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 265 毫秒
1.
Hauptmann  W.  Drochner  A.  Vogel  H.  Votsmeier  M.  Gieshoff  J. 《Topics in Catalysis》2007,42(1-4):157-160
Different global kinetics for the oxidation of NO on platinum are compared in this work. The general form of the equation for the reaction rate is with . Furthermore steady-state and temperature-programmed experiments were performed. The best results of simulation coupled with parameter estimation were obtained using α = 0.28, β = 0.49, γ = 0 and δ = 1, along with an activation energy of 47.5 kJ mol−1.  相似文献   

2.
Empirical correlations of flow properties of poly(vinyl chloride) were made using data reported by a number of investigators. Correlation was made by plotting the reduced variable viscosity η/η0 versus \documentclass{article}\pagestyle{empty}\begin{document}$ (\eta _0 \dot \gamma \bar M_w )/(_\rho RT) $\end{document} or \documentclass{article}\pagestyle{empty}\begin{document}$ (\eta _0 \dot \gamma \bar M_w ^{0.5} )/(_\rho RT) $\end{document} for unplasticized PVC and versus \documentclass{article}\pagestyle{empty}\begin{document}$ (\eta _0 \dot \gamma \bar M_w ^{0.5} )/(_\rho RTW_2 ^a ) $\end{document} with polymer concentration, W2, for PVC containing plasticizer.  相似文献   

3.
Summary The electrochemical reduction of NO 3 - in 0.1 M K2SO4 and 0.05 M KNO3 solution was studied on various electrodes in two different cell configurations, a divided and an undivided one. The products in all cases were NO 2 - , NH3, N2 and small amounts of NO2 and NO. The more efficient cathodes as regards the conversion of NO 3 - to N2 were Al and the alloy Sn85Cu15, where the selectivity for nitrogen formation was 43 and 35.3% at –1.8 and –2.0 V, respectively. The kinetic analysis of the experimental results was carried out by numerical solution of the resulted differential equations according to the scheme: The rate constants on Sn85Cu15 at –2.0 V for the above reactions were found to be k1=4.9 × 10–4 s–1, k2=1.76 × 10–5 s–1 and k3=7.66 × 10–3 l mol–1 s–1. At more negative potential more NO 2 - ions reduced and converted either to N2 or NH3. The rate constant of reduction of nitrate was almost the same in the region between –1.7 and –2.0 V, because the reaction is limited by the diffusion. In order to oxidize a part of the undesirable byproducts NO 2 - and NH3 at the anode of the cell to nitrate and nitrogen respectively, an undivided cell was used. Comparison between the two cell configurations indicated that, although in the undivided cell the % removal efficiency of nitrate was lower than that in the divided one, the selectivities of NO 2 - and NH3 were 4.8 and 2.2 times lower, respectively.  相似文献   

4.
p-methoxyphenoxy and p-chlorophenoxy group containing methacrylate based monomer 2-(p-methoxyphenoxy)-2-oxo-ethyl methacrylate (pMPOEMA) and 2-(p-chlorophenoxy)-2-oxo-ethyl methacrylate (p-ClPOEMA) were synthesized by reacting p-methoxyphenyl chloroacetate (pMPClAcO) and p-chlorophenyl chloroacetate (pClPClAcO) with sodium methacrylate in acetonitrile respectively. (pMPClAcO) and (pClPClAcO) were prepared by reacting p-methoxyphenol and p-chlorophenol dissolved in benzene with chloroacetylchloride. The free-radical-initiated copolymerization of (pMPOEMA) and (pClPOEMA) with acrylonitrile (AN) were carried out in 1,4-dioxane solution at 65 C using 2,2′-azobisisobutyronitrile (AIBN) as an initiator with different monomer-to-monomer ratios in the feed. The monomers and copolymers were characterized by FTIR, 1H- and 13C-NMR spectral studies. The copolymer compositions were evaluated by nitrogen content in polymers. The reactivity ratios of the monomers were determined by the application of Fineman–Ross and Kelen–Tüdös methods. The analysis of reactivity ratios revealed that pMPOEMA and pClPOEMA are more reactive than AN, and copolymers formed are statistically in nature. The molecular weights ( and ) and polydispersity index of the polymers were determined using gel permation chromagtography. Thermogravimetric analysis of the polymers reveal that the thermal stability of the copolymers increases with an increase in the mole fraction of AN in the copolymers. Glass transition temperatures of the copolymers were found to decrease with an increase in the mole fraction of AN in the copolymers.  相似文献   

