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 31 毫秒
1.
Pulse radiolysis of acetonitrile solutions of tetra-n-butyl ammonium salts of 2- and 4-carboxybenzophenones [BP-COO···N+(C4H9)4] were performed in order to generate directly the reduced forms of the benzophenone moieties within pre-formed ion pairs. In earlier studies on photochemical electron transfer reactions, ion pairs containing a tetraalkyl ammonium cation and a benzophenone radical anion were formed in an electron transfer to the triplet BP from a quencher consisting of a tetraalkyl ammonium salt of (phenylthio)acetic acid. In the current work, the [BP•−COO···N+(C4H9)4] ion pairs were formed by direct reduction of the salts without the complication of a third moiety, i.e., the (phenylthio)acetic anion. The spectra and kinetic parameters of the radiolytically-reduced salts were compared to the behavior of reduced forms of the 2- and 4-COOH substituted benzophenones. The results from the pulse radiolysis and photochemistry were compared and explained in terms of the different structures of the ion pairs.  相似文献   

2.
The mechanism of the photoinduced reaction of the lowest excited singlet state of the 10-methylacridinium (AcrMe+) cation with benzyltrimethylsilane (BTMSi) in acetonitrile has been investigated by means of steady-state and time-resolved methods. A variety of stable products was found after irradiation (365 nm) of the reaction mixture under aerobic and oxygen-free conditions. The stable products were identified and analyzed using UV–Vis spectrophotometry, high performance liquid chromatography (HPLC), and mass spectrometry (MS). Based on Stern–Volmer plots of the AcrMe+ fluorescence quenching by BTMSi (using fluorescence intensity and lifetime measurements), the rate constants were determined to be k q = 1.24 (± 0.02) × 1010 M−1 s−1 and k q = 1.23 (± 0.02) × 1010 M−1 s−1, i.e., close to the diffusion-controlled limit in acetonitrile, indicating the dynamic quenching mechanism. The quenching process was shown to occur via an electron-transfer reaction leading to the formation of acridinyl radicals (AcrMe) and C6H5CH2Si(CH3)3 •+ radical cations. Based on stationary and flash photolysis experiments, a detailed mechanism of the secondary reactions is proposed and discussed. The AcrMe radical was shown to decay by two processes. The fast decay, observed on the nanosecond timescale, was attributed to the back-electron transfer occurring within the initial radical ion pair. The slow decay on the microsecond timescale was explained by recombination reactions of radicals which escaped from the radical pair, including benzyl radicals formed via C–Si bond cleavage in the C6H5CH2Si(CH3)3 •+ radical cation.  相似文献   

3.
Chemiluminescence activated by the chelates Eu(tta)3phen and Eu(dbm)3phen in the thermal decomposition of diphenyldiazomethane in benzene in the presence of oxygen was examined at 333 K. The following photophysical characteristics of the luminescence emitter, excited triplet benzophenone (BP), were found: the phosphorescence quantum yield ϕ BP 0 = 1.0 × 10−5, the true lifetime τ BP 0 = 7 × 10−8 s, and the rate constant of radiative deactivationk r, BP = ϕ BP 0 τ BP 0 −1 = 160 s−1.  相似文献   

4.
The reactions of ytterbium naphthalene complex C10H8Yb(THF)2 with 2-cyclopentadienylethanol, 1-cyclopentadienylpropan-2-ol, 3-cyclopentadienyl-1-butoxypropan-2-ol, and cyclopentadienyldimethylsilyl-tert-butylamine were studied. The bivalent ytterbium complexes with chelate bifunctional cyclopentadienyl ligands [(η5−C5H5)CH2CH21−O)]Yb(THF), [(η5−C5H5)CH2CH21−O)]Yb(DME). [(η5−C5H5)CH2CH(Me)(η1−O)]Yb(THF), [(η5−C5H5)CH2CH(CH2OC4H9)(η1−O)]Yb(THF), and [(η5−C5H5)SiMe21−N(Bu1))]Yb(THF) were obtained and characterized. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 4, pp. 742–745, April, 2000.  相似文献   

