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.
On irradiation in the presence of Fe(CO)5, the allenecarboxylates 1 afforded binuclear carbonyliron complexex 6 (Scheme 3), whereas the allenic lactone 7 under similar conditions gave a mixture of one binuclear and two mononuclear carbonyliron complexes ( 9 , 8 , and 10 ; Scheme 4). The structure of the complexes has been elucidated by X-ray crystallography. The structure of the binuclear complex 9 corresponds to that of 6 , while 8 has been shown to be a 1,3-butadiene(tricabonyl)iron complex. The unique structure of the 10 represents a new type of allenic complex. A stepwise formation of the complexes via intermediate allene(tetracarbonyl) iron complexes type 11 and 13 is suggested. Treatment of the binuclear complex 6b with FeCl3 led to the formation of the free ligand and a mixture of mononuclear complexes 13 and 14 (Scheme 5). On heating, the 1,3-diene complex 8 yielded the free ligand 15 , the prouduct of a (1,3) H shift in the allene 7 ; the complex 10 on the other hand liberates 7 on treatment with ethylenetracarbonitrile (TCNE) (Scheme 6).  相似文献   

2.
The bis‐C‐glucoside 2 has been synthesised as the first representative of a series of templated glucosides and cellooligosaccharides that mimick part of the unit cell of cellulose I. As expected, there are, at best, weakly persistent H‐bonds between the two glucosyl residues in (D6)DMSO and (D7)DMF solution. The acetylated oct‐1‐ynitol 7 and deca‐1,3‐diynitol 12 were prepared from the gluconolactone 5 (Scheme 1). Coupling of 12 to PhI and 2‐iodothiophene yielded 13 and 14 , respectively, while dimerisation of the benzylated and acetylated deca‐1,3‐diynitols 10 and 12 afforded the bis‐C‐glucosyloctatetrayne 15 and the less stable 16 , respectively. The 2‐glucosylthiophene 17 was obtained by treating the C‐silylated deca‐1,3‐diynitol 9 with Na2S. Cross‐coupling of (trimethylsilyl)acetylene (TMSA) with 1,8‐bis(triflyloxy)‐9,10‐anthraquinone ( 20 ) at elevated temperature gave the dialkynylated 21 ; its structure was established by X‐ray analysis (Scheme 2). Sequential coupling of 6 or 7 and TMSA to 20 gave the symmetric dialkyne 21 , the mixed dialkynes 23 (from 6 ) and 25 (from 7 ), and the symmetric diglucoside 36 (from 7 ) in modest yields; a stepwise coupling to the acetylated monotriflate 28 proved advantageous. It led to the oct‐1‐ynitol 29 and the deca‐1,3‐diynitol 33 that were transformed into the triflates 30 and 34 , respectively. Coupling of the triflate 34 to the oct‐1‐ynitol 7 gave the unsymmetric bis‐C‐glucoside 35 ; this was obtained in higher yields by coupling the triflate 30 to the deca‐1,3‐diynitol 12 . Coupling of the bistriflate 20 with either 7 or 12 afforded the symmetric bis‐C‐glucosides 36 and 37 , respectively. Deacetylation (KCN in MeOH) of 35 – 37 provided the unsymmetric bis‐C‐glucoside 2 and the symmetric analogues 3 and 4 .  相似文献   

