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.
Quantum yields for the formation of superoxide ions, O2?1, and singlet oxygen, 1O2, were determined during the photolyses of gilvocarcin M (GM) in air-saturated dry dimethylsulfoxide (DMSO) and in 45:55 (vol/vol) DMSO-water mixtures. The quantum yield for the photoreduction of methyl viologen by GM in nitrogen-saturated dry DMSO was also determined. These values are not different, within experimental error, from those corresponding to gilvocarcin V (GV). Because GV is a strong photocytotoxic agent and GM is not, these results imply that Type I and Type II mechanisms are not important pathways in the cytotoxicity of GV.  相似文献   

2.
Abstract— The transient absorption spectra of the intermediates produced by the 355 nm laser excitation of gilvocarcin derivatives have been investigated in various solvents. The spectra consist of a triplet-triplet absorption in the visible region and a residual absorption observed between 340 and 700 nm due to a long-lived species, assigned to the radical cation. A broad-fast decaying band with a maximum at around 700 nm attributed to the solvated electron is also seen in solutions containing a low DMSO/water volume ratio and at 266 nm irradiation of a 50% methanol/water solvent mixture. The molar absorption coefficient of the triplet state of gilvocarcin V (GV) and gilvocarcin M (GM), determined by the energy transfer method, is independent of the solvent properties and has a value of 3.0 × 104/ M cm. The triplet decay rate constants for both drugs are between 1 and 5 × 104/s. A similar initial yield and triplet decay rate constant of GV were observed in the presence of 3.4 m M thymine. Thus, a quenching rate constant of the GV's triplet state by thymine is estimated to be lower than 106/Ms. The triplet quantum yields of both antibiotics determined by using the comparative method are higher in dimethylsulfoxide (DMSO) (0.18) than are those corresponding to 25% DMSO/water (0.06). The decrease in φT in the presence of water could be attributed to an enhanced internal conversion rate constant from the S1 state or to an increase in the photoionization yield. The similarity of the transient intermediates and their yields for GV and GM suggest that their photobiological differences are due to other factors such as DNA binding constants, preferential localization of the drugs in the cell or the enhanced reactivity of the vinyl group toward cellular components.  相似文献   

3.
Nanosecond laser photolytic studies of 4-nitro-N,N-dimethylnaphthylamine (4-NDMNA) in nonpolar and polar solvents at room temperature show a transient species with an absorption maximum in the 500-510-nm range. This species is assigned to the lowest triplet excited state of 4-NDMNA. The absorption maximum of this state is independent of solvent polarity, and its lifetime is a function of the hydrogen donor efficiency of the solvent. In n-hexane the lifetime 1/k of the triplet state is 9.1 × 10?6 sec, while in acetonitrile 1/k is 2.0 × 10?7 sec. The hydrogen abstraction rate constant kH of the triplet state with tributyl tin hydride (Bu3SnH) in n-hexane is 1.7 × 107M?1·sec?1, while in the case of isopropyl alcohol as hydrogen donor, kH is 4.0 × 107M?1·sec?1. The activation energy for the hydrogen abstraction by the triplet state from Bu3SnH in deaerated n-hexane is 0.6 kcal/mol. The lack of spectral shift with increasing solvent polarity, and the appreciable hydrogen abstraction reactivity of the triplet state, also independent of solvent polarity, seem to indicate that this excited state is an n-π* state which retains its n-π* character even in polar media.  相似文献   

4.
The E ? Z photoisomerization of the title compound (UA) (a naturally occurring sunscreen) has been studied in aqueous solution. At a UA concentration of 6mM and using 313nm excitation, φE→z= 0.52, φZ→E= 0.47 and the photostationary state is 34% E. Under these conditions, loss of UA is minimal. Low energy triplet quenchers fail to impede the isomerization, but the reaction can be induced by several triplet sensitizers. The ET for UA is estimated to be approximately 55 kcal/mol.  相似文献   

