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 140 毫秒
1.
Abstract— The effect of u.v. irradiation (2537 Å) on the RNA bacteriophage f2 has been studied with respect to the adsorption of f2 to E. coli K12 (male strain), the penetration of f2-RN A into the host cell and the conversion of the phage nucleic acid to the double-stranded replicative intermediate. The biological parameter most sensitive to u.v. was the plaque-forming ability of the phage. Its loss could be attributed to several factors. (1). A binding of capsid protein to phage nucleic acid interfering with host penetration by the f2-RNA. (2). Desorption of some irradiated phage at 37° from their attachment sites on the host. (3). Molecular alterations in the RNA preventing formation of the replicative intermediate within the host. The relationship of these factors to the lack of photoreactivability of irradiated f2 is discussed.  相似文献   

2.
The photolysis of SO2 at 3712 Å in the presence of the 1,2-dichloroethylenes has been investigated at 22deg;C. The data are consistent with the SO2(3B1) photosensitized isomerization of the 1,2-dichloroethylene isomer. A kinetic treatment of the initial quantum yield data was consistent with the formation of a polarized charge-transfer intermediate whenever SO2(3B1) molecules and one of the 1,2-dichloroethylene isomers collide which ultimately decays unimolecularly to the cis-isomer with a probability of 0.70 ± 0.26 and to the trans-isomer with a 0.37 ± 0.16 probability. Quenching rate constants for removal of SO2(3B1) molecules by cis- and trans-1,2-dichloroethylene have been estimated from quantum yield data and from laser excited phosphorescence lifetimes using an excitation wavelength of 3130 Å. Estimates of the quenching rate constant (units of 1./mole ± sec) are for the cis-isomer, (1.63 ± 0.71) × 1010, quantum yield data, and (2.44 ± 0.11) × 1010, lifetime data; and for the trans-isomer,(2.59 ± 0.09)×1010, lifetime data, and (2.35 ±0.89) × 1010, quantum yield data. An experimentally determined photostationary composition,[cis-C2Cl2H2]/[trans-C2Cl2H2] = 1.8 - 0.1, was in good agreement with a value of 2.00 - 1.15 which was predicted from rate constants derived in this study.  相似文献   

3.
It was found that the quantum yield of the fluorescence ofmeso-tetraferrocenylporphyrin (TFcP) is at most 3.0·10−5, and that of the triplet state of FTcP is at least 200 times lower than the quantum yield ofmeso-tetraphenylporphyrin (TPP). Excitation of TFcP in CCl4 by light with λ>410 nm results in the oxidation of TFcP. The singlet and triplet excited states of TPP in toluene and acetonitrile are quenched by ferrocene with rate constants of 1.2·1010 and 1.7·1010, (4.6±0.5)·108 and (1.37±0.21)·109 L mol−1 s−1, respectively. The quenching mechanisms are discussed. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 10, pp. 1924–1927, October, 1999.  相似文献   

4.
Synthesis and photochemistry of acridin-9-ylmethoxycarbonyl(Amoc)as a new photochemically removableprotecting group for alcohols were described.Three carbonates of alcohols 1—3 were synthesized throughcondensation of 9-hydroxymethylacridine and chloroformates of alcohols,including benzyl alcohol,phenethylalcohol and one galactose derivative.The photolysis of protected alcohols can efficiently release the correspondingalcohol in the efficiencies(Q_(u1)ε)of 100—200(quantum yield Q_(u1)=0.011—0.023,and molar absorptivity ε=9.1×10~3—9.8×10~3 mol~(-1)·L·cm~(-1))under 360 nm light.  相似文献   

5.
The maximum photosteady state fraction of K, xKmax, and the ratio of the quantum yields of the forward and back light reactions, trans-bacteriorhodopsin (bR) hArr; K, φbRK, were obtained by measuring the absorption changes produced by illumination of frozen water-glycerol (1:2) suspensions of light-adapted purple membrane at different wavelengths at -165°C. An independent method based on the second derivative of the absorption spectrum in the region of the β-bands was also used. It was found that The quantum yield ratio (0.66 ± 0.06) was found to be independent of excitation wavelength within experimental error in the range510–610 nm. The calculated absorption spectrum of K has its maximum at603–606 nm and an extinction 0.85 ± 0.03 that of bR. At shorter wavelengths there are P-bands at 410, 354 and 336 rim. Using the data of Hurley et al. (Nature 270,540–542, 1977) on relative rates of rhodopsin bleaching and K formation, the quantum yield of K formation was determined to be 0.66 ± 0.04 at low temperature. The quantum efficiency of the back reaction was estimated to be 0.93 ± 0.07. These values of quantum efficiencies of the forward and back light reactions of bR at - 165°C coincide with those recently obtained at room temperature. This indicates that the quantum efficiencies of both forward and back light reactions of bacteriorhodopsin are temperature independent down to -165°C.  相似文献   

