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 937 毫秒
1.
Iodoperfluooralkylation of terminal alkenes and alkynes is effectively photo‐promoted by benzophenone 2 (BP) or the photoreducible copper(II) complex 1 . In particular, BP at 1 mol% in methanol upon 365 nm irradiation with a low‐pressure mercury lamp (type TLC=thin layer chromatography, 6 W) results in a fast reaction with excellent reaction yields. Complex 1 and BP 2 exhibited very similar reactivity, suggesting that the reactions involving 1 are likely to be governed by the benzophenone photoactivation processes, rather than copper(I)/(II) redox processes. Mechanistic investigations using transient absorption spectroscopy revealed that a deactivation pathway of the benzophenone triplet (3BP*) is via its reaction with the methanol solvent. We propose that the generated radicals, in particular .CH2OH, play a key role in the initiation step forming Rf. by reacting with RfI, Rf. then entering a radical chain cycle. 1H NMR studies provided evidence that a substantial amount (∼7% NMR yield) of the hemiacetal CH3OCH2OH is formed, i.e., the possible by‐product of the reaction between .CH2OH and RfI. Finally, DFT calculations indicate that a triplet‐triplet energy transfer (TTET) process from 3BP* to perfluorooctyl iodide (C8F17I) is unlikely or should be rather slow under the reaction conditions, consistent with the transient absorption studies.

  相似文献   


2.
Synthesis and Reactions of Chiral Dithiocarbamates Derived from (R)-(−)- or (S)-(+)-2- The synthesis, diastereoselective alkylation reactions, dithiocarboxylation, and aldol condensation of several substituted methyl (R1 CH2) (S)-2-(methoxymethyl)-pyrrolidine-1-dithiocarboxylates (S)- 2 and of the corresponding (R)-derivatives (R)- 2 are described. The new enantiomeric dithiocarbamates (S) -2a – e , and (R) -2a – d are obtained by reaction of (S)-(+)-[(S) -1 ] or (R)-(−)-2-(methoxymethyl)-pyrrolidine [(R) -1 ], respectively, with carbon disulfide in dry methanol/anhydrous sodium acetate and the appropriate alkylating agent. The cyclic ketene dithioacetals (S) -3 and (R) -3 are formed by dithiocarboxylation procedure of (S) -2a and (R) -2a whereas (S) -6 and (R) -6 are obtained by aldol reaction with isobutyraldehyde. (S)- 2c , d and (R) -2c , d react in a diastereoselective manner after deprotonation with n-BuLi or LiTMP/LiBr at −78°C in THF with alkyl halides to the enantiomeric compounds 4a /ent -4a, 4b /ent -4b and 5 /ent -5 , respectively.  相似文献   

3.
Reaction of anthracenide A ·‐ with N‐benzoylaziridines 1a,b forms charged radicals 3a,b by single electron transfer and homolytic ring opening. Reactions follow that are known or expected as e.g. coupling with position 9 of A ·‐ forming dihydroanthracene anions 9a,b that yield amidoethylated dihydroanthracenes 10a,b , or react with 1a,b giving finally 9,10‐bis‐amidoethylated dihydroanthracenes 11a,b . Results depend on experimental conditions and on the counter ions Na+ or Li+. Coupling is not regiospecific: contributions by positions 2 and 1 reach 29% or 4%, respectively, of total coupling with the primary radical 3a ; much higher contributions are possible with Li. Product 21s (probably 3,3′‐disubstituted tetrahydrobianthryl) may arise by hydrogen detachment from the first intermediate ( 29 ) of coupling with position 2 and dimerization of the formed 2‐substituted A ·‐ ( 30 ). Coupling products may be fully aromatized or may be hydroxylated in one of the benzylic positions. With counter ion Li+ a non‐SET reaction of 1a with the dimer of A ·‐ is indicated by the isolation of 9‐benzoyl‐dihydroanthracene 15 and by 19% yield of 16a (aromatized 10a ). Reaction of 3b with anthracene is indicated by 10,10′‐disubstituted tetrahydrobianthryl 37 .  相似文献   

