into 1 which was then dehydrated to 2 . The conversion of the ‘cis’-ester 7b to either 3 (Scheme 6) or 4 (Scheme 7) was accomplished by transforming firstly the carbethoxy group to an isopropyl group via 7b → 18 → 19 → 20 , oxidation of 20 to 21 , then alkylative 1,2-enone transposition 21 → 22 → 23 → 3 . By regioselective hydroboration and oxidation, the same precursor 20 gave a single ketone 25 which was subjected to the regioselective sulfenylation-alkylation-desulfenylation sequence 25 → 26 → 27 → 4 .  相似文献   

3.
1,5-Dipolar Electrocyclization of Acyl-Substituted ‘Thiocarbonyl-ylides’ to 1,3-Oxathioles The reaction of α-diazoketones 15a, b with 4,4-disubstituted 1,3-thiazole-5(4H)-thiones 6 (Scheme 3), adamantanethione ( 17 ), 2,2,4,4-tetramethyl-3-thioxocyclobutanone ( 19 ; Scheme 4), and thiobenzophenone ( 22 ; Scheme 5), respectively, at 50–90° gave the corresponding 1,3-oxathiole derivatives as the sole products in high yields. This reaction opens a convenient access to this type of five-membered heterocycles. The structures of three of the products, namely 16c, 16f , and 20b , were established by X-ray crystallography. The key-step of the proposed reaction mechanism is a 1,5-dipolar electrocyclization of an acyl-substituted ‘thiocarbonyl-ylide’ (cf. Scheme 6). The analogous reaction of 15a, b with 9H-xanthen-9-thione ( 24a ) and 9H-thioxanthen-9-thione ( 24b ) yielded α,β-unsaturated ketones of type 25 (Scheme 5). The structures of 25a and 25c were also established by X-ray crystallography. The formation of 25 proceeds via a 1,3-dipolar electrocyclization to a thiirane intermediate (Scheme 6) and desulfurization. From the reaction of 15a with 24b in THF at 50°, the intermediate 26 (Scheme 5) was isolated. In the crude mixtures of the reactions of 15a with 17 and 19 , a minor product containing a CHO group was observed by IR and NMR spectroscopy. In the case of 19 , this side product could be isolated and was characterized by X-ray crystallography to be 21 (Scheme 4). It was shown that 21 is formed – in relatively low yield – from 20a . Formally, the transformation is an oxidative cleavage of the C?C bond, but the reaction mechanism is still not known.  相似文献   