6.
Abstract— Holding complexes of u.v.-irradiated (254 nm) T3 phage in E. coli B/r cells for several hours at 37°C in buffer, or broth with chloramphenicol, affects the phage survival in at least two different ways: (1) by enhancing excision repair, resulting under certain conditions in liquid-holding recovery (LHR), and (2) by destroying the phage (holding inactivation). LHR is most apparent in buffer containing 20 μg ml-1 chloramphenicol (CAP). It is expressed by as much as a 10–fold increase in the fraction of complexes that display host-cell reactivation (resulting from excision repair), but the percentage of u.v. lesions repaired within repair-proficient complexes is slightly decreased. LHR is not observed if T3 infects the repair-deficient strain Bs-1. Holding inactivation is readily observed with unirradiated phage complexes in broth containing CAP. The response of irradiated-phage complexes to liquid-holding conditions is more complex: holding inactivation is less effective for irradiated than for unirradiated phage DNA (i.e. the irradiated DNA is to some extent ‘protected’), and processes leading to LHR are superimposed. Thus under certain holding conditions one observes the paradoxical phenomenon that the viable titer of irradiated phage is several times higher than that of unirradiated phage. The nature of holding inactivation is not known, nor is the mechanism by which irradiated DNA is partially protected against it. Holding inactivation does not require protein synthesis; it is rather enhanced at high CAP concentration and seems to be favored by otherwise active cell metabolism. At high CAP concentrations (200–400 μg ml-1, as compared to 20 μg ml-1) irradiated-phage complexes show neither LHR nor protection against holding inactivation. Likewise they fail to undergo some step by which the phage DNA becomes insensitive to repair inhibition by caffeine.  相似文献   

7.
Formic acid is considered a promising energy carrier and hydrogen storage material for a carbon‐neutral economy. We present an inexpensive system for the selective room‐temperature photocatalytic conversion of formic acid into either hydrogen or carbon monoxide. Under visible‐light irradiation (λ>420 nm, 1 sun), suspensions of ligand‐capped cadmium sulfide nanocrystals in formic acid/sodium formate release up to 116±14 mmol H2 gcat?1 h?1 with >99 % selectivity when combined with a cobalt co‐catalyst; the quantum yield at λ=460 nm was 21.2±2.7 %. In the absence of capping ligands, suspensions of the same photocatalyst in aqueous sodium formate generate up to 102±13 mmol CO gcat?1 h?1 with >95 % selectivity and 19.7±2.7 % quantum yield. H2 and CO production was sustained for more than one week with turnover numbers greater than 6×105 and 3×106, respectively.  相似文献   

8.
n-C3H7ONO was photolyzed with 366 nm radiation at ?26, ?3, 23, 55, 88, and 120°C in a static system in the presence of NO, O2, and N2. The quantum yields of C2H5CHO, C2H5ONO, and CH3CHO were measured as a function of reaction conditions. The primary photochemical act is and it proceeds with a quantum yield ?1 = 0.38 ± 0.04 independent of temperature. The n-C3H7O radicals can react with NO by two routes The n-C3H7O radical can decompose via or react with O2 via Values of k4/k2 ? k4b/k2 were determined to be (2.0 ± 0.2) × 1014, (3.1 ± 0.6) × 1014, and (1.4 ± 0.1) × 1015 molec/cm3 at 55, 88, and 120°C, respectively, at 150-torr total pressure of N2. Values of k6/k2 were determined from ?26 to 88°C. They fit the Arrhenius expression: For k2 ? 4.4 × 10?11 cm3/s, k6 becomes (2.9 ± 1.7) × 10?13 exp{?(879 ± 117)/T} cm3/s. The reaction scheme also provides k4b/k6 = 1.58 × 1018 molec/cm3 at 120°C and k8a/k8 = 0.56 ± 0.24 independent of temperature, where   相似文献   