4.
Benzoin (B), benzoinacetate (BA), benzoinmethylether (BME) and benzoinisopropylether (BIPE) were irradiated at room temperature in benzene solution in the presence of styrene (St), methyl methacrylate (MMA), vinylacetate (VAc) or acrylonitrile (AN). Flash photolysis experiments at λ=347 nm yielded (a) rate constants kq (in 1 mol-1s-1) of the reaction between excited sensitizers and monomers: 8·109 (B/St), 5·108 (B/MMA), 5·109 (BA/St), 8·108 (BA/MMA); (b) rate constants KR.+M (in 1mol-1 s-1) of the reaction between sensitizer radicals and monomers: about 1.5·105 (BME/St, BME/VAc, BA/VAc, B/VAc), 9· 104 (BME/MMA), 2·104 (BME/AN). The reaction R·+M caused in certain cases (B/St, B/VAc, BME/St) the formation of an additional optical absorption after the flash. Stationary irradiations at λ>320 nm of monomer solutions (5mol/1) showed that BA is least effective. Rates of polymerization increased in the series BA<B<BIPE<BME. For the systems containing St or MMA it was found that ?i=i+0.6αR (?i=quantum yield for the initiation of kinetic chains, αR =fraction of triplets converted to radicals). The fraction of radicals starting kinetic chains is ca. 0.3 in these cases.  相似文献   

5.
The reduction of selenious ions to its elemental state has been found to be a function of selenious ion concentration, solution acidity, gas flow rate and temperature. The rate equation can be written as: where α = ?1/2, β = 1, γ = 0.9, E = ?53 411 kJ kg?1mol?1, R = 8314 kJ kg?1 mol?1 K?1 and k0 = 194. The reaction temperature starts to decrease when the conversion rate reaches a maximum. This observation can be used to determine the end of the reaction.  相似文献   

6.
Synthesis of Cyclic N,O- and N,S-Heterocyclic Compounds by Condensation of Fatty Acid Alkylolamides with Carbonyl Compounds and Phosphorous Pentasulfide Respectively Fatty acid alkylolamides of the type RCONHCHR1 (CH2)n CHR2OH (R: alkyl rest with 2 to 17 carbon atoms, cycloalkyl or aryl rest; R1 and R2: alkyl rests with 1 to 3 carbon atoms or hydrogen; and n = 0–2) are converted by acidic condensation with lower aliphatic or aromatic aldehydes and ketones respectively to mono- and disubstituted N-acyl-oxazolidines, N-acyl-1,3-tetrahydrooxazines and N-acyl-1,3-hexahydrooxazepines. Furthermore, new types of 2-alkyl-4,5-dihydro-6H-1,3-thiazine (n = 1) besides monoalkyl substituted Δ-2-thiazolines (n = 0) were obtained from fatty acid alkylolamides by condensation with P2S5. A large number of the compounds prepared show excellent bacteriostatic and fungistatic action.  相似文献   

7.
Reaction of β-diketiminate lithium salts, nacnacRLi(THF) (R = 2,6-xylyl (1a), R-CH(Me)Ph (1b)), with ZrCl4 or ZrCl4·THF yielded the mono-diketiminate complexes nacnacRZrCl3(THF), 2a and 2b. Complex 2a did not react further with a second equivalent of 1a, but with CpLi or IndLi to form pseudo-tetrahedral complexes 3 and 4. No product could be isolated from reaction of 2b with CpLi. Coordination of the diketiminate ligand in 2a is best described as a distorted κ2-, in-plane coordination. Complexes 3 and 4, however, display a haptotropic shift of the diketiminate ligand and its coordination is best described as κ22-coordination. Although not previously described, analyses of available structures in the literature reveal that non-symmetric κ22-coordination is an independent coordination mode, which exists next to the previously described “η5-like” coordination.  相似文献   

8.
An additive‐ and transition metal‐free perfluoroalkanesulfonylation of alkynyl(phenyl)iodonium tosylates with sodium perfluoroalkanesulfinates (RfnSO2Na) is described. The poorly nucleophilic RfnSO2Na reacted with alkynyl(phenyl)iodonium salts in dichloromethane at room temperature under a nitrogen atmosphere for 5–60 minutes to afford a variety of acetylenic triflones and alkynyl perfluoroalkyl sulfones in good to quantitative yields. The position of substituents on the phenyl rings of the arylethynyl moiety in the iodonium salts had a big influence on the reaction. The formation of five‐membered cyclic vinyl sulfones suggested that the reaction proceeds via an alkylidene carbene intermediate. Furthermore, successful scaling‐up of the reaction demonstrates the practicality of the new method. Advantages of the method include short reaction times, mild conditions, and the easy access to perfluoroalkanesulfonylation reagents (RfnSO2Na).