4.
The thermal reaction of 7-isopropyl-1,3,4-trimethylazulene (3-methylguaiazulene; 2 ) with excess dimethyl acetylenedicarboxylate (ADM) in decalin at 200° leads to the formation of the corresponding heptalene- ( 5a/5b and 6a/6b ; cf. Scheme 3) and azulene-1,2-dicarboxylates ( 7 and 8 , respectively). Together with small amounts of a corresponding tetracyclic compound (‘anti’- 13 ) these compounds are obtained via rearrangement (→ 5a/5b and 6a/6b ), retro-Diels-Alder reaction (→ 7 and 8 ), and Diels-Alder reaction with ADM (→ ‘anti’- 13 ) from the two primary tricyclic intermediates ( 14 and 15 ; cf. Scheme 5) which are formed by site-selective addition of ADM to the five-membered ring of 2 . In a competing Diels-Alder reaction, ADM is also added to the seven-membered ring of 2 , leading to the formation of the tricyclic compounds 9 and 10 and of the Diels-Alder adducts ‘anti’- 11 and ‘anti’- 12 , respectively of 9 and of a third tricyclic intermediate 16 which is at 200° in thermal equilibrium with 9 and 10 (cf. Scheme 6). The heptalenedicarboxylates 5a and 5b as well as 6a and 6b are interconverting slowly already at ambient temperature (Scheme 4). The thermal reaction of guaiazulene ( 1 ) with excess ADM in decalin at 190° leads alongside with the known heptalene- ( 3a ) and azulene-1,2-dicarboxylates ( 4 ; cf. Schemes 2 and 7) to the formation of six tetracyclic compounds ‘anti’- 17 to ‘anti’- 21 as well as ‘syn’- 19 and small amounts of a 4:1 mixture of the tricyclic tetracarboxylates 22 and 23 . The structure of the tetracyclic compounds can be traced back by a retro-Diels-Alder reaction to the corresponding structures of tricyclic compounds ( 24--29 ; cf. Scheme 8) which are thermally interconverting by [1,5]-C shifts at 190°. The tricyclic tetracarboxylates 22 and 23 , which are slowly equilibrating already at ambient temperature, are formed by thermal addition of ADM to the seven-membered ring of dimethyl 5-isopropyl-3,8-dimethylazulene-1,2-dicarboxylate ( 7 ; cf. Scheme 10). Azulene 7 which is electronically deactivated by the two MeOCO groups at C(1) and C(2) shows no more thermal reactivity in the presence of ADM at the five-membered ring (cf. Scheme 11). The tricyclic tetracarboxylates 22 and 23 react with excess ADM at 200° in a slow Diels-Alder reaction to form the tetracyclic hexacarboxylates 32 , ‘anti’- 33 , and ‘anti’- 34 (cf. Schemes 10–12 as well as Scheme 13). A structural correlation of the tri- and tetracyclic compounds is only feasible if thermal equilibration via [1,5]-C shifts between all six possible tricyclic tetracarboxylates ( 22, 23 , and 35–38 ; cf. Scheme 13) is assumed. The tetracyclic hexacarboxylates 32 , ‘anti’- 33 , and ‘anti’- 34 seem to arise from the most strained tricyclic intermediates ( 36–38 ) by the Diels-Alder reaction with ADM.  相似文献   

5.
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   

6.
Formation of 1,2,4-Trithiolanes in Three-Component Reactions of Phenyl Azide, Aromatic Thiones, and 2,2,4,4-Tetramethylcyclobutanethiones: A Sulfur-Transfer Reaction to ‘Thiocarbonyl-thiolates’ ((Alkylidenesulfonio)-thiolates) as Reactive Intermediates The reaction of PhN3 and aromatic thioketones 18 (two-component reaction) at 80° yields only the corresponding imines 22 , S, and N2. Under similar conditions, in the presence of sterically crowded 2,2,4,4-tetramethyl-cyclobutanethiones 19 (three-component reaction), 1,2,4-trithiolanes of type 20 are formed in good yields in addition to imines 22 (Scheme 4). In case of 19a and 19c (X = CO, CS), the symmetrical trithiolanes 21a and 21b , respectively, are also isolated. With 4,4-dimethyl-2-phenyl-1,3-thiazole-5(4H)-thione ( 24 ) instead of aromatic thioketone 18 , imine 25 , trithiolane 21a , and 1,4,2-dithiazolidine 26 are formed (Scheme 5). A reaction mechanism for the formation of 1,2,4-trithiolanes 20 and 21 , including an S-transfer to generate ‘thiocarbonyl-thiolates’ 2b and/or 2c and 1,3-dipolar cycloaddition with a thioketone, is proposed in Scheme 7.  相似文献   

