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 828 毫秒
1.
The lowest electronic excited state of the complexes [Ru(2,2′-bipyridine)3]2+, fac-[ClRe (CO)3(2,2′-bipyridine)], and fac-[(pyridine) Re (CO)3(2,2′-bipyridine)]+ can be quenched by methyl viologen, MV2+, N,N′-dimethyl-4,4′-bipyridinium, in fluid solutions. The quenching obeys Stern—Volmer kinetics as deduced from plots of relative luminescence quantum yield vs [MV2+], and the data are consistent with a quenching process that is essentially diffusion controlled. Pulsed laser excitation (18 ns, 354.7 nm frequency tripled Nd: YAG) of the metal complexes in the presence of MV2+ shows that a detectable fraction of the quenching results in net electron transfer to form MV+. The MV+ is detectable by resonance Raman scattering from the trailing portion of the excitation pulse. Excited state electron transfer to MV2+ from a photo-excited complex anchored to SiO2 has also been detected by transient Raman spectroscopy. High surface area SiO2 was functionalized by reaction with 4-[2-(trimethoxysilyl)ethyl]pyridine to give [SiO2]-SiEtpyr. Reaction of [SiO2]-SiEtpyr with [(CH3CN)Re(CO)3(2,2′-bipyridine)]+ then yields [SiO2]-[(SiEtpyr) Re (CO)3 (2,2′-bipyridine)]+. Electron transfer quenching of the photo-excited immobilized Re complex occurs when suspended in CH3CN solutions of MV2+ to yield MV+ as detected by resonance Raman scattering and by lifetime attenuation in the presence of MV2+.  相似文献   

2.
A novel rhenium(I) bipyridyl complex 1a, [(4,4’-di-COOEt-bpy)Re(CO)3(py-NHCO-PTZ)PF6] and a model 1b, [(4,4’-di-COOEt-bpy)Re(CO)3(py-PTZ)PF6] (bpy is 2, 2’-bipyridine, py-NHCO-PTZ is phenothiazine-(10-carbonyl amide) pyridine and py-PTZ is 10-(4-picolyl) phenothiazine) were synthesized. Their photo-induced electron transfer (ET) reaction with electron acceptor methyl viologen (MV2+) in acetonitrile was studied by nanosecond laser flash photolysis at room temperature. Photoexcitation of 1 in the presence of MV2+ led to ET from the Re moiety to MV2+ generating Re(II) and methyl viologen radical (MV·+). Then Re(II) was reduced either by the charge recombination with MV·+ or by intramolecular ET from the attached PTZ, regenerating the photosensitizer Re(I) and forming the PTZ radical at 510 nm. In the case of 1b, the absorption for PTZ radical can be observed distinctly accompanied intermolecular ET, whereas not much difference at 510 nm can be detected for 1a on the time scale of the experiments. This demonstrates that the linking bridge plays a key role on the intramolecular ET in complex 1.  相似文献   

3.
Electron transfer from photoexcited tetrasulfonated Zn(II)-tetraphenylporphyrin (ZnTSTPP) to methyviologen (MV2+) was studied. From the investigation of relative fluorescence intensity and emission lifetime against the MV2+ concentration, it was concluded that the electron transfer takes place by a static mechanism. Based on the analysis of the quenching behavior, it was concluded that the static reaction did not follow an ordinary Perrin model, but interaction of the donor (photoexcited Zn-TSTPP) and the acceptor (MV2+) molecules, ionic interaction in the present case, is responsible. The analysis of the quenching gave the equilibrium constant for the interaction to be K = 6.5 × 104 M−1. A two-dimensional selfassembled macromolecular ionic complex between ZnTSTPP and MV2+ is proposed.  相似文献   