  相似文献   


9.
The effects of temperature, structure and sulphuric acid concentration on the selectivity of the oxidation of aliphatic ketones (R1COR2) ( 1a – g ) by thallic sulphate have been investigated. With increasing temperature the quantity of internal α-hydroxyketone ( 3a – d ) decreases and the quantity of 1-hydroxyketone ( 2a – d ) increases in the oxidation of methyl alkyl ketones (R2>CH3) ( 1a – d ). The same concerns to the oxidation products of hexan-3-one ( 1e ): 2-hydroxy-hexan-3-one ( 2e ) and 4-hydroxy-hexan-3-one ( 3e ), respectively. “The inverse selectivity temperature” (IST) for oxidation of linear methyl alkyl ketones ( 1a – c ) and hexan-3-one ( 1e ) has been found. With the use of the linear free-energy relationship it has been found that the selectivity of the reaction decreases with increasing the polar and steric effects of substituents R1, R2. With increasing the sulphuric acid concentration the selectivity of the oxidation of methyl alkyl ketones ( 1a , 1d ) increases.  相似文献   

10.
Free Radical Reactions of N-Heterocyclic Compounds. XI. Reaction of 3-Methyl-pyrazolin-5-ones with Phenoxy Radicals Pyrazolin-5-ones ( 3a–i ) were oxidized with 2,4,6-trisubstituted phenoxy radicals ( 2a–d ) to the corresponding radicals ( 4a–i ), which dimerised or combined with phenoxy radical ( 2a ) depending on the R1- and R4-substituents in ( 3 ). In the case of 3-methyl-1-phenyl-pyrazolin-5-one ( 3f ) the primary radical combination products were not found, but the corresponding quinone methide ( 17 ) and the o-phenol derivative ( 18 ) were isolated. Products and yields have been investigated as a function of mol ratio substrate: oxidant and solvent. The radical combination products ( 7–10 ) could de-tertbutylated in the presence of aluminium chloride or in the presence of trifluoroarcetic acid, forming heterocyclic substituted phenols ( 21 ) and ( 22 ).  相似文献   

11.
The electrochemical oxidation of various substituted aryl cycloheptatrienes and cyclohepta‐1,3,5‐triene, the parent compound, in acetonitrile is investigated with the aid of cyclic voltammetry (CV), experiments at the rotating ring disk electrode (RRDE) and controlled potential electrolysis. 1‐(p‐methoxyphenyl)‐cycloheptatriene 2c , 1‐ and 3‐(p‐di‐methylaminophenyl)‐cycloheptatriene 2d , 3d are converted by anodic oxidation into the radical cations the main reaction of which is not the deprotonation. According to the proposed mechanism the electrochemical oxidation proceeds along the radical cation dimerization. The dimer cations ( 1 22+) decompose under deprotonation into tropylium ions 4 and starting compound. In the case of 1‐(p‐dimethylaminophenyl)‐cycloheptatriene 2d the deprotonation reaction can compete with the dimerization. The oxidation of the cycloheptatriene compounds is kinetically controlled by the homogeneous chemical steps rather than by the initial electron transfer.  相似文献   

12.
Photochemistry and Photophysics of 3-(2-Isoxazolinyl)-phenylketones 3-Benzoyl-Δ2-1,2-oxazolines ( 1–6 ) are formed by 1,3-dipolar cycloaddition between benzoylnitril-N-oxide ( 8 ) and dihydrofurane 9 or 1,3-dioxep-5-enes ( 10a–c ). The preparative yields are small due to the competitive dimerization of the dipole 8 . Two stereoisomers are obtained by using 2-substituted 1,3-dioxep-5-enes as dipolarophiles. The different steric position of the substituents in 3–6 gives rise to different spectral data. The synthesized ketones possess triplet states with a high degree of charge transfer character. Therefore, the ability to H-abstraction reaction from alcohols is small. For ketone 2 and methanol as H-donor a rate constant of k = 4,1 · 102 M−1s−1 is determined. Also by electron transfer reactions with triethylamine and some onium compounds the reactivity of the T1 of the ketones 1–6 is less compared to those of nπ* excited ketones. The photolysis of the ketones takes place very unselectively and leads to a product mixture. The quantum yields for the decay of the ketones are 10−2 to 10−3.  相似文献   

