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.
Regioselective 1,3-dipolar cycloaddition of Cyanogen N,N′-dioxide ( 2 ) to trimethylsilyl enol ethers 3a-d, 6 and 7 gave the corresponding 5,5′-bis(trimethylsilyloxy)-3,3′-Δ2-biisoxazolines which upon short heating with 10% hydrochloric acid afforded 3,3′-biisoxazoles 5a-d , 8 and 9. Only the intermediate 5,5′-bis(trimethylsilyloxy)-derivative 4a was isolated and studied. Reaction of 2 with vinyl methyl ketone ( 10 ) gave biisoxazoline 11 which by oxidation with γ-manganese dioxide gave biisoxazole 12.  相似文献   

2.
The reaction of 3-(2-oxocycloalkylidene)indol-2-one 1 with thiourea and urea derivatives has been investigated. Reaction of 1 with thiourea and urea in ethanolic potassium hydroxide media leads to the formation of spiro-2-indolinones 2a-f in 40–50% yield and a novel tetracyclic ring system 4,5-cycloalkyl-1,3-diazepino-[4,5-b]indole-2-thione/one 3a-f in 30–35% yield. 3-(2-Oxocyclopentylidene)indol-2-one afforded 5′,6′-cyclopenta-2′-thioxo/ oxospiro[3H-indole-3,4′(3′H)pyrimidin]-2(1H)-ones 2a,b and 3-(2-oxocyclohexylidene)indol-2-one gave 2′,4′a,5′,6′,7′,8′- hexahydro-2′-thioxo/oxospiro[3H-indole-3,4′ (3′H)-quinazolin]-2(1H)-ones 2c-f . Under exactly similar conditions, reaction of 1 with fluorinated phenylthiourea/cyclohexylthiourea/phenylurea gave exclusively spiro products 2g-1 in 60–75% yield. The products have been characterized by elemental analyses, ir pmr. 19F nmr and mass spectral studies.  相似文献   

3.
A variety of nitrogen‐containing heterocycles were synthesized by passing vapors of aromatic amines over calcium oxide at 450–650 °C under nitrogen carrier gas. Reaction of 2‐aminobiphenyl 3a at 560 °C gave carbazole 4 in 80% yield. Reaction of 2, 2′‐diaminobiphenyl 3b afforded a mixture of carbazole 4 and 4‐aminocarbozole 6b. In the case of 2‐amino‐2′‐nitrobiphenyl 3c, benzo[c]cinnoline 7 was obtained along with carbazole 4. Reaction of 2‐amino‐2′‐methoxybiphenyl 3d gave four products of carbazole 4,4‐hydroxycarbazole 6e, phenanthridine 8 and dibenzofuran 9. Reaction of 2‐aminodiphenylmethane 5a afforded acridine 10. In the case of 2‐aminobenzophenone 5b, acridone 11 was obtained as a major product. Reaction of 2‐aminobenzhydrol 5c gave acridine 10. When 2‐aminodiphenylamine 5d was reacted, phenazine 12 was obtained in good yield. In contrast, reaction of 2‐aminodiphenyl ether 5e produced only 2‐hydroxydiphenylamine 13. Reaction of 4‐aminophenanthrene 14 produced 4H‐benzo[def]carbazole 15 in 61% yield.  相似文献   

4.
Bromination of 3-bromofuro[2,3-b]- 1a , -[3,2-b]- 1b and - [3,2-c]pyridine 1d afforded the 2,3-dibromo derivatives 2a, 2b and 2d , while the -[2,3-c]- compound 1c did not give the dibromo derivative. Nitration of 1a-d gave the 2-nitro-3-bromo compounds 3a-d . The N-oxides 4a-d of 1a-d were submitted to the cyanation with trimethylsilyl cyanide to yield the corresponding α-cyanopyridine compound 6a-d . Chlorination of 4a and 4d with phosphorus oxychloride gave mainly the chloropyridine derivatives 7a, 7′a and 7d , while 4b and 4c gave mainly the chlorofuran derivatives 7′b and 7′c accompanying formation of the chloropyridine derivatives 7b, 7′b and 7c . Acetoxylation of 4a and 4b with acetic anhydride yielded the acetoxypyridine compounds 8a, 8′a and 8b , while 4c and 4d gave the acetoxypyridine 8′c, 8′d and 8′d , pyridone 8c and 8d , acetoxyfuran 8′c and dibromo compound 9c and 9′c.  相似文献   