7.
It is shown that, upon irradiation in CDCl3 solution, 5,6,8,10-tetramethylheptalene-1,2-dicarboxylic anhydride ( 6 ) rearranges to its double-bond-shift (DBS) isomer 7 in an equilibrium reaction (Scheme 2). The isomer 7 is DBS stable at ?50°. At ca. 30°, a thermal equilibrium with 97.8% of 6 and 2.2% of 7 is rapidly established. Similarly, the ‘ortho’-anhydrides 9 and 11 (Schemes 4 and 5) can be rearranged to their corresponding DBS isomers 12 and 13 , respectively. Whereas 12 is DBS stable at 30° (at 100° in tetralin, 94.0% of 9 are in equilibrium with 6.0% of 12 ), the i-Pr-substituted isomer 13 is already at 30° in thermal equilibrium with 11 leading to 98.7% of 11 and 1.3% of 13 . It is shown by rearrangement of diasteroisomeric ‘ortho’-anhydrides of known relative and absolute configuration (Scheme 6) that the DBS in such five-ring-annelated heptalenes occurs with retention of the configuration of the heptalene skeleton as already established for other heptalene compounds. It is found that the DBS process may also take place under acid catalysis (e.g. HCl/CH3OH), thus yielding 9 from 12 (Scheme 9). The ‘ortho’-anhydrides 21 and 23 (Scheme 10) which are isomeric with 9 and 11 (Scheme 3) undergo rapid DBS' already at room temperature. The thermal equilibrium 21?22 consists of 18% of 21 and 82% of 22 at 30° and that of 23?24 of 17% of 23 and 83% of 24 at ?30°. From these equilibrium mixtures, the pure DBS isomer 22 can be obtained by crystallization. Again, these rapid DBS' occur with retention of configuration of the heptalene skeleton (Fig. 4).  相似文献   

8.
Dimethyl diazomalonate ( 4 ) and thiobenzophenone ( 2a ) do not react in toluene even after warming to 50°. After addition of catalytic amounts of Rh2(OAc)4, a smooth reaction under N2 evolution afforded a mixture of thiiranedicarboxylate 5 and (diphenylmethylidene)malonate 6 (Scheme 2). A reaction mechanism via an intermediate ‘thiocarbonyl ylide’ 7 , formed by the addition of the carbenoid species 8 to the S-atom of 2a , is plausible. Similar reactions were carried out with 9H-xanthene-9-thione ( 2b ), 9H-thioxanthene-9-thione ( 2c , Scheme 4), and 1,3-thiazole-5(4H)-thione 18 (Scheme 6). In the cases of 2b and 2c , spirocyclic 1,3-dithiolanetetracarboxylates 14a and 14b , respectively, were obtained as the third product. Reaction mechanisms for their formation are proposed in Scheme 5: S-transfer from intermediate thiirane 12 to the carbenoid species yielded thioxomalonate 15 which underwent a 1,3-dipolar cycloaddition with ‘thiocarbonyl ylide’ 16 . An alternative is the formation of ‘thiocarbonyl ylide’ 17 via carbene addition to 15 , followed by 1,3-dipolar cycloaddition with 2b and 2c , respectively.  相似文献   

9.
Regioselective 1,3-Dipolar Cycloadditions of a ‘Thiocarbonyl-methanide’ ((Alkylidenesulfonio)methanide) with Aromatic Sulfines Reaction of the spirocyclic 2,5-dihydro-1,3,4-thiadiazole 7 and thiobenzophenone S-oxide ( 6a ) in THF at 45° yielded the spirocyclic 1,3-dithiolane 1-oxide 8 , thiirane 9 , and the diazane derivative 10 in a ratio of 61:15:23 (Scheme 2). The formation of 8 is rationalized by a 1,3-dipolar cycloaddition of ‘thiocarbonyl-methanide’ 1 , generated from 7 by thermal elimination of N2, and the C?S bond of sulfine 6a . Cyclization of intermediate 1 leads to thiirane 9 . Under the same conditions, 7 and adamantane-2-thione S-oxide ( 6b ) or 2,2,4,4-tetramethyl-3-thioxocyclobutanone S-oxide ( 4 ) reacted to give only 9 and 10 but no cycloadduct of type 8 (Scheme 4). With the aim to favor the formation of 8 , a mixture of 6a and 1.1 equiv. of 7 was heated to 45° without any solvent in a sealed tube. The ratio of products was only slightly different from that of the thermolysis in THF. An analogous experiment with 7 and 9H-fluorene-9-thione S-oxide ( 6c ) yielded cycloadduct 13 and 9 (Scheme 5). It is most interesting that the 1,3-dipolar cycloadditions of 1 and the sulfines 6a and 6c proceeded with different regioselectivity. A reaction mechanism for the unexpected formation of 10 is proposed in Scheme 7. The key step is the base-catalyzed ring opening of 7 and the nucleophilic addition of the thereby formed thiolate 21 onto the sulfonium ion 19 .  相似文献   

