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 matching of ring and cap orbitals for overlap is used to arrive at the best carborynes among the many possibilities. Accordingly, 1,2-carboranes, 1,2-silaboranes (C2BnHn+2, and Si2BnHn+2, n = 4, 5, 8, and 10), and their dehydrogeno derivatives were studied with use of the Density Functional Theory (B3LYP/6-311+G*). The dehydrogenation of 2,3-C2B5H7 (6a) to 2,3-C2B5H5 (13a) is estimated to be even less endothermic than those of benzene and 1,2- C2B10H12 (1a) to benzyne and 1,2-C2B10H10 (8a) by more than 21 kcal/mol. This is due to the extra stabilization gained through better overlap of the C2B3H3 ring with the 2 BH caps. The relatively larger size of the Si atom leads to overlap requirements in silaboranes that are different from those in carboranes. The lower Si-Si single bond energy and the preference of Si for lower coordination result in unusual structures in dehydrogenosilaboranes. One of the Si atoms moves away from the surface in Si2B10H10 (15), Si2B8H8 (16, 17, and 18), and 1,2-Si2B5H5 (19). One Si atom forms a bridge to a trigonal surface in 2,3-Si2B5H5 (20) and 1,2-Si2B4H4 (21). Estimates of three-dimensional aromaticity with NICS calculations show that the exohedral double bond does not influence three-dimensional aromaticity.  相似文献   

2.
1INTRODUCTIONBenzoheteroazepinecompoundsareimportantpharmaceuticalagents〔1,2〕.Inre centyears,ithasbeenfoundthatbenzoheteroaze...  相似文献   

3.
孔繁祚 《有机化学》1997,17(1):38-42
吡喃糖的1,2-及1,3-缩水内醚苄醚由相应的吡喃糖的C-2氧负离子(对1,3-缩水内醚是C-3氧负离子)与连有氯原子的C-1经分子内关环反应而制备, 而有些吡喃糖的1,2-缩水内醚苄醚是由相应的吡喃糖的C-1氧负离子与连有对甲苯磺酰氧基的C-2经分子内关环(倒关环)反应而得。呋喃糖的1,2-缩水内醚苄醚只能用倒关环法合成。1,2-(或1,3-)缩水内醚糖的开环反应通常给出1,2-反式连接(对1,3缩水内醚是1,3反式连接)的糖苷键。1,2-及1,3-缩水内醚糖的构象分析是通过^1H NMR测定及分子力学计算的方法而完成的。  相似文献   

4.
The hexa-coordinated rhenium(I) tricarbonyl complexes of 5,10,15,20-tetraphenyl-21-thiaporphyrin 1 and 5,10,15,20-tetraphenyl-21-oxaporphyrin 2 have been synthesized by thermal reaction of corresponding free base porphyrin with Re(2)(CO)(10). The compounds 1 and 2 are characterized by HR-MS mass, (1)H, and (13)C NMR, FTIR, UV-vis, and electrochemical techniques and the structures are determined by X-ray single crystal analysis. The X-ray analysis revealed different coordination behavior of Re(CO)(3) with 21-thiaporphyrin and 21-oxaporphyrin. Interestingly, the Re(CO)(3) coordinates to two of the three inner nitrogens and one sulfur instead of three inner nitrogens as generally expected to produce unique compound 1 whereas it coordinates to three inner nitrogens but not with furan oxygen to form compound 2. The 21-thiaporphyrin ring is more distorted in compound 1 compared to 21-oxaporphyrin ring in compound 2 on complexation with Re(CO)(3). The presence of three carbonyl groups in compounds 1 and 2 are verified by (13)C NMR and IR spectroscopy. The absorption spectra of compounds 1 and 2 showed ill-defined Q-bands along with broad Soret band and the extinction coefficients are much lower than their corresponding free base porphyrins. The compounds 1 and 2 showed two reversible porphyrin ring based reductions supporting their electron deficient nature. The compound 1 is very stable under protonation conditions, and the protonation occurs at the uncoordinated pyrrole ring whereas the compound 2 undergoes decomplexation under same conditions. Furthermore, compound 1 showed the fluxional behavior in coordination mode of binding in solution.  相似文献   