5.
The photooxygenation of (4R,4aS,7R)-4,4a,5,6,7,8-hexahydro-4,7-dimethyl-3H-2-benzopyran ( 16 ) was performed in (i) MeOH, (ii) acetaldehyde, and (iii) acetone at ?78°. The products obtained respectively were (i) (2R)-2-[(1S,4R)-4-methyl-2-oxocyclohexyl]propyl formate ( 17 ; 72% yield), (ii) 17 (54.5%), (1R,4R,4aS,7R)-3,4,4a,5,6,7-hexahydro-4,7-dimethyl-1H-2-benzopyran-2-yl hydroperoxide ( 19 ; 16.7%), a 12:1 ratio of (3R,4aR,7R,7aS,10R,11aR)-7,7a,8,9,10,11-hexahydro-3,7,10-trimethyl-6H-[2]benzopyrano[1,8a-e]-1,2,4-trioxane ( 20 ) and its C(3)-epimer 21 (17%), together with evidence for the 1,2-dioxetane ( 22 ) originating from the addition of dioxygen to the re-re face of the double bond of 16 , and iii) unidentified products and traces of 22 . Addition of trimethylsilyl trifluoromethanesulfonate (Me3SiOTf) to the acetone solution of 16 after photooxygenation afforded (4aR,7R,7aS,10R,11aR)-7,7a,8,9,10,11-hexahydro-3,3,7,10-tetramethyl-6H-[2]benzopyrano[1,8a-e]-1,2,4,-trioxane ( 23 , 40%). The photooxygenation of 16 in CH2Cl2 at ?78° followed by addition of acetone and Me3SiOTf afforded 17 (11%), 23 (59%), and (4aR,7R,7aS,10R,11aR)-7,7a,8,9,10,11-hexahydro-3,3,7,10-tetramethyl-6H-[2]benzopyrano[8a,1-e]-1,2,4-trioxane ( 24 ; 5%. Repetition of the last experiment, but replacing acetone by cyclopentanone, gave 17 (16%), (4′aR,7′R,7′aS,10′R,11′aR)-7′,7′a,8′,9′,10′,11′-hexahydro-7′,10′-dimethylspiro[cyclopentane-1,3′-6′H-[2]benzopyrano[1,8a-e]-1,2,4-trixane] ( 25 ; 61%), and (4′aR,7′R,7′aS,10′R,11′aR)-7′,7′a,8′,9′,10′,11′-hexahydro-7′,10′-dimethylspiro[cyclopentane-1,3′-6′H-[2]benzopyrano[8a,1-e]-1,2,4-trixane] ( 26 , 4%). The X-ray analysis of 23 was performed, which together with the NMR data, established the structure of the trioxanes 20, 21, 24, 25 , and 26 . Mechanistic and synthesis aspects of these reactions were discussed in relation to the construction of the 1,2,4-trioxane ring in arteannuin and similar molecules.  相似文献   

6.
A simple and high-yield synthesis of biologically significant 2′-deoxy-6-thioguanosine ( 11 ), ara-6-thioguanine ( 16 ) and araG ( 17 ) has been accomplished employing the Stereospecific sodium salt glycosylation method. Glycosylation of the sodium salt of 6-chloro- and 2-amino-6-chloropurine ( 1 and 2 , respectively) with 1-chloro-2-deoxy-3,5-di-O-(p-toluoyl)-α-D-erythro-pentofuranose ( 3 ) gave the corresponding N-9 substituted nucleosides as major products with the β-anomeric configuration ( 4 and 5 , respectively) along with a minor amount of the N-7 positional isomers ( 6 and 7 ). Treatment of 4 with hydrogen sulfide in methanol containing sodium methoxide gave 2′-deoxy-6-thioinosine ( 10 ) in 93% yield. Similarly, 5 was transformed into 2′-deoxy-6-thioguanosine (β-TGdR, 11 ) in 71 % yield. Reaction of the sodium salt of 2 with 1-chloro-2,3,5-tri-O-benzyl-α-D-arabinofuranose ( 8 ) gave N-7 and N-9 glycosylated products 13 and 9 , respectively. Debenzylation of 9 with boron trichloride at ?78° gave the versatile intermediate 2-amino-6-chloro-9-β-D-arabinofuranosyl-purine ( 14 ) in 62% yield. Direct treatment of 14 with sodium hydrosulfide furnished ara-6-thioguanine ( 16 ). Alkaline hydrolysis of 14 readily gave 9-β-D-arabinofuranosylguanine (araG, 17 ), which on subsequent phosphorylation with phosphorus oxychloride in trimethyl phosphate afforded araG 5′-monophosphate ( 18 ).  相似文献   