13.
Nine new cerebrosides 1a–d , 2a , 2b , 3a–c were found in the extract of a Far‐Eastern glass sponge Aulosaccus sp. (class Hexactinellida). These β‐d ‐glucopyranosyl‐(1 → 1)‐ceramides contain sphingoid bases (2S,3S,4R,11Z)‐2‐aminoeicos‐11‐ene‐1,3,4‐triol (in 1a – d ), (2S,3S,4R,13Z)‐2‐aminoeicos‐13‐ene‐1,3,4‐triol (in 2a , b ) and (2S,3S,4R,13S*,14R*)‐2‐amino‐13,14‐methylene‐eicosane‐1,3,4‐triol (in 3a – c ), which are N‐acylated by (2R,15Z)‐2‐hydroxydocos‐15‐enoic (in 1a , 2a , 3a ), (2R,16Z)‐2‐hydroxytricos‐16‐enoic (in 1b , 2b , 3b ), (2R,17Z)‐2‐hydroxytetracos‐17‐enoic (in 1d ) and (2R)‐2‐hydroxydocosanoic (in 1c , 3c ) acids. The monoenoic and cyclopropane‐containing sphingoid bases of compounds 1a–d , 2a , 2b , 3a–c have not been found previously in any sphingolipids. The structures of the cerebrosides were elucidated on the basis of 1H‐, 13C‐NMR spectroscopy, mass spectrometry, optical rotation data and chemical transformations. A simplified method for the assignment of the absolute configuration of 2‐hydroxy fatty acids by GC analysis of their (2R)‐ and (2S)‐oct‐2‐yl esters was proposed.  相似文献   

14.
The ability of dibenzothiophene S-oxide (1) to photochemically induce strand breaks in plasmid DNA was studied under anaerobic conditions. DNA cleavage is monitored by the conversion of closed circular pUC19 DNA (form I) to the nicked (form II) and linear forms (form III) using densitometer digital imaging of ethidium stained gels. In buffered aqueous–acetonitrile (9:1) solutions the single-strand cleavage is efficient and does not require an alkaline reaction workup. Photodeoxygenation of 1 in buffered aqueous–acetonitrile (9{:}1) solutions containing benzene led to the production of phenol. The effect of solvent deuteration does not support the involvement of 1O2 from a sulfoxide dimerization reaction nor a sensitized photooxygenation reaction. The results are interpreted in terms of a sulfoxide photodeoxygenation via oxygen atoms [O(3P)] where oxidation of DNA can lead to single-strand breaks. Since the reaction of O(3P) atoms with water itself is endoergic, we propose that hydroxyl radicals do not intervene in the DNA cleaving reaction.  相似文献   

15.
The synthesis of 2-substituted cis-octahydroindolones from the reaction of cis-2-amino-1-alkenylcyclopentanols with aldehydes was studied to examine whether stereoselection in the aza-Cope–Mannich reaction could be controlled by the nature of the nitrogen substituent. 2-Alkylamino-1-(1-phenylethenyl)cyclopentanols 7 and 8 , which contain nitrogen substituents of widely differing size (Me and CHPh2), were condensed with four aldehydes to give oxazolidines 9a–d and 10a-c. Rearrangement of these intermediates at 23–60 °C, in the presence of 0.9 equiv of (±)-10-camphorsulfonic acid in acetonitrile, gave cis-octahydroindolones 11a-d and 12a–c in yields of 77–95%. Using a combination of single-crystal X-ray crystallography, 1H nOe measurements, and comparisons with known materials it was established that the N-methyl oxazolidines 9a–d provided exclusively cis-octahydroindolones having the 2-substituent trans to the angular substituents, while N-benzhydryl analogs 10a–c provided exclusively the all-cis products 12a–c. These results are interpreted to mean: (1) When the nitrogen substituent is small (Me), the stereochemistry-determining [3,3]-sigmatropic rearrangement occurs preferentially through a transition-state topography having the R2 substituent oriented quasiequatorially ( 14 → 15 → 16 ); (2) When this substituent is large (CHPh2), destabilizing steric interactions between the vicinal R1 and R2 substituents causes the rearrangement to occur preferentially through the alternate iminium ion stereoisomer ( 17 → 18 → 19 ).  相似文献   