4.
Results are reported for the reaction of methylviologen radical cation, MV+ with platinum colloidal particles, studied by stopped flow spectrophotometry. The rate of the reaction depends on the gaseous pretreatment of the particles. For particles reduced by hydrogen, the kinetics are usually first order with respect to MV+. The reaction is also first order in the concentration of platinum, and is inhibited in a first order manner by the product MV2+. This inhibition suggests that MV2+ is adsorbed on the particle surfaces, and this has been confirmed by ac, ring—disc electrode studies on macroscopic platinum electrodes. At high concentrations of MV+ some deviation from first order kinetics is observed. These results are all explained by a kinetic model in which either the desorption of MV2+ or the adsorption of MV+ is the rate limiting process. The rate of consumption of MV+ on an oxidised surface is an order of magnitude faster than that on the reduced surface. Ring—disc studies show that this is because the MV+ is not producing H2 but is reducing the surface oxide. The results are shown to fit a simple model which takes into account this titration of the oxide layer. The model also explains why the rate on partially oxidised surfaces will appear to have an order greater than one in [Pt].  相似文献   

5.
-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   

6.
A fast and sensitive photokinetic method for the determination of paraquat (MV2+) (1?27×10?5M) is described, based on the rate of photoreduction of MV2+ by EDTA, sensitized by acridine yellow in the absence of oxygen. The rate of photoreduction, which is a linear function of the concentration of MV2+ is monitored polarographically by recording the limiting current of p- benzoquinone, which is reduced by the radical monocation MV generated in the photochemical reaction. The results obtained by the application of the fixed-time, fixed-concentration change and initial-rate kinetic methods are evaluated. An alternative method for monitoring the rate of the process is by measuring the time necessary for the total reduction of p-benzoquinone. The end-point is detected with two platinum electrodes at an applied voltage of 100 mV. The procedure has been successfully applied to the determination of paraquat in commercial herbicides, waters, and flowers and in spiked soils and blood sera.  相似文献   

7.
The ultrafast charge separation at the quantum dot (QD)/molecular acceptor interface was investigated in terms of acceptor concentration and the size of the QD. Time‐resolved experiments revealed that the electron transfer (ET) from the photoexcited QD to the molecular acceptor methylviologen (MV2+) occurs on the fs time scale for large acceptor concentrations and that the ET rate is strongly reduced for low concentrations. The increase in the acceptor concentration is accompanied with a growth in the overlap of donor and acceptor wavefunctions, resulting in a faster reaction until the MV2+ concentration reaches a saturation limit of 0.3–0.4 MV2+ nm?2. Moreover, we found significant QD size dependence of the ET reaction, which is explained by a change of the free energy (ΔG).  相似文献   

8.
The methyl viologen dication, used under the name Paraquat as an agricultural reagent, is a well‐known electron‐acceptor species that can participate in charge‐transfer (CT) interactions. The determination of the crystal structure of this species is important for accessing the CT interaction and CT‐based properties. The title hydrated salt, bis(1,1′‐dimethyl‐4,4′‐bipyridine‐1,1′‐diium) hexacyanidoferrate(II) octahydrate, (C12H14N2)2[Fe(CN)6]·8H2O or (MV)2[Fe(CN)6]·8H2O [MV2+ is the 1,1′‐dimethyl‐4,4′‐bipyridine‐1,1′‐diium (methyl viologen) dication], crystallizes in the space group P 21/c with one MV2+ cation, half of an [Fe(CN)6]4− anion and four water molecules in the asymmetric unit. The FeII atom of the [Fe(CN)6]4− anion lies on an inversion centre and has an octahedral coordination sphere defined by six cyanide ligands. The MV2+ cation is located on a general position and adopts a noncoplanar structure, with a dihedral angle of 40.32 (7)° between the planes of the pyridine rings. In the crystal, layers of electron‐donor [Fe(CN)6]4− anions and layers of electron‐acceptor MV2+ cations are formed and are stacked in an alternating manner parallel to the direction of the −2a + c axis, resulting in an alternate layered structure.  相似文献   

