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.
The influence of 20 aromatic amines or carbazoles D and 14 carbonyl compounds A on the cathodic luminescence of 9,10-dichloro-9,10-dihydro-9,10-diphenylanthracene, DPACl2, was investigated in DMF at a rotating platinum disk electrode. The cathodically generated oxidizing agent in the energy sufficient electron transfer luminescence mechanism is intercepted by D if E1/10x(D) is more negative than +1.3 V (SCE). The emission of DPA is also observed if DPACl2 is reduced by the anion radicals of the non-luminescent compounds A. This result proves the participation of DPA+ in the luminescence process. A one-step fractionation of DPACl2 by A? into DPA+ and 2 Cl? is proposed as a mechanism of the formation of this strong oxidizing intermediate in a reduction process. The threshold reduction potential of A in this reaction is about ?1.15 V.  相似文献   

2.
Roberto Martín 《Tetrahedron》2008,64(27):6270-6274
A 9,10-diarylanthracene having two ionophilic imidazolium tags on peripheral positions has been synthesized through a radical chain addition of 1,2-dimethyl-3-(3-mercaptopropyl)imidazolium to 9,10-distyrylanthracene. OLED cells prepared with this derivative exhibits 13.6-fold efficiency enhancement as compared to that prepared with 9,10-diphenylanthracene at the same concentration. Based on electrical conductivity measurements, this efficiency enhancement has been attributed to the blocking of spurious current flow due to ion relocation near the electrodes. The advantage of covalent attachment compared to physical mixtures of 9,10-diphenylanthracene and 1,2-dimethyl-3-(3-mercaptopropyl)imidazolium ionic liquid derives from the experimental difficulty of having an active layer with the target anthracene-to-ionic liquid ratio due to the remarkable differences in viscosity of the two components of the mixture.  相似文献   

3.
During the photolysis of the endoperoxide of 9,10-diphenylanthracene, two different reactions are observed, depending on the irradiation wavelength: (i) Excitation of the S1 band causes a homolytic cleavage of the peroxide bridge with a quantum yield Q2 = 0.08. (ii) Irradiation of the S2 band leads to an adiabatic photocleavage of the endoperoxide into 9,10-diphenylanthracene and singlet molecular oxygen with a quantum yield Q1 = 0.28. Both reaction pathways confirm the theory of Kearns and Khan concerning the photolysis of endoperoxides  相似文献   

4.
Laser photolysis techniques have been employed show that the quenching of the excited electronic states of 9,10-di-phenylanthracene involve an energy transfer mechanism resulting in generation of singlet oxygen from both the lowest singlet and triplet states of the hydrocarbon. Internal conversion from the excited singlet of the 9,10-diphenylanthracene is enhanced by oxygen as well. This process is attributed to formation of sterically hindered conformers. The temperature dependence of internal conversion of 9,10-diphenylanthracene was also examined.  相似文献   

5.
Quenching of fluorescence of polycyclic aromatic hydrocarbons (PAH), namely, naphthalene, anthracene, 9,10-diphenylanthracene, 9,10-dibromoanthracene by C60 fullerene in ethylbenzene at 293 K was found and investigated. The phenomenon is characterized by abnormally high values of bimolecular rate constants for quenching (k bim = (0.18–6.78)·1012 L mol−1 s−1) determined from the Stern—Volmer dependence of the PAH fluorescence intensity on the C60 concentration and occurs through the inductive-resonance (dominant channel) and exchange-resonance (minor channel) energy transfer from 1PAH* to C60. The overlap integrals of the PAH fluorescence spectra with the C60 absorption spectrum and the critical energy transfer distances were calculated. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 3, pp. 432–436, March, 2007.  相似文献   

6.
Measurements of the quantum yield for photoperoxidation of 9,10-diphenylanthracene as a function of dissolved oxygen concentration in toluene at 20°C. are consistent with generation of 1O2 from both the lowest singlet and triplet states of the hydrocarbon.  相似文献   

7.
Reaction of the Reissert anion with the carbonyl group of 1,3,4,6,7,11b-hexahydro-9,10-dimethoxybenzo[a]quinolizin-2-one and with 3-ethyl-1,2,3,4,6,7-hexahydro-9,10-dimethoxy-11bH-benzo[a]quinolizin-2-carboxaldehyde give emetine analogs. This anion does not react with the carbonyl group of 3-aIkyl-1,3,4,6,7,11b-hexahydro-9,10-dimethoxybenzo[a]quinolizin-2-one but instead gives a rearrangement product and the benzoquinolizinone cyanohydrin.  相似文献   

