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.
Fluorescence quenching of 1, 1, 4, 4-tetraphenyl-1, 3-butadiene (TPB) by aniline has been carried out at room temperature (298 K) to understand the role of quenching mechanisms. The study has been carried out by both steady state (in different solvents) and by transient method (in cyclohexane). The Stern-Volmer plot has been found to be linear for all the solvents studied. The probability of quenching per encounter ‘p’ is determined in all the solvents and is found to be less than unity. It is found that, the activation energy Ea (Ea) is greater than the activation energy of diffusion, Ed. The results obtained by the transient method infer that the thermally assisted intersystem crossing, a non-radiative deactivation process from S1 to T2 is responsible for observed decrease in quantum yield and lifetime. Hence, from both the methods it can be concluded that quenching mechanism is not solely due to the material diffusion, but there is also contribution from the activation energy.  相似文献   

2.
The electrical resistivity of dilute Zn-Mn alloys (c=0–0.6 at% Mn) has been measured in a temperature range from room temperature (RT) down to 0.4 °K. Three different series of samples are investigated: 1. annealed 72 h, 400 °C, 2. annealed and aged 1 year atRT, 3. coldworked atRT. All samples show a minimum in the residual resistance atT minc 1/5 followed by an increase of the resistivityρk s · lnT (Kondo-effect). For some alloys the resistivity minimum is followed by a resistivity maximum atT max.T max mainly depends on the distribution of the Mnatoms in the Zn-matrix. The coefficientk s is also strongly influenced by changes of the Mn-distribution. For alloys withc>0.02 at% (max. solubility of Mn in Zn atRT) the slopek s =?ρ/? lnT decreases with increasing degree of precipitation of the manganese. Samples withc<0.02 at% however, show the opposite effect, increasingk s with increasing Mn-precipitation. Moreover, during the aging process, we observe an increase in the “residual resistance” (resistance ratioR T/R 293 atT min) for these alloys. The effect can be explained by assuming that the precipitation leads to the formation of superparamagnetic clusters.  相似文献   

3.
The fluorescence quenching of 2,2″-dimethyl-p-terphenyl (DMT) by carbon tetrachloride by steady state in different solvents, and by transient method in benzene has been carried out at room temperature. The Stern-Volmer (SV) plot has been found to be non-linear with a positive deviation for all the solvents studied. In order to interpret these results we have invoked the ground state complex and sphere of action static quenching models. Using these models various rate parameters have been determined. The magnitudes of these parameters imply that sphere of action static quenching model agrees well with the experimental results. Hence the positive deviation in the SV plots is attributed to the static and dynamic quenching. Further, from the studies of temperature dependence of rate parameters and lifetime measurements, it could be explained that the positive deviation is due to the presence of a small static quenching component in the overall dynamic quenching. With the use of finite sink approximation model, it was possible to check whether these bimolecular reactions as diffusion limited and to estimate independently distance parameter R′ and mutual diffusion coefficient D. Finally an effort has been made to correlate the values of R′ and D with the values of the encounter distance R and the mutual diffusion coefficient D determined using the Edward's empirical relation and Stokes-Einstein relation.  相似文献   

4.
Thermoelectric power was studied in semimagnetic (diluted magnetic) Pb1−x−ySnyMnxTe (x=0.12,y=0.72) semiconductor in the temperature range T=10–100 K covering both ferromagnetic and paramagnetic range of temperatures of this material. In addition to standard diffusion contribution to the thermoelectric power (αDT), we found in PbSnMnTe an additional “magnetic” contribution (αFM). Employing the effect of carrier-concentration-controlled ferromagnetism in PbSnMnTe, we demonstrate that the temperature dependence of αFM(T) shows a clear maximum at the ferromagnetic Curie temperature. This experimental finding is discussed in terms of Kasuya model for the thermoelectric properties of diluted magnetic metallic systems with sd-exchange interaction.  相似文献   

5.
Fluorescence quenching of 1,3‐diphenyl benzene (m‐terphenyl) by carbon tetrachloride (CCl4) at steady state in different solvents, namely n‐hexane, n‐heptane, cyclohexane, toluene, benzene acetonitrile, 1,4‐dioxane, and with a transient method in benzene has been done at room temperature to understand the role of quenching mechanisms. The Stern–Volmer plot was found to be linear for all the solvents studied. The probability of quenching per encounter p was determined in all the solvents and was found to be less than unity. Further, from the studies of rate parameters and lifetime measurements in benzene at different temperatures (30–60°C), it was shown that the phenomenon of quenching is generally governed by the well‐known Stern–Volmer (S‐V) plot. The activation energy E a (Ea) of quenching was determined using literature values of activation energy of diffusion E d, and it was found to be greater than E d, which confirms the fact that the quenching mechanism is not solely due to material diffusion but there is also contribution from activation energy.  相似文献   