9.
李波  吕功煊 《物理化学学报》2013,29(8):1778-1784
以曙红Y(EY)敏化Pt/TiO2(EY-Pt/TiO2)光催化产氢体系为模型, 研究了电子传递剂甲基紫精(MV2+)的加入对该体系产氢活性和稳定性的影响, 并通过紫外-可见光(UV-Vis)吸收光谱、荧光光谱和光电化学表征手段对MV2+的作用机制进行了研究. 结果表明, 当以三乙醇胺(TEOA)为电子给体时, MV2+可使EY激发态发生氧化性和还原性淬灭, 有效降低了不稳定中间体EY3-·的形成和积累, 促进了电子由染料分子向产氢活性位点的有效传递, 从而提高了产氢体系的活性和稳定性. 两种敏化体系瞬态光电流以及产氢活性受EY浓度影响的差异进一步证明, MV2+作为电子传递剂有效提高了光生电子的传递和利用效率.  相似文献   

10.
Feasibility was demonstrated for the catalysis of the sodium sulfide reduction of methylviologene (MV2+) in aqueous solution using cupric sulfide nanoparticles. The catalytic activity of the nanoparticles depends on their size. The basic features were found for the formation of the MV radical–cation in the reduction of MV2+ by HS anions in the presence of CuS nanoparticles as the catalyst. This is an equilibrium reaction.  相似文献   

11.
The behavior of viologen polymer (P-V2+) as an electron transfer catalyst in the reaction of hydrogen generation was studied. In the photoirradiation system, which contains triethanolamine (TEA), Ru(bpy)3+3, and P-V2+, the amount of hydrogen evolution was less than methyl viologen (MV2+); P-V2+, however, was more effective in sodium dithionite as the electron donor and showed higher initial rates than MV2+.  相似文献   

12.
A dosimeter of poly(vinyl alcohol) (PVA) film containing methyl viologen dichloride (MV2+ (Cl-)2) was characterized by means of ESR and u.v. spectrometries. γ-irradiation of the MV2+-PVA dosimeter induced one-electron reduction of MV2+· to thecation radical (MV+), thus giving rise to blue coloration. The resulting MV showed an ESR signal with a g-factor of 2.0031. The yield of MV at a given radiation dose was estimated from duplicate integral of the ESR first-derivative spectra by reference to 1,1'-diphenyl-2-picrylhydrazyl (DPPH). The yield of MV thus estimated increased linearly with increasing the radiation dose up to about 1.4 Mrad. The ESR spectrometry of MV showed a linear correlation with the u.v. spectrometric method reported previously.  相似文献   

13.
Abstract— Fluorescence quenching of amphiphilic copolymers, poly(9-vinylphenanthrene-co-sodium 2-acrylamido-2-methylpropanesulfonate) (APh) and poly(9-vinylphenanthrene-co-3-methacrylamidopropyltrimethylammonium methyl sulfate) (QPh), in aqueous solution, was studied using methyl viologen (MV2+) or 4,4'-bipyridinium-1, 1'-bis(trimethylenesulfonate) (SPV) as oxidative quenchers. The fluorescence of the excited phenanthrene groups in APh was found to be efficiently quenched by MV2+. The apparent second-order rate constant for the quenching, kq, ranged in the magnitude of 1011 -1012M-1 s-1, which are well beyond the diffusion-controlled limit. This is presumably due to an increase of the effective concentration of MV2+ around the fluorophore in the copolymer resulting from electrostatic attraction between MV2+ and anionic segments of APh. This strong electrostatic interaction also favors the formation of ground-state EDA (electron donor acceptor) complex between the phenanthrene residue and MV2+. Such striking behaviors were not observed with the related model compound. Unexpectedly, the quenching with SPV, a zwitterionic quencher, was also enhanced in the polymer system (kq= 2–6 × 1010M-1 s-1), suggesting the presence of some attractive interaction between APh and SPV. Contrary to the APh system, the fluorescence quenching of the corresponding cationic polymer (QPh) with MV2+ was strongly diminished (kq= 5 × 108M-1 s-1). This indicates that the polycation of QPh effectively prevents the access of MV2+ to the polymer.  相似文献   