8.
Two CrIII–picolinato complexes were obtained and characterized in solution. The [Cr(C2O4)(pyac)2] and [Cr(C2O4)2(pyac)]2– ions (pyac = picolinic acid anion) in acidic solutions undergo a reversible one-end CrIII–picolinato chelate ring opening via CrIII—N bond breaking. The reaction rate was determined spectrophotometrically in the 0.1–1.0 M HClO4 range at I = 1.0 M. The observed pseudo-first order rate constant depends on [H+] according to the equation: k obs = a + b[H+] + c/[H+]. A reaction mechanism, which assumes participation of the protonated and unprotonated forms of the reactants, has been proposed. The kinetic parameters a, b, c have been defined as a = k 1, b = k 2 Q 1, c = k –1/Q 2, where k 1, k –1,k 2 are rate constants for the forward and reverse processes and Q 1, Q 2 are the protolytic equilibrium constants in the term of the proposed mechanism. The activation parameters have been determined and discussed.  相似文献   

9.
10.
Applying thek 0 standardization method to prompt-gamma activation analysis (PGAA) offers similar benefits as in instrumental neutron activation analysis. It has been demonstrated that under constant flux conditionsk 0-factors obtained by normalizing to a titanium comparator, measured separately, yield consistent analytical sensitivity ratios. The ratio method has been generalized by using stoichiometric compounds for the determination ofk 0-factors. Since chlorine forms compounds with essentially everyelement and it also serves as a detector efficiency standard,k 0 values have been determined relative to chlorine as an internal standard for several analytically important elements in two reactor facilities: the thermal guided beam at the BRR in Budapest and the cold-neutron beams at the NBSR at NIST.  相似文献   

11.
Introduction     
According to the maximum overlap method and the concept that the overlap integral may be used as an indicator of bond strength, a further discussion of bond strength F, an universal formula of F, and the correlation between the bond strength discussed here and that defined by Pauling have been given in this paper. For the type of molecule MLk, under the approximations of k = 2 and using the projection method of the angular part of bond orbitals, the same results as those derived by Pauling can be obtained from the universal formula. From the same formula, two more equations containing g and h atomic orbitals have been derived.  相似文献   

12.
By exposure of α-fluorinated benzophenones to McMurry reaction conditions, we have observed the remarkable formation of 9,10-diphenylanthracene derivatives. This unexpected transformation necessitates the cleavage of the exceptionally stable aromatic C−F bond under mild McMurry conditions. In this work, the condensation of several related fluorinated benzo- and acetophenones has been investigated, which allow us to propose a domino-like fusion mechanism for this unusual transformation. The scope and limitations of the fluorine-promoted benzophenone fusion are subsequently discussed.  相似文献   

13.
An electroanalytical technique has been utilized in a new method for the study of reactive intermediates in polymerization reactions. A ring-disk electrode system generated persistent carbocation radicals whose stability decreased in the order: 1,3,6,8-tetraphenylpyrene (TPP), rubrene (Ru), 9,10-diphenylanthracene (DPA), and 9,10-dimethylanthracene (DMA). Radical cations from these parent compounds flowed to a collecting ring which was controlled potentiostatically to reduce unreacted cations. When styrene or isobytyl vinyl ether was added to the solution, the concentration of carbocation radicals reaching the electrode was reduced. Current collection efficiencies N were determined as a function of rotation speed ω for each monomer concentration. Plots of N?1 as ω?1 in the absence of monomer show no dependence on ω (indicative of stable intermediates), but a linear dependence is found with each concentration of monomer. This indicates a first-order dependence on radical cation concentration. The rate constants show a trend in cation reactivities which is in agreement with that obtained by other methods. This technique, however, extends the range of investigation to a much shorter time scale.  相似文献   

14.
The lifetimes of 9,10-diphenylanthracene in dilute solutions of cyclohexane and benzene at 25°C have been found to be 7.58 ± 0.04 and 6.95 ± 0.04 ns respectively. Measurements of the relative quantum yields show that the dependence on the solvent is caused by an increased probability for non-radiative decay in benzene compared with cyclohexane. This behaviour is shown to partly reconcile previous conflicting data on the radiative properties of this molecule.  相似文献   

15.
A polyimide (6F-THP) with a tetrahydropyranyl group (THP) in its side chain has been synthesized. The THP group exhibits a high acidolysis rate in this polymer's film. This rate was faster than that of a tertbutoxycarbonyl group (t-BOC), which has been previously reported [1]. Furthermore, the deprotected fluorinated polyimide (6FDA-AHHFP) became soluble in an aqueous base due to the presence of a hydroxyl group attached to the phenyl group of the diamine segment. The polyimide thus provides high performance as a photopolymer when used in conjunction with a photoacid generator after the post-exposure baking process (PEB). The photoacid generators used in this study were p-nitrobenzyl-9,10-dimethoxyanthoracene-2-sulfonate (NBAS) and diphenyliodonium-9,10-dimethoxyanthoracene-2-sulfonate (DIAS). The quantum yields of photodissociation and photoacid generation were also measured. The photoacid-generating quantum yields closely corresponded to the photosensitivities of the photoreactive polyimide system. It was confirmed that the THP group was easily deprotected even in the 6F-THP film with p-toluenesulfonic acid as a model acid catalyst. The activation energy of the THP deprotection reaction was determined to be 12.8 kcal/mol (19.5 kcal/mol in the case of t-BOC). The relationships between the THP deprotecting rate constant (kd) and acid molecular size and between kd and polyimide structure were further investigated.  相似文献   