16.
Electrochemically induced synthesis of β-lactams, by cyclization (via C3C4 bond formation) of haloamides XCHR1CONR2CHR3R4 (X: Br, Cl), has been achieved in suitable solvent-supporting electrolyte solutions previously electrolyzed under galvanostatic control. The yields and the stereochemistry of the process are affected by the nature of substituents R1-R4 and of solvent-supporting electrolyte solutions and by the electrolysis conditions.  相似文献   

17.
A highly efficient head‐to‐tail dimerization of a styrene was developed using a cationic palladium(II)‐catalyzed selective C C bond forming reaction. The complex [AllylPd(PPh3)]+OTf, which is believed to generate ‘palladium hydride’ (Pd H), catalyzed the dimerization of various styrenes in excellent yields as single isomers. This Pd(II)‐catalyzed reaction provides a new economical C C bond forming method.

  相似文献   


18.
Electron transfer reactions between the photoexcited triplets of several porphyrin and chlorophyll precursors, 3P (electron donors), and neutral duroquinone, DQ (electron acceptor), were studied by selective laser excitation-Fourier transform (FT)-EPR spectroscopy. Four different P, which are related to the photosynthetic apparatus, were examined in ethanol solutions at −243 K: pyrochlorophyll a (pChla), zinc-pyrochlorophyll a (ZnpChla), zinc-tetraphenylporphyrin (ZnTPP), and magnesium-tetraphenylporphyrin (MgTPP). This highly time-resolved EPR spectroscopy (∼10 ns delay time, ẗd, between the laser excitation and the detection pulse) permits differentiation between the various CIDEP mechanisms, i.e., triplet mechanism (TM) and radical-pair mechanism (RPM), associated with the electron transfer process. The spectra of the DQ.− generated by the different triplet precursors strongly depend on two inherent parameters characteristic of the molecular system: (1) the zero-field splitting (ZFS) and the selective singlet-triplet (Ti l̊ S1) population rates, Ai (i = x, y, z); (2) the difference in g-values and hyperfine components between the donor and acceptor radicals. Since the differences in g-values and hyperfines are very nearly the same in all four P.+-DQ.− systems examined, the different results, particularly in the time range 10 ns < ẗ<d < 300 ns, are attributed to triplet spin memory transfer via TM. A quantitative spectral analysis is presented to describe the time evolution FT-EPR spectra. This method is used to determine the following dynamic processes: (1) spin lattice relaxation rates of the triplet precursors in solution; (2) donor-acceptor electron transfer reaction rates; (3) spin lattice relaxation rates of the acceptor radical.  相似文献   

19.
A series of diorganotin dicarboxylates of the general formula (CH3)2Sn(OCOCHR3CHR2GeR1)2 where R1=(C6H5)3, (P-CH3C6H4)3, N(CH2CH2O)3, R2=C6H5, H, CH3, P-CH3OC6H4, P-ClC6H4, P-CH3C6H4, R3=CH3 and H, have been synthesized by the reaction of dimethyltin oxide with germanium substituted propionic acid in 1:2 molar ratio in toluene. The H2O formed was removed azeotropically using a Dean and Stark apparatus. All the compounds have been characterized by IR, multinuclear (1H, 13C, 119Sn) NMR, mass and Mössbauer spectroscopies. All compounds were found to have potential activity against bacteria.  相似文献   

20.
The tandem nucleophilic addition‐cyclization reaction of o‐alkynylbenzaldehydes or o‐alkynylacetophenones 2 with dialkyl phosphites or dialkyl phosphonothioates 1 took place very smoothly in the presence of 1,8‐diazabicyclo[5.4.0]undec‐7‐ene (DBU) in THF at room temperature. In all cases, the reaction proceeded in a regioselective manner leading to the 5‐exo‐dig products 3 in excellent yields. The phenomenon of a 1,5‐sigmatropic hydrogen shift or a 1,5‐sigmatropic methyl shift was observed in this reaction depending on the different substituent groups such as R3 in the o‐alkynylbenzaldehyde or o‐alkynylacetophenone 2 substrates.  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