3.
Cyclic ketene N,X‐acetals 1 are electron‐rich dipolarophiles that undergo 1,3‐dipolar cycloaddition reactions with organic azides 2 ranging from alkyl to strongly electron‐deficient azides, e.g., picryl azide ( 2L ; R1=2,4,6‐(NO2)3C6H2) and sulfonyl azides 2M – O (R1=XSO2; cf. Scheme 1). Reactions of the latter with the most‐nucleophilic ketene N,N‐acetals 1A provided the first examples for two‐step HOMO(dipolarophile)–LUMO(1,3‐dipole)‐controlled 1,3‐dipolar cycloadditions via intermediate zwitterions 3 . To set the stage for an exploration of the frontier between concerted and two‐step 1,3‐dipolar cycloadditions of this type, we first describe the scope and limitations of concerted cycloadditions of 2 to 1 and delineate a number of zwitterions 3 . Alkyl azides 2A – C add exclusively to ketene N,N‐acetals that are derived from 1H‐tetrazole (see 1A ) and 1H‐imidazole (see 1B , C ), while almost all aryl azides yield cycloadducts 4 with the ketene N,X‐acetals (X=NR, O, S) employed, except for the case of extreme steric hindrance of the 1,3‐dipole (see 2E ; R1=2,4,6‐(tBu)3C6H2). The most electron‐deficient paradigm, 2L , affords zwitterions 16D , E in the reactions with 1A , while ketene N,O‐ and N,S‐acetals furnish products of unstable intermediate cycloadducts. By tuning the electronic and steric demands of aryl azides to those of ketene N,N‐acetals 1A , we discovered new borderlines between concerted and two‐step 1,3‐dipolar cycloadditions that involve similar pairs of dipoles and dipolarophiles: 4‐Nitrophenyl azide ( 2G ) and the 2,2‐dimethylpropylidene dipolarophile 1A (R, R=H, tBu) gave a cycloadduct 13 H , while 2‐nitrophenyl azide ( 2 H ) and the same dipolarophile afforded a zwitterion 16A . Isopropylidene dipolarophile 1A (R=Me) reacted with both 2G and 2 H to afford cycloadducts 13G , J ) but furnished a zwitterion 16B with 2,4‐dinitrophenyl azide ( 2I) . Likewise, 1A (R=Me) reacted with the isomeric encumbered nitrophenyl azides 2J and 2K to yield a cycloadduct 13L and a zwitterion 16C , respectively. These examples suggest that, in principle, a host of such borderlines exist which can be crossed by means of small structural variations of the reactants. Eventually, we use 15N‐NMR spectroscopy for the first time to characterize spirocyclic cycloadducts 10 – 14 and 17 (Table 6), and zwitterions 16 (Table 7).  相似文献   

4.
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   

5.
The switch from a concerted to a two‐step pathway of 1,3‐dipolar cycloadditions was recently established for the reactions of sterically hindered ‘thiocarbonyl ylides' with acceptor ethylenes. This mechanism via zwitterionic intermediates is studied here for 1,3‐dipoles 5A and 5B , which are derived from 2,2,5,5‐tetramethylcyclopentanethione and 1,1,3,3‐tetramethylindan‐2‐thione, respectively, and contain a highly screened reaction center. In the reactions of 8A and 8B (the precursors of 5A and 5B ) with dimethyl 2,3‐dicyanofumarate ( 15 ) and 2,3‐dicyanomaleate ( 16 ), virtually identical ratios of cis‐ and trans‐thiolanes were observed ( 17 / 18 93 : 7 for 5a and 94 : 6 for 5B ). Thus, full equilibration of rotameric zwitterions precedes cyclization; an anteceding disturbing isomerization 15 ⇌ 16 had to be circumvented. The cis,trans assignment of the cycloadducts rests on three X‐ray analyses. The kinetically favored cis‐thiolanes 17 isomerize at >80° to 18 (trans), and irreversible cleavage leads to thione 7 and trans,cis isomeric dimethyl 1,2‐dicyanocyclopropane‐1,2‐dicarboxylates ( 27 and 28 , resp.). Furthermore, the zwitterionic intermediates equilibrate with the cyclic seven‐membered ketene imine 21 , which was intercepted under conditions where the solvent contained 2 vol‐% of H2O or MeOH. Lactams 22 were obtained with H2O in high yields, and the primary products of capturing by MeOH were the cyclic ketene O,N‐acetals 23 , which subsequently tautomerized to the lactim methyl ethers 24 . When 5B was reacted with ethenetetracarbonitrile in CDCl3/MeOH (98 : 2 vol‐%), the analogous cyclic ketene imine 13B was trapped to the extent of 93%.  相似文献   