14.
We have developed a new intermediate monomer, 2,7‐[bis(4,4,5,5‐tetramethyl‐1,3,2‐dioxaborolan‐2‐yl)‐9,9‐bis(3‐(tert‐butyl propanoate))]fluorene, that allows the easy synthesis of water‐soluble carboxylated polyfluorenes. As an example, poly[9,9′‐bis(3′′‐propanoate)fluoren‐2,7‐yl] sodium salt was synthesized by the Suzuki coupling reaction, and the properties of the polymer were studied in aqueous solutions of different pH. Fluorescence quenching of the polymer by different cationic quenchers (MV2+, MV4+, and NO2MV2+; MV=methyl viologen) was studied, and the quenching constants were found to be dependent on the charge and electron affinity of the quencher molecule and the pH of the medium. The largest quenching constant was observed to be 1.39×108 M ?1 for NO2MV2+ at pH 7. The change in polymer fluorescence upon interaction with different proteins was also studied. Strong fluorescence quenching of the polymer was observed in the presence of cytochrome c, whereas weak quenching was observed in the presence of myoglobin and bovine serum albumin. Lysozyme quenched the polymer emission at low protein concentrations, and the quenching became saturated at high protein concentrations. Under similar experimental conditions, the polymer showed improved quenching efficiencies toward cationic quenchers and a more selective response to proteins relative to other carboxylated conjugated polymers.  相似文献   

15.
The kinetics and mechanism of the uncatalyzed and Ru(III)‐catalyzed oxidation of methylene violet (3‐amino‐7‐diethylamino‐5‐phenyl phenazinium chloride) (MV+) by acidic chlorite is reported. With excess concentrations of other reactants, both uncatalyzed and catalyzed reactions had pseudo‐first‐order kinetics with respect to MV+. The uncatalyzed reaction had first‐order dependence on chlorite and H+ concentrations, but the catalyzed reaction had first‐order dependence on both chlorite and catalyst, and a fractional order with respect to [H+]. The rate coefficient of the uncatalyzed reaction is (5.72 ± 0.19) M?2 s?1, while the catalytic constant for the catalyzed reaction is (22.4 ± 0.3) × 103 M?1 s?1. The basic stoichiometric equation is as follows: 2MV+ + 7ClO2? + 2H+ = 2P + CH3COOH + 4ClO2 + 3Cl?, where P+ = 3‐amino‐7‐ethylamino‐5‐phenyl phenazinium‐10‐N‐oxide. Stoichiometry is dependent on the initial concentration of chlorite present. Consistent with the experimental results, pertinent mechanisms are proposed. The proposed 15‐step mechanism is simulated using literature; experimental and estimated rate coefficients and the simulated plots agreed well with the experimental curves. © 2003 Wiley Periodicals, Inc. Int J Chem Kinet 35: 294–303, 2003  相似文献   

16.
In DMSO/water (4:1), photolysis of the dihydroxy-Sn (IV)-mesoporphyrin dimethyl ester (SnP)/methyl viologen (MV2+)/ethylene diamine tetraacetic acid (EDTA) ternary system produces methyl viologen cation radical with a quantum yield of 0.67, much higher than that of systems with other metal complexes of mesoporphyrin dimethyl ester. Neither EDTA nor MV2+ quenches the stationary fluorescence of SnP, implying that the reaction does not take place at the singlet state. With flash photolysis we obtain the T-T absorption spectrum of SnP (λmax 440 nm). By following the decay of this absorption, the triplet life time of SnP is estimated to be 41 μs. The life time is related to the concentration of either MV2+ or EDTA. Good linear relationships are obtained by plotting τ0τ vs. the concentration of MV2+ or EDTA (Stern-Volmer plot), from which we determine the quenching constants: kq(MV2+) =5.5 × 107 mol?1, s?1; kq (EDTA) =2.7 × 107 mol?1, s?1. The data suggests that upon photolysis of the above ternary system, both oxidative quenching and reductive quenching of the triplet state of the sensitizer are occurring. From the measured phosphorescence spectrum (λmax 704 nm) and the ground state redox, potentials (Ered1/2?-0.84V, Eox1/2?+1.43 V, vs. Ag/AgCl, KCl (sat.)), we obtain the redox potential of triplet SnP to be E(P+/P*T)?-0.33 V, E(P*T+/P?)?+0.92 V. Matching this data with the redox potential of MV2+ and EDTA, we establish the fact that during the photolysis of the SnP/MV2+/EDTA ternary system, both oxidative and reductive quenching are thermodynamically favorable processes. This is also the reason why the SnP sensitized reaction is much more efficient relative to other mesoporphyrin derivatives.  相似文献   