7.
The regiospecific reaction of 5-vinyl-3′,5′-di-O-acetyl-2′-deoxyuridine ( 2 ) with HOX (X = Cl, Br, I) yielded the corresponding 5-(1-hydroxy-2-haloethyl)-3′,5′-di-O-acetyl-2′-deoxyuridines 3a-c . Alternatively, reaction of 2 with iodine monochloride in aqueous acetonitrile also afforded 5-(1-hydroxy-2-iodoethyl)-3′,5′-di-O-acetyl-2′-deoxyuridine ( 3c ). Treatment of 5-(1-hydroxy-2-chloroethyl)- ( 3a ) and 5-(1-hydroxy-2-bromoethyl)-3′,5′-di-O-acetyl-2′-deoxyuridine ( 3b ) with DAST (Et2NSF3) in methylene chloride at -40° gave the respective 5-(1-fluoro-2-chloroethyl)- ( 6a , 74%) and 5-(1-fluoro-2-bromoethyl)-3′,5′-di-O-acetyl-2′-deoxyuridine ( 6b , 65%). In contrast, 5-(1-fluoro-2-iodoethyl)-3′,5′-di-O-acetyl-2′-deoxyuridine ( 6e ) could not be isolated due to its facile reaction with methanol, ethanol or water to yield the corresponding 5-(1-methoxy-2-iodoethyl)- ( 6c ), 5-(1-ethoxy-2-iodoethyl)- ( 6d ) and 5-(1-hydroxy-2-iodoethyl)-3′,5′-di-O-acetyl-2′-deoxyuridine ( 3c ). Treatment of 5-(1-hydroxy-2-chloroethyl)- ( 3a ) and 5-(1-hydroxy-2-bromoethyl)-3′,5′-di-O-acetyl-2′-deoxyuridine ( 3b ) with thionyl chloride yielded the respective 5-(1,2-dichloroethyl)- ( 6f , 85%) and 5-(1-chloro-2-bromoethyl)-3′,5′-di-O-acetyl-2′-deoxyuridine ( 6g , 50%), whereas a similar reaction employing the 5-(1-hydroxy-2-iodoethyl)- compound 3c afforded 5-(1-methoxy-2-iodoethyl)-3′,5′-di-O-acetyl-2′-deoxyuridine ( 6c ), possibly via the unstable 5-(1-chloro-2-iodoethyl)-3′,5′-di-O-acetyl-2′-deoxyuridine intermediate 6h . The 5-(1-bromo-2-chloroethyl)- ( 6i ) and 5-(1,2-dibromoethyl)-3′,5′-di-O-acetyl-2′-deoxyuridine ( 6j ) could not be isolated due to their facile conversion to the corresponding 5-(1-ethoxy-2-chloroethyl)- ( 6k ) and 5-(1-ethoxy-2-bromoethyl)-3′,5′-di-O-acetyl-2′-deoxyuridine ( 61 ). Reaction of 5-(1-hydroxy-2-bromoethyl)-3′,5′-di-O-acetyl-2′-deoxyuridine ( 3b ) with methanolic ammonia, to remove the 3′,5′-di-O-acetyl groups, gave 2,3-dihydro-3-hydroxy-5-(2′-deoxy-β-D-ribofuranosyl)-furano[2,3-d]pyrimidine-6(5H)-one ( 8 ). In contrast, a similar reaction of 5-(1-fluoro-2-chloroethyl)-3′,5′-di-O-acetyl-2′-deoxyuridine ( 6a ) yielded (E)-5-(2-chlorovinyl)-2′-deoxyuridine ( 1b , 23%) and 5-(2′-deoxy-β-D-ribofuranosyl)furano[2,3-d]pyrimidin-6(5H)-one ( 9 , 13%). The mechanisms of the substitution and elimination reactions observed for these 5-(1,2-dihaloethyl)-3′,5′-di-O-acetyl-2′-deoxyuridines are described.  相似文献   