10.
The 3‐allyl‐2‐methylquinazolin‐4(3H)‐one ( 1 ), a model functionalized terminal olefin, was submitted to hydroformylation and reductive amination under optimized reaction conditions. The catalytic carbonylation of 1 in the presence of Rh catalysts complexed with phosphorus ligands under different reaction conditions afforded a mixture of 2‐methyl‐4‐oxoquinazoline‐3(4H)‐butanal ( 2 ) and α,2‐dimethyl‐4‐oxoquinazoline‐3(4H)‐propanal ( 3 ) as products of ‘linear’ and ‘branched’ hydroformylation, respectively (Scheme 2). The hydroaminomethylation of quinazolinone 1 with arylhydrazine derivatives gave the expected mixture of [(arylhydrazinyl)alkyl]quinazolinones 5 and 6 , besides a small amount of 2 and 3 (Scheme 3). The tandem hydroformylation/reductive amination reaction of 1 with different amines gave the quinazolinone derivatives 7 – 10 . Compound 10 was used to prepare the chalcones 11a and 11b and pyrazoloquinazolinones 12a and 12b (Scheme 4).  相似文献   

11.
1,3-Dipoles with a Central S-Atom from the Reaction of Azides and Thiocarbonyl Compounds: An Unexpected MeS Migration in the Trapping Product of a ‘Thiocarbonyl-aminide’ with Methyl Dithiobenzoate Reaction of PhN3 with O-methyl thiobenzoate ( 11a ) and thioacetate ( 11c ) as well as with the dithio esters 11b,d at 80° yields the corresponding imidates and thioimidates 12 (Scheme 3). The formation of 12 is rationalized by a 1,3-dipolar cycloaddition of the azide and the C?S group followed by successive elimination of N2 and S. In the three-component reaction of 11b , PhN3, and the sterically crowded thioketone 1a , 1,2,4-trithiolane 13a and 1,4,2-dithiazolidine 3a are formed in addition to 12b (Scheme 4). The heterocycles 13a and 3a are trapping products of 1a and ‘thiocarbonyl-thiolate’ 5a and ‘thiocarbonyl-aminide’ 2a (Ar?Ph), respectively (Scheme 6). These 1,3-dipoles are formed as reactive intermediates. Surprisingly, in the presence of catalytic amounts of acids, the major product is the (methyldithio)cyclobutyl thioimidate of type 14 (Scheme 5), formed by an acid-catalyzed MeS migration in dithiazolidine 17 . A reaction mechanism is proposed in Scheme 7.  相似文献   