17.
The pulse radiolysis of FA and FA:water solutions was studied in the absence and presence of redox indicator 1,1′-dimethyl-4,4′-bipyridinium dichloride (methyl viologen, MV2+). The experiments performed in the presence of MV2+ have provided strong support to the idea that the first species obtained from the reaction of esol and OH with FA produces radicals that show reactivity towards the MV2+. Both the radicals on reaction with MV2+ results in the appearance of the well-known intense blue MV•+ radical absorption signal (λmax = 395 nm, λmax = 605 nm). The intermediate radicals formed during radiolysis were used to generate silver nanoparticles.  相似文献   

18.
Photochemical properties of Ru(bpy)2(poly-4-methyl-4′-vinyl-2,2′-bipyridine)Cl2 ( 2 ) were studied and compared with that of Ru(bpy)3Cl2. Continuous irradiation of a solution, which contains polymer 2 as a photosensitizer, methylviologen (MV2+) or 4,4′-bipyridinium-1,1′-bis(trimethylenesulfonate) (SPV) as an electron acceptor and triethanolamine (TEOA) as a sacrificial donor, resulted in the formation of viologen radical ion (MV+ or SPV?). The rate of formation of MV+ or SPV? for the polymer 2 system was smaller than that for the Ru(bpy)3 Cl2 systems. The reason for this fact was kinetically analyzed by quenching experiments of excited Ru(II) complexes by MV2+ or SPV, the photosensitized reactions of the TEOA–Ru(II) complex–MV2+ or -SPV systems, and the dye laser photolysis of the Ru(II) complex–MV2+ or -SPV systems.  相似文献   

19.
《中国化学快报》2023,34(3):107346
A cadmium tetracyanoplatinate host clathrate, (MV)[Cd2{Pt(CN)4}3]?2(H2O) (1), including a methylviologen dication (MV2+) was synthesized, and the crystal structures, photochromic and photoluminescence properties were investigated. In 1, the alternatively parallel stacking between the MV2+ dications as electron acceptors in the channels and the electron donors [Pt1(CN)4]2– units in the host frameworks give a unique donor-acceptor (DA) system. Under UV irradiation, the electron transfer between MV2+ and [Pt(CN)4]2– ions generates MV·+ radicals with a photochromic behavior from pale-yellow to blue. This process occurs through single-crystal-to-single-crystal (SCSC) transformation and obvious structure variation of viologen cations is successfully observed. Moreover, the spectral overlap between the emission bands of 1 and the absorption around 623 nm for the MV·+ radicals leads to a modulation of the photoluminescence.  相似文献   

20.
Tert-Butyl hydroperoxide and hydrogen peroxide readily react with the radical cation derived from 2,2′-azinobis(3-ethylbenzothiazoline-6-sulfonic acid) (ABTS). The reaction is inhibited by ABTS and protons, and can be interpreted in terms of a mechanism comprising a partially reversible electron transfer ROOH+ABTS•+↔ ROO · + ABTS + H+ (1) followed by the self-reactions of the hydroperoxide derived radicals and reactions between them and another ABTS derived radical. A complete kinetic analysis allows an evaluation of the rate constant for reaction (1). A value of 0.2 M−1 s−1 was obtained for both compounds. The back reaction of process (1) is more relevant when tert-butyl hydroperoxide is employed. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet 30: 565–570, 1998  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