8.
Three‐ and five‐membered rings that bear the (Si‐C‐S ) and (Si‐C‐C‐C‐S ) unit have been synthesized by the reactions of L SiCl ( 1 ; L =PhC(NtBu)2) and L′ Si ( 2 ; L′ =CH{(C?CH2)(CMe)(2,6‐iPr2C6H3N)2}) with the thioketone 4,4′‐bis(dimethylamino)thiobenzophenone. Treatment of 4,4′‐bis(dimethylamino)thiobenzophenone with L SiCl at room temperature furnished the [1+2]‐cycloaddition product silathiacyclopropane 3 . However, reaction of 4,4′‐bis(dimethylamino)thiobenzophenone with L′ Si at low temperature afforded a [1+4]‐cycloaddition to yield the five‐membered ring product 4 . Compounds 3 and 4 were characterized by NMR spectroscopy, EIMS, and elemental analysis. The molecular structures of 3 and 4 were unambiguously established by single‐crystal X‐ray structural analysis. The room‐temperature reaction of 4,4′‐bis(dimethylamino)thiobenzophenone with L′ Si resulted in products 4 and 5 , in which 4 is the dearomatized product and 5 is formed under the 1,3‐migration of a hydrogen atom from the aromatic phenyl ring to the carbon atom of the C? S unit. Furthermore, the optimized structures of probable products were investigated by using DFT calculations.  相似文献   

9.
Thermal decomposition of the diazonium sulfate derived from N-methyl-(1-phenyl-3-methylpyrazol-5-yl)-2-aminobenzamide afforded products formulated as 1-phenyl-3-methyl[2]benzopyrano[4,3-c]pyrazol-5-one (yield 10%), 1,4-dimethyl-3-phenylpyrazolo[3,4-c]isoquinolin-5-one (yield 10%), N-methyl-(1-phenyl-3-methylpyrazol-5-yl)-2-hydroxybenzamide (yield 8%) and 4′-hydroxy-2,3′-dimethyl-1′-phenylspiro[isoindoline-1,5′-[2]-pyrazolin]-3-one (yield 20%). Decomposition of the diazonium sulfate derived from N-methyl-(1,3-diphenylpyrazol-5-yl)-2-aminobenzamide gave products formulated as 7,9-dimethyldibenzo[e,g]pyrazolo[1,5-a][1,3]-diazocin-10-(9H)one (yield 8%), 4-methyl-1,3-diphenylpyrazolo[3,4-c]isoquinolin-5-one (yield 7%) and 4′-hydroxy-2-methyl-1′,3′-diphenylspiro[isoindoline-1,5′-[2]pyrazolin]3-one (yield 10%). The spiro compounds 6a,b underwent thermal and acid-catalysed conversion into the hitherto unknown 2-benzopyrano[4,3-c]pyrazole ring system 7a,b in good yield. Analytical and spectral data are presented which supported the structures proposed.  相似文献   

10.
Upon irradiation in 2-propanol, p-nitroacetophenone 1 was reduced via the triplet state to p-hydroxyaminoacetophenone 5 which was further reduced to p-aminoacelophenone 2 and 4,4′-diacetylazobenzene 4 . Similar irradiation of 5 also gave 2 and 4 , and its oxidation by oxygen gave 4,4′-diacetylazoxybenzene 3 . Photolysis of monomeric p-nitrosoacetophenone 6 afforded acetophenone and 3 that were not produced during the irradiation of 1 . Possible photoreaction pathways were discussed on the basis of published mechanisms.  相似文献   