5.
The photophysical properties of benzoporphyrin derivative monoacid ring A (BPD-MA), a second-generation photosensitizer currently in phase II clinical trials, were investigated in homogeneous solution. Absorption, fluorescence, triplet-state, singlet oxygen (O2(1Δg)) sensitization studies and photobleaching experiments are reported. The ground state of this chlorin-type molecule shows a strong absorbance in the red (λ≈ 688 nm, ?≈ 33 000 M?1 cm?1 in organic solvents). For the singlet excited state the following data were determined in methanol: energy level, Es= 42.1 kcal mol?1, lifetime, Φf= 5.2 ns and fluorescence quantum yield, Φf= 0.05 in air-saturated solution. The triplet state of BPD-MA has a lifetime, τf >. 25 ns, an energy level, ET= 26.9 kcal mol?1 and the molar absorption coefficient is ?T= 26 650 M?1 cm?1 at 720 nm. A dramatic effect of oxygen on the fluorescence (φf) and intersystem crossing (φT) quantum yields has been observed. The BPD-MA presents rather high triplet (φT= 0.68 under N2-saturated conditions) and singlet oxygen (φΔ= 0.78) quantum yields. On the other hand, the presence of oxygen does not significantly modify the photobleaching of this photostable compound, the photodegradation quantum yield (φPb) of which was found to be on the order of 5 × 10?5 in organic solvents.  相似文献   

6.
The low-lying singlet and triplet states of H2CBe and HCBeH are examined using ab inito molecular orbital theory. In agreement with earlier results, the lowest-lying structure of H2CBe has C2v symmetry and is a triplet with one π electron (3 B1). The results presented here suggest that the lowest-energy singlet structure is the (1B1) open-shell singlet, also with C2v symmetry, at least 2.5 kcal/mol higher in energy. The singlet C2v structure with two π electrons (1A1) is 15.9 kcal/mol higher than 3B1. All of these structures are bound with respect to the ground state of methylene and the beryllium atom. In HCBeH, linear equilibrium geometries are found for the triplet (3Σ) and singlet (1Δ) states. The triplet is more stable than the singlet (1Δ) by 35.4 kcal/mol, and is only 2.9 kcal/mol higher in energy than triplet H2 CBe. Since the transition structure connecting these two triplet molecules is found to be 50.2 kcal/mol higher in energy than H2 CBe, both triplet equilibrium species might exist independently. The harmonic vibrational frequencies of all structures are also reported.  相似文献   

7.
The photochemistry of 1,2,3-indanetrione (1) has been examined in solution at room temperature by steady state and laser flash photolysis. The triplet state of 1 (T = 6.5 μs, δmux, = 360 and 570 nm, in dry acetonitrile) reacts preferentially via an a-cleavage process followed by a considerably slower loss of carbon monoxide. Triplet 1 shows a remarkably fast hydrogen abstraction rate constant when in the presence of 1,4-cyclohexadiene (kr= 1.4 times 106M?1s?1) in spite of its low excitation energy (ET= 42 kcal/mol). This behavior can be explained by assuming that the vicinal carbonyls coplanar to the ketyl radical play an important role in its stabilization.  相似文献   

8.
The photophysical properties of rufloxacin, 9-fluoro-2r3-dihydro-10-(4-methyl-l-pyrazinyl)-7-oxo-7-H-pyri-do[l,2,3-de]-l,4-benzothiazin-6-carboxylic acid, a fluoroquinolone antibacterial drug exhibiting photosensitizing action toward biological substrates, were studied in aqueous solutions at neutral pH. The lowest excited electronic states of the zwitterion were characterized by both experimental techniques and theoretical methods. Steady-state and time-resolved emission, triplet-state absorption and singlet oxygen production were investigated. The results indicate that the lowest excited singlet is a fluorescent, relatively long-lived state (φr= 0.075, Tr? 4.5 ns) with an efficient intersystem crossing to the triplet manifold (φisc? 0-7)- The lowest triplet is a long-lived state (TT? 10 μs at 295 K in 0.01 M phosphate buffer), with properties that make it a good candidate for being the precursor of the photodecarboxylation of the drug. It is quenched by oxygen at a rate of 1.7 times 109M-1 s-1 and singlet oxygen is formed with a quantum yield of 0.32 in air-saturated solutions.  相似文献   

