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 15 毫秒
1.
The structures and energies of the 1:1 acetylene/cyanoacetylene, acetylene/dicyanoacetylene and cyanoacetylene/dicyanoacetylene complexes in solid argon matrices have been investigated using FT-IR spectroscopy and ab initio calculations, at the B3LYP/6-31G** level of theory. For the three complexes, predicted frequency shifts for the L shaped structures, characterized by a hydrogen bond between the nitrogen of the cyano group and the acetylenic proton, were found to be in good agreement with those experimental. Only in the case of acetylene/cyanoacetylene complex, we obtained a second minimum with a T shaped structure characterized by an interaction between the proton of cyanoacetylene and the Π system of acetylene. It appears clearly that HC3N acts as an electrophile or as a nucleophile in these complexes.  相似文献   

2.
The reactivity of intermolecular complexes cyanoacetylene:acetylene and dicyanoacetylene:acetylene, trapped in solid argon matrixes at 10 K and irradiated with vacuum UV, has been studied. FTIR measurements, together with (2)H, (15)N, (13)C labeling experiments and with density functional theory (DFT) calculations (B3LYP/aug-cc-pVTZ), pointed to the formation of cyanodiacetylene HC(5)N (cyanobutadiyne). This synthetic route is potentially important for chemical models of the Titan's atmosphere.  相似文献   

3.
Upon irradiation with 300-nm UV light, the photolysis of diazomalonates in benzene unexpectedly affords 2,6-dicarboxylate bicyclo[3.2.0]hepta-2,6-dienes in low yields. These products are proven to be derived from cyclohepta-1,3,5-triene intermediates presumably via a tandem 1,5-carboxylate migration/[2+2] cycloaddition sequence.  相似文献   

4.
The UV absorption spectra of CF(3)CHO, C(2)F(5)CHO, C(3)F(7)CHO, C(4)F(9)CHO, CF(3)CH(2)CHO, and C(6)F(13)CH(2)CHO were recorded over the range 225-400 nm at 249-297 K. C(x)F(2)(x)(+1)CHO and C(x)F(2)(x)(+1)CH(2)CHO have broad absorption features centered at 300-310 and 290-300 nm, respectively. The strength of the absorption increases with the size of the C(x)F(2)(x)(+1) group. There was no discernible (<5%) effect of temperature on the UV spectra. Quantum yields for photolysis at 254 and 308 nm were measured. Quantum yields at 254 nm were 0.79 +/- 0.09 (CF(3)CHO), 0.81 +/- 0.09 (C(2)F(5)CHO), 0.63 +/- 0.09 (C(3)F(7)CHO), 0.60 +/- 0.09 (C(4)F(9)CHO), 0.74 +/- 0.08 (CF(3)CH(2)CHO), and 0.55 +/- 0.09 (C(6)F(13)CH(2)CHO). Quantum yields at 308 nm were 0.17 +/- 0.03 (CF(3)CHO), 0.08 +/- 0.02 (C(4)F(9)CHO), and 0.04 +/- 0.01 (CF(3)CH(2)CHO). The quantum yields decrease with increasing size of the C(x)F(2)(x)(+1) group and with increasing wavelength of the photolysis light. The photolysis quantum yield at 308 nm for CF(3)CHO measured here is a factor of at least 8 greater than that reported previously. Photolysis is probably the dominant atmospheric fate of C(x)F(2)(x)(+1)CHO (x = 1-4) and is an important fate of C(x)F(2)(x)(+1)CH(2)CHO (x = 1 and 6). These results have important ramifications concerning the yield of perfluorocarboxylic acids in the atmospheric oxidation of fluorotelomer alcohols.  相似文献   

5.
本文研究了由偏二氯乙烯-丙烯酸甲酯共聚合物(VdCl2-MA)和五甲氧基红指示剂组成的酸敏变色记录材料的光谱特性,测定了其光分解量子产率.通过添加光敏剂使原来只有在254nm紫外光下曝光才能发生光分解反应的酸敏变色材料在35nm紫外光下曝光也可发生光分解而变色.酸敏变色材料在254nm处紫外光曝光量子产率为0.032-0.020;在35nm处紫外光曝光量子产率为0.110-0.034.光敏剂的加入使酸敏变色材料量子产率提高了2-4倍.  相似文献   

6.
Photolysis of 2,5-dichloro-3,6-bis(dialkylamino)-[1,4]benzoquinone (1) at 254 nm or 300–500 nm produces the corresponding cyclized products (2) via the electron transfer-proton transfer sequence in a variety of solvents in quantitative yields within 10 min. In contrast, photolysis of 2,5-dichloro-3,6-bis(dialkylamino)-[1,4]benzoquinone (1) at >500 nm light source affords the elimination products (3) in good yields via the δ-hydrogen abstraction.  相似文献   