9.
trans-3-Methyl-4-(p-anisyl)-1,2-dioxetane 1, trans-3-methyl-4-(o-anisyl)-1,2-dioxetane 2 , 3-methyl-3-benzyl-1,2-dioxetane 3 , and 3-methyl-3-p-methoxybenzyl-1,2-dioxetane 4 were synthesized in low yield by the β-bromo hydroperoxide method. The activation parameters were determined by the chemiluminescence method (for 1 ΔG≠ = 22.8 ± 0.3 kcal/mol, Δ≠ = 22.2, ΔS≠ = −1.7 e.u., k60 = 7.6 × 10−3s−1; for 2 ΔG≠ + 23.6 ± 0.3 kcal/mol, ΔH≠ = 22.8, ΔS≠ = −2.2 e.u., k60 = 2.5 × 10−3S−1; for 3 ΔG≠ = 24.0 ± 0.4 kcal/mol, ΔH≠ = 23.1, ΔS≠ = −2.7 e.u., k60 = 1.2 × 10−3S−1; for 4 ΔG≠ = 24.0 ± 0.2 kcal/mol, ΔH≠, = 23.2, ΔS≠, = −2.4 e.u., k60 = 1.2 × 10−3s−1). Thermolysis of 1–4 produced excited carbonyl fragments (direct production of high yields of triplets relative to excited singlets) [chemiexcitation yields ϕT, ϕS, respectively: for 1 0.02, 0.0001; for 2 0.02, 0.0001; for 3 0.03, 0.0002; for 4 0.02, 0.0001]. The effect of paramethoxyaryl substitution was consistent with electronic effects. The ortho substitution in 2 resulted in an increase in stability of the dioxetane, opposite that observed for an electronic effect. The results are discussed in relation to a diradical-like mechanism.  相似文献   

10.
The dipole (), quadrupole (C), and dipole-quadrupole (B) polarizabilities and the dipole hyperpolarizability () of the chloride ion have been calculated by using the many-body perturbation theory approach and a series of large polarized GTO/CGTO basis sets. The complete fourth-order treatment of the electron correlation effects with a basis set comprising the s, p, d, f, and g functions gives: =38.01 a.u., C=211.5 a.u., B=–5.14×103 a.u., and =128. 5×103 a.u. as compared to the corresponding SCF values (=31.49 a.u., C=158.9 a.u., B=–2.92×103 a.u., =57.7×103 a.u.). The quenching of polarizabilities of the Cl ion in solutions and ionic crystals is discussed.  相似文献   

11.
Quantum yield measurements for the SO2(3B1) photosensitized isomerization of cis-1,2-difluoroethylene have been made at 3712 Å and 22°C. The [SO2]/[cis-C2F2H2] ratio was varied from 47.4 to 455 and the quantum yield measurements over this variation of concentration ratios were consistent with a mechanism in which SO2(3B1) molecules and the cis isomer form a collision intermediate which decomposes with a probability of 0.42 ± 0.17 and 0.58 ± 0.17 of producing trans- and cis-1,2-difluoroethylene, respectively. When SO2 was subjected to prolonged irradiations in the presence of initially either pure cis- or pure trans-1,2-difluoroethylene, a photostationary composition, [cis]/[trans] = 1.0 ± 0.2, was obtained. The rate constant at 22°C for removal of SO2(3B1) molecules by cis-1,2-difluoroethylene was estimated to be (1.72 ± 0.72) × 1010 1./mole · sec.  相似文献   

12.
Abstract— To determine the maximum range of coupling between side-chain photochromism and polypeptide conformation change, we modified the carboxylate side chains of succinylated poly(l -lysine) with a spiropyran to form polypeptide I. The extent of modification was determined to be 35.5%. The spacer group length between the polypeptide a-carbon and the dye was 12 atoms, providing minimum polypeptide-dye interaction. Conformation changes were monitored by circular dichroism as a function of light adaptation and solvent composition (hexafluoroisopropanol [HFIP] vs trifluoroethanol [TFE]). Under all solvent compositions, the dark-adapted dye was in the merocyanine form. Light adaptation by visible light converted the dye to the spiropyran form. When dissolved in TFE, I adopted a helical conformation insensitive to light adaptation. With increasing percentage HFIP, a solvent-induced helix-to-coil transition was observed around 80% (vol/vol) HFIP. At 100% HFIP, both light- and dark-adapted forms of I were in the coil state. Near the midpoint of the solvent-induced helix-to-coil transition, light adaptation caused conformation changes. Applying helix-to-coil transition theory, we measured a statistically significant difference in coil segment-HFIP binding constant for light- v.v dark-adapted solutions (6.38 ± 0.03 M-1vs 6.56 ± 0.03 M-1), but not for the nucleation parameter σ (1.2 ± 0.4 10-3 v.v 1.3 ± 0.3 × 10-3). The small binding constant difference translated to a light-induced binding energy difference of 17 cal/mol/monomer. Near the midpoint of the helix-to-coil transition, collective interactions between monomer units made possible the translation of a small energy difference (less than RT) into large macromolecular conformation changes. This work parallels similar behavior observed in poly(isocyanate) (Green, M. M. et al. J. Am. Chem. Soc. 115 , 4941–4942, 1993). The subtle differences in dye-backbone interaction in I suggested a maximum coupling distance (12 atoms) beyond which polypeptide conformation and dye state are uncoupled.  相似文献   