5.
Summary The basic structural properties of xerogels of crosslinked poly(acrylic acid) were defined and determined: xerogel density (ρxg), xerogel volume fraction in the equilibrium-swollen state (v2), the number average molar mass between network crosslinks , the crosslink density (ρc) and the distance between macromolecular chains (d). A crosslinking ratio (X) increase leads to a linear increase in the values for ρxg and ρc, while the values and d decrease. The isothermal swelling kinetic curves of four samples of structurally different poly(acrylic acid) xerogels in bidistilled water at different temperatures ranging from 25 to 45 °C were determined. It is shown that isothermal kinetic swelling curves could not be described with the model of first-order reaction kinetics in entire. It was found that these curves could be described by the Johanson-Mampel-Avrami (JMA) equation. For all of the investigated xerogel samples, the initial swelling rate (vin), effective reaction rate constant (k) and equilibrium swelling degree increased with swelling temperature increase. Based on the determined values of the vin and k, the activation energy (Ea) and pre-exponential factor (lnA) were determined. It was concluded that the activation energy linearly increased with increasing distance between macromolecular chains (d) and molar mass between the network crosslinks . The relationship between the activation energy changes with pre-exponential factor (compensation effect) caused by xerogel structural properties was established. Isothermal swelling kinetics could be completely described by the kinetics of phase transition of the xerogel transformation from glassy to rubbery state, i.e. with the JMA kinetic equation.  相似文献   

6.
The influence of electrolytes, which are dissolved in the aqueous absorbent and do not react with nitrogen oxides, on the absorption kinetics of both these components was investigated experimentally. In addition to demineralized water, various salt solutions of different concentrations as well as sodium hydroxide solution were used as absorbents. The term H \documentclass{article}\pagestyle{empty}\begin{document}$ H\sqrt {k_1 D} $\end{document} for N2O4 and N2O3, which is important for the design of industrial absorbers, was determined as a function of composition and concentration of the absorbents. In the case of N2O4, the chosen measuring and evaluation methods permitted a separate determination of the rate constant k of the pseudo first order reaction and of the solubility H. The diffusion coefficient D of the gas in the absorbent can be obtained only by calculation. Experimental results showed that \documentclass{article}\pagestyle{empty}\begin{document}$(H\sqrt {k_1 D} )\,_{{\rm N}_{\rm 2} {\rm O}_{\rm 4} } $\end{document} decreases with increasing ionic strength I, however, without a clear indication of any ion-specific effects. This decrease does not appear to be caused simply by a reduction in solubility (salting out effect), or in diffusion coefficient, but at least, to the same extent, through a decrease of the rate constant k with increasing electrolyte content in the absorbent. The measurements permitted the determination of the gas-based salting out parameter for N2O4. The investigations on the absorption of N2O3 in water and in an Na2SO4 solution showed no experimentally detectable influence of dissolved salts on \documentclass{article}\pagestyle{empty}\begin{document}$(H\sqrt {k_1 D} )\,_{{\rm N}_{\rm 2} {\rm O}_{\rm 3} } $\end{document}. The numerical value of \documentclass{article}\pagestyle{empty}\begin{document}$(H\sqrt {k_1 D} )\,_{{\rm N}_{\rm 2} {\rm O}_{\rm 3} } $\end{document} is six times that of \documentclass{article}\pagestyle{empty}\begin{document}$(H\sqrt {k_1 D} )\,_{{\rm N}_{\rm 2} {\rm O}_{\rm 4} } $\end{document}.  相似文献   

7.
  The concept of absolute electrode potential in aqueous and solid electrochemistry is discussed in light of the first experimental investigation utilizing a two Kelvin probe system which allows for direct in situ measurement of the work functions of both the, emersed or spillover modified, working and reference electrodes. In both cases, i.e. emersed electrodes in aqueous electrochemistry and spillover-modified electrodes in solid electrochemistry, it is found that the following two equations relate the working–reference electrode potential difference, U WR, and the work functions, ΦW and ΦR of the emersed or spillover-modified working and reference electrodes:
where U WR is varied either by varying the gaseous composition or via a potentiostat. These equations show that the work function of emersed electrodes in aqueous electrochemistry or of spillover-modified electrodes in solid state electrochemistry is the natural choice of the absolute electrode potential:
The value was obtained as the absolute potential value of the H2/H+ electrode in aqueous solutions at , pH = 0 and T = 298 K, while the value of was measured as the absolute potential value of the O2/O2− electrode in YSZ (8 mol% Y2O3-stabilized-ZrO2) at and T = 673 K.  相似文献   

8.
Simulation and optimal design of the reactor for the seeded continous emulsion polymerization process have been done in this work. An internal mixer (Toray Hi-Mixer) as seeder connected with a stirred tank is designed to correlate conversion, molecular weight, and MWD with the model simulation proposed. An optimal mean residence time of seeder \documentclass{article}\pagestyle{empty}\begin{document}$(\bar \theta _s)_c$\end{document} is found to lie between \documentclass{article}\pagestyle{empty}\begin{document}$(\bar \theta _1)$\end{document} and tin, where \documentclass{article}\pagestyle{empty}\begin{document}$(\bar \theta _1)_{{\rm opt}} = (3aS_0 /2r\eta N_a \alpha)^{3/5}$\end{document} and tin = 1.57(aS0/riηNA)3.5. The optimal design of the process is performed according to the above relations under several polymerization conditions. In general, the increase in number of stages inside the seeder can reduced the volume of CSTR for a required production. Molecular weight of products is increased by increasing the number of stages inside the seeder, by decrasing the concentration of the initiator, and by increasing the concentration of the emulsifier under the optimal conditons.  相似文献   