5.
The hydrazones of benzophenone, benzil, and acetophenone were allowed to react with acetoacetanilide to give azinoamides 18 , and the reaction of 18 with Appel's dehydrathon contiditons (triphenylphosphine/carbon tetra-chloride/triethylatnine) led to the corresponding azinoketimines 19 , which underwent elctrocyclic ring closure under the reaction conditions to give pyrazolo-fused heterocycles. Azinoamide 18a gave a 4,9-dihydropyazolo-[5,1-b]quinazoine 21 , while 18b yielded 2,3-dihydro-1H-imidazo[1,2-b]pyrazol-2-one 26 and 1H-imidazo[1,2-b]-pyrazole 29 . Compound 18c gave a monocyclic N-α-styryl-5-(phenylamino)pyrazole 32 .  相似文献   

6.
The syntheses of selenenate/seleninate esters and related derivatives by aromatic nucleophilic substitution (SNAr) reactions of 2‐bromo‐3‐nitrobenzylalcohol ( 13 ) and 2‐bromo‐3‐nitrobenzaldehyde ( 17 ) with Na2Se2/nBuSeNa are described. The reaction of 13 with Na2Se2 at room temperature afforded 7‐nitro‐1,2‐benzisoselenole(3 H) ( 15 ) instead of the desired diaryl diselenide 14 . Oxidation of selenenate ester 15 with hydrogen peroxide afforded the corresponding selenium(IV) derivative, 7‐nitro‐1,2‐benzisoselenole(3 H) selenium oxide ( 18 ). 2‐(Butylselanyl)‐3‐nitrobenzaldehyde ( 19 ) was synthesized by treating compound 17 with in situ generated nBuSeNa. The bromination reaction of selenide 19 did not afford the expected arylselenenyl bromide 20 , instead, it resulted in the formation of the unexpected 7‐nitro‐1,2‐benzisoselenol(3 H)‐3‐ol ( 21 ) and 3,3′‐oxybis(7‐nitro‐1,2‐benzisoselenole(3 H)) ( 22 ), respectively. The facile formation of heterocycles 21 and 22 is rationalized in terms of the aromatic ring strain in selenenyl bromide 20 . The presence of intramolecular secondary Se⋅⋅⋅O interactions in esters 15 , 18 , 21 , 22 , and selenenic anhydride 29 has been confirmed by single‐crystal X‐ray diffraction studies as well as computational studies. The presence of an intramolecular Se⋅⋅⋅O interaction in esters 4b , 8 , 15 , 18 , 21 , and 22 has been further proved by natural bond orbital (NBO) and atoms in molecules (AIM) calculations. Glutathione peroxidase‐like (GPx) antioxidant activities of 15 , 18 , 21 , 22 , and related heterocycles such as 7‐nitro‐1,2‐benzisoselenol(2 H)‐3‐one selenium oxide ( 4b ), 7‐nitro‐1,2‐benzisoselenol(2 H)‐3‐one ( 8 ), and 29 have been determined by the coupled reductase assay.  相似文献   

7.
The title compounds were prepared starting from the dihydropyrrolones 4 – 6 . Nucleophilic displacement and ring closure yielded the 1H‐pyrrolo[3,2‐c]isothiazol‐5(4H)‐ones 8 and 10 . The fused systems formed salts with strong acids and electrophiles ( 15, 16 ), as well as with bases. Oxidation led either to S(2)‐oxides ( 18a, 20a ) or to the corresponding bicyclic sultams ( 18b, 20b ), depending on the reaction conditions. The sulfinamide 18a was also obtained from the known 1,2‐dithiolopyrrolone S‐oxide 21 by a ring‐opening/ring‐closure reaction sequence. O‐Methylation of 8 furnished the ‘azafulvene' 17 . The oxidative addition of [Pt(η2‐C2H4)L2] ( 24a : L=Ph3P, 24b : L=1/2 dppf, 24c : L=1/2 (R,R)‐diop) to 18a and 20a led to the cis‐amido‐sulfenato Pt complexes 25 and 26a – c , respectively.  相似文献   