16.
Diazodiphenylmethane ( DDM ) undergoes cycloadditions to 1‐substituted buta‐1,3‐dienes exclusively at the C(3)?C(4) bond. At room temperature, the N2 loss from the initially formed 4,5‐dihydro‐3H‐pyrazoles 2 is faster than the cycloaddition and furnishes the vinylcyclopropane derivatives 7 and 9 with structural retention at the C(1)?C(2) bond. 2‐Substituted butadienes react with DDM at the C(3)?C(4) bond to give 12 ; isoprene, however, affords 3,4/1,2 products in the ratio of 86 : 14. DDM is a nucleophilic 1,3‐dipole: 1‐Cyanobutadiene reacts 400 times faster than 1‐methoxybuta‐1,3‐diene (DMF, 40°). The log k2 for the additions to six 1‐substituted butadienes show a linear correlation with σp (Hammett) and ?=+2.9; the log k2 of five 2‐substituted butadienes are linearly related to Taft's σI (?=+1.7). The structures of the vinylcyclopropanes 7, 9 , and 12 are established by NMR spectra and oxidation. A cyclopropyl carbinyl cation is made responsible for the isomerization of 12 , R=Ph, Me, by acetic acid to 4‐substituted 1,1‐diphenylpenta‐1,3‐dienes 25 and 29 ; TsOH at 200° converts 25 further to 9,10‐dihydro‐9‐methyl‐10‐phenyl‐9,10‐ethanoanthracene ( 27 ). Thermal rearrangement of 7, 9 , and 12 at 200–300° produces the 3‐ or 1‐substituted 4,4‐diphenylcyclopentenes 30 and 31 . These give the same mass spectra as the vinylcyclopropanes, and an open‐chain distonic radical cation is suggested as common intermediate. Besides spectroscopic evidence for the cyclopentene structures, hydrogenation and epoxidation are described; NMR data support the trans‐attack by perbenzoic acid.  相似文献   

17.
The one-electron transfer to large π-delocalized hydrocarbons provides an interesting possibility to crystallize solvent-separated ion-pair salts containing optimally solvated cations. Accordingly, the reduction of 9,10-diphenylanthracene in aprotic THF solution at a sodium metal mirror allows to grow dark-blue prismatic crystals of its radical anion and sixfold THF-solvated sodium cation. The structure of the radical anion is very similar to that recently published for the neutral molecule. According to AM1 hypersurface calculations based on the structural data, the phenyl twist angles obviously must be determined by lattice packing, and the negative charge is delocalized predominantly within the anthracene π system. The counter cation [Na(THF)6], reported ordered for the first time, shows nearly octahedral coordination within a rather densily packed solvent shell. Due to the strong repulsions between the solvent molecules, its isodesmically calculated solvation enthalpy is smaller than that of the analogous dimethoxyethane complex [Na(DME)3].  相似文献   

18.
The electrochemical reduction of eight quinones, 9,10-anthraquinone (1), duroquinone (2), 2,6-di-tert-butyl-1,4-benzoquinone (3), 2,6-dimethoxy-1,4-benzoquinone (4), 9,10-phenanthrenequinone (5), tetrachloro-1,2-benzoquinone (6), tetrabromo-1,2-benzoquinone (7) and 3,5-di-tert-butyl-1,2-benzoquinone (8), have been studied in acetonitrile. In every case it was found that cyclic voltammograms differed in significant ways from those expected for simple stepwise reduction of the quinone to its radical anion and dianion. The various types of deviations for the eight quinones have been cataloged and some speculation is offered concerning their origins.  相似文献   

19.
The first examples of the slippage formation of rotaxane‐like structures in the presence of an anion template are reported between a macrocycle, synthesised by exploiting Eglinton coupling, and stoppered pyridinium axle components. The role of the anion template in the slippage process has been explored by kinetic studies. 1H NMR spectroscopic investigations reveal the slippage species formed are not rotaxanes but pseudorotaxanes with some rotaxane character. The anion template significantly influences the amount of rotaxane character and the rate of slippage. Importantly, the fastest slippage rates, kon, are achieved with the non‐coordinating hexafluorophosphate anion, whereas the slowest slippage off rates, koff, are observed in the presence of coordinating anions, such as chloride. Since the koff rates are significantly smaller than the kon rates in the presence of coordinating anions, these anions act as templates favouring formation of the slippage species thermodynamically. Consequently, the resulting pseudorotaxanes with coordinating anions have greater rotaxane character. Two strategies for converting the slippage pseudorotaxanes into rotaxanes using hydrogenation or complexation with cobalt carbonyl are investigated.  相似文献   

20.
Thermal reaction of 1,3-diphenylisobenzofuran and tetramethylcyclopentadienone with PdLO2 complex (L = PPh3) gives compounds identical to those produced by singlet molecular oxygen. Photochemical reaction of 1,9-diphenylanthracene with PdLO2 or PdL3 in the presence of oxygen gives the 9,10-endoperoxide adduct.  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