9.
Many electrochemical processes suffer in varying degrees from mass transfer limitations. These limitations may require operation at considerably less than economic optimum current densities. Mass transfer to a surface may be considerably enhanced by insertion of turbulence promoters in the fluid flow path near the affected surface.An instrument was developed to measure local current densities in the hydrodynamically very difficult region near the turbulence promoter. A general method for the relative evaluation of hydrodynamic conditions has been developed. Generalization of the data permits optimization of hydrodynamic cell design using the promoter shapes investigated.

Notation

Symbols A Coefficient for cell power costs, $ m2 (As)–1 - A c Cell area, m2 - a Constant in Equation 4 - B Coefficient for area-proportional costs, $ A (m2 s)–1 - C Coefficient for pumping power costs, $ A (m2 s)–1 - C b Bulk concentration, kg mol m–3 - C bi Inlet bulk concentration, kg mol m–3 - C e Energy cost, $ (Ws)–1 - C i Interfacial concentration, kg mol m–3 - ¯C s Amortized area cost, $ (m2 s)–1 - D Current—density-insensitive costs, $ s–1 - D e Equivalent diameter, m - D Diffusion constant, m2 s–1 - e Current efficiency - F d Cell feed rate, m3 s–1 - F 96.5×106 A s kg eq–1 - g Channel width, m - h Channel height, m - i Current density, A m–2 - i opt Economic optimum current density, A m–2 - K Total costs of running cell, $ s–1 - (K–D)ideal Total sensitive costs under hydrodynamically ideal conditions, $ s–1 - k c Convective mass transfer coefficient, m s–1 - L Total length of flow path, m - l Promoter spacing, m - N Mass flow rate to surface due to convection, kg mol m2 s–1 - n e Number of electrons transferred in electrode reaction - P c Power required by cell, W - P/L Average pressure gradient in channel, N m–3 - R av Effective cell resistance, m2 - S Open channel cross-section, m2 - S 0 Minimum channel cross-section at promoter, m2 - s i Stoichiometric coefficient of species i - t i Transport number of species i in solution - ¯t i Effective tranport number of species at polarized surface - V Average fluid velocity, m s–1 - x Distance from inception of concentration disturbance, m - 1 Electrical power conversion efficiency - 2 Pumping power conversion efficiency - Solution viscosity, kg (m s)–1 - Solution density, kg m–3 Dimemionless groups Fanning friction factor - Reynolds number - R h/g Channel aspect ratio - D e/l Promoter frequency - S/S 0 Contraction coefficient - Sherwood number - Degree of reaction - Dimensionless total sensitive - Dimensionless current density - Energy cost ratio  相似文献   

10.
11.
The heterogeneous bulk polymerization of acrylonitrile initiated by AIBN has been studied by means of an improved dilatometric technique and a new method of analysis, where the initial reaction rate (vw)0 results from the intercept of a straight line in a \documentclass{article}\pagestyle{empty}\begin{document}$ \frac {\ln \left( 1 \hbox{---} {\rm U} \right)} {{\rm e}^{{- 0,5} {\rm k}_{\rm s}{\rm t} \hbox{---} 1}}$\end{document} versus t plot. It has been found that the initial reaction rate is proportional to the square root of the initial catalyst concentration S0. The ratio of the rate coefficients of propagation and termination\documentclass{article}\pagestyle{empty}\begin{document}$\frac { {\rm k}_{\rm a} } { {\rm k}_{ {\rm w}^{2} } } $\end{document} could be calculated from the slope of a straight line passing through the origin in a plot of (vw)0 versus \documentclass{article}\pagestyle{empty}\begin{document}$\sqrt { {\rm S}_{0} }$\end{document} and yielded a value of 280 mol 1?1.  相似文献   