6.
The 3,3‐dichloro‐2,2,4,4‐tetramethylcyclobutanethione ( 4b ) was prepared from the parent diketone by successive reaction with PCl5 and Lawesson reagent in pyridine. This new thioketone 4b was transformed into 1‐chlorocyclobutanesulfanyl chloride 5 and chloro 1‐chlorocyclobutyl disulfide 9 by treatment with PCl5 and SCl2, respectively, in chlorinated solvents (Schemes 1 and 2). These products reacted with S‐ and P‐nucleophiles by substitution of Cl at the S‐atom; e.g., the reaction with 4b yielded the di‐ and trisulfides 6b and 11 , respectively. Surprisingly, only pentasulfide 12 was formed in the reaction of 9 with thiobenzophenone (Scheme 3). In contrast to 5 and 9 , the corresponding chloro 1‐chlorocyclobutyl trisulfide 13 could not be detected, but reacted immediately with the starting thioketone 4b to give the tetrasulfide 14 (Scheme 4). Oxidation of 4b with 3‐chloroperbenzoic acid (mCPBA) yielded the corresponding thione oxides (= sulfine) 15 , which underwent 1,3‐dipolar cycloadditions with thioketones 3a and 4b (Scheme 5). Furthermore, 4b was shown to be a good dipolarophile in reactions with thiocarbonylium methanides (Scheme 6) and iminium ylides (= azomethine ylides; Scheme 7). In the case of phenyl azide, the reaction with 4b gave the symmetrical trithiolane 25 (Scheme 8).  相似文献   

7.
Treatment of 2,2,4,4‐tetramethylcyclobutane‐1,3‐dione ( 6 ) in THF with CF3SiMe3 in the presence of tetrabutylammonium fluoride (TBAF) yielded the corresponding 3‐(trifluoromethyl)‐3‐[(trimethylsilyl)oxy]cyclobutanone 7 (Scheme 1) via nucleophilic addition of a CF anion at the CO group and subsequent silylation of the alcoholate. Under similar conditions, the ‘monothione' 1 reacted to give thietane derivative 8 (Scheme 2), whereas in the case of ‘dithione' 2 only the dispirodithietane 9 , the dimer of 2 , was formed (Scheme 3). A conceivable mechanism for the formation of 8 is the ring opening of the primarily formed CF3 adduct A followed by ring closure via the S‐atom (Scheme 2). In the case of thiobenzophenones 4 , complex mixtures of products were obtained including diarylmethyl trifluoromethyl sulfide 10 and 1,1‐diaryl‐2,2‐difluoroethene 11 (Scheme 4). Obviously, competing thiophilic and carbophilic addition of the CF anion took place. The reaction with 9H‐fluorene‐9‐thione ( 5 ) yielded only 9,9′‐bifluorenylidene ( 14 ; Scheme 6); this product was also formed when 5 was treated with TBAF alone. Treatment of 4a with TBAF in THF gave dibenzhydryl disulfide ( 15 ; Scheme 7), whereas, under similar conditions, 1 yielded the 3‐oxopentanedithioate 17 (Scheme 9). The reaction of dithione 2 with TBAF led to the isomeric dithiolactone 16 (Scheme 8), and 3 was transformed into 1,2,4‐trithiolane 18 (Scheme 10).  相似文献   

8.
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   

9.
The reactions of α-diazo ketones 1a,b with 9H-fluorene-9-thione ( 2f ) in THF at room temperature yielded the symmetrical 1,3-dithiolanes 7a,b , whereas 1b and 2,2,4,4-tetramethylcyclobutane-1,3-dithione ( 2d ) in THF at 60° led to a mixture of two stereoisomeric 1,3-oxathiole derivatives cis- and trans- 9a (Scheme 2). With 2-diazo-1,2-diphenylethanone ( 1c ), thio ketones 2a–d as well as 1,3-thiazole-5(4H)-thione 2g reacted to give 1,3-oxathiole derivatives exclusively (Schemes 3 and 4). As the reactions with 1c were more sluggish than those with 1a,b , they were catalyzed either by the addition of LiClO4 or by Rh2(OAc)4. In the case of 2d in THF/LiClO4 at room temperature, a mixture of the monoadduct 4d and the stereoisomeric bis-adducts cis- and trans- 9b was formed. Monoadduct 4d could be transformed to cis- and trans- 9b by treatment with 1c in the presence of Rh2(OAc)4 (Scheme 4). Xanthione ( 2e ) and 1c in THF at room temperature reacted only when catalyzed with Rh2(OAc)4, and, in contrast to the previous reactions, the benzoyl-substituted thiirane derivative 5a was the sole product (Scheme 4). Both types of reaction were observed with α-diazo amides 1d,e (Schemes 5–7). It is worth mentioning that formation of 1,3-oxathiole or thiirane is not only dependent on the type of the carbonyl compound 2 but also on the α-diazo amide. In the case of 1d and thioxocyclobutanone 2c in THF at room temperature, the primary cycloadduct 12 was the main product. Heating the mixture to 60°, 1,3-oxathiole 10d as well as the spirocyclic thiirane-carboxamide 11b were formed. Thiirane-carboxamides 11d–g were desulfurized with (Me2N)3P in THF at 60°, yielding the corresponding acrylamide derivatives (Scheme 7). All reactions are rationalized by a mechanism via initial formation of acyl-substituted thiocarbonyl ylides which undergo either a 1,5-dipolar electrocyclization to give 1,3-oxathiole derivatives or a 1,3-dipolar electrocyclization to yield thiiranes. Only in the case of the most reactive 9H-fluorene-9-thione ( 2f ) is the thiocarbonyl ylide trapped by a second molecule of 2f to give 1,3-dithiolane derivatives by a 1,3-dipolar cycloaddition.  相似文献   