12.
A new approach to ketose-derived nucteosides is described. It is based upon a chain elongation of 1-deoxy-1-nitroaldoses, followed by activation of the nitro group as a leaving group, and introduction of a pyrimidine or purine base. Thus, the nitroaldose 7 was prepared from 3 by pivaloylation (→ 4 ), synthesis of the anomeric nitrones 5/6 , and ozonolysis of 6 (Scheme 1). Partial hydrolysis of 4 yielded 8/9 , which were characterized as the acetates 10/11 and transformed into the nitrones 12/13 . Ozonolysis of 12/13 gave 14/15 , which were acetylated to 16/17 . Henry reaction of 7 lead to 19 and 20 , which were acetylated to 21 and 22 (Scheme 2). Michael addition of 7 to acrylonitrile and to methyl propynoate yielded the anomers 23/24 and 25/26 , respectively. Similar reactions of 16/17 were prevented by a facile β-elimination. Therefore, the nitrodiol 15 was transformed into the orthoesters 27 and then, by Henry reaction, partial hydrolysis, and acetylation, into 28 and 29 (Scheme 2). The structure of 19 was established by X-ray analysis. It was the major product of the kinetically controlled Henry reaction of 7 . Similarly, the β-D-configurated nitroaldoses 23 and 25 were the major products of the Michael addition. This indicates a preferred ‘endo’-attack on the nitronate anion derived from 7 . AMI calculations for this anion indicate a strong pyramidalization at C(1), in agreement with an ‘endo’-attack. Nucleosidation of 21 by 31 afforded 32 and 33 . Yields depended strongly upon the nature and the amount of the promoter and reached 77% for 33 , which was transformed into 34 , 35 , and the known ‘psicouridine’ ( 36 ; Scheme 3). To probe the mechanism, the trityl-protected 30 was nucleosidated yielding 37 , or 37 and 38 , depending upon the amount of FeCl3. Nucleosidation of the nitroacetate 28 was more difficult, required SnCl2 as a promoter, and yielded 39 and 40 . The β-D-anomer 40 was transformed into 36 . Nucleosidation of 23 (SnCl4) yielded the anomers 41 and 42 , which were transformed into 43 and 44 , and hence into 45 and 46 (Scheme 4). Similarly, nucleosidation of 25 yielded 47 and 48 , which were deprotected to 49 and 50 , respectively. The nucleoside 49 was saponified to 51 . Nucleosidation of 21 by 52 (SnCl2) afforded the adenine nucleosides 53 and 54 (Scheme 5). The adenine nucleoside 53 was deprotected (→ 55 → 56 ) to ‘psicofuranine’ (1), which was also obtained from 58 , formed along with 57 by nucleosidation of 28 . The structure and particularly the conformation of the nitroaldoses, nitroketoses, and nucleosides are examined.  相似文献   

13.
Oxidations of 5α‐hydroxy‐B‐norcholestan‐3β‐yl acetate ( 8 ) with Pb(OAc)4 under thermal or photolytic conditions or in the presence of iodine afforded only complex mixtures of compounds. However, the HgO/I2 version of the hypoiodite reaction gave as the primary products the stereoisomeric (Z)‐ and (E)‐1(10)‐unsaturated 5,10‐seco B‐nor‐derivatives 10 and 11 , and the stereoisomeric (5R,10R)‐ and (5S,10S)‐acetals 14 and 15 (Scheme 4). Further reaction of these compounds under conditions of their formation afforded, in addition, the A‐nor 1,5‐cyclization products 13 and 16 (from 10 ) and 12 (from 11 ) (see also Scheme 6) and the 6‐iodo‐5,6‐secolactones 17 and 19 (from 14 and 15 , resp.) and 4‐iodo‐4,5‐secolactone 18 (from 15 ) (see also Scheme 7). Oxidations of 5β‐hydroxy‐B‐norcholestan‐3β‐yl acetate ( 9 ) with both hypoiodite‐forming reagents (Pb(OAc)4/I2 and HgO/I2) proceeded similarly to the HgO/I2 reaction of the corresponding 5α‐hydroxy analogue 8 . Photolytic Pb(OAc)4 oxidation of 9 afforded, in addition to the (Z)‐ and (E)‐5,10‐seco 1(10)‐unsaturated ketones 10 and 11 , their isomeric 5,10‐seco 10(19)‐unsaturated ketone 22 , the acetal 5‐acetate 21 , and 5β,19‐epoxy derivative 23 (Scheme 9). Exceptionally, in the thermal Pb(OAc)4 oxidation of 9 , the 5,10‐seco ketones 10, 11 , and 22 were not formed, the only reaction being the stereoselective formation of the 5,10‐ethers with the β‐oriented epoxy bridge, i.e. the (10R)‐enol ether 20 and (5S,10R)‐acetal 5‐acetate 21 (Scheme 8). Possible mechanistic interpretations of the above transformations are discussed.  相似文献   

