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.
We have extended our study of the magnetic field dependence of monomer (MM) and excimer (ID) delayed fluorescence of aromatics in fluid solution. IM and ID show a monotonic decrease with increasing field strength at room temperature in agreement with theoretical predictions. At lower temperatures both IM and ID show an initial increase followed by a monotonic decrease. This marked temperature dependence can be attributed to a “freezing-out” of the effect of molecular motion on the field effect. At even lower temperatures the field effect curves of IM and ID start to diverge. Two models have been proposed to explain the mechanism of excimer formation in triplet—triplet annihilation. Qualitatively at least the divergence can be rationalized in terms of both models.  相似文献   

2.
Intramolecular excimer formation in 2,4-diphenylpentanes has been examined in a homologous series of alkanes, in ethanol and in mixtures of ethanol, ethylene glycol and glycerol. The ratio of the emission intensities of dimer and monomer (ID/IM) is not affected in low viscosity solvents but, above 4 cP, viscosity effects are discernible and a relationship of the form ID/IM = Aη?2 is obeyed. In methylene chloride, only the dl molecule exhibits a decrease of the efficiency of excimer sampling. The temperature dependence of ID/IM in isooctane and methylene chloride has been interpreted in terms of the activation energy of the excimer sampling.  相似文献   

3.
Polyvinylcarbazole (PVCa) films and solutions emit normal and excimer fluorescence between 77 and 425 K. The absolute (IM and ID) and relative (IM/ID) intensities emitted strongly depend on temperature. The usual U-shaped curve is obtained for log IM/ID as a function of 1/T in the case of polymer films. In solution, two minima corresponding to two different excimers are observed. The formation and dissociation of PVCa excimers in films and solutions have been interpreted according to the usual kinetic scheme. The binding energies for the excimer in films and for the high temperature excimer in solutions are respectively 4 and 2·8 kcal mole−1.  相似文献   

4.
Steady-state fluorescence measurements and molecular dynamics simulations have been used to study the intramolecular formation of excimers in five model compounds for polyesters containing naphthalene groups separated by flexible spacers. The model compounds are derived from 2-hydroxynaphthalene and HOOC (CH2)n COOH, n = 2–6. The ratio of the intensity of excimer and monomer emissions, ID/IM, is nearly independent of the viscosity of the medium, η, over the range covered in dilute solution. Although ID/IM is always very small, it shows an odd–even effect for the first four members of the series, with maxima when n is odd. Molecular dynamics simulations provide an explanation for the small values of ID/IM, their weak dependence on η, and the trend of ID/IM with n. The results for the present series of model compounds are compared with previous work, which reported larger values of ID/IM, and a stronger dependence of ID/IM on η, for bichromophoric compounds derived from 2-naphthoic acid and aliphatic glycols, where the direction of the ester groups is reversed. The origin of the difference in the behavior of ID/IM in the two series is identified. © 1997 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 35: 1127–1133, 1997  相似文献   

5.
The fluorescence of a series of copolymers of 2-naphthyl methacrylate (2-NM) and methyl methacrylate (MMA) with various contents of 2-NM (obtained in chloroform, carbon tetrachloride and acetonitrile) was investigated. A linear dependence between the ratio of the excimer to monomer emission intensities (ID/IM) and the diad fraction (fnn) of 2-NM monomer units was established. The relationship between ID/IM and fnn · In (In = the mean sequence length of 2-NM units) fits a logarithmic curve. The results indicate that the excimer emission is determined mainly by the nearest neighbour naphthalene-containing monomer units in the copolymer chain. The copolymers obtained in acetonitrile have higher values of ID/IM than those obtained in chloroform and carbon tetrachloride. This difference is due to the higher content of mm-triads in copolymers from acetonitrile, confirmed by 1H-NMR analysis of the samples of poly(methyl methacrylate) formed from copolymers of 2-NM and MMA.  相似文献   

6.
A criterion was proposed to estimate the necessity of the derivatization of organic substances for their determination on conventional nonpolar phases, based on such characteristic of analytes as molecular weight (M r), normal boiling point (T bp), and molar refraction (MR D). All these constants can be presented as indices relative to nonpolar n-alkanes (similarly to chromatographic retention indices), I(M), I(T), and I(MR D), which can be compared to each other as differences ΔT − M = I(T) − I(M) and ΔT − M R D = I(T) − I(MR D). Substances do not require derivatization if ΔT − M < 400 and ΔT − M R D < 600, while at ΔT − M > 600 and ΔT − MRD > 800, derivatization is necessary.  相似文献   