10.
The synthesis of bolaamphiphiles from unusual β‐amino acids or an alcohol and C12 or C20 spacers is described. Unusual β‐amino acids such as a sugar amino acid, an AZT‐derived amino acid, a norbornene amino acid, and an AZT‐derived amino alcohol were coupled with spacers under standard conditions to get the novel bolaamphiphiles 5 – 8 (Scheme 1), 12 and 13 (Scheme 2), and 17 and 20 (Scheme 3). Some of these compounds, on precipitation from MeOH/H2O, self‐assembled into organized molecular structures.  相似文献   

11.
Base‐promoted (KOH or MeONa in MeOH, or NaH in THF) cycloisomerisation of partially benzylated, 1‐substituted (R = Ph CC, pyridin‐2‐yl, or Br) ald‐1‐ynitols leads to (Z)‐configured five‐, six‐, and seven‐membered exo‐glycals. The reactivity of the ald‐1‐ynitols depends upon their configuration. The ald‐1‐ynitols were derived from 2,3,5‐tri‐O‐benzyl‐D ‐ribofuranose 1 , and the corresponding, partially O‐benzylated galactose, glucose, and mannose hemiacetals by ethynylation. The hex‐1‐ynitol 2 derived from 1 (61%) was transformed via the 1‐phenylbuta‐1,3‐diyne 3 and the 1‐(pyridin‐2‐yl)acetylene 5 into the five‐membered exo‐glycals 4 and 6 (in 66 and 72% yields, resp., from 2 ). The analoguous ethynylation of 2,3,4,6‐tetra‐O‐benzyl‐D ‐galactose 8 was accompanied by elimination of one benzyloxy (BnO) group to the hept‐3‐en‐1‐ynitol 9 (71%), which was transformed into the non‐5‐ene‐1,3‐diynitol 10 and further into the six‐membered exo‐glycal 11 (50% from 9 ). Addition of Me3SiCCH to the galactose 8 and to the gluco‐ and manno‐analogues 16 and 24 gave epimeric mixtures of the silylated oct‐1‐ynitols (86% of 12L / 12D 45 : 55, 94% of 17L / 17D 7 : 3, and 86% of 25L / 25D 55 : 45), which were separated by flash chromatography, and individually transformed into the corresponding 1‐bromooct‐1‐ynitols. Upon treatment with NaH in THF, only the minor epimers 13L, 18D , and 26D cyclised readily to form the seven‐membered hydroxy exo‐glycals. They were acetylated to the more stable monoacetates 14L, 23D , and 28D (82–89% overall yield). Under the same conditions, the epimers 13D, 18L , and 26L decomposed within 12 h mostly to polar products. The difference of reactivity was rationalised by analysing the consequences of an intramolecular C(3)O H ⋅⋅⋅ OC(7) H‐bond of the intermediate alkoxides on the orientation of O C(7) of 13L, 18D , and 26D and its proximity to the ethynyl group.  相似文献   