14.
Stereoselective Syntheses of (Z)-(10-Methoxy-4H-benzo[4,5]cyclohepta[1,2-b]thiophen-4-ylidene)acetic Acid Two stereoselective syntheses for the antiinflammatory compound 1 ((Z)-isomer) are described. In the first approach (Strategy A, Scheme 1) the stereoselective synthesis of 1 was realized via the bicyclic compound 11 under thermodynamic conditions, followed by a thiophene annelation with retention of the double-bond geometry (Schemes 2–4). Optimized conditions were necessary to avoid (E/Z)-isomerization during annelation. In the second approach (Strategy B, Scheme 1), diastereoisomer 17b was obtained selectively from a mixture of the diastereoisomers 17b and 18b by combining thermodynamic epimerization and solubility differences (Scheme 5). Diastereoisomer 17b was converted into the tricyclic compound 23 using a novel thiophene annelation method which we described recently (Scheme 6). In a final step, a stereospecific ‘syn’-elimination transformed the sulfoxide 24 into the target compound 1 (Scheme 7). To avoid (E/Z)-isomerization, it was necessary to trap the sulfenic acid liberated during the reaction. The key reactions of both approaches are highly stereoselective (> 97:3).  相似文献   

15.
Synthesis of ‘Push-Pull’-OligoAcetylenes ‘Push-pull’ triacetylenes 11a , b , c , as well as ‘push-pull’ tetraacetylene 13b have been prepared by reaction of the corresponding trichloroene(oligoinyl)amines 9 and 10 with 2 mol-equiv. of BuLi followed by acylation. The sequences (Schemes 3 and 4) are very simple and straightforward, they could in principle be applied to the synthesis of ‘push-pull’ pentaAcetylenes 15 and hexaacetylenes 17 (Scheme 5). Main limitations are the moderate yields as well as the low thermal stability of push-pull oligoacetylenes.  相似文献   

16.
The application of the ‘direct amide cyclization’ conditions to the linear δ‐hydroxy diamide 11 is described (Scheme 3). Instead of the cyclization to the expected nine‐membered cyclodepsipeptide, only the chloro acid 12 was obtained. Its formation could be explained by consecutive formation of the 1,3‐oxazol‐5(4H)‐one 16 and the six‐membered imino lactone 17 as intermediates (Scheme 4). The spontaneous isomerization of the latter gave 12 in a good yield.  相似文献   

17.
The synthesis of α‐benzamido‐α‐benzyl lactones 23 of various ring size was achieved either via ‘direct amide cyclization’ by treatment of 2‐benzamido‐2‐benzyl‐ω‐hydroxy‐N,N‐dimethylalkanamides 21 in toluene at 90 – 110° with HCl gas or by ‘ring transformation’ of 4‐benzyl‐4‐(ω‐hydroxyalkyl)‐2‐phenyl‐1,3‐oxazol‐5(4H)‐ones under the same conditions. The precursors were obtained by C‐alkylations of 4‐benzyl‐2‐phenyl‐1,3‐oxazol‐5(4H)‐one ( 15 ) with THP‐ or TBDMS‐protected ω‐hydroxyalkyl iodides. Ring opening of the THP‐protected oxazolones by treatment with Me2NH followed by deprotection of the OH group gave the diamides 21 , whereas deprotection of the TBDMS series of oxazolones 25 with TBAF followed by treatment with HCl gas led to the corresponding lactones 23 in a one‐pot reaction.  相似文献   