13.
Abstract— The quantum yields for the photohydration of dimethyluracil were determined for concentrations in the range 5 × 10--1--1 × 10--3M by use of 240–280 nm irradiation. The average quantum yield (0.0139 f 0.0005) was independent of both concentration and irradiation wavelength.  相似文献   

14.
The photophysical properties of bonellin, a free-base chlorin, were studied in ethanolic solution. For the singlet excited state the following data were determined: an energy level, EBS= 187 ± 2kJ mol-1, a lifetime, τf= 6.3± 0.1ns at 298 K, and fluorescence quantum yields, φr= 0.07 ± 0.02 (298 K) and 0.20 ± 0.04 (77 K). The S1→ T intersystem crossing quantum yield was φisc= 0.85 ± 0.1. No phosphorescence was observed at 298 K and 77 K. Based on quenching experiments the triplet state energy level was determined to be EBT= 180 ± 20 kJ mol-1. A unimolecular decay rate constant, k1= (2.3 ± 0.5)· 103 s-1 at room temperature, and a molar absorption coefficient, εT443= 9500 ± 500 M-1 cm-1, were obtained for the triplet state. This species was quenched by O2 with ko2= (1.7 ±0.3)· 108M-1 s-1, and by benzoquinone with kq= (5.2 ± 0.3)-109M-1 s-1. The latter value, as well as the high value determined for the triplet annihilation rate constant, k2= (2 ± 0.5)· 109M-1 s-1, might reflect an electron transfer mechanism. Copper bonellin had a shorter triplet lifetime (>20 ns), which offers a possible explanation for its lack of photodynamic action.  相似文献   

15.
Riboflavin was irradiated anaerobically in aqueous EDTA solutions over the pH range 2.5–10. In other dye systems (Bonneau and Pereyre, 1975), only the trivalent anion of EDTA was found to have significant reactivity for photoreduction. For riboflavin, the reactivity begins with monoanionic EDTA, and the reactivity is markedly increased as the charge increases. This suggests that the charge on the reductant is more important to the electron transfer process for riboflavin than the formation of a nonhydrogen bonded nitrogen site on EDTA. At high concentrations of EDTA in the pH range 4–8, quenching of the photoreduction occurs, which can be explained by an energy transfer between the excited singlet state of riboflavin and trianionic EDTA, possibly as an association complex. The rate constants for the photoreduction of riboflavin by the monovalent, divalent, and trivalent anions of EDTA are 1.0 times 107M-1 s-l, 4.8 times 10′M-1 s-l, and 2.0 times 108M-1s-1, respectively. The rate constant for the singlet state quenching by trianionic EDTA is 3 times 109M-l s-1, and the limiting quantum yield for intersystem crossing for riboflavin in aqueous solution is 0.50 ± 0.05.  相似文献   

16.
3-Methyl-3-(o-tolyl)-1,2-dioxetane 1 and 3-methyl-4-(o-bromophenyl)-1,2-dioxetane 2 were synthesized in low yield by the β-bromo hydroperoxide method. The activation parameters were determined by the chemilumin-escence method (for 1 ΔG? = 24.7 ± 0.3 kcal/mol, ΔH? = 25.4, ΔS? = + 1.9 e.u., k60 = 3.4 × 10?4s?1; for 2 ΔG? = 24.7 ± 0.4 kcal/mol, ΔH? = 24.7, ΔS? = 0.0 e.u., k60 = 4.1 × 10?4s?1). Thermolysis of 1–2 directly produced high yields of excited triplets as expected for this type of dioxetane [triplet chemiexcitation yields (?7) for 1 0.03; for 2 0.02; the ?T/?S ratios were estimated to be approximately 200 for both compounds]. The effect of ortho-aryl substituents was inconsistent with electronic effects. The ortho substitution in 1–2 resulted in a marked increase in stability of the dioxetanes. The results are discussed in relation to a diradical-like mechanism.  相似文献   