9.
The base hydrolysis of (αβS) (salicylato) (tetraethylenepentamine)cobalt(III) has been investigated in MeOH + water and DMSO + water media (0–70% (v/v) cosolvents) at 20.0 ? t°C ? 35.0 and I = 0.10 mol dm?3 (ClO4?). The phenoxide species [(tetren)CoO2CC6H4O]+ undergoes both OH?-independent and OH?-catalyzed hydrolysis via SN1ICB and SN1CB mechanism, respectively. The OH?-independent hydrolysis of the phenoxide species is catalyzed by both DMSO + water and MeOH + water media, the former exerting a much stronger rate accelerating effect than the latter. The OH?-catalyzed reaction is strongly accelerated by DMSO + water medium but insensitive to the composition of MeOH + water medium up to 40% (v/v) MeOH beyond which it was not detectable under the experimental conditions. Data analysis has been attempted on the basis of the solvent stabilizing and destabilizing effects on the initial state and transition state of the concerned reactions. The nonlinear variation of the activation parameters, ΔH and ΔS, with solvent compositions presumably indicates that the solvent structural effects mediate the energetics of solvation of the initial state and transition state of the concerned reactions. The linearity in ΔH vs. ΔS plot accomodating all data for k1 and k2 paths in DMSO + water and MeOH + water further suggests that the solvent effects on these parameters are mutually compensatory.  相似文献   

10.
The structure of the lithium and potassium salts of φSCH3,φSOCH3,φSO2CH3,φSO(NCH3)CH3 has been studied by 13C NMR in different solvents. The results show that the metalated carbon is nearly pyramidal in and nearly planar in φSOCH2?M+, whatever the solvent are cation are φSO2CH2? φSCH2?M+ and φSO(NCH3)CH2?M+ are in an intermediate hydridization state, cation and solvent dependent. For the sulfoxide, a four-center chelate is proposed, stable to strong solvating agents and only disrupted by cryptands. It is very likely responsible for the planar configuration of the anionic carbon.The low temperature study of φSOCH2Li shows the existence of aggregates in THF. HMPA or external lithium salts disrupt these associations, giving rise to other species.The 13C NMR parameters of the whole series of sulfur-stabilized carbanions are quite consistent with the date reported for phosphorous and arsenic ylids: the
coupling constants appear to be a good probe of the geometry of the anionic carbon, whereas the chemical shifts are rather insensitive to its hybridization state.  相似文献   

11.
The dication C2H has been investigated by ab initio molecular orbital theory. It is found to have a linear (Dh), structure with a triplet (3σ?g) ground state. Deprotonation to C2H+ is exothermic by 9.8 kcal/mol, but this process is hindered by a large barrier of 65 kcal/mol.  相似文献   

12.
The analysis of the activation parameters for the formal H‐atom transfer reaction between 2,2,5,7,8‐pentamethyl‐6‐chromanol (ChrOH) and 2,2‐diphenyl‐1‐picrylhydrazyl (dpph?) reveals that these parameters are effective probes of the actual reaction mechanism. Indeed, the A factors measured in various polar and apolar solvents are localized in three distinct domains according to whether the reaction occurs via outer‐sphere electron transfer (ET) from the anion ChrO? or hydrogen atom transfer (HAT). For instance, A = 5.9 × 105 M?1 s?1 and Ea = 2.5 kcal mol?1 in cyclohexane where the reaction proceeds by HAT, whereas in methanol, ethanol, and their mixtures with water where there is a substantial ET contribution A > 109 M?1s?1 and Ea > 7 kcal mol?1. Interestingly, in nonhydroxylic polar solvents, A~ 107 M?1s?1 and the Ea values reflect the H‐bond accepting ability of the solvent in agreement with the “standard” kinetic solvent effects on HAT reactions. Addition of small quantities of pyridine accelerates the reaction rates in these solvents. This suggests that the H‐bonded complex (ChrOH···Py) is able to react via intermolecular ET with dpph?. It is known, in fact, that pyridine lowers the oxidation potential of phenols by ~0.5 V and the ΔGET of ChrOH + dpph? consequently decreases by about 10 kcal mol?1. © 2012 Wiley Periodicals, Inc. Int J Chem Kinet 44: 524–531, 2012  相似文献   

13.
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   

14.
We used semi-empirical and ab initio calculations to investigate the nucleophilic attack of the OH? ion on the β-lactam carbonyl group. Both allowed us to detect reaction intermediates pertaining to proton-transfer reactions rather than the studied reaction. We also used the PM3 semi-empirical method to investigate the influence of the solvent on the process. The AMSOL method predicts the occurrence of a potential barrier of 20.7 kcal/mol due to the desolvation of the OH? ion in approaching the β-lactam carbonyl group. Using the supermolecular approach and a H2O solvation sphere of 20 molecules around the solute, the potential barrier is lowered to 17.5 kcal/mol, which is very close to the experimental value (16.7 kcal/mol).  相似文献   