5.
The CIDEP spectra of transient radicals during photolysis of the duroquinone (DQ)/ethylene glycol (EG) system in acid, basic, and micellar environments were measured with a home-made highly time-resolved ESR spectrometer. In the DQ/EG homogeneous solution, the enhanced emissive CIDEP signal of the neutral durosemiquinone radical DQH was observed. When the DQ/EG solution at pH 9 or the DQ/EG/TX-100/H2O micelle system was photolyzed, the CIDEP signal of the duroquinone anion radical (DQ•−) was obtained. When the DQ/EG solution at pH 2.5 was irradiated, the CIDEP signal of DQH appeared. These experimental results indicate that the neutral radical DQH was formed by proton transfer from EG to 3DQ*, and that DQ•− was formed by dissociation of DQH accompanying polarization transfer.  相似文献   

6.
Two stable thiazolylazo anion radical complexes of ruthenium(II), [Ru(L1•−)(Cl)(CO)(PPh3)2] (1) and [Ru(L2•−)(Cl)(CO)(PPh3)2] (2) (where L1 = 2′-Thiazolylazo-2-imidazole and L2 = 4-(2′-Thiazolylazo)-1-n-hexadecyloxy-naphthalene), have been synthesized and characterized by spectroscopic and electrochemical techniques. The radical nature of the complexes has been confirmed from their room temperature magnetic moments and X-band ESR spectra. The radical complexes display a moderately intense (ε ~ 104 M−1 cm−1) and relatively broad band in 430–460 nm region. In the microcrystalline state, complexes (1) and (2) display strong ESR signals at g = 1.951 and g = 1.988, respectively. In CH2Cl2 solution, complexes (1) and (2) show a quasireversible one-electron response near −0.64 and −0.59 V, respectively, versus Ag/AgCl due to the radical redox couple [RuII(L)(Cl)(CO)(PPh3)2]/[RuII (L•−)(Cl)(CO)(PPh3)2].  相似文献   

7.
Magnetic exchange couplings in bis(ketimide) binuclear UIV/UIV complexes [Cp′2UCl]2(μ-ketimide) diuranium(IV) and [(C5H5)2(Cl)An]2(μ-ketimide) (Cp′ = C5Me4Et; ketimide = N=CMe-(C6H4)-MeC=N) have been investigated computationally using relativistic density functional theory (DFT) combined with the broken symmetry (BS) approach. Using the B3LYP hybrid functional, the BS ground state of these UIV/UIV 5f 2–5f 2 complexes has been found of lower energy than the high spin (HS) quintet state, indicating an antiferromagnetic character (estimated coupling constant |J| < 5 cm−1) which has not yet been evidenced unambiguously experimentally. On the contrary, the BP86 GGA functional overestimates greatly the antiferromagnetic character of the complexes (|J| > 100 cm−1). As recently reported for para-bis(imido) [(C5H5)3U]2(μ-imido) uranium(V) complex, spin polarization is mainly responsible for the antiferromagnetic coupling through the π-network orbital pathway within the bis(ketimide) bridge. Furthermore, spin polarization is exalted by the combined roles of the 5f metal orbitals and of the π-conjugated ketimide bridging ligand which permit electronic communication between the two uranium atoms albeit separated by a distance of the order of 10 ?. The MO analysis clarifies which MOs contribute to the antiferromagnetic coupling in the binuclear complexes under consideration and brings to light the 5f orbitals driving contribution.  相似文献   

8.
The reaction of aniline with hydrogen atom is investigated herein using the hybrid meta-DFT functional of BB1 K. Hydrogen atom is found to preferentially add at an ortho position. However, the fate of the o-(C6H5NH2)H adduct is found to be solely the deactivation of the initial addition channel. The rate constant for the abstraction channel (C6H5NH2 + H → C6H5NH + H2) is fitted by the expression 1.10 × 10−11 exp(−4,200/T) cm3 molecule−1 s−1. Our calculated rate constant for the abstraction channel agrees very well with the available experimental measurements. Satisfactory agreement is found between calculated and experimental measurements for the displacement channel (C6H5NH2 + H → C6H6 + NH2). Our detailed analysis for the corresponding displacements in toluene and phenol suggests that the three systems exhibit similar behavior with regard to the relative importance of abstraction and displacement channels.  相似文献   