11.
Reactions of 2-aminobenzenesulfonamide ( 1 ) with allyl, methyl, 2-chloroethyl aor 3-chloropropyl isocyanates gave 2-(methylureido)-, 2-(allylureido)-, 2-(2′-chloroethylureido)- and 2-(3′-chloropropylureido)-benzene sulfonamides 3a,b and 7a,b in excellent yields. Treatment of 3a,b at refluxing temperature of DMF afforded 2H-1,2,4-benzothiadiazin-3(4H)-one 1,1-dioxide ( 4 ) in good yield. However, when compounds 7a,b were refluxed in 2-propanol, 3-(2′-aminoethoxy)-2H-1,2,4-benzothiadiazine 1,1-dioxide ( 11a ) and 3-(3′-aminopropoxy)-2H-1,2,4-benzothiadiazine 1,1-dioxide ( 11b ) were obtained in a form of the hydrochloride salts 10a,b in 87% and 78% yields respectively. Heating 11b in ethanol gave a dimeric form of 2H-1,2,4-benzothiadiazin-3(4H)-one 1,1-dioxide and 3-(3′-aminopropoxy)-2H-1,2,4-benzothiadiazine 1,1-dioxide ( 12 ) in 55% yield. Treating of 7a,b or 11a,b with triethylamine at the refluxing temperature of 2-propanol afforded 3-(2′-hydroxyethylamino)-2H-1,2,4-benzothiadiazine 1,1-dioxide ( 2a ) and 3-(3′-hydroxypropylamine)-2H-1,2,4-benzothiadiazine 1,1-dioxide ( 2b ) via a Smiles rearrangement.  相似文献   

12.
Condensation of 2-nitrothiophenol with ethyl propiolate afforded 3-(2-nitrophenylthio)propenoate. Oxidation of sulfur atom to sulfone group gave ethyl 3-(2-nitrophenylsulfonyl)propenoate, which underwent condensation with tosyl methylisocyanide (TosMIC) to yield ethyl 4-(2-nitrophenylsulfonyl)pyrrole-3-carboxylate. Reduction of nitro group afforded ethyl 4-(2-aminophenylsulfonyl)-1H-pyrrole-3-carboxylate, which was cyclized to 2H-pyrrolo[3,4-b][1,5] benzothiazepin-10(9H)-one 4,4-dioxide. Similar procedure was used for the synthesis of 9,10-dihydro-10-methyl-2H-pyrrolo[3,4-b][1,5]benzothiazepine 4,4-dioxide.  相似文献   

13.
The reaction of 1-O-hexadecyl-2-O-methyl-sn-glycerol with 2,3,6,2′,3′,4′,6′-hepta-O-acetyl-α-lactosylphosphoramidate or α-maltosylphos-phoramidate in the presence of trimethylsilyl triflate and molecular sieves afforded 1-O-hexadecyl-2-O-methyl-3-O-(2,3,6,2′,3′,4′,6′-hepta-O-acetyl-β-lactosyl)-sn-glycerolipid or β-maltosyl-sn-glycerolipid stereoselectively in moderate yields after column chromatography. Alkaline hydrolysis of the O-peracetyl glycerolipids gave the desired β-glycolipids 1 and 2.  相似文献   

14.
A series of CN-bridged trinuclear Ru complexes of the general structure [RuL2(μ-(CN)Ru(CN)L2′)2] where L is 2,2′-bipyridine-4,4′-dicarboxylic acid and L′ is 2,2′-bipyridine ( 1 )2,2′-bipyridine-4,4′-dicarboxylic acid ( 2 ), 4,4′-dimethyl-2,2′-bipyridine ( 3 ), 4,4′-diphenyl-2,2′-bipyridine ( 4 ), 1,10-phenanthroline ( 5 ), and bathophenanthrolinedisulfonic acid ( 6 ) have been synthesized, and their spectral and electrochemical properties investigated. The two carboxylic functions on the 2,2′-bipyridine ligand L serve as interlocking groups through which the dye is attached at the surface of TiO2 films having a specific surface texture. The role of these interlocking groups is to provide strong electronic coupling between the π* orbital of the 2,2′-bipyridine and the 3d-wave-function manifold of the conduction band of the TiO2, allowing the charge injection to proceed at quantum yields close to 100 %. The charge injection and recombination dynamics have been studied with colloidal TiO2, using laser photolysis technique in conjunction with time-resolved optical spectroscopy. Photocurrent action spectra obtained from photo-electrochemical experiments with these trinuclear complexes cover a very broad range in the visible, making them attractive candidates for solar light harvesting. Monochromatic incident photon-to-current conversion efficiencies are strikingly high exceeding 80% in some cases. Performance characteristics of regenerative cells operating with these trinuclear complexes and ethanolic triiodide/iodide redox electrolyte have been investigated. Optimal results were obtained with complex 1 which gave a fill factor of 75 % and a power conversion efficiency of 11.3% at 520 nm.  相似文献   