15.
Density functional theory (DFT) has been used to study the solvolysis process of the organophosphorus compound P-[2-(dimethylamino)ethyl]-N,N-dimethylphosphonamidic fluoride (GV) with simple nucleophile [hydroxide (HO?)] and α-nucleophiles [hydroperoxide (HOO?) and hydroxylamine anion (NH2O?)]. The lowest energy conformer of GV used for the solvolysis process was identified with Monte Carlo conformational search (MCMM) algorithm employing MMFFs force field followed by DFT calculations. The profound effect was found for α-nucleophiles toward the solvolysis of GV compared to normal alkaline hydrolysis. Incorporation of solvent (water) employing SCRF (PCM) model at B3LYP/6-31+G* showed that solvolysis of GV with hydroperoxide (activation energy = 7.6 kcal/mol) is kinetically more favored compared to hydroxide and hydroxylamine anion (activation energy = 11.0 and 9.2 kcal/mol, respectively). The faster solvolysis of GV with hydroperoxide is achieved due to strong intermolecular hydrogen bonding in the transition state geometry compared to similar α-nucleophile hydroxylamine anion. Assistance of a water molecule in solvolysis of GV affects the activation barriers; however, the hydroperoxidolysis remains the preferential process. The topological properties of electron density distributions for (–X–H···O, X = O, N) intermolecular hydrogen bonding bridges have been analyzed in terms of Bader theory of atoms in molecules (AIM). Further, the analysis was extended by natural bond orbital (NBO) methods for the strength of intermolecular hydrogen bonding in the transition state geometries. This study showed that the reactivity of these α-nucleophiles toward the solvolysis of GV is a delicate balance between the nucleophilicity and hydrogen-bond strength. Solvation governs the overall thermodynamics for the destruction of GV, which otherwise is unfavored in the gas phase studies.  相似文献   

16.
Kinetics of polymerization of acrylamide initiated by Thallium(III) perchlorate was investigated in aqueous perchloric acid medium in the temperature range of 55–70°C. The rates of polymerization were measured varying the concentration of the monomer, initiator, and perchloric acid. The rate of polymerization was found to increase with increase of temperature, monomer concentration, initiator concentration, and perchloric acid concentration. The effect of additives like different solvents, surfactants, and retarders on the rate of polymerization was studied. Molecular weights of the polymer were determined by viscometry. The chain transfer constants for the monomer (CM) and that for the solvent dioxan (Cs) were calculated to be 7.33 × 10?3 and 6.66 × 10?3, respectively. From the Arrhenius plot, the overall activation energy (Ea) was calculated to be 10.68 kcal/mol. The energy of initiation was calculated to be 12.36 kcal/mol. Depending on the results obtained, a suitable reaction mechanism has been suggested and a rate equation has been derived.  相似文献   

17.
Fluorescent nucleoside analogs, commonly used to explore nucleic acid dynamics, recognition and damage, frequently respond to a single environmental parameter. Herein we address the development of chromophores that can simultaneously probe more than one environmental factor while having each associated with a unique spectroscopic signature. We demonstrate that an isomorphic emissive pyridine‐modified 2‐deoxy‐uridine 1 , containing multiple sensory elements, responds to changes in acidity, viscosity, and polarity. Protonation of the pyridine moiety (pKa 4.4) leads to enhanced emission (λem=388 nm) and red‐shifted absorption spectra (λabs=319 nm), suggesting the formation of an intramolecular hydrogen bond with the neighboring pyrimidine carbonyl. This “locked” conformation can also be mimicked by increasing solvent viscosity, resulting in a stark enhancement of emission quantum yield. Finally, increasing solvent polarity substantially impacts the chromophore’s Stokes shift [from 5.8×103 cm?1 at ET(30)=36.4 kcal mol?1 to 9.3 ×103 cm?1 at ET(30)=63.1 kcal mol?1]. The opposite effect is seen for the impact of solvent polarity of the protonated form. The characteristic photophysical signature induced by each parameter facilitates the exploration of these environmental factors both individually and simultaneously.  相似文献   