12.
Thermal decomposition of thiones of selected N‐, O‐ and S‐heterocycles under flash vacuum thermolysis conditions was investigated. In the case of six‐membered 4H‐3,1‐benzoxathiin‐4‐thione 6 , the course of the reaction depended on the substitution pattern at C(2) (Scheme 3). Thus, the 2‐unsubstituted derivative 6a led to the unstable product 2 , which upon treatment with MeOH was converted quantitatively into methyl 2‐mercaptobenzoate ( 7 ). The analogous thermolysis of the 2,2‐dimethyl derivative 6b yielded 2‐methyl‐4H‐1‐benzothiopyran‐4‐thione ( 8 ) as a sole product. In the case of thiophthalide derivatives 15 , a thermal rearrangement in the gas phase leading to the corresponding benzo[c]thiophen‐1(3H)‐ones 16 in high yields was observed (Scheme 6). Unexpectedly, thionation of 1,3‐oxathiolan‐5‐one 17 with Lawesson's reagent under standard conditions led to 1,2‐dithietane derivative 19 , which, after the gas‐phase thermolysis, underwent a ring enlargement to yield 3H‐1,2‐dithiole 20 (Scheme 7). The six‐membered 4H‐1,3‐benzothiazine‐4‐thione 21 was shown to give three products: phenanthro[9,10‐c]‐1,2‐dithiete ( 22 ), 3H‐1,3‐benzodithiole‐3‐thione ( 23 ), and N‐(3H‐1,2‐benzodithiol‐3‐ylidene)prop‐2‐en‐1‐amine ( 24 ) (Scheme 8). The latter is the product of the initial reaction, whereas 22 and 23 are postulated to be formed as secondary products of the conversion of the intermediate 6‐(thioxomethylene)cyclohexa‐2,4‐diene‐1‐thione ( 26 ) (Schemes 9 and 10).  相似文献   

13.
Reaction of Di(tert-butyl)- and Diphenyldiazomethane and 1,3-Thiazole-5(4H)-thiones: Isolation and Crystal Structure of the Primary Cycloadduct Reactions of diazo compounds with C?S bonds proceed via the formation of thiocarbonyl ylides, which, under the reaction conditions, undergo either 1,3-dipolar cycloadditions or electrocyclic ring closer to thiiranes (Scheme 1). With the sterically hindered di(tert-butyl)diazomethane ( 2c ), 1,3-thiazole-5(4H)-thiones 1 react to give spirocyclic 2,5-dihydro-1,3,4-thiadiazoles 3 (Scheme 2). These adducts are stable in solution at ?20°, and they could be isolated in crystalline form. The structure of 3c was established by X-ray crystallography. In CDCl3 solution at room temperature, a cycloreversion occurs, and the adducts of type 3 are in an equilibrium with 1 and 2c . In contrast, the reaction of 1 with diphenyldiazomethane ( 2d ) gave spirocyclic thiiranes 4 as the only product in high yield (Scheme 3). The crystal structure of 4b was also determined by X-ray analysis. The desulfurization of compounds 4 to 4,5-dihydro-5-(diphenylmethylidene)-1,3-thiazoles 5 was achieved by treating 4 with triphenylphosphine in boiling THF. The crystal structure of 5f is shown.  相似文献   

14.
The reactions of several thioketones containing a conjugated C?C bond with diazo compounds were investigated. All of the selected compounds reacted via a 1,3‐dipolar cycloaddition with the C?S group and subsequent N2 elimination to yield thiocarbonyl ylides as intermediates, which underwent a 1,3‐dipolar electrocyclization to give the corresponding thiirane 25 , or, by a subsequent desulfurization, to give the olefins 33a and 33b . None of the intermediate thiocarbonyl ylides reacted via 1,5‐dipolar electrocyclization. If the α,β‐unsaturated thiocarbonyl compound bears an amino group in the β‐position, the reactions with diazo compounds led to the 2,5‐dihydrothiophenes 40a – 40d . In these cases, the proposed mechanism of the reactions led once more to the thiocarbonyl ylides 36 and thiiranes 38 , respectively. The thiiranes reacted via an SNi′‐like mechanism to give the corresponding thiolate/ammonium zwitterion 39 , which underwent a ring closure to yield the 2,5‐dihydrothiophenes 40 . Also in these cases, no 1,5‐dipolar electrocyclization could be observed. The structures of several key products were established by X‐ray crystallography.  相似文献   