9.
Gamma radiolysis of oxygenated 1–10 mM azide solutions was carried out at various pH values. In oxygenated 10 mM azide solutions, H2O2 and NO 2 were observed as radiolytic products while NH3 was not. The concentration of H2O2 reached its maximum level at a dose of 1 kGy, whereas NO 2 yield increased non-linearly beyond 2 kGy in this system. Both in aerated and oxygenated systems, G(NO 2 ) and G(H2O2) were found to vary with N 3 concentration. The yield of NO 2 was found to be dependent on both dose rate and pH. On pulse radiolysis, NO 2 was found as a radiolytic product in aerated 1 mM azide solution at pH 6.8. In this system the intermediate generated exhibits absorbance around 250 nm. The overall results obtained during the present study reveal that in presence of both reducing radical (mainly e aq ) and oxygen, N 3 produced an intermediate possibly NH2O 2 radical, which is the prime source for NO 2 generation.  相似文献   

10.
Laser ablation of titanium oxides at 355 nm and ion–molecule reactions between [(TiO2)x]–• cluster anions and H2O or O2 were investigated by Fourier transform ion cyclotron resonance mass spectrometry (FTICR MS) with an external ion source. The detected anions correspond to [(TiO2)x(H2O)yOH] and [(TiO2)x(H2O)yO2]–• oxy-hydroxide species with x = 1 to 25 and y = 1, 2, or 3 and were formed by a two step process: (1) laser ablation, which leads to the formation of [(TiO2)x]–• cluster anions as was previously reported, and (2) ion–molecule reactions during ion storage. Reactions of some [(TiO2)x]–• cluster anions with water and dioxygen conducted in the FTICR cell confirm this assessment. Tandem mass spectrometry experiments were also performed in sustained off-resonance irradiation collision-induced dissociation (SORI-CID) mode. Three fragmentation pathways were observed: (1) elimination of water molecules, (2) O2 loss for radical anions, and (3) fission of the cluster. Density functional theory (DFT) calculations were performed to explain the experimental data.  相似文献   

11.
Reactions of radiolytically generated CO3 •− with some ferric heme proteins, catalase, cytochrome c, and horseradish peroxidase (HRP), were studied. Carbonate radical anion oxidized amino acid residues of these proteins, but did not react directly with heme iron. HRP and catalase lost about 30% and 20% of their activity, respectively, after the reaction with 100 μM of CO3 •−. The rate constants of the reactions of CO3 •− with the investigated proteins measured by the pulse radiolysis method at pH 8–8.4 and 10 varied from 1.0 × 108 M−1 s−1 (for cytochrome c) to 3.7 × 109 M−1 s−1 (for catalase).  相似文献   

12.
The OH and the NO2 radicals generated pulse radiolytically in N2O-saturated aqueous solution at pH 8–8.5 oxidize Mesna to form the corresponding thiyl radicals which on reaction with thiolate ions form an RSSR type of transient with λmax = 420 nm. The rate constants for the formation of these transients were determined. In the absence of O2 at pH=6, the RS radicals formed show an absorption maximum at 360 nm and an ε=200±50 dm3 mol−1 cm−1. The rate constant k (OH+RSH) was 6×109 dm3 mol−1 s−1 as determined from competition kinetics. In the presence of O2 the Mesna thiyl radical was seen to rapidly add oxygen to form an RSOO type of species with λmax = 535 nm, ε=700±50 dm3 mol−1 cm−1 and k (RS+O2)=1.3×108 dm3 mol−1 s−1. Both the RS and the RSOO radicals formed by the oxidation of Mesna were able to abstract H-atoms from ascorbate ions and k(RS +AH)=~k(RSOO+AH)=~6−7×108 dm3 mol−1 s−1-. Moderately strong oxidants like CCl3OO and the (CH3)3CO radicals, having a reduction potential of +1.4−1.6 V vs NHE were unable to oxidize Mesna. The results thus reflect on the pro- and anti-oxidant properties of Mesna.  相似文献   

13.