7.
The ratio of the intensities of excimer and monomer emission, ID/IM, has been measured for Py? COO? (CH2)m? OOC? Py (m = 2–6, Py = 1-pyrenyl) in solvents of different viscosity, η. These compounds are dimeric models for the corresponding polymers, selected to eliminate the influence of energy migration on the population of an excimer. The solvents are H(CH2)nOH, n = 1–7, and ethylene glycol. The values of ID/IM decrease as η increases, and the rate of this decrease is a function of m. Therefore the qualitative dependence of ID/IM on m is a function of η. In the limit as η → ∞, ID/IM is an odd–even function of m for m = 2–5, and the values at m = 5 and 6 are nearly identical. A rotational-isomeric state analysis of the conformations of the diesters can rationalize the appearance of the odd–even effect. This result is insensitive to assumptions about the extent of overlap of the two pyrene ring systems that is required for the production of emission in the excimer band. © 1993 John Wiley & Sons, Inc.  相似文献   

8.
Emission spectra of pyrene in hexane have been obtained over a temperature range (from 130 to 260 K) that has not been explored before for concentrations ranging from 10?4 mole/? to 2 × 10?2 mole/?. A conventional and new approximation which does not depend on the experimental set-up response functions has been used for evaluating pyrene excimer association energy WDM and other pyrene parameters. In both methods WDM agrees quite satisfactorlly, at all concentrations used, with that reported in the literature and obtained by other techniques. However, both approximation fail to yield the right values of the rate parameters at concentrations ? 2 g/?. This set the upper limit of sample solubility to be at C = 2g/? for our range of temperature. Furthermore, the new approximation can probably be used at higher concentration (for higher range of temperature) and even may be used for other organic molecules. There was also no difference in the ratio of the excimer (ID) to monomer (IM) quantum yields when an intense laser beam was used as a source of excitation rather than a super-pressure Hg-lamp. This suggested that the concentrations we used might not be large enough. As a result, the laser beam would not create enough density of excited molecules to affect the ratio of ID/IM.  相似文献   

9.
The hydrodynamic characteristics and weight-average molecular weights of some samples of poly-(4,4′-oxydiphenylene) pyromellitamic acid (I) and poly (4,4′-phenylene) pyromellitamic acid (II) in DMF have been obtained by sedimentation, viscometry and light scattering. For (I) the dependence of S0 on M was obtained and it was found that the hydrodynamic parameters F13P?1 depend on M up to M = 105. The experimental results were treated by using two models, viz. a Yamakawa-Fujii persistent chain and a rotational isomeric chain with free rotation about valence bonds. It was shown that the hydrodynamic behaviour of macromolecules of (I) with a pin-joint oxygen atom in the main chain is satisfactorily described by the rotational isomeric chain model with free rotation about valence bonds. The dependence of [η] on M was obtained and the constants K and a in an equation of the Mark-Kuhn-Houwink type were determined for (II). The flexibility of the chain was estimated and the Kuhn segment was found to be 200 Å using the Yamakawa-Fujii theory. The flexibilities of the chains of (I) and (II) are compared; it is shown that the rigidity of the chain of (II) without a pin-joint oxygen atom in the monomer unit is much higher than that of the chain of (I). The dependence of S0 on M obtained for (II) suggests that its macromolecules show a more pronounced tendency for formation of aggregates than those of (I).  相似文献   

10.
Although molecular amorphous materials represent an important area of research in solid-state chemistry, studies pertaining to these systems have been restricted almost exclusively to amorphous solids based on a single molecule. In this study, we found that, while the 2,4,6-bis(4-tert-butylphenyl)phenoxyl radical (2M) and its dimer (2D) did not give single-component amorphous solids, they rapidly formed the corresponding binary amorphous solid IIa following their condensation from benzene, dichloromethane, chloroform, and ethyl acetate solutions. The formation of IIa could be attributed to the good solubilities of 2M and 2D in these solvents and the high packing efficiencies of these amorphous solids. IIa was also obtained when crystals of 2D (IIb) were ground together. The solid-state formation of IIa would not only involve the locational exchange of 2M and 2D, but would also involve chemical exchanges.  相似文献   

