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 328 毫秒
1.
介质阻挡放电等离子体中·OH和HO2·自由基的数值模拟计算   总被引:4,自引:0,他引:4  
在介质阻挡放电等离子体N2/O2/H2O/HCHO体系中通过解Boitzmann方程,得到电子能量分布函数,利用得到的电子能量分布函数计算电子-分子碰撞反应速率常数.然后把有关的反应速率常数带入速率方程,计算得到该体系在介质阻挡放电时,·OH、HO2·和电子的浓度随时间的演变以及·OH、HO2·浓度随H2O、O2摩尔分数的变化,并将模拟结果与实验值进行了对比,两者符合得较好.  相似文献   

2.
介质阻挡放电等离子体中·OH和HO2·自由基的数值模拟计算   总被引:1,自引:0,他引:1  
在介质阻挡放电等离子体N2/O2/H2O/HCHO体系中通过解Boltzmann方程, 得到电子能量分布函数, 利用得到的电子能量分布函数计算电子-分子碰撞反应速率常数. 然后把有关的反应速率常数带入速率方程, 计算得到该体系在介质阻挡放电时,·OH、HO2·和电子的浓度随时间的演变以及·OH、HO2·浓度随H2O、O2摩尔分数的变化, 并将模拟结果与实验值进行了对比, 两者符合得较好.  相似文献   

3.
在化学反应动力学的实验数据处理中常常遇到在已知反应速率的微分方程组(简称速率方程,REQ),或其解所示各组元浓度对时间依赖的函数关系(简称动力学方程,KEQ)的形式时,如何由实验数据确定这些方程中的化学反应动力学参数(主要是一些反应速率常数)。关于这一问题的数学原理,过去文献上已有报  相似文献   

4.
利用电子自旋共振波谱(ESR)研究了在N2气中γ射线辐射诱导聚碳硅烷(PCS)自由基的产生和演变行为.ESR谱图分析结果表明,γ射线辐射诱导PCS产生的自由基为硅自由基(≡Si·).低剂量辐照时硅自由基的浓度随吸收剂量的增加而线性增加,硅自由基的辐射化学产额G值约为9,吸收剂量达到200 k Gy后,硅自由基的浓度趋于饱和.室温下硅自由基的浓度随存储时间的延长而逐渐降低,在N2气中存储时硅自由基的半衰期约23 d,在空气中存储时硅自由基的氧化反应导致衰减速率加快,半衰期仅为8 h.温度升高硅自由基衰减速率加快,在N2气中250℃加热处理可以完全清除硅自由基.  相似文献   

5.
本文以氮分子激光器为激发光源,研究了激发铀酰离子在高氯酸溶液中的发光衰减和时间分辨发光光谱.铀酰离子的发光衰减与铀浓度以及溶液的pH有关,当铀浓度Cu<10^-^3mol.dn^-^3时,溶液pH为1.5-4.0范围内,发光呈单一指数衰减;当Cu≥10^3^-mol.dm^-^3时,发现发光呈双指数衰减,即除了上述发光组份外,还观察到另一发光寿命较长的组份.用时间分辨方法测得的铀酰溶液的发光光谱表明,上述现象与激发态水合铀酰离子及其水解产物形成的发光体有关.  相似文献   

6.
采用部分偕胺肟化的聚丙烯腈纤维与铁离子反应,形成偕胺肟合铁(Ⅲ)纤维,以此为吸附材料,吸附水溶液中的还原棕染料. 研究了其吸附反应条件、吸附规律及吸附反应动力学. 结果表明,pH值11.5~12.5、温度60 ℃和吸附时间60 min为最佳吸附反应条件. 偕胺肟合铁(Ⅲ)纤维对还原棕的吸附反应符合Langmuir方程和Freundlich等温吸附经验式. 采用不同初始浓度研究吸附时间与溶液浓度的关系,用微分法确定了反应级数和反应速率常数. 由不同温度下的速率常数,并结合Arrhenius方程求出了反应的活化能. 结果表明,偕胺肟合铁(Ⅲ)纤维对还原棕的吸附符合一级反应动力学特征,速率方程c=c_0e~(-kt),速率常数k=32.01e~(-E_a/RT),活化能E_a=11.55 kJ/mol.  相似文献   