6.
The intermolecular photoinduced reactions between triplet ketone molecules and aliphatic amines and pyridine are studied by the quenching of delayed fluorescence of anthraquinone and benzophenone vapors by diethylamine, dibutylamine, cyclohexylamine, triethylamine, and pyridine. In the temperature range 423–573 K, the delayed fluorescence quenching rate constants k q are estimated from changes in the decay rate constant and the intensity of delayed fluorescence upon increasing pressure of bath gases. It is ascertained that, in the gas phase, the mixtures under study exhibit both a negative and a positive dependence of k q on temperature, which indicates that some photoinduced reactions do not have activation barriers. The rate constant k q is shown to increase with decreasing ionization potential of the electron donors. This points to the importance of interactions with charge transfer in the photoreaction of triplet ketone molecules with aliphatic amines and pyridine in the gas phase. The relationship between k q and the change in the free energy ΔG upon the photoinduced intermolecular electron transfer, which is the primary stage of the photochemical reaction, is studied. It is shown that the dependence k q G) for the donor-acceptor pairs under study is described well by the Marcus equation, in which the average vibrational energies of the donor and acceptor are taken into account for the estimate of ΔG.  相似文献   

7.
We present an analytic model of jet quenching, based on the (D)GLV energy loss formalism, to describe the system size dependence of QGP-induced parton absorption in relativistic heavy ion collisions. Numerical simulations of the transverse momentum dependence of jet quenching are given for central Au+Au and Cu+Cu reactions. Low p Tdijet correlations are shown to be sensitive to the reappearance of the lost energy as soft hadrons. At high p Twe find that the attenuation of dihadrons is similar to that of single inclusive particles. Comparison to recent data from PHENIX and STAR is given as a test of the jet quenching theory.  相似文献   

8.
We present preliminary results on Knight shift T1 and T2 measurements between 300 and 1700 K. We have observed the motional narrowing due to the self diffusion as well as a strong quadrupolar contribution to T1 due to the diffusion of interstitial gases. K varies from 0.58 to 0.63%. This temperature dependence is shown to be mainly due to an intrinsic temperature effect. The high temperature Korringa like behaviour of T1 with T1T = 1.05 sec. K is explained in terms of the temperature dependence of the d spin contribution.  相似文献   

9.
The existence of two temperatures: Tp-Cooper pairing and Tc-Bose–Einstein condensation in high temperature superconductors has been stipulated in a lightly potassium-doped C60 by Magnetically Modulated Microwave Absorption. This doping level corresponds to the carrier density greater than the critical one: x>x1. In case of rubidium lightly doped C60, where the carrier density x was smaller than the critical one: x<x1, anomalous EPR temperature dependence was observed. The characteristic temperature of bound electron pair formation Tp≈65 K and the energy gap 2Δ/k=30 K were estimated from the temperature dependence of the EPR signal intensity in non-superconducting state. These results suggest that the liquid fermions–liquid bosons transition can be observed as the opening of the spin gap at temperature Tp postulated in Micnas–Ranninger–Robaszkiewicz theory.  相似文献   

10.
The electron spin-lattice relaxation times (T 1) of a variety of semiquinone ions in hydrogen bonding solvents have been measured by the pulsed saturation recovery technique as a function of temperature (T) and viscosity (η) of the solvent. Also linewidths (ΔH) have been measured in suitable cases in such solvents at low radical concentrations (~10?4 M). It is observed that (i) the temperature and viscosity dependence ofT 1 can be fitted to an equation of the form 1/T 1=A(T/η)+Bexp(-ΔE/RT) whereA andB are constants and ΔE is an activation energy of the order of 1 kcal mole?1 for these systems; (ii)T 1 is essentially independent of the radical concentration within the range 10?3 to 5×10?2 M; (iii) the concentration independent part of the linewidth (ΔH) increases linearly with (η/T) at sufficiently low temperatures, and (iv) the (η/T) dependent part ofT 1 is sensitive to the size of the semiquinone as well as that of the solvent molecule, whereas the linewidth which is proportional to (η/T) at high viscosity, low temperature region is not sensitive to the size of the semiquinone and that of the solvent. Based on these observations, it is postulated that in hydrogen bonding solvents, three types of motion contribute significantly to electron spin relaxation:
  1. A restricted small step diffusional motion, not involving large changes in the orientation of the molecule, leading to the dominant viscosity dependent contributions toT 1 and ΔH, due to spin rotation interaction;
  2. a large amplitude reorientation of the semiquinone, coupled to translational diffusion, resulting in viscosity dependent contributions toT 1 and ΔH, throughg-modulation;
  3. a hindred rotation of the semiquinone within the solvent cage, contributing toT 1 due to spin rotation interaction.
The fact thatT 1 is not sensitive to the concentration of the radicals, is ascribed to the formation of the solvent cage that prevents the close approach of radicals, thereby rendering radical-radical interactions to be weak mechanisms for relaxation, even at relatively high radical concentrations.  相似文献   