15.
A series of side chain reactions starting from the 6‐ and 7‐styryl‐substituted 1,3‐dimethyllumazines 1 and 21 as well as from the 6‐ and 7‐[2‐(methoxycarbonyl)ethenyl]‐substituted 1,3‐dimethyllumazine 2 and 22 were performed first by addition of Br2 to the C?C bond forming the 1′,2′‐dibromo derivatives 3, 4, 24 , and 26 in high yields (Schemes 1 and 3) (lumazine=pteridine‐2,4(1H,3H)‐dione). Treatment of 3 with various nucleophiles gave rise to an unexpected tele‐substitution in 7‐position and elimination of the Br‐atoms generating 7‐alkoxy‐ (see 5 and 6 ), 7‐hydroxy‐ (see 7 ) and 7‐amino‐6‐styryl‐1,3‐dimethyllumazines (see 8 – 11 ) (Scheme 1). On the other hand, 4 underwent, with dilute DBU (1,8‐diazabicyclo[5.4.0]undec‐2‐ene), a normal HBr elimination in the side chain leading to 18 , whereas treatment with MeONa afforded a more severe structural change to 19 . Similarly, 24 and 26 reacted to 27, 32 , and 33 under mild conditions, whereas in boiling NaOMe/MeOH, 24 gave 7‐(2‐dimethoxy‐2‐phenylethyl)‐1,3‐dimethyllumazine ( 30 ) which was hydrolyzed to give 31 (Scheme 3). From the reactions of 4 and 24 with DBU resulted the dark violet substance 20 and 25 , respectively, in which DBU was added to the side chain (Scheme 2). The styryl derivatives 1 and 21 could be converted, by a Sharpless dihydroxylation reaction, into the corresponding stereoisomeric 6‐ and 7‐(1,2‐dihydroxy‐2‐phenylethyl)‐1,3‐dimethyllumazines 34 – 37 (Scheme 4). The dihydroxy compounds 34 and 35 were also acetylated to 38 and 39 which, on catalytic reduction followed by formylation, yielded the diastereoisomer mixtures 40 and 41 . Deacetylation to 42 and 45 allowed the chromatographic separation of the diastereoisomers resulting in the isolation of 43 and 44 as well as 46 and 47 , respectively. Introduction of a 6‐ or 7‐ethynyl side chains proceeded well by a Sonogashira reaction with 6‐ ( 48 ) or 7‐chloro‐1,3‐dimethyllumazine ( 55 ) yielding 49 – 51 and 56 – 58 (Scheme 5). The direction of H2O addition to the triple bond is depending on the substituents since the 6‐ ( 49 ) and 7‐(phenylethynyl)‐1,3‐dimethyllumazine ( 56 ) showed attack at the 2′‐position yielding 53 and 60 , in contrast to the 6‐ ( 51 ) and 7‐ethynyl‐1,3‐dimethyllumazine ( 58 ) favoring attack at C(1′) and formation of 6‐ ( 52 ) and 7‐acetyl‐1,3‐dimethyllumazine ( 59 ).  相似文献   

16.
The synthesis and characterization of new amidinate‐stabilized germatrisilacyclobutadiene ylides [L3Si3GeL′] (L=PhC(NtBu)2; L′=ËL; Ë=Ge ( 3 ), Si ( 7 )) are described. Compound 3 was prepared by the reaction of [LSi? SiL] ( 1 ) with one equivalent of [LGe? GeL] ( 2 ) in THF. Compound 7 was synthesized by the reaction of 2 with excess 1 in THF. The bisamidinate germylene [L2Ge:] ( 4 ) is a by‐product in both reactions. Moreover, compound 7 was prepared by the reaction of 3 with one equivalent of 1 in THF. Compounds 3 and 7 have been characterized by NMR spectroscopy, X‐ray crystallography, and theoretical studies. The results show that compounds 3 and 7 are not antiaromatic. The puckered Si3Ge four‐membered rings in 3 and 7 have a ylide structure, which is stabilized by amidinate ligands and the electron delocalization within the Si3Ge four‐membered ring.  相似文献   

17.
The migratory insertions of cis or trans olefins CH(X)?CH(Me) (X = Ph, Br, or Et) into the metal–acyl bond of the complex [Pd(Me)(CO)(iPr2dab)]+ [B{3,5‐(CF3)2C6H3}4]? ( 1 ) (iPr2dab = 1,4‐diisopropyl‐1,4‐diazabuta‐1,3‐diene = N,N′‐(ethane‐1,2‐diylidene)bis[1‐methylethanamine]) are described (Scheme 1). The resulting five‐membered palladacycles were characterized by NMR spectroscopy and X‐ray analysis. Experimental data reveal some important aspects concerning the regio‐ and stereochemistry of the insertion process. In particular, the presence of a Ph or Br substituent at the alkene leads to the formation of highly regiospecific products. Moreover, in all cases, the geometry of the substituents in the formed palladacycle was the same as in the starting olefin, as a consequence of a cis addition of the Pd–acyl fragment to the C?C bond. Reaction with CO and MeOH of the five‐membered complex derived from trans‐β‐methylstyrene (= [(1E)‐prop‐1‐enyl]benzene) insertion, yielded the 2,3‐substituted γ‐keto ester 9 with an (2RS,3SR)‐configuration (Scheme 3).  相似文献   