7.
周斌  白晨曦  何仁 《分子催化》2002,16(5):387-389
研究了用钴配合物催化的乙烯与三异丁基铝的置换反应动力学. 置换反应速率相对于烷基铝浓度、配合物浓度及乙烯压力呈一级关系. 置换反应速率的动力学方程为r=1.4 ×106[R3Al][Cat.][pC2H4]. 反应的活化能为71.9 kJ/mol.  相似文献   

8.
用光子相关光谱测定了苯乙烯-二乙烯基苯(St-DVB)共聚超微粒在良溶剂(甲苯)与θ溶剂(环己烷)中不同温度下的分子扩散系数,与流体力学方程结合给出了该微粒在无限稀时的等效流体力学半径。微粒分子扩散与温度的关系符合Arrhenius方程。不同温度的测试结果表明,St-DVB共聚超微粒的微交联网络结构使其分子形态不易改变。在θ溶剂中,T<θ时,微粒发生聚集,表现出大分子的协同扩散;T>θ时,在较高浓度,扩散系数偏离与浓度的线性关系。流体力学作用和热运动决定了微粒分子的扩散行为。  相似文献   

9.
采用紫外光谱法研究了腈水合酶催化丙烯腈水合的过程,在不同丙烯腈初始浓度下,测定了催化过程中275nm紫外吸光度的变化,计算出丙烯酰胺的生成速率.用Michaelis-Menten方程对不同丙烯腈浓度下的Nocardiasp.腈水合酶催化速率进行了拟合,得到该酶以丙烯腈为底物的米氏常数(Km)为8.46mmol/L,单位质量腈水合酶的催化速率常数(kcat)为2398μmol/(min·mg).  相似文献   

10.
本文利用皮秒时间分辨条纹相机技术检测了3种染料在立方颗粒溴化银上吸附后形成聚集体的荧光光谱,分析了3种染料在不同染料浓度下对染料聚集体到溴化银导带的超快电子转移过程的影响,进而分析其对增感效率的影响关系,并探讨了增感过程的微观机理.实验结果表明,荧光衰减的动力学曲线与一个双指数函数拟合得相当好,存在一快一慢两个衰减成分,快衰减成分占拟合较大比例,表明其源于与荧光衰减相竞争的从激发态染料聚集体到AgBr导带的电子转移.光致电子转移的速率及增感效率随着染料相对浓度的增加表现出一定的变化趋势,染料浓度增加,增感效率减小.  相似文献   

11.
用动态光散射方法研究了酚酞型聚醚砜(PES-C)在良溶剂DMF中的稀溶液性质。在稀溶液中,PES-C分子由于内旋转发生链折叠,整体上表现出柔性链的性质,较好的符合球形模型;而分子链的局部刚性结构又使分子尺寸稳定;它的扩散行为随温度的变化符合Arrehnius方程。表征了PES-C分子在稀溶液中的形态结构,且给出了PES-C分子在DMF中的扩散系数、扩散活化能、无限稀时的扩散系数和流体力学半径等重要特征参数。  相似文献   

12.
The friction and diffusion coefficients of a massive Brownian particle in a mesoscopic solvent are computed from the force and the velocity autocorrelation functions. The mesoscopic solvent is described in terms of free streaming of the solvent molecules, interrupted at discrete time intervals by multiparticle collisions that conserve mass, momentum, and energy. The Brownian particle interacts with the solvent molecules through repulsive Lennard-Jones forces. The decays of the force and velocity autocorrelation functions are analyzed in the microcanonical ensemble as a function of the number N of solvent molecules and Brownian particle mass and diameter. The simulations are carried out for large system sizes and long times to assess the N dependence of the friction coefficient. The decay rates of these correlations are confirmed to vary as N(-1) in accord with earlier predictions. Hydrodynamic effects on the velocity autocorrelation function and diffusion coefficient are studied as a function of Brownian particle mass and diameter.  相似文献   