8.
[reaction: see text] Pentacoordinate 1,2-oxastibetanes 14a-d, which are formal [2 + 2]-cycloadducts of the reactions of stibonium ylides with carbonyl compounds, were successfully synthesized by the reactions of the corresponding bromo-2-hydroxyalkylstiboranes with NaH. The crystal structures of 14a and 14c were established by X-ray crystallographic analyses, showing their distorted trigonal bipyramidal structures and smaller C-Sb-O angles of the four-membered ring around antimony than the C-P-O angle of pentacoordinate 1,2-oxaphosphetane 3. The 1H, 13C, and 19F NMR spectra of 14a-d are consistent with the trigonal bipyramidal structure in the solution state. Although 14a did not decompose at all at 220 degrees C in o-xylene-d(10), the thermolyses of 3-phenyl-1,2-oxastibetane 14c were carried out at 220 degrees C in o-xylene-d(10) and at 140 degrees C in acetonitrile-d(3) to give the corresponding oxirane 28 with retention of configuration and cyclic stibinite 25. The formation of 28 is explained by apical-equatorial ligand coupling around antimony via a polar transition state, which is more favorable than olefin formation. In contrast, the thermolyses of 14c in the presence of LiBr and LiBPh4 gave oxirane 29 with inversion of configuration and the olefin 30, respectively. The formation of 29 and 30 is considered to proceed via an anti-betaine-type intermediate and hexacoordinate 1,2-oxastibetanide 36, respectively. Selective formation of 28, 29, and 30 in the thermolyses of 14c, which is regarded as an intermediate in the reaction of an alpha-phenyl-substituted stibonium ylide with a carbonyl compound, showed that the change of the reaction conditions controls the reactivity of a 1,2-oxastibetane compound.  相似文献   

9.
An efficient syntheses of 2-substituted 1,2-dihydro-1-naphthols and 2-substituted 1-naphthols has been developed that involves the sequential palladium-catalyzed ring opening of oxabicyclic alkenes with aryl and vinyl halides followed by oxidation of with IBX. In the first step of the sequence, a combination of Pd(OAc)2, PPh3, Zn, and PMP in dry DMF was employed to catalyze the ring opening of 7-oxabenzonorbornadienes with aryl and vinyl halides to afford the corresponding cis-2-substituted 1,2-dihydronaphthols in good to excellent yields. These reactions occurred under very mild conditions with a variety of aryl halides bearing electron-withdrawing or -donating groups. Similarly, a 7-azabenzonorbornadiene substituted with an electron-withdrawing group on the nitrogen atom underwent facile ring-opening reaction with aryl halides to provide cis-2-substituted (1,2-dihydro-1-naphthyl)carbamates in excellent yields. Oxidation of the intermediate 1,2-dihydro-1-naphthols using IBX yielded the corresponding 2-substituted 1-naphthols in good to excellent yields.  相似文献   

10.
The minimum energy reaction pathway (MERP) of the toluene-cycloheptatriene radical cation rearrangement (TOL/CHT-rearrangement) has been calculated by the UHF and DFT model at the level UHF/6-311+G(3df,2p)//UHF/6-31G(d) and B3LYP/6-311+G(3df,2p)//B3LYp/6-31G(d), respectively, including the ring walk of the substituent by a 1,2-shift around the aromatic ring. This ring walk corresponds to interconversion of distonic ions and norcaradiene radical cations (the two intermediates of the TOL/CHT-rearrangement) by making and breaking of the external C-C bonds of the cyclopropane moiety of the intermediate norcaradiene structure. For toluene radical cation 1, UHF calculations adequately reproduce earlier results(4) and show, that the ring walk of the CH(3)-substituents requires slightly more energy than formation of the cycloheptatriene radical cation. By the DFT model, the distonic ion, which is formed initially by a 1,2-H shift from CH(3) to the benzene ring, is not stable but the transition state of an interconversion of norcaradiene radical cations along a ring walk of the CH(3) substituent. The activation energy for this ring walk exceeds that for formation of the cycloheptatriene radical cation by c. 30 kJ mol(-1). Thus, isomerization of 1 by a ring walk of the CH(3)-substituent competes with the TOL/CHT-rearrangement likely only for excited 1. The calculation was repeated for the MERPs of a TOL/CHT-rearrangement of para-xylene radical cation 5 and ethylbenzene radical cation 2, yielding basically the same results as for 1. According to the calculation, polar substituents alter significantly the relative energies of the competing routes of isomerization. For benzylcyanide 3 (X = CN), the activation energy for a ring walk of the NC-CH(2)-substituent is distinctly below that of a ring enlargement. For benzyl methyl ether 4 (X = OCH(3)), the distonic intermediate along the UHF-MERP is unusually stable. Further, the 7-methoxy-norcaradiene radical ion is unstable and corresponds to a transition state between isomeric distonic intermediates differing by a 1,2-shift of the side chain. In contrast, the 7-methoxy-norcaradiene radical ion is the only intermediate of the DFT-MERP, and the distonic ion is the transition state for a 1,2-shift of the cyclopropane ring. A ring walk of the CH(3)OCH(2)-substituent is much more favorable than formation of a 7-methoxy-cycloheptatriene radical cation in both MERPs. The findings of the theoretical calculation are substantiated by the mass spectrometric fragmentations of meta- and para-methoxymethylated 1-phenylethanols 8 and 9 and of para-methoxymethyl substituted benzyl ethyl ether 10 and benzyl n-propyl ether 11. Important fragmentation routes of metastable molecular ions of these compounds correspond to elimination of alcohols. Use of deuterated derivatives shows that the elimination occurs by a "false" ortho-effect which requires migration of a ROCH(2)-substituent around the benzene ring. Results of particular interest are obtained for the asymmetric bis-ethers 10 and 11. Here, the MIKE spectra of the molecular ions of deuterated analogs reveal a selective ring walk of the C(2)H(5)OCH(2)- and n-C(3)H(7)OCH(2)-side chain, respectively.  相似文献   