Abstract  

O-Tolyl/benzyl dithiocarbonates, ROCS2Na (R = o-, m-, or p-CH3C6H4–, and –CH2C6H5), were synthesized and characterized. These new ligands reacted with PCl3/POCl3 in refluxing toluene which resulted in the formation of phosphorus(III) and phosphorus(V) tolyl/benzyl dithiocarbonates corresponding to [(ROCS2) n PCl3−n ] and [(ROCS2) n POCl3−n ] (R = o-, m-, or p-CH3C6H4–, and –CH2C6H5; n = 1, 2, 3). These pale yellow liquid compounds were characterized by IR, mass, and NMR (1H, 13C, and 31P) spectral studies, which suggest the dithiocarbonate ligands bind in a monodentate mode leading to P–S–C linkages in these derivatives.  相似文献   

14.
The decomposition studies of S-nitrosothiols (RSNO) are important due to their potential role in vivo in connection with the storage and transport of nitric oxide (NO) within the body. Reactions of hydroxyl radicals (OH) with a number of RSNOs (S-nitroso derivatives of N-acetyl-dl-penicillamine, l-cysteinemethylester, N-acetylcysteamine, and dl-penicillamine) in aqueous medium at neutral and acidic pH have been reported in the present study. Radiation chemical technique (steady state and pulse radiolysis) has been utilized for the determination of the reaction rate constants, the end product analyses, and the transient intermediate species. The rate constants for the reaction of OH with the selected RSNOs were determined using a competition kinetic method with 2′-deoxy-d-ribose as the competitor. All the rate constants were found to be of the order of diffusion controlled (1010 M−1 s−1). The degradation yield of RSNOs was found to be quantitative (i.e., G(–RSNO) ≈ G(OH)) at neutral and acidic pH. The major products of decomposition were the respective disulfide (RSSR) and nitrite (NO2 ). A good material balance is also obtained between the degradation yield and the formation of the products (i.e., G(–RSNO) ≈ G(RSSR) + G(NO2 )). The major transient intermediate was the thiyl radical (RS). Its intermediacy was confirmed by making use of the electron transfer reaction of 2,2′-azinobis(3-ethylbenzothiazoline-6-sulfonate) (ABTS2−) to RS, which results in the formation of ABTS•− having a transient absorption spectrum with λmax at 410 nm. Based on these results, a generalized reaction mechanism is deduced for the reaction of OH with RSNO.  相似文献   

15.
Triphenylmethyl salts of the very weakly-coordinating borate anions [CN{B(C6F5)3}2] (1), [H2N{(C6F5)3}2] and [M{CNB(C6F5)3}4]2− (M = Ni, Pd) have been prepared in simple one-pot reactions. Mixtures of (SBI)ZrMe2/1/AlBu 3 i (SBI = rac-Me2Si(Ind)2) are 30–40 times more active in ethylene polymerizations at 60–100°C than (SBI)ZrCl2/MAO. The quantification of anion effects on propene polymerization activity at 20°C gives the order [CN{B(C6F5)3}2] > [H2N{(C6F5)3}2] ≈ B(C6F5) 4 ≫ [MeB(C6F5)3]. The highest productivities were of the order of ca. 3.0 × 108 g PP (mol Zr)−1 h−1 [C3H6]−1, about 1.3–1.5 times higher than with B(C6F5) 4 as the counter anion. The titanium system CGCTiMe2/1/AlBu 3 i gave activities that were very similar to the zirconocene catalyst. The concentration of active species [C*] as determined by quenched-flow kinetic techniques indicates typical values of around 10%, independent of the counter anion, for both the borate and MAO systems. Pulsed field-gradient spin echo and nuclear Overhauser effect NMR experiments on systems designed to be more realistic models for active species with longer polymeryl chains, (SBI)M(CH2SiMe3)(μ-Me)B(C6F5)3 and [(SBI)MCH2SiMe 3 + ...B(C6F5) 4 ] (M = Zr, Hf), demonstrated the influence of bulky alkyl chains on the ion pair solution structures: while the MeB(C6F5)3 compound exists as a simple inner-sphere ion-pair, the B(C6F5) 4 compound is an outer-sphere ion pair (OSIP), a consequence of the relegation of the anion into the second coordination sphere by the γ-agostic interaction with the alkyl ligand. The OSIP aggregates to ion hextuples (10 mM) or quadruples (2 mM). Implications for the polymerization mechanism are discussed; the process follows an associative interchange (I a) pathway. The text was submitted by the authors in English.  相似文献   