11.
The ratio of excimer to monomer emission intensities, denoted by ID/IM, was measured for Py–COO(CH2CH2O)mCO ? Py, where Py denotes the 1-pyrenyl group and m = 1–4, in solvents of different viscosity, η. Three different systems were used to change the viscosity of the medium: (a) Mixtures of methanol and ethylene glycol at 25°C, (b) linear aliphatic alcohols, H(CH2)nOH, where n =1–6, also at 25°C, and (c) ethylene glycol over the range 6.6–35°C. The ratio ID/IM decreases sharply as η increases, and the rate of the decrease in ID/IM is a function of m. Quantitatively, the dependence of ID/IM on η at high viscosity, i.e., the slope [d(ID/IM)/d(1/η)], is larger in the present work than in another series of 1-pyrenyl diesters in which the flexible spacer is an oligomer of polyethylene, instead of an oligomer of polyoxyethylene. In the limit where η → ∞, the ratio ID/IM assumes its largest value in the bichromophoric compound with m = 2. However, as η decreases the compound with m = 3 becomes the one with the largest ID/IM. A complete rotational isomeric state analysis (for the compounds with m = 1–3) and a Monte Carlo simulation (for the compound m = 4) of the conformations of the diesters can rationalize the behavior of ID/IM in the high viscosity limit. ©1995 John Wiley & Sons, Inc.  相似文献   

12.
Systems of the type MIMIIIS2 (chalcopyrite)-CdS (wurtzite) where MI = Ag, Cu and MIII = Al, Ga, In were investigated to determine the regions of mutual solid solubility. It was found that the chalcopyrite structure could not tolerate extensive CdS substitution. When MIII was Al or Ga the solubility of MIMIIIS2 in CdS was also very limited. However, when MIII = In (rIn3+ ? rGa3+ > rAl3+), the solubility of MIInS2 in CdS was quite extensive (~50%). These results are consistent with a prior study on systems of the type MIMIIIS2ZnS which indicated that in sulfides, larger cations tend to result in the formation of new quaternary, wurtzite phases.  相似文献   

13.
The electrochemical Peltier effect was studied at a gold electrode in solutions containing some Fe(II)/Fe(III) redox couples by measuring the local temperature change in the electrode/solution interphase under controlled-potential and controlled-current polarization. Relative values of the electrochemical Peltier coefficient for the cathodic process at equilibrium potential, which is denoted by (Πc)I=0, were determined by analyzing the observed temperature change as a function of current. The values of (Πc)I=0 were found to be positive for the Fe(H2O)62+/Fe(H2O)63+ systems in HClO4 (1 M), HNO3 (1 M), H2SO4 (0.5 M), and HCl (1 M), their magnitudes being very similar in the first three acid solutions, but smaller in the HCl solution. On the other hand, a negative value of (Πc)I=0 was obtained in the case of a Fe(CN)64?/Fe(CN)63? couple in a H2SO4 (0.5 M) solution. Such a difference in the Peltier coefficient is considered to be due to the difference in the ionic species of iron involved in the electrode reaction.  相似文献   

14.
The thermal decomposition of the complexes M 2 I Cu(SO4)2 · 6 H2O and M2Ni(SO4)2 · · 6 H2O (MI=NH4, K, Rb, Tl) containing the complex cation MII(H2O)6 2+ (MIl = =Cu, Ni) was studied. The values of the experimental activation energyE obtained for the dehydration reactions of both complex cations were found to be influenced in different ways by the outer-sphere cations present. It was therefore concluded that the activation energy of the decomposition of Cu(H2O)6 2+ depends on the degree of tetragonal distortion of this cation, which increases with the ionic radius of cation MI. TheΔH values of the studied reactions depend less on the structures of the coordination polyhedra.  相似文献   

15.
LCAC-SW method has been extended to study the reaction dynamics for ion-pair formation processes.M+X2M++X-2 reaction system involves two potential energy surfaces,i.e.,the covalence state(M+X2) and the ionic state(M++X-2) and their crossing effect.The working equations for calculating state-to-state probability have been derived based on the above two-state model.The selected-state reaction probabilities of collinear ion-pair formation process M+I2M++I-2(M=Na,K,Cs) on Aten-Lanting-Los two-state potential energy surface have been calculated.The results show that the reaction probabilities are of resonance effect.  相似文献   