13.
The diffusion process of a single spherical nanoparticle immersed in a fluid solvent is studied by molecular dynamics simulations. When the nanoparticle mass stays constant, it is shown that, at short times, the decay of the nanoparticle velocity autocorrelation function is strongly modified when the particle diameter increases. It is also shown that, at large times, the characteristic algebraic decay induced by the hydrodynamic correlations between the solvated particle and the solvent presents a scaling behavior depending on the particle diameter.  相似文献   

14.
The effects of three different variables (initiator concentration, polarity of the solvent and reaction temperature) on the rate of dispersion polymerization of styrene in alcohols have been investigated. It was found that the rate of polymerization increases with the initiator (AIBN) concentration at the 0% conversion level and becomes independent of it at higher monomer conversions. More significant was the result that the rate was also found to increase with solvent polarity. This is consistent with thermodynamic equilibrium calculations which account for the partitioning behavior of monomer and solvent in both the solution and the particle phases. The results further suggest the existence of two different kinetic regions: one at low conversions, where the reaction takes place primarily in the solution phase, and one at high conversions, where the reaction takes place primarily in the particle phase. © 1997 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 35 : 2907–2915, 1997  相似文献   

15.
We report a molecular dynamics study of a binary mixture consisting of a large (host) particle and a smaller (guest) particle whose radius is varied over a range. These simulations investigate the possible existence of a diffusion anomaly or levitation effect in dense fluids, previously seen for guest molecules diffusing within porous solids. The voids in the larger component have been characterized in terms of void and neck distributions by means of Voronoi polyhedral analysis. Four different mixtures with differing ratios of guest to host diffusivities (D) have been studied. The results suggest that the diffusion anomaly is seen in both close-packed solids with disorder and dense fluids. In the latter, the void network is constantly and dynamically changing and possesses a considerable degree of disorder. The two regimes, viz., the linear regime (LR) and the anomalous regime (AR), found for porous solids are shown to exist for a dense medium as well. The linear regime is characterized by D(g) proportional to 1/sigma(gg)(2), where sigma(gg) is the diameter of the guest. The anomalous regime exhibits a maximum in D up to rather high temperatures (T = 1.663), even though in porous solids the maximum disappears at higher temperatures. In agreement with previous studies on porous solids, a particle in the AR is associated with lower activation energy, lower friction, and less backscattering in the velocity autocorrelation function when compared to a particle in the LR. Wavevector dependent self-diffusivity, Delta, and decay of the intermediate scattering function, F(s)(k, t), exhibit contrasting behaviors for the LR and AR. For LR, Delta exhibits a minimum at values of k at which there are spatial correlations in S(k) while a smooth decrease with k is seen for AR. For LR, F(s)(k, t) shows a biexponential decay corresponding to two different time scales of motion. Probably, the fast decay is associated with motion within the first shell of solvent neighbors and the slow decay with motion past these shells. For AR, a single-exponential decay is seen. The results indicate a breakdown of the Stokes-Einstein (SE) relationship. The relevant quantity that determines the validity of the SE relationship is the levitation parameter which is indirectly related to the solute/solvent radius ratio and not either the size of the solute or the solvent alone.  相似文献   

16.
Biodegradable polyesteramide copolymer P(CL/AU) based on -caprolactone and 11-aminoundecanoic acid was synthesized by the melt polycondensation method. Polyesteramide (PEA) microspheres were prepared by a simple O/W emulsion solvent evaporation method. The effects of variations in preparation parameters (such as emulsifier concentration, polymer concentration, polymer solution adding rate, stirring rate, and whether vacuum was applied) were studied in detail. The obtained microsphere morphologies were observed using an optical microscope and via scanning electron microscopy (SEM). The particle size distribution was determined using a Malvern laser particle sizer. When the PEA microspheres were incubated in PBS saline, the particle size increased at first, and then decreased after a longer time period; the theory that this behavior was due to degradation of the microspheres was confirmed by SEM.  相似文献   