16.
The structure, stability, and potential existence of chromium complexes with triallylborane B(CH2—CH=CH2)3 and its analogs with coordination of the Cr atom to three double bonds are discussed. Complexes LCr(CO)3 are studied where L = (C3H5)3CH, (C3H5)3B, (C3H5)3BNMe3, [(C3H5)3BF]−, (C2H4)3, C6H6, (C3H5)3CPh, (C3H5)3B(NC5H5). The calculations are carried out in DFT terms in approximations PBE/3z and BP86/TZ2P. The calculated dissociation energies of the studied complexes into the fragments Cr(CO)3 and L range from 48.3 to 63.0 kcal mol−1 (BP86/TZ2P) and from 54.2 to 66.9 kcal mol−1 (PBE/3z). Ligands (C3H5)3ER (ER = CPh, B(NC5H5)) can form two isomeric complexes A and B due to coordination of tricarbonylchromium to either three double bonds or benzene or pyridine ring. In the former case, isomer A is less stable as compared to isomer B, while in the latter case isomer Ais more stable. The possibility of selective synthesis of one of these isomers, namely, tris(η2-allyl)pyridineborane complex, is predicted.  相似文献   

17.
Studies on photo-catalytic degradation of benzene using TiO2 photo-catalyst as a suspension in water is reported. Degradation studies have been carried out using 350 nm UV light. Phenol, a photo-catalytic product of benzene, was monitored under varying experimental conditions such as amount of TiO2, concentration of benzene, photolysis time, ambient (air, O2, Ar, N2O and N2O–O2 mixture), etc. The phenol yields in both aerated and O2-purged systems increased with the photolysis time. In contrast to oxygenated systems, the yields of phenol in deoxygenated (viz. Ar-purged and N2O-purged) systems were quite low (~30 μM) and remained steady. H2O2 yields in all these systems were also monitored, and found lower than an order of magnitude as compared to phenol yields for the respective systems. The rate of phenol production in aerated 1 mM benzene solution containing 0.05% TiO2 suspension was evaluated at 12.3 μM min−1 which is lower than the rate obtained in an O2-saturated system (22.4 μM min−1). The low yields of phenol in both Ar- and N2O-purged systems, and also the increasing trends in oxygenated systems, together reveal that, for the phenol formation with an enhanced rate, oxygen is essential. In the present study, it is implied that the photo-generated hole, which is mainly an OH radical, is either freely available in the aqueous phase or migrates to the aqueous phase from the catalyst surface, to react with benzene to produce HO-adduct radical. Later, following reaction with oxygen, the adduct produces phenol. On the other hand, h+ and surface adsorbed OH radical, being trapped/bonded due to rigid association with the catalyst surface, were not able to generate phenol under similar experimental conditions. The mechanism of phenol formation with TiO2 photolysis in an aqueous system is rechecked, on the basis of present results on h+/surface adsorbed OH radical/unbound OH radical scavenging by benzene, collectively with previous reports on various systems.  相似文献   