7.
Abstract— Photolysis of tritium-labelled thymine-derived photoproducts by 254-nm ultraviolet radiation (u.v.) in conidia of Streptomyces griseus was measured by chromatography of cell hydrolysates. The relative photolysis cross-sections of uracilthymine dimer (UT○) at various wavelengths are the same as those of thymine-thymine dimer (TT○), and their ratios at 313, 365, 405 and 436 nm are 2:1:2:3. Except at 436 nm, these relative values agree very well with cross-sections previously reported for photoreactivation of u.v. killing in this organism, leading to the conclusion that photoreactivation in the wild type is due to repair of cyclobutane-type pyrimidine dimers. In a mutant showing restricted photoreactivation (S. griseus PHR-1), post-u.v. treatments at the above wavelengths did not affect UT○ and TT○ in the conidia, supporting the earlier suggestion that this organism does not contain active PR enzyme. Another u.v. photoproduct, the precursor of a pyrimidine adduct (PO-T) that appears in cell hydrolysates, was removed from both wild-type and mutant cells very efficiently at 313 nm. This is presumably a direct photochemical reaction. In addition, in wild-type cells, the precursor of PO-T appeared to be inefficiently removed photoenzymatically at all wavelengths. Removal of the precursor of PO-T appears to be biologically significant, however, only in the mutant.  相似文献   

8.
Cyclobutane pyrimidine dimers and pyrimidine(6-4)pyrimidone adducts represent the two major classes of far-UV-induced DNA photoproducts. Because of the lack of appropriate detection methods for each individual photoproduct, little is known about the effect of the sequence on their formation. In the present work, the photoproduct distribution obtained upon exposure of a series of dinucleoside monophosphates to 254 nm light was determined. In the latter model compounds, the presence of a cytosine, located at either the 5′- or the 3′- side of a thymine moiety, led to the preferential formation of (6-4) adducts, whereas the cis-syn cyclobutane dimer was the main thymine-thymine photoproduct. In contrast, the yield of dimeric photoproducts, and particularly of (6-4) photoadducts, was very low upon irradiation of the cytosine–cytosine dinucleoside monophosphate. However, substitution of cytosine by uracil led to an increase in the yield of (6-4) photoproduct. It was also shown that the presence of a phosphate group at the 5′- end of a thymine-thymine dinucleoside monophosphate does not modify the photoproduct distribution. As an extension of the studies on dinucleoside monophosphates, the trinucleotide TpdCpT was used as a more relevant DNA model. The yields of formation of the thymine-cytosine and cytosine–thymine (6-4) photoproducts were in a 5:1 ratio, very close to the value obtained upon photolysis of the related dinucleoside monophosphates. The characterization of the two TpdCpT (6-4) adducts was based on H NMR, UV and mass spectroscopy analyses. Additional evidence for the structures was inferred from the analysis of the enzymatic digestion products of the (6-4) adducts of TpdCpT with phosphodiesterases. The latter enzymes were shown to induce the quantitative release of the photoproduct as a modified dinucleoside monophosphate in a highly sequence-specific manner.  相似文献   

9.
Singlet oxygen-initiated decomposition of urocanic acid (UCA) (3-(1H-imidazol-4(5)-yl)-2-propenoic acid) was used to successfully confirm the report that UCA generates singlet oxygen when irradiated with ultraviolet A light (UVA). The UCA-generated singlet oxygen converts UCA to one or more products that then catalyze the further destruction of the UCA with UVA light by singlet oxygen formation. Some nicking of the phiX-174 supercoiled plasmid DNA was observed when UCA was irradiated with UVA to complete destruction of the starting material, and the product mixture was then mixed with the plasmid in the dark. More extensive nicking was seen when the photoproduct mixture and the plasmid were irradiated with UVA light. An "aged" (4 days) solution of UCA photoproduct no longer caused nicking in the dark but retained the capability to nick the plasmid when irradiated. There is evidence for the presence of hydroperoxides in the UCA photolysis product mixture, and the quenching studies with 2-propanol indicate that free radicals are involved in the plasmid-nicking photochemistry. Singlet oxygen does not appear to play a role in the nicking of the plasmid.  相似文献   

10.
A substituted anthraquinone (AQ), previously shown to photochemically generate benzaldehyde in methanol solution, was attached to a commercially available resin via an 11 carbon tether and an amide bond. Photolysis of the polymer-bound AQ with visible or 350 nm UV light resulted in the formation of benzaldehyde in yields of 50-55% as determined by HPLC. The phenolic positions in the polymer were then alkylated using benzyl bromide and 1-iodo-3-(4-nitrophenyl)propane in a coupling reaction with K(2)CO(3) as a base and a solution-phase proton shuttle. Photolysis of these alkylated polymers resulted in the formation of benzaldehyde (54-89%) and 3-(4-nitrophenyl)-propanal (58-67%). The yields of both aldehydes dropped considerably with subsequent realkylation and photolysis, and the polymer beads began to show signs of deterioration. This is the first time that aldehydes have been made photochemically on a solid-supported phase.  相似文献   