17.
The aims of this study were to identify how the solvent selection affects particle formation and to examine the effect of the initial drug solution concentration on mean particle size and particle size distribution in the supercritical antisolvent (SAS) process. Amorphous atorvastatin calcium was precipitated from seven different solvents using the SAS process. Particles with mean particle size ranging between 62.6 and 1493.7 nm were obtained by varying organic solvent type and solution concentration. By changing the solvent, we observed large variations in particle size and particle size distribution, accompanied by different particle morphologies. Particles obtained from acetone and tetrahydrofuran (THF) were compact and spherical fine particles, whereas those from N-methylpyrrolidone (NMP) and dimethylsulfoxide (DMSO) were agglomerated, with rough surfaces and relatively larger particle sizes. Interestingly, the mean particle size of atorvastatin calcium increased with an increase in the boiling point of the organic solvent used. Thus, for atorvastatin particle formation via the SAS process, particle size was determined mainly by evaporation of the organic solvent into the antisolvent phase. In addition, the mean particle size was increased with increasing drug solution concentration. In this study, from the aspects of particle size and solvent toxicity, acetone was the better organic solvent for controlling nanoparticle formation of atorvastatin calcium.  相似文献   

18.
Abstract We report direct femtosecond measurements of the excited state dynamics of hematoporphyrin derivative (HpD) in solution. The dynamics are found to be very sensitive to the solvent and pH of aqueous solutions. The decay of the excited singlet states is much faster in acidic and pH 7 buffer aqueous solutions (<230 ps) than in basic aqueous solutions or organic solvents (> 10 ns). The dynamical results show strong correlation with static fluorescence measurements: weaker fluorescence in acidic and pH 7 buffer solutions corresponding to shorter-lived excited states. A new fast decay component with a time constant around 5 ps is identified both in acidic aqueous solutions and in organic solvents such as acetone and attributed to internal conversion from the second to the first excited singlet state of aggregates or certain oligomers in HpD, in accord with the observation that the fast decay component is larger at a higher concentration. Oxygen is found to have no effect on the dynamics on the time scale investigated, 1 ns, indicating that oxygen quenching of the singlet excited states is insignificant on this time scale. The sensitive solvent and pH dependence of the excited state dynamics has important clinical implications in the use of HpD as a photosensitizing agent.  相似文献   

19.
 The kinetics of free radical decay in the polymerization of MMA initiated by AIBN was studied by means of ESR spectroscopy. It was found that the curves of radical decay are strongly associated with the reaction temperature, the initiator concentration and the solvent. In the case of the radical polymerization carried out at high temperature or in solution, the radical concentration first reached a maximum, then declined monotonously with reaction time. It was also found that the greater the amount of initiator or the higher the temperature, the more rapidly the radicals decay. When the bulk polymerization was implemented at a relatively low temperature, the curves of radical decay became more complicated, i.e.,the radical concentration rapidly rose to a maximum, then dropped to a minimum, finally increased again with reaction time.Taking into account the diffusion effect, a semi-empirical equation is suggested to describe the kinetics of propagating radical decay.  相似文献   

20.
Porous polylactide (PLA) microspheres were fabricated by an emulsion‐solvent evaporation method based on solution induced phase separation. Scanning electron microscopy (SEM) observations confirmed the porous structure of the microspheres with good connectivity. The pore size was in the range of decade micrometers. Besides large cavities as similarly existed on non‐porous microspheres, small pores were found on surfaces of the porous microspheres. The apparent density of the porous microspheres was much smaller than that of non‐porous microspheres. Fabrication conditions such as stirring rate, good solvent/non‐solvent ratio, PLA concentration and dispersant (polyvinyl alcohol, PVA) concentration had an important influence on both the particle size and size distribution and the pore size within the microspheres. A larger pore size was achieved at a slower stirring rate, lower good solvent/non‐solvent ratio or lower PLA concentration due to longer coalescence time. Copyright © 2005 John Wiley & Sons, Ltd.  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