11.
In this paper we present and discuss experimental methods to determine the positron diffusion coefficient from slow positron beam measurements. We also evaluate the use of the annihilation line Doppler-broadening technique in positron diffusion measurements, as compared on the more commonly used method of positronium fraction. The effects of incomplete positron thermalization and uncertainties of the positron implantation profile at low-positron incident energies to the measured data are discussed. We apply the presented methods to the model case of A1(110) system in the temperature range from 20 to 500 K. This data shows that Doppler-broadening and positronium fraction measurements give consistent results for the positron diffusion coefficient in A1(110), where D+(300 K) = 1.7(2) cm2/s with the temperature dependence D+ T –0.62(3).  相似文献   

12.
Temperature effects on negative-ion formation in positive-ion-surface scattering are studied within the framework of the time-dependent Anderson-Newns model. It is shown that the negative-ion formation is significantly enhanced at finite temperature T, provided kBT is not less than the Anderson correlation energy U, where kB is the Boltzmann constant. In the transient region (femtosecond timescale), temperature effects are, however, masked by large energy fluctuations.  相似文献   

13.
Positronium quenching through its collisions with gaseous atoms and molecules is discussed. It is found that the values of the normalized pick-off quenching parameter, 1Zeff, at room temperature are roughly proportional to the geometric collisional cross-sections estimated from the radius of the positronium and those of the atoms and molecules derived from the viscosity. This suggests that the probabilities of the pick-off quenching of the thermalized positronium per collision with various atoms and molecules are almost constant.  相似文献   

14.
非晶态离子导体Li2B2O4晶化前期的离子导电性   总被引:2,自引:0,他引:2       下载免费PDF全文
陈立泉  王连忠  车广灿  王刚 《物理学报》1983,32(9):1177-1182
本文研究了非晶态离子导体Li2B2O4的离子电导率与温度的关系,特别着重于晶化前期的离子迁移特性。当温度低于TK(≈310℃)时,离子电导率遵从Arrhenius关系。当高于晶化温度(≈411℃)时,以晶态中的离子迁移为主。在Tkc时,电导率偏离热激活机制呈反常增高。我们把这一过程称为晶化前期过程。可以用自由体积模型进行描述。晶化前期又可分为两部分:当温度低于、Tp(≈380℃)时,由于自由体积的重新分布,导致了电导率的增高;当T>Tp时,出现了少量微晶,但晶化量小于5%,由于非晶母体与微晶之间的界面效应使得离子导电性显著增强。可以通过室温淬火,把晶化前期非晶态的状态保持到室温,从而有可能制备出离子电导率高于纯非晶态的材料。 关键词:  相似文献   

15.
Kinetics and mechanisms for reactions of OH with methanol and ethanol have been investigated at the CCSD(T)/6-311 + G(3df2p)//MP2/6-311 + G(3df2p) level of theory. The total and individual rate constants, and product branching ratios for the reactions have been computed in the temperature range 200-3000 K with variational transition state theory by including the effects of multiple reflections above the wells of their pre-reaction complexes, quantum-mechanical tunneling and hindered internal rotations. The predicted results can be represented by the expressions k1 = 4.65 × 10−20 × T2.68 exp(414/T) and k2 = 9.11 × 10−20 × T2.58 exp(748/T) cm3 molecule−1 s−1 for the CH3OH and C2H5OH reactions, respectively. These results are in reasonable agreements with available experimental data except that of OH + C2H5OH in the high temperature range. The former reaction produces 96-89% of the H2O + CH2OH products, whereas the latter process produces 98-70% of H2O + CH3CHOH and 2-21% of the H2O + CH2CH2OH products in the temperature range computed (200-3000 K).  相似文献   

16.
A recently reported anomalous behaviour of the positronium annihilation rate with temperature in certain super-cooled organic liquids is explained in terms of a simple free-volume model modified in the presence of molecular cluster formation. This model apart from showing the entropic origin of the phenomenon accounts for the existence of the transition temperature “Tr” much above the glass-transition temperature Tg. It also predicts for the threshold temperature “Tr” fo clusters to commence formation a value of 310-7.0+8.0 K as against the experimental value of 304 K for ortho-terphenyl.  相似文献   