18.
The kinetics of self-termination of benzophenone oxide (BPO) in the liquid phase was studied by flash photolysis. The extinction coefficient of BPO (ε) was found to be virtually independent of the solvent nature, ε=(1.9±0.1)·103 L mol−1 cm−1. The rate constant of the BPO self-temination increases from 1.8·107 (MeCN) and 7.4·107 (C6H6) to 1.5·109 (n-decane) and 2.0·109 L mol−1 s−1 (n-pentane) at 293±2 K. Solvation of BPO promotes a polar state of the molecule in MeCN and C6H6. In nonpolar hydrocarbons, a great contribution is made by the biradical structure resulting in an increase in the rate constant and a shift of the absorption maximum to the long-wave region (from 410 nm in MeCN to 425 nm inn-pentane). In solutions of benzene and acetonitrile, benzophenone oxide reacts with the parent diazo compound with a rate constant of (2–4)·105 L mol−1 s−1 (293±2 K) along with the self-termination. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 7, pp. 1329–1332, July, 1998.  相似文献   

19.
Abstract  Two organic–inorganic hybrid compounds based on the Anderson-type clusters [TeMo6O24]6−, [(H2O)2Co(TeMo6O24)][(C10N2H10)2] · 9.5H2O (1), [(C10N2H9)Ni(H2O)3]2[TeMo6O24] · 8.5H2O (2), have been synthesized by hydrothermal reactions and characterized by elemental analyses, IR spectra, thermal stability analyses, and single-crystal X-ray diffraction. Compound 1 displays a 1D chain structure constructed from alternating [TeMo6O24]6−clusters and Co2+ along the a axis with two pendant ligands 4,4′-bpy (4,4′-bipyridine). Compound 2 is composed of [TeMo6O24]6− clusters coordinated by [Ni(bpy)(H2O)3]2+ moieties, and a supramolecular architecture is further formed through extensive hydrogen bonds interactions. Graphical Abstract  Two organic–inorganic hybrid compounds based on the Anderson-type clusters [TeMo6O24]6−and the unit [M(4,4′-bpy)] have been synthesized under the hydrothermal conditions. Compound 1 displays a 1D chain structure constructed covalently from alternating polyoxoanions [TeMo6O24]6− and Co2+ along the a axis with two pendant ligands 4,4′-bipyridine. Compound 2 is composed of [TeMo6O24]6− polyoxoanion coordinated by [Ni(bpy)(H2O)3]2+ moieties and shows a 1D chain structure through the hydrogen bonds interactions. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

20.
In this study, we observed unprecedented cleavages of the Cβ–Cγ bonds of tryptophan residue side chains in a series of hydrogen-deficient tryptophan-containing peptide radical cations (M•+) during low-energy collision-induced dissociation (CID). We used CID experiments and theoretical density functional theory (DFT) calculations to study the mechanism of this bond cleavage, which forms [M – 116]+ ions. The formation of an α-carbon radical intermediate at the tryptophan residue for the subsequent Cβ–Cγ bond cleavage is analogous to that occurring at leucine residues, producing the same product ions; this hypothesis was supported by the identical product ion spectra of [LGGGH – 43]+ and [WGGGH – 116]+, obtained from the CID of [LGGGH]•+ and [WGGGH]•+, respectively. Elimination of the neutral 116-Da radical requires inevitable dehydrogenation of the indole nitrogen atom, leaving the radical centered formally on the indole nitrogen atom ([Ind]-2), in agreement with the CID data for [WGGGH]•+ and [W1-CH3GGGH]•+; replacing the tryptophan residue with a 1-methyltryptophan residue results in a change of the base peak from that arising from a neutral radical loss (116 Da) to that arising from a molecule loss (131 Da), both originating from Cβ–Cγ bond cleavage. Hydrogen atom transfer or proton transfer to the γ-carbon atom of the tryptophan residue weakens the Cβ–Cγ bond and, therefore, decreases the dissociation energy barrier dramatically.  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