17.
C2H5ONO was photolyzed with 366 nm radiation at ?48, ?22, ?2.5, 23, 55, 88, and 120°C in a static system in the presence of NO, O2, and N2. The quantum yield of CH3CHO, Φ{CH3CHO}, was measured as a function of reaction conditions. The primary photochemical act is and it proceeds with a quantum yield ?1a = 0.29 ± 0.03 independent of temperature. The C2H5O radicals can react with NO by two routes The C2H5O radical can also react with O2 via Values of k6/k2 were determined at each temperature. They fit the Arrhenius expression: Log(k6/k2) = ?2.17 ± 0.14 ? (924 ± 94)/2.303 T. For k2 ? 4.4 × 10?11 cm3/s, k6 becomes (3.0 ± 1.0) × 10?13 exp{?(924 ± 94)/T} cm3/s. The reaction scheme also provides k8a/k8 = 0.43 ± 0.13, where   相似文献   

18.
i-C4H9ONO was photolyzed with 366-nm radiation at ?8, 23, 55, 88, and 120°C in a static system in the presence of NO, O2, and N2. The quantum yield of i-C3H7CHO, Φ{i-C3H7CHO}, was measured as a function of reaction of reaction conditions. The primary photochemical act is and it proceeds with a quantum yield ?1 = 0.24 ± 0.02 independent of temperature. The i-C4H9O radicals can react with NO by two routes The i-C4H9O radical can decompose via or react with O2 via Values of k4/k2 ? k4b/k2 were determined to be (2.8 ± 0.6) × 1014, (1.7 ± 0.2) × 1015, and (3.5 ± 1.3) × 1015 molec/cm3 at 23 55, and 88°C, respectively, at 150-torr total pressure of N2. Values of k6/k2 were determined from ?8 to 120°C. They fit the Arrhenius expression: For k2 ? 4.4 × 1011 cm3/s, k6 becomes (3.2 ± 2.0) × 10?13 exp{?(836 ± 159)/T} cm3/s. The reaction scheme also provides k4b/k6 = 3.59 × 1018 and 5.17 × 1018 molec/cm3 at 55 and 88°C, respectively, and k8b/k8 = 0.66 ± 0.12 independent of temperature, where   相似文献   

19.
Abstract— Protection by acridine orange against ultraviolet light effects in resting cells of E. coli B/r/1, try- was studied with special reference to a possible oxygen effect. Dose-response relationships were described by the function S= 1–(1 - e-kD)n where S is the surviving fraction and D is the u.v. dose in ergs/mm2. For cells suspended in 5 × 10--6M acridine orange (AO) in air, the radiation sensitivity k was reduced from 0.010 (ergs/mm2))-1 in the absence of the dye to 0.0053 (ergs/mm2)-1 in the presence of the dye. Under anoxia at this AO concentration, k was further reduced to 0.0015 (ergs/mm2)-1. The oxygen effect ratio, kO2/kN2, was 3.5 at this concentration of AO. Greater protection was observed in cells suspended in 2 × 10--5M AO, the oxygen effect ratio was unchanged. No oxygen effect was detected in the u.v. response in the absence of the dye. The value of n was reduced from about 12 with no dye to about 5 at dye concentrations of 5 × 10--6M AO or more when oxygen was present. Under anoxia, in the presence of AO, n was further reduced to about 1.3. Atebrin, an efficient u.v. protective agent but an inefficient photodynamic agent, had no oxygen effect for protection against u.v. inactivation. Acridine orange protected against u.v.-induced reversion to tryptophan indepence in E. coli WP2 to about the same extent as it did for inactivation. A similar oxygen effect was observed for both inactivation and mutagenesis.  相似文献   

20.
Chlorophyll-a was incorporated into cellulose acetate films and the triplet state decay kinetics and electron transfer from triplet to p-benzoquinone in aqueous solution was studied using laser flash photolysis and EPR. The triplet was found to decay by first order kinetics with a rate constant which was independent of Chl concentration. The triplet yield, however, was concentration dependent. These properties are due to quenching which occurs only at the singlet state level. In the presence of quinone, the triplet is quenched and, when the quinone is in an aqueous solution in contact with the film, Chl cation radical (C±) as well as the semiquinone anion radical (Q±) can be observed. The C decays by second order kinetics with a rate constant of 1.5 × 106M-1 s-1. Although triplet conversion to radicals is slightly lower in the films as compared to fluid solutions (? 3 times), the lifetimes of the radicals are greatly increased (? 103 times).  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