11.
Irradiation of 1-alkyl-substituted 1,2-dihydronaphthalenes ( 10, 11, 12 ) with a lowpressure mercury lamp yields by ring opening ω-vinyl-o-quinodimethanes, which undergo [1, 7] H-shifts to give 1,2-divinyl-benzenes ( 8, 18, 23 ; cf. schemes 2, 3 and 4). In a further photoreaction of the divinylbenzenes, benzobicyclo [3.1.0]hex-2-enes ( 17, 19, 22 ) are formed. 2-Alkyl-substituted 1,2-dihydronaphthalenes ( 13, 14, 15, 16 ) are transformed by irradiation into ω-vinyl-o-quinodimethanes, which show [1, 7] H-shifts to yield in this case 2-(buta-1′, 3′-dienyl)-toluenes ( 9, 25, 26, 27 ; cf. schemes 6 and 7). The irradiation of 1-methyl- ( 10 ) and 1-ethyl-1, 2-dihydronaphthalene ( 11 ) with a high-pressure mercury lamp produces, besides the products of irradiation using the lowpressure lamp, 2-ethyl-allenylbenzene ( 24 ), and (from 11 ) 4-exo-ethyl-benzobicyclo[3.1.0]hex-2-ene (exo- 20 ) and 2-propyl-allenylbenzene ( 21 ), respectively (cf. scheme 5). Obviously, these products arise from a photreaction of the primarily formed ω-vinyl-o-quinodimethanes a .  相似文献   

12.
1INTRODUCTIONInrecentyears,ithasfoundthatbenzothia/diazepinetricyclicderivativesarepo tentialpharmaceuticalagents〔12〕.Thus,we...  相似文献   

13.
Two conceptually different routes to transient 1,2-diselones are reported: 1) via ring fragmentation of the 1,4,2-diselenazine system 6, and 2) by the tributylphosphane-induced depolymerisation of the shelf-stable organoselenium polymer 15. Evidence for the intermediacy of 1,2-diselone species 7 and 16 is provided in both cases by in situ trapping with dimethyl acetylenedicarboxylate (DMAD) to yield 1,4-diselenin derivatives. The route via 15 is especially expedient and trapping of 16 is efficient. Subsequent reactions of adduct 17 afford [1,2-ethanediylbis(diphenylphosphane)] [5,6-bis(methoxycarbonyl)-1,4-diselenin-2,3-dithiolato]nickel(IV) (20). Theoretical calculations at Hartree-Fock (HF) and Moller-Plesset electron-correlated levels (MP2) suggest that the cyclic 1,2-diselenete structure 7c is significantly more stable than the acyclic 1,2-diselone structure 7a. For the bicyclic system 16, the difference in energy between the cyclic and acyclic structures is considerably reduced due to the conformational rigidity imposed by the fused 1,3-dithiole ring. In contrast, the acyclic structures of the 1,2-dithione analogues 13a and 25a are computed to be more stable than their corresponding cyclic 1,2-dithiete structures 13c and 25c.  相似文献   

14.
Condensation of 2-nitrobenzenesulfonyl chloride with 2-ethoxycarbonyl-1H-pyrrole in the presence of potassium tert-butoxide and 18-crown-6 furnished 2-ethoxycarbonyl-1-(2-nitrobenzenesulfonyl)-1H-pyrrole. Reduction of nitro group to amino and subsequent cyclization by heating the aminoester in the presence of 2-hydroxypyridine as a bifuctional catalyst led to 11-oxo(10H)-pyrrolo[1,2-b] [1,2,5]benzothiadiazepine 5,5-dioxide. Treatment of the latter compound with di-4-morpholinylphosphinic chloride gave the corresponding phosphinyloxyimine, which on reacting with formylhydrazine underwent intramolecular cyclization to afford the title tetracyclic ring.  相似文献   