12.
13.
Experimental results on the rate of lateral flame spread and time for piloted ignition under an externally imposed radiant flux were analyzed with a simple theroretical model. The data were developed from a radiant panel apparatus that considers a wall mounted sample with a flux distribution \documentclass{article}\pagestyle{empty}\begin{document}$ (\dot q_{\rm e} ^{\prime \prime } ) $\end{document} of 5 W cm?2 at the ignited end to 0.2 W cm?2 at the other end. It is shown that after an appropriate preheating time (flux exposure time before sample is ignited) the rate of flame spread (Vf) results can be correlated by \documentclass{article}\pagestyle{empty}\begin{document}$ V_{\rm f} - {\textstyle{1 \over 2}} = C\left( {\dot q''_{{\rm o,ig}} - \dot q_{\rm e} ^{\prime \prime } } \right) $\end{document} where C is a material ‘constant’ and \documentclass{article}\pagestyle{empty}\begin{document}$ \dot q''{\rm }_{{\rm o,ig}} $\end{document} is minimum flux for piloted ignition—also a material (and configuration) constant. An extension of this model demonstrates that Vf can also be expressed in terms of an ‘ignition temperature’ and the surface temperature of the material. Both correlations are derivable from a single flame spread experiment. Results are presented for a number of typical wood and plastic materials.  相似文献   

14.
$\begin{array}{l}{\hbox{R}^1\hbox{R}^2\hbox{CHOH}} \\ {\hbox{RCH}_2\hbox{OH} }\end{array} \dynrightarrow{Oxone}{\hbox{CH}_3\hbox{CN/H}_2\hbox{O}, 70^{\circ}\hbox{C}} \begin{array}{l}{\hbox{R}^1\hbox{R}^2\hbox{CO}} \\ {\hbox{RCOOH}} \end{array} A simple and environmentally friendly procedure for the oxidation of alcohols is presented utilizing Oxone? (2KHSO5 · KHSO4 · K2 SO4) as oxidant and polymer-supported 2-iodobenzamide as catalyst in CH3CN/H2O mixed solvents.  相似文献   

15.
A mathematical model is presented for the optimization of the hydrogen-chlorine energy storage system. Numerical calculations have been made for a 20 MW plant being operated with a cycle of 10 h charge and 10h discharge. Optimal operating parameters, such as electrolyte concentration, cell temperature and current densities, are determined to minimize the investment of capital equipment.Nomenclature A ex design heat transfer area of heat exchanger (m2) - a F electrode area (m2) - heat capacity of liquid chlorine (J kg–1K–1) - heat capacity of hydrogen gas at constant volume (J kg–1 K–1) - c p,hcl heat capacity of aqueous HCl (J kg–1 K–1) - C $acid cost coefficient of HCl/Cl2 storage ($ m–1.4) - C $ex cost coefficient of heat exchanger ($ m–1.9) - C $F cost coefficient of cell stack ($ m–2) - cost coefficient of H2 storage ($ m–1.6) - C $j cost coefficient of equipmentj ($/unit capacity) - C $pipe cost coefficient of pipe ($ m–1) - C $pump cost coefficient of pump ($ J–0.98 s–0.98) - E cell voltage (V) - F Faraday constant (9.65 × 107 C kg-equiv–1) - F j design capacity of equipmentj (unit capacity) - G D design electrolyte flow rate (m3 h–1) - heat of formation of liquid chlorine (J kg-mol–1 C12) - H f 0 ,HCl heat of formation of aqueous HCl (J kg-mol–1HCl) - H m total mechanical energy losses (J) - I total current flow through cell (A) - i operating current density of cell stack (A m–2) - L length of pipeline (m) - N number of parallel pipelines - nHCl change in the amount of HCl (kg-mole) - P pressure of HCl/Cl2 storage (kPa) - p 1 H2 storage pressure at the beginning of charge (kPa) - p 2 H2 storage pressure at the end of charge (kPa) - –Q ex heat removed through the heat exchanger (J) - R universal gas constant (8314 J kg-mol–1 K–1) - the solubility of chlorine in aqueous HCl (kg-mole Cl2 m–3 solution) - T electrolyte temperature (K) - T 2 electrolyte temperature at the end of charge (K) - T max maximum electrolyte temperature (K) - T min minimum electrolyte temperature (K) - t final time (h) - t ex the length of time for the heat exchanger operation (h) - Uit ex overall heat transfer coefficient (J h–1 m–2 K–1) - V acid volume of HCl/Cl2 storage (m3) - } volume of H2 storage (m3) - v design linear velocity of electrolyte (m s–1) - amount of liquid chloride at timet (kg) - amount of liquid chlorine at timet 0 (kg) - w hcl amount of aqueous HCl solution at timet (kg) - W p design brake power of pump (J s–1) - X electrolyte concentration of HCl at timet (wt fraction) - X f electrolyte concentration of HCl at the end of charge (wt fraction) - X i electrolyte concentration of HCl at the beginning of charge (wt fraction) - X 0 electrolyte concentration of HCl at timet 0 (wt fraction) - Y objective function to be minimized ($ kW–1 h–1) - j the scale-up exponent of equipmentj - overall electric-to-electric efficiency (%) - acid safety factor of HCl/Cl2 storage - fractional excess of liquid chlorine - p pump efficiency - average density of HCl solution over the discharge period