18.
Reversible deprotonation of fascaplysin ( 1 ) was achieved with non‐nucleophilic bases (Scheme 1). Under basic aqueous conditions, opening of ring D of 1 occurred, yielding zwitter‐ionic reticulatine 2a , whereas, in a methoxide‐containing MeOH solution, an unexpected addition of three molecules of MeOH to the pyridinium ring produced an isomer mixture 3 of a trimethoxy‐substituted compound (Scheme 2). Transformation of the keto group of 1 to the oxime 4A took place in the presence of pyridine as base (Scheme 3). Grignard and alkyllithium reagents added as expected to the keto group of 1 , providing tertiary alcohols 5 and 6 (Scheme 4).  相似文献   

19.
In the title compounds, C12H20O6, (I), and C9H16O6, (II), the five‐membered furanose ring adopts a 4T3 conformation and the five‐membered 1,3‐dioxolane ring adopts an E3 conformation. The six‐membered 1,3‐dioxane ring in (I) adopts an almost ideal OC3 conformation. The hydrogen‐bonding patterns for these compounds differ substantially: (I) features just one intramolecular O—H...O hydrogen bond [O...O = 2.933 (3) Å], whereas (II) exhibits, apart from the corresponding intramolecular O—H...O hydrogen bond [O...O = 2.7638 (13) Å], two intermolecular bonds of this type [O...O = 2.7708 (13) and 2.7730 (12) Å]. This study illustrates both the similarity between the conformations of furanose, 1,3‐dioxolane and 1,3‐dioxane rings in analogous isopropylidene‐substituted carbohydrate structures and the only negligible influence of the presence of a 1,3‐dioxane ring on the conformations of furanose and 1,3‐dioxolane rings. In addition, in comparison with reported analogs, replacement of the –CH2OH group at the C1‐furanose position by another group can considerably affect the conformation of the 1,3‐dioxolane ring.  相似文献   

20.
Diphenyl phosphorazidate (DPPA) was used as the azide source in a one-pot synthesis of 2,2-disubstituted 3-amino-2H-azirines 1 (Scheme 1). The reaction with lithium enolates of amides of type 2 , bearing two substituents at C(2), proceeded smoothly in THF at 0°; keteniminium azides C and azidoenamines D are likely intermediates. Under analogous reaction conditions, DPPA and amides of type 3 with only one substituent at C(2) gave 2-diazoamides 5 in fair-to-good yield (Scheme 2). The corresponding 2-diazo derivatives 6–8 were formed in low yield by treatment of the lithium enolates of N,N-dimethyl-2-phenylacetamide, methyl 2-phenylacetate, and benzyl phenyl ketone, respectively, with DPPA. Thermolysis of 2-diazo-N-methyl-N-phenylcarboxamides 5a and 5b yielded 3-substituted 1,3-dihydro-N-methyl-2H-indol-2-ones 9a and 9b , respectively (Scheme 3). The diazo compounds 5–8 reacted with 1,3-thiazole-5 (4H)-thiones 10 and thiobenzophenone ( 13 ) to give 6-oxa-1,9-dithia-3-azaspiro[4.4]nona-2,7-dienes 11 (Scheme 4) and thiirane-2-carboxylic acid derivatives 14 (Scheme 5), respectively. In analogy to previously described reactions, a mechanism via 1,3-dipolar cycloaddition, leading to 2,5-dihydro-1,3,4-thiadiazoles, and elimination of N2 to give the ‘thiocarbonyl ylides’ of type H or K is proposed. These dipolar intermediates with a conjugated C?O group then undergo either a 1,5-dipolar electrocyclization to give spirohetrocycles 11 or a 1,3-dipolar electrocyclization to thiiranes 14 .  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