11.
Abstract— Photolysis of naphthalene on the surface of SiO2 under an atmosphere of air produces phthalic acid as the only major photoproduct, accounting for 49%o of the consumed naphthalene. Photolysis on Al2O3 also produces phthalic acid, in 31% yield. Photolysis of 1 -methylnaphthalene on SiO2 proceeds under similar conditions to produce 2-acetylbenzoic acid (35%) as the major photoproduct with the production of a small amount of I-naphthaldchyde (6%). 1-Cyanonaphthalene does not photooxidize under similar conditions. The presence of oxygen is necessary for the photodecomposition of naphthalene and 1-methylnaphthalene to proceed. Superoxide formed from the photolysis of naphthalene at the SiO2/air interface is readily observed by electron paramagnetic resonance spectroscopy. In the absence of naphthalene no superoxide is observed. A mechanism involving electron transfer from the S1 state of the naphthalene to O2 is proposed on the basis of these observations and related literature precedent.  相似文献   

12.
The cross-coupling reaction of 2,3-dibromo-5,10,15,20-tetraphenylporphyrin with corresponding organostannanes in the presence of a Pd0 catalyst in THF at reflux temperature yields free base 2,3-dialkynylporphyrins 1a,c-e. The subsequent deprotection of trimethylsilyl group of 1a with TBAF in THF under aqueous conditions produces the 2,3-diethynyl-5,10,15,20-tetraphenylporphyrins 1b in 87% yield. Compounds 1a-d undergo zinc insertion upon treatment with Zn(OAc)2.2H2O in CHCl3/MeOH to give zinc(II) 2,3-dialkynyl-5,10,15,20-tetraphenylporphyrins (2a-d) in 70-92% yields. Thermal Bergman cyclization of 1a-e and 2a-d was studied in chlorobenzene and approximately 35-fold 1,4-cyclohexadiene at 120-210 degrees C. Compounds 1b and 2b with R = H react at lower temperature (120 degrees C) and produce cyclized products 3b and 4b in higher yields (65-70%) than their propyl, isopropyl, and phenyl analogues, with R = Ph being the most stable. Continuing in this trend, the -TMS derivatives 1a and 2a exhibit no reactivity even after heating at 190 degrees C in chlorobenzene/CHD for 24 h. Photolysis (at lambda >/= 395 nm) of 1b and 2b at 10 degrees C leads the formation of isolable picenoporphyrin products in 15 and 35% yields, respectively, in 72 h, whereas these compounds are stable in solution under same reaction conditions at 25 degrees C in the dark. Unlike thermolysis at 125 degrees C, which did not yield Bergman cyclized product for R = Ph, photolysis generated very small amounts of picenoporphyrin products (3c: 5%; 4c: 8% based on 1H NMR) as well as a mixture of reduced porphyrin products that were not separable. Thus, trends in the barrier to Bergman cyclization in the excited state exhibit the same trend as those observed in the ground state as a function of R-group. Finally, photolysis of 2b at 10 degrees C with lambda >/= 515 or 590 nm in benzene/iPrOH (4:1, 72 h) produces 4b in 15 and 6% isolated yields, indicating that conjugation of the enediyne unit into the porphyrin electronic transitions leads to sufficient distortion to generate photoproduct even with long wavelength excitation.  相似文献   

13.
Photocleavage of Conjugated π,π-Epoxyenones. UV.-Irradiation of 3-(1′,2′-Epoxy-2′-methyl-prop-1′-yl)-5,5-dimethyl-2-cyclohexene-1-one On 1π,π*-excitation (δ = 254 nm) 9 undergoes cleavage of the C(δ), C(δ)-bond yielding 17 and a , which gives 18 by photofragmentation. In presence of maleinic ester the photolysis of 9 yields 20 , in presence of methanol 21 and 22 are obtained. By photocleavage of the C(δ), O-bond 9 is converted into b giving 14 . Photolysis of 14 yields 15 ( A + B ) and 16 . On 1n,π*-excitation (δ λ? 347 nm) of 9 cleavage of the C(δ), O-bond ( 9 → b ) seems to be the preferred reaction, whereas products of a are formed in traces, only.  相似文献   