15.
The reactions of (2S)-2-amino-2-substituted-N-(4-nitrophenyl)acetamides 16a-c, succindialdehyde (13), and benzotriazole afforded enantiopure (3S,5R,7aR)-5-(1H-1,2,3-benzotriazol-1-yl)-3-substituted-1-(4-nitrophenyl)tetrahydro-1H-pyrrolo[1,2-a]imidazol-2-ones 17a-c, which were converted by sodium borohydride into (3S,7aR)-3-substituted-1-(4-nitrophenyl)tetrahydro-1H-pyrrolo[1,2-a]imidazol-2-ones 18a-c. Chiral (2S)-2-amino-2-substituted-N-(4-methylphenyl)acetamides 12a-d, easily prepared in two steps from N-Boc-alpha-amino acids 10a-d, similarly reacted with glutaraldehyde (20) and benzotriazole to generate 5-benzotriazolyl-3-substituted-hexahydroimidazo[1,2-a]pyridin-2(3H)-ones 21a-d, which were converted by sodium borohydride directly into optically active 3-substituted-hexahydroimidazo[1,2-a]pyridin-2(3H)-ones 22a-d.  相似文献   

16.
Five 1-(p-substituted phenyl)-1,4-dihydronicotinamides (GPNAH-1,4-H(2)) and five 1-(p-substituted phenyl)-1,2-dihydronicotinamides (GPNAH-1,2-H(2)) were synthesized, which were used to mimic NAD(P)H coenzyme and its 1,2-dihydroisomer reductions, respectively. When the 1,4-dihydropyridine (GPNAH-1,4-H(2)) and the 1,2-dihydroisomer (GPNAH-1,2-H(2)) were treated with p-trifluoromethylbenzylidenemalononitrile (S) as a hydride acceptor, both reactions gave the same products: pyridinium derivative (GPNA(+)) and carbanion SH(-) by a hydride one-step transfer. Thermodynamic analysis on the two reactions shows that the hydride transfer from the 1,2-dihydropyridine is much more favorable than the hydride transfer from the corresponding 1,4-dihydroisomer, but the kinetic examination displays that the former reaction is remarkably slower than the latter reaction, which is mainly due to much more negative activation entropy for the former reaction. When the formed pyridinium derivative (GPNA(+)) was treated with SH(-), the major reduced product was the corresponding 1,4-dihydropyridine along with a trace of the 1,2-dihydroisomer. Thermodynamic and kinetic analyses on the hydride transfer from SH(-) to GPNA(+) all suggest that the 4-position on the pyridinium ring in GPNA(+) is much easier to accept the hydride than the 2-position, which indicates that when the 1,4-dihydropyridine is used the hydride donor to react with S, the formed pyridinium derivative GPNA(+) may return to the 1,4-dihydropyridine by a hydride transfer cycle; but when the 1,2-dihydropyridine is used as the hydride donor, the formed pyridinium derivative can not return to the 1,2-dihydropyridine by the hydride reverse transfer from SH(-) to GPNA(+). These results clearly show that the hydride-transfer cycle is favorable for the 1,4-dihydronicotinamides, but unfavorable for the corresponding 1,2-dihydroisomers.  相似文献   

17.
Reactions of Ph(2)C(3) dianion, prepared from 1,3-diphenylpropyne and n-butyllithium, with alkyl thiocyanates or alkane dithiocyanates gave 1,3-bis(alkylthio)allenes 1 or tetrathiacyclic bisallenes 2, respectively. Thermal reactions of 1 gave thiophenes 4 and 7, benzothiepin 5, 1,2-bis(benzylidene)cyclobutane 6, thiete 8, and alpha,beta-unsaturated ketone 9, and the reactions of tetrathiacyclic bisallenes 2a gave a cyclic dimer, 1,2-bis(benzylidene)cyclobutane derivative 10, quantitatively. Irradiation of 1,3-bis(alkylthio)allenes 1 and tetrathiacyclic bisallenes 2a caused rearrangement to give alkynes 18, 20, and 21. In the irradiation of the cyclic bisallenes 2a, isomerizations from dl to meso and meso to dl isomers were also found. In the reactions of allenes 1 and cyclic bisallenes 2a with diphenyl diazomethane, the diazomethane reacted selectively with the double bond rather than with the sulfur atom.  相似文献   