15.
The palladium(0)‐catalyzed polyaddition of bifunctional vinyloxiranes [1,4‐bis(2‐vinylepoxyethyl)benzene ( 1a ) and 1,4‐bis(1‐methyl‐2‐vinylepoxyethyl)benzene ( 1b )] with oxygen nucleophiles such as hydroquinone and bisphenol A gave new unsaturated polyethers containing an allyl aryl ether moiety and pendant hydroxy groups. The polyaddition with 1a was largely affected by the phosphine ligands employed and the reaction temperature. The polyaddition with hydroquinone and bisphenol A was conducted at room temperature for 24 h in tetrahydrofuran in the presence of PPh3 and gave the desired polyethers in good yields, the number‐average molecular weights (Mn) of which were 5700 and 7700, respectively. 1,2‐Bis(diphenylphosphino)ethane (dppe) was not effective in the polyaddition with 1a . The polyaddition of 1b with hydroquinone and bisphenol A gave the corresponding polyethers despite the kinds of ligands employed (PPh3 and dppe), contrary to the polyaddition with 1a . The polyaddition of 1b with 4,4′‐biphenol was also carried out in the presence of Pd2(dba)3 · CHCl3/dppe as a catalyst (where dba is dibenzylideneacetone) and afforded the expected polyether with a high Mn value (Mn = 24,900). In addition, vinyloxirane 1b could reacted with racemic 1,1′‐bi‐2‐naphthol to give the corresponding polyether in a good yield. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 476–482, 2003  相似文献   

16.
In the presence of H2SO4, the sydnone rings were condensed with formaldehyde to give diarylation in 4-position of the sydnone rings: 3,3′-diphenyl-4,4′-methylenedisydnone is obtained from 3-phenylsydnone in 61% yield, and 3,3′-di-p-methylphenyl-4, 4′-methylenedisydnone is obtained from 3-(4′-methylphenyl)sydnone in 70% yield. And the possible reaction mechanism is discussed.  相似文献   

17.
A reaction of 2-aminobenzenesulfonamide ( 1 ) with 2-chloroethyl or 3-chloropropyl isothiocyanate in isopropanol afforded 2-(2′-chloroethylthioureido)- and 2-(3′-chloropropylthioureido)benzenesulfonamides ( 2a,b ) in 67% and 55% yield respectively. Treatment of 2a,b with triethylamine in methanol at room temperature furnished 3-(2′-aminoethylthio)- and 3-(3′-aminopropylthio)-2H-1,2,4-benzothiadiazine 1,1-dioxides ( 9a,b ) in quantitative yield. Heating 2b to reflux in methanol under neutral conditions gave 9b but in the form of the hydrochloride 8b which could be converted into the free base 9b by treating with ammonia water. When compounds 2a,b were treated with triethylamine in methanol at elevated temperature, 3-(2′-mercaptoethylamino)- and 3-(3′-mercaptopropylamino)-2H-1,2,4-benzothiadiazine 1,1-dioxides ( 10a,b ) were obtained in good yield. Alternatively, 10a,b could also be prepared from 9a,b in 95% and 77% yield respectively.  相似文献   

18.
Synthesis of Bromosubstituted Butenolides II . Methyl 4,4′-dibromosenecioate ( 2 ) was prepared by double N-bromosuccinimide bromination of methyl senecioate ( 1 ) and converted to methyl 4,4′-diiodo-senecioate ( 3 ) with sodium iodide and to 3-bromomethyl-2-buten-4-olide ( 4 ) with aqueous hydrobromic acid. A mixture of methyl (Z)- and (E)-4-bromosenecioate ( 8 and 9 ) yielded 3-methyl-2-butenolide ( 5 ) with aqueous hydrobromic acid and a mixture of (Z)-and (E)-4-methoxy-senecioic acid ( 10 and 11 ) with methanolic potassium hydroxide. N-Bromosuccinimide treatment of the butenolide 5 afforded 4-bromo-3-methyl-2-buten-4-olide ( 6 ) and 4,4-dibromo-3-methyl-2-buten-4-olide ( 7 ).  相似文献   