14.
Abstract— Ultraviolet photolysis of riboflavin in pyridine leads to the formation of a red photoproduct. This photoreaction is apparently attributed to the photochemical interaction of pyridine with riboflavin. It is found that the rate of the photoreaction follows 1/2 order kinetics with respect to pyridine. ESR spectrum of the irradiated reaction mixture appears to be identical with that of riboflavin semiquinone obtained from visible light irradiation. Whether or not such a flavin radical is the responsible intermediate in the photoreaction is not established. Possible modes of the reaction are discussed. Absorption, excitation, and emission spectra of the red photoproduct strongly suggest that the structure of the photoproduct is considerably different from that of riboflavin or pyridine, indicating an extensive photolytic process involving covalent bond alterations.  相似文献   

15.
The MCD spectra of phloroglucin, 1,3,5-trimethoxybenzene, 1,3,5-tricyanobenzene, 1,3,5-trinitrobenzene, 1,3,5-benzenetricarboxylic acid and 1,3,5-benzenetricarbonyl chloride were measured. For the 1E′ ← 1A′1 transition of phloroglucin or 1,3,5-tricyanobenzene and the 1E′ ← 1A′ transition of 1,3,5-trimethoxybenzene, the Faraday A term was observed and the A/D value was extracted. However, 1,3,5-trinitrobenzene, 1,3,5-benzenetricarboxylic acid and 1,3,5-benzenetricarbonyl chloride showed no magnetic circular dichroism in this spectral region. The magnetic moments in the 1E′ states of these molecules seem to be quenched by the effects of substituents. The magnetic moments in the 1E′ states of benzene derivatives are sensitive to substitution in the benzene ring.  相似文献   

16.
C. Von Sonntag 《Tetrahedron》1969,25(24):5853-5861
The UV photolysis (λ = 185 nm) of liquid methanol yields hydrogen, glycol, formaldehyde, methane and traces of ethane in quantum yields of 0·83, 0·78, 0·058, 0·05 and 0·002 resp. (related to φ(H2) = 0·4 of the ethanol-actinometer (5 mole/1 in water)). The isotopic distribution of the hydrogen (85% HD) formed in the photolysis of CH3OD shows, that as in the gasphase2 the scission of the O---H-bond (1) is the major process. CH3OH + hv (λ = 185 nm) → CH3O + H (1)

In methanoi-water mixtures (nearly all the light of the wavelength λ = 185 nm is absorbed by methanol) the quantum yields of hydrogen, glycol, methane and ethane are greatly reduced, while the formaldehyde yield remains unaffected. In 1 molar solution φ(H2) = 0·42, φ(glycol) = 0·32 and φ(CH4) = 6 x 10−4 is obtained. Ethane cannot be detected.  相似文献   


17.
Photolysis of 1,2-dithiane (1) in acetonitrile with single walled carbon nanotubes (SWCNTs) was earlier reported to form thiol-functionalized SWCNTs via the butane-1,4-dithiyl diradical (2). The present study shows that 2 instead undergoes a facile rearrangement to thiophane-2-thiol (6). This photoreaction is clean, rapid, and irreversible under 313 nm irradiation. The secondary photolysis of 6 with SWCNTs at a shorter wavelength (254 nm) leads to 2-thiophanyl radicals 8, which derivatize SWCNTs by covalent attachment. Pyrolysis of the resulting "sulfurized SWCNTs" affords a mixture of organosulfur compounds, including thiophene formed by dehydrogenation. An unknown additional mechanism causes high TGA weight loss and a large incorporation of sulfur.  相似文献   

18.
The time-resolved electronic absorption spectra of CN radical, resulting from the laser photolysis of dicyanoacetylene (DCA) at 193 and 248 nm, were analyzed. Detection of the other probable photodissociation product—C3N radical—has not been possible. Photolysis at 193 nm produces CN both in the X2+ and in the A2Πi manifolds, the latter—probably populated via a two-photon process—being revealed by a delayed (collision-induced) transfer to the higher vibrational levels of the ground electronic state. Photolysis at 248 nm is an efficient two-photon process. A simplified kinetic model for the decay of CN radical has been proposed and the rate constant for the CN + DCA reaction was derived. Semiempirical INDO/S CI-1 calculations of the DCA valence shell electronic transitions were performed.  相似文献   

19.
Wavelength dependent photoadditions of 6-cyano-1,3-dimethyluracil to phenylacetylenes are described. Photolysis with Prex-filtered light gave the fused cyclobutenes (3, 9), whereas irradiation with 254-nm light afforded the rearranged adducts (5, 11). The structures of the adducts were established by X-ray analyses.  相似文献   

20.
9-Trimethylstannylanthracene (1) was synthesized and its photolysis by 365-nm light was studied. In aprotic solvents, the dimerization of 1 involves positions 9 and 10 and yields a head-to-tail dimer. The main route of the photolysis of 1 in alcohols is the cleavage of the C-Sn bond with the formation of anthracene. The quantum yields of the photoreaction and the lifetimes and quantum yields of 1 fluorescence were determined.  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