18.
陈艳  屠树江 《结构化学》2007,26(9):1039-1042
The title compound 5-cyclopropyl-10-(4-fluorophenyl)-7,7-dimethyl-7,8-dihydro- 5H-indeno[1,2-b]quinolne-9,11(6H,10H)-dione was obtained by the reaction of 4-fluorobenzal- dehyde, 2H-indene-1,3-dione and 3-(cyclopropylamino)-5,5-dimethylcyclohex-2-enone in the presence of acetic acid under microwave irradiation. Its structure was confirmed by IR and 1H- NMR spectra. The crystal is of monoclinic, space group P21/c with a = 14.138(3), b = 8.952(2), c = 17.140(3) , β = 102.253(3)o, C27H24FNO2, Mr = 413.47, Z = 4, V = 2119.9(8) 3, Dc = 1.296 g/cm3, μ(MoKα) = 0.087 mm-1, F(000) = 872, the final R = 0.0425 and wR = 0.0905 for 2336 observed reflections (I > 2σ(I)). X-ray analysis revealed that the pyridine ring adopts a boat conformation and the six-membered ring fused with it assumes a twist boat conformation.  相似文献   

19.
2-quinolylcarbene 23 and 1-isoquinolylcarbene 33 are generated by flash vacuum thermolysis (FVT) of the corresponding triazolo[1,5-a]quinoline and triazolo[5,1-a]isoquinoline 19 and 29, as well as 2-(5-tetrazolyl)quinoline and 1-(5-tetrazolyl)isoquinoline 20 and 30, respectively. These carbenes rearrange to 1- and 2-naphthylnitrene 21 and 31, respectively, and the nitrenes are also generated by FVT of 1- and 2-naphthyl azides 18 and 28. The products of FVT of both the nitrene and carbene precursors are the 2- and 3-cyanoindenes 26 and 27 together with the nitrene dimers, viz. azonaphthalenes 25 and 35, and the H-abstraction products, aminonaphthalenes 24 and 34. All the azide, triazole, and tetrazole precursors yield 3-cyanoindene 26 as the principal ring contraction product under conditions of low FVT temperature (340-400 degrees C) and high pressure (1 Torr N(2) as carrier gas for the purpose of collisional deactivation). This ring contraction reaction is strongly subject to chemical activation, which caused extensive isomerization of 3-cyanoindene to 2-cyanoindene under conditions of low pressure (10(-3) Torr). 2-Cyanoindene is calculated to be ca. 1.7 kcal/mol below 3-cyanoindene in energy; accordingly, high-temperature FVT of these cyanoindenes always gives mixtures of the two compounds with the 2-cyano isomer dominating. Photolysis of trizolo[1,5-a]quinoline 19 and triazolo[5,1-a]isoquinoline 29 in Ar matrixes causes partial ring opening to the corresponding 2-diazomethylquinoline 19' and 1-diazomethylisoquinoline 29'. The photolysis of the former gives rise to a small amount of the cyclic ketenimine 22, the intermediate connecting 2-quinolylcarbene and 1-naphthylnitrene.  相似文献   

20.
Flash vacuum pyrolysis (FVP) of indol-1-ylacrylate derivatives 11 and 15 or the isomeric indol-3-ylacrylates 21, 22, and 24 at 925 degrees C (0.05 Torr) provides pyrrolo[1,2-a]indol-3-ones 2, 18, 28, and 29 in 53-90% yield by a cascade mechanism that involves a sigmatropic migration, elimination, electrocyclization sequence. Pyrrolo[1,2-a]imidazol-5-ones 3 and pyrrolo[1,2-c]imidazol-5-ones 4 were similarly obtained by FVP of corresponding 2,5-unsubstituted imidazol-1-ylacrylates (e.g., 33), with the former isomer predominating in ca. 80:20 ratio. Migration to the 2-position is therefore favored in the initial sigmatropic shift. FVP of 2-substituted imidazol-1-ylacrylates 35, 37, and 51 (825-875 degrees C) instead give pyrrolo[1,2-c]imidazol-5-ones 56-58 only (88-91%), and that of 4,5-disubstituted imidazol-1-ylacrylates 39 and 41 (825-850 degrees C) provide pyrrolo[1,2-a]imidazol-5-ones 59 and 60 exclusively (93-95%), and thus the selectivity of the initial shift can be controlled by the presence of substituents on the imidazole 2- and 5-positions. FVP of the benzimidazole analogues 61 and 62 at 950 degrees C gave the pyrrolo[1,2-a]benzimidazol-1-ones 6 (71%) and 63 (36%), respectively.  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