17.
Results of step fluctuation experiments for Mo(0 1 1), using low-energy electron microscopy, are re-examined using recently developed procedures that offer accurate coefficients of surface mass diffusion. By these means, surface diffusion Ds is documented at T/Tm ∼ 0.5, while the crossover to relaxation driven by bulk vacancy diffusion is inferred for T/Tm ∼ 0.6. Here, Tm is the melting temperature Tm = 2896 K. We obtain Ds = 4 × 10−4 exp(−1.13 eV/kBT) cm2/s for the temperature interval 1080-1680 K. Possible indications of diffusion along step edges appear for T/Tm ∼ 0.4. The same measurements of step fluctuation amplitudes determine also the step stiffness, which by symmetry is anisotropic on Mo(0 1 1). It is shown that three independent procedures yield mutually consistent step stiffness anisotropies. These are (1) step fluctuation amplitudes; (2) step relaxation rate anisotropies; and (3) the observed anisotropies of islands in equilibrium on the Mo(0 1 1) surface. The magnitude of the step stiffness obtained from step edge relaxation is consistent with earlier measurements that determine diffusion from grain boundary grooving.  相似文献   

18.
The oxygen quenching rate constants k T O2 of the triplet state T 1 of vapors of polycyclic aromatic hydrocarbons (PAHs) with strongly different oxidation potentials 0.44 eV < E OX < 1.61 eV and energies of the triplet levels 14800 cm?1 < E T < 24500 cm?1 (anthracene, 2-aminoanthracene, 9-nitroanthracene, chrysene, phenanthrene, fluoranthene, and carbazole) are estimated from the measured dependences of the decay rates and intensities of delayed fluorescence on the oxygen pressure P O2. It is found that the rate constants k T O2 vary from 4 × 103 (9-nitroanthracene) to 4 × 105 s?1 Torr?1 (2-aminoanthracene) and increase with decreasing oxidation potentials E OX of PAHs. The rate constants k T O2 for vapors and solutions are compared. The dependences of k T O2 on the free energy of two intermolecular processes, namely, triplet energy transfer to oxygen and electron transfer, are analyzed. It is shown that the rate constants k T O2 increase with decreasing electron transfer free energy, which proves that, along with energy transfer, charge-transfer interactions contribute to the quenching of the triplet states of PAH vapors.  相似文献   

19.
Rotational and translational movements of 1-oxyl-2,2,6,6-tetramethyl-4-oxypiperidine (TEMPOL) spin probe in the room temperature ionic liquid (RTIL) 1-octyl-3-methylimidazolium tetrafluoroborate (omimBF4) and in two molecular solvents, 1-propanol and isopropyl benzene (cumene), have been studied by X-band electron paramagnetic resonance (EPR) spectroscopy. Rotational correlation times τ c of spin probes and the intermolecular spin exchange rate constants k e were measured from EPR spectra at different temperatures and TEMPOL concentrations, and compared with the published data. The τ c values were calculated both by known equations and from the EPR spectra simulation. Rotation movements of TEMPOL in omimBF4 cannot be described by the model of the isotropic Brownian diffusion, which is valid for conventional solvents. The correct modeling of EPR spectra in RTIL can be achieved with the assumption of different rotational mobility of the spin probe around different molecular axes. The rotational, D rot, and translational, D tr, diffusion coefficients were calculated from τ c and k e values. The Debye–Stokes–Einstein law is valid in all three solvents while the dependence of D tr on T/η is not linear in Stokes–Einstein coordinates. The effective activation energy E rota of the rotational movements in omimBF4 is noticeably higher than the corresponding values for conventional solvents, while the effective activation energies E tra of the translational movements are comparable in all solvents studied.  相似文献   

20.
The temperature evolution of the magnetic structure of multiferroics Tb1 ? x Ce x Mn2O5 (x = 0, 0.20, 0.25) has been investigated using the neutron scattering methods. It has been found that, despite the qualitative similarity of the magnetic states and the series of phase transitions for pure TbMn2O5 (TMO) and doped crystals Tb1 ? x Ce x Mn2O5 (TCMO, x = 0.20 and 0.25), there are significant differences in their properties. In contrast to TMO, where there are three magnetic phases, TCMO can include two magnetic phases that coexist in a wide temperature range and exhibit a rather wide temperature hysteresis. One of these phases with wave vector k 1 = (0.5, 0, k z1), k z1 = 0.25, is commensurate and arises at temperatures below T N ~ 39 K (for x = 0.2) and T N ~ 38 K (x = 0.25). The second phase is incommensurate with wave vector k 2 = (1/2, 0, k z2), k z2 = 0.256(2), and appears upon cooling at T = 21 K (x = 0.2) and T = 19 K (x = 0.25). Upon further cooling to 16 K, the component k z2 increases to 0.292(2) and then remains constant. The component k z1 increases to the value of 0.280(2) upon cooling in the range from 15 to 10 K and then remains constant down to 1.5 K. With an increase in the temperature, the components k z1 and k z2 undergo reverse changes to their initial values, but these changes occur at temperatures 7 K higher than those observed with a decrease in the temperature. For TMO, two phases also coexist, but the temperature hysteresis in this case is considerably smaller than for TCMO. This is explained by different densities of domain walls and different sizes of domains in pure and doped crystals.  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