18.
(S)‐5‐Benzoyloxymethyl‐3‐[(E)‐(dimethylamino)methylidene]tetrahydrofuran‐2‐one (6), prepared in 5 steps from L‐glutamic acid (1), was used as precursor in a one step ‘ring switching’ synthesis of (S)‐2‐hydroxy‐3‐heteroaryl‐l‐propyl benzoates 13‐18, 23, 24. In the reaction of 6 with 2‐aminopyridine (21) and 2‐amino‐4,6‐dimethylpyrimidine (22) the corresponding dimethylamine substitution products (25, 26) were obtained.  相似文献   

19.
It is shown that azulene ( 1 ) and dimethyl acetylenedicarboxylate (ADM) in a fourfold molar excess react at 200° in decalin to yield, beside the known heptalene- ( 5 ) and azulene-1,2-dicarboxylates ( 6 ), in an amount of 1.6% tetramethyl (1RS,2RS,5SR,8RS)-tetracyclo[6.2.2.22,501,5]tetradeca-3,6,9,11,13-pentaene-3,4,9,10-tetracarboxylate(‘anti’-7) as a result of a SHOMO (azulene)/LUMO(ADM)-controlled addition of ADM to the seven-membered ring of 1 followed by a Diels-Alder reaction of the so formed tricyclic intermediate 16 (cf. Scheme 3) with a second molecule of ADM. The structure of ‘anti’-7 was confirmed by an X-ray diffraction analysis. Similarly, the thermal reaction of 5,7-dimehtylazulene ( 3 ) with excess ADM in decalin at 120° led to the formation of ca. 1% of ‘anti’- 12 , the 7,12-dimethyl derivative of‘anti’-7, beside of the corresponding heptalene- 10 and azulene-1,2-dicaboxylated (cf Scheme 2). The introduction of Me groups at C(1)and C(3)of azulene ( 1 ) and its 5,7-dimethyl derivative 3 strongly enhance the thermal formation of the corresponding tetracyclic compound. Thus, 1,3-dimethylazulene ( 2 ) in the presence of a sevenfold molar excess of ADM at 200° yielded 20% of ‘anti’- 9 beside an equal amount of dimethyl 3-mehtylazulene-1,2-dicarboxylate ( 8 ;cf. Scheme 1), and 1,3,5,7-tetramethylazulene ( 4 ) with a fourfold molar excess of ADM AT 200° gave a yield of 37% of‘anti’- 15 beside small amount of the corresponding heptalene- 13 and azulene-1,2-dicarboxylates 14 (cf.Scheme 2).  相似文献   

20.
Earlier phytochemical work on Plectranthus ambiguus (Lamiaceae) afforded a series of tetracyclic phyllocladane‐type (=13β‐kaurane) diterpenoids (see 1a – f ). In the course of investigations concerning the reaction behavior of this rare natural‐products, a new constituent of P. ambiguus was isolated, (2S,3R,16R)‐phyllocladane‐2,3,16,17‐tetrol 2,3‐diacetate ( 1g ), and another eighteen new phyllocladanes were prepared by chemical transformations and characterized. The main constituent 1b of P. ambiguus was chemically transformed to the known natural diterpenoid calliterpenone (=(16R)‐16,17‐dihydroxyphyllocladan‐3‐one; 2 ) thus unambiguously establishing its structure (Scheme 1). Epimerization at C(16) via the epoxy derivative 20 yielded 16‐epicalliterpenone ( 21 ), 17‐hydroxyphylloclad‐15‐ene‐3‐one ( 22 ), and (16R)‐3‐oxophyllocladan‐17‐al ( 23 ) (Scheme 6). Comparing this reaction sequence with the corresponding one starting from diastereoisomeric (16R)‐16,17‐dihydroxy‐ent‐kauran‐3‐one (=abbeokutone; 27 ) showed basically the same outcome (Scheme 7). Furthermore, three new C(16)‐substituted ent‐kauran‐3‐ones were characterized. Reliable spectroscopic arguments for the determination of the configuration at C(16) in phyllocladanes and kauranes as well as for the differentiation of the diastereoisomeric skeletons are presented.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号