18.
Second‐order rate constants have been measured spectrophotometrically for the reactions of Op‐nitrophenyl thionobenzoate ( 1 , PNPTB) with HO?, butan‐2,3‐dione monoximate (Ox?, α‐nucleophile), and p‐chlorophenoxide (p‐ClPhO?, normal nucleophile) in DMSO/H2O of varying mixtures at (25.0±0.1) °C. Reactivity of these nucleophiles significantly increases with increasing DMSO content. HO? is less reactive than p‐ClPhO? toward 1 up to 70 mol % DMSO although HO? is over six pKa units more basic in these media. Ox? is more reactive than p‐ClPhO? in all media studied, indicating that the α‐effect is in effect. The magnitude of the α‐effect (i.e., k/kp) increases with the DMSO content up to 50 mol % DMSO and decreases beyond that point. However, the dependency of the α‐effect profile on the solvent for reactions of 1 contrasts to that reported previously for the corresponding reactions of p‐nitrophenyl benzoate ( 2 , PNPB); reactions of 1 result in much smaller α‐effects than those of 2 . Breakdown of the α‐effect into ground‐state (GS) and transition‐state (TS) effects shows that the GS effect is not responsible for the α‐effect across the solvent mixtures. The role of the solvent has been discussed on the basis of the bell‐shaped α‐effect profiles found in the current study as well as in our previous studies, that is, a GS effect in the H2O‐rich region through H‐bonding interactions and a TS effect in the DMSO‐rich media through mutual polarizability interactions.  相似文献   

19.
-We have carried out a very detailed study, using fluorescence and optical flash photolysis techniques, of the photoreduction of methyl viologen (MV2+) by the electron donor ethylene diamine tetraacetic acid (EDTA) in aqueous solution sensitized by the dye acridine orange (AOH+). A complete mechanism has been proposed which accounts for virtually all of the known observations on this reaction. This reaction is novel in that both the triplet and the singlet state of AOH+ appear to be active photochemically. We have shown that mechanisms previously proposed for this reaction are probably incorrect due to an artifact. At pH 7 the fluorescence quantum yield φs of AOH+ is 0.26 ± 0.02 and the fluorescence lifetime is 1.8 ± 0.2 ns. φs is pH dependent and reaches a maximum of 0.56 at pH 4. The fluorescence of AOH+ is quenched by MV2+ at concentrations above 1 mM and the quenching obeys Stern-Volmer kinetics with a quenching rate constant of (1.0 ± 0.1) × 1010M?1 s?1. The quenching of the AOH+ excited singlet state by MV2+ almost certainly returns the AOH+ to its ground state with no photochemistry occurring. EDTA also quenches the fluorescence of AOH· with Stern-Volmer kinetics but with a smaller rate constant (6.4 ± 0.5) × 108M?1s?1 at pH 7. In this case the quenching is reactive resulting in the formation of semireduced AOH. In the presence of MV2+, flash irradiation of AOH+ does result in the reversible formation of the semireduced MV? which absorbs at 603 nm. We attribute this to a photochemical reaction of the triplet state of AOH+ with MV2+. The initial quantum yield for formation of MV? (φMV:)0 was found to be constant at 0.10 ± 0.05 for [MV2+] from 5 × 10?5 to 1.0 × 10?3 with [AOH+] = 8 × 10?6M. Previous workers had found that (φMV:)0 appears to decrease with decreasing [AOH+]; however, on careful investigation, we found this was most probably due to quenching of the triplet state of AOH+ by trace amounts of oxygen. When EDTA is added to a mixture of AOH + and MV2+ at pH 7, the photochemical formation of MV? becomes irreversible as the [EDTA] is increased. The quantum yield for the irreversible formation of MV? exceeds 0.10 becoming as large as 0.16 for [EDTA] = 0.014M. This fact requires that an alternative photochemical process must be operative and we present evidence that this is a reaction of EDTA with the excited singlet state of AOH+ to produce the semi-reduced AOH- which then reacts with MV2+ to produce MV?. The full kinetic scheme was tested by computer simulation and found to be totally consistent. This also enabled the processing of a full set of rate constants. When colloidal PtO2 was added to the optimal mixture [EDTA] = 3.4 × 10?2M; [MV2+] = 5 × 10?4M; [AOH+] = 4 × 10?5M; pH6 H2 gas was produced at a rate of 0.2μmol H2h?1. Thus, acridine orange should serve as an effective sensitizer in reactions designed to use solar energy to photolyze water.  相似文献   

20.
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   

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