16.
Steady‐state fluorescence was used to measure the ratio of emission intensities, denoted ID/IM, for excited state complexes and excited monomers of five trichromophoric compounds, 2‐naphthyl‐COO‐(CH2)m‐OOC‐2,6‐dinaphthyl‐COO‐(CH2)m‐OOC‐2‐naphthyl, m = 2–6. The linear aliphatic alcohols H(CH2)nOH, n = 1–7, as well as mixtures of ethylene glycol and methanol, were used to change the viscosity of the medium, η. The values of ID/IM depend on η and m. A Rotational Isomeric State model and Molecular Dynamics simulations were used for interpretation of the experimental results. © 1999 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 37: 253–266, 1999  相似文献   

17.
A sample of Na2UO2(C2O4)2 · 5D2O (I) was studied by powder neutron diffraction. The compound crystallizes in the triclinic system, space group P1, unit cell parameters: a = 6.934(1) ?, b = 7.566(1) ?, c = 15.409(2) ?, ?? = 94.720(6)°, ?? = 96.281(6)°, ?? = 111.765(5)°, Z = 2, R F = 5.35, R I = 6.73 and ??2 = 2.89. The hydrogen bonds and non-valence contacts involved in the formation and binding of the {Na2[UO2(C2O4)2(D2O)](D2O)4} layers in I were analyzed using the Voronoi-Dirichlet polyhedra.  相似文献   

18.
The dual effect of the microabsorption heterogeneity (MAH) of emitters on the fluorescence intensity (I fl) is shown on the basis of the literature data. The duality is due to the covering of large fluorescent grains by small filler particles (II type effect) and its absence (I type effect). For some materials, the effect of MAH can change not only in magnitude but also in sign. It is shown experimentally that the prediction of the covering effect is difficult. It was found that, upon crushing multicomponent materials, different granulometric fractions differ in their chemical composition. This makes the theoretical account of the effect of the MAH of the emitter on the results x-ray fluorescence analysis hardly probable. It is shown that the experimental account by means of radiation scattered (I σ) or absorbed (I ab) by the sample possible only in a particular case, because the dependences I fl,I σ, and I ab on the particle size differ in nature.  相似文献   

19.
The equilibrium quotients for the formation of Co(NH3)5Cl2+ from Co(NH3)5OH23+ and Cl? were 3.74±0.25 M?1 and 6.07±0.54 M?1 at 45.0°C in 10:1 mole ratio water: dimethyl sulfoxide and in 25 w/w % aqueous ethanol, respectively, and those forthe formation of the ion pair Co(NH3)5OH23+ . Cl? were 1.21±0.20 M?1 and 1.58±0.17 M?1, respectively, in the same solvents. The aquation and anation rateconstants were determined at 45.0°C for these two solvents over the range of chloride-ion concentrations 0.0 ≤ [Cl?] ≤ 0.9 M. The aquation rate constant was essentially independent of chloride-ion concentration in each solvent over this range. The inverse of the pseudo-first-order anation rate constant was linearly dependent on the inverse of the chloride-ion concentration in each solvent. The least squares relationships between (1/kan) and (1/[Cl?]) gave intercepts and ratios of intercept to slope which were analyzed interms of Id and D mechanisms. It was concluded that the data were not satisfied by a D mechanism, but that they were consistent with an Id mechanism.  相似文献   

20.
It is established that krypton difluoride (KrF2) interacts with metallic gold in anhydrous hydrogen fluoride (HF) in the presence of alkaline and alkaline-earth metal fluorides with the formation of complex hexafluoroaurates with the general formula MIAuVF6 (MI = Li, Na, K, Rb, Cs) and bis-hexafluoroaurates MII(AuVF6)2 (MII = Mg, Ca, Sr, Ba). As a result of the subsequent interaction of bis-hexafluoroaurates with krypton difluoride, new coordination compounds with the composition MII(AuVF6)2 · nKrF2 (MII = Ca, Sr, Ba; n = 0–4) were synthesized. The Raman spectra were studied and a comparative analysis of the Raman spectra of the obtained compounds was performed. The possible variants of the structural transformation of the AuF 6 ? anions in compounds and the position and character of the chemical bond of the “guest” (KrF2) in these compounds are discussed.  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