19.
Herein, we provide a new protocol to reduce various transition-metal complexes by using organosilicon compounds in a salt-free fashion with the great advantage of generating pure low-valent metal species and metallic(0) nanoparticles, in sharp contrast to reductant-derived salt contaminants obtained by reduction with metal reductants. The organosilicon derivatives 1,4-bis(trimethylsilyl)-2,5-cyclohexadiene ( 1 a ), 1-methyl-3,6-bis(trimethylsilyl)-1,4-cyclohexadiene ( 1 b ), 1,4-bis(trimethylsilyl)-1,4-diaza-2,5-cyclohexadiene ( 2 a ), 2,5-dimethyl-1,4-bis(trimethylsilyl)-1,4-diaza-2,5-cyclohexadiene ( 2 b ), 2,3,5,6-tetramethyl-1,4-bis(trimethylsilyl)-1,4-diaza-2,5-cyclohexadiene ( 2 c ), and 1,1′-bis(trimethylsilyl)-1H,1′H-4,4′-bipyridinylidene ( 3 ) all served as versatile reductants for early transition-metal complexes and produced only easy-to-remove organic compounds, such as trimethylsilylated compounds and the corresponding aromatics, for example, benzene, toluene, pyrazine, and 4,4′-bipyridyl, as the byproducts. The high solubility of the reductants in organic solvents enabled us to monitor the catalytic reactions directly and to detect any catalytically active species so that we could elucidate the reaction mechanism.  相似文献   

20.
1,4-Diphenyl-2,3-dioxabicyclo[2.2.1]hept-5-ene ( 2 ), on treatment with a catalytic amount of trimethylsilyl trifluoromethanesulfonate (Me3SiOTf) in CH2Cl2 at ?78°, reacts with excess (?)-menthone ( 10 ) to give (1S,2S,4′aS,5R,7′aS)-4′a,7′a-dihydro-2-isopropyl-5-methyl-6′,7′-diphenylspiro[cyclohexane-1,3′-[7′H]cyclopenta-[1,2,4]trioxine] ( 11 ) and its (1R,2S,4′aR,5R,7′aR)-diastereoisomer 12 in a 1:1 ratio and in 21% yield. Repeating the reaction with 1.1 equiv. of Me3SiOTf with respect to 2 affords 11 , 12 , and (1S,2S,3′a.R,5R,6′aS)-3′a,6′a-dihydro-2-isopropyl-5-methyl-3′a-phenoxy-5′-phenylspiro[cyclohexane-l,2′-[4′H]cyclopenta[1,3]dioxole] ( 13 ) together with its(1R,2S,3′aS,5R,6′aR)-diastereoisomer 14 in a ratio of 3:3:3:1 and in 56% yield. (+)-Nopinone( 15 ) in excess reacts with 2 in the presence of 1.1 equiv. of Me3SiOTf to give a pair of 1,2,4-trioxanes ( 16 and 17 ) analogous to 11 and 12 , and a pair of 1,3-dioxolanes ( 18 and 19 ) analogous to 13 and 14 , in a ratio of 8:2:3:3 and in 85% yield. (?)-Carvone and racemic 2-(tert-butyl)cyclohexanone under the same conditions behave like 15 and deliver pairs of diastereoisomeric trioxanes and dioxolanes. In general, catalytic amounts of Me3SiOTf give rise to trioxanes, whereas 1.5 equiv. overwhelmingly engender dioxolanes. Adamantan-2-one combines with 2 giving only (4′aRS,7′aRS)-4′a,7′a-dihydro-6′.7′a-diphenylspiro[adamantane-2,3′-[7′H]cyclopenta[1,2,4]trioxine] in 98% yield regardless of the amount of Me3SiOTf used. The reaction of 1,4-dipheny 1-2,3-dioxabicyclo[2.2.2]oct-5-ene ( 32 ) with 10 and 1.1 equiv. of Me3SiOTf produces only the pair of trioxanes 33 and 34 homologous to 11 and 12 . Treatment of the (S,S)-diastereoisomer 33 with Zn and AcOH furnishes (1S,2S)-1,4-diphenylcyclohex-3-ene-1,2-diol. The crystal structures of 11 – 13 and 16 are obtained by X-ray analysis. The reaction courses of 10 and the other chiral cyclohexanones with prochiral endoperoxides 2 and 32 to give trioxanes are rationalized in terms of the respective enantiomeric silylperoxy cations which are completely differentiated by the si and re faces of the ketone function. The origin of the 1,3-dioxolanes is ascribed to 1,2 rearrangement of the corresponding trioxanes, which occurs with retention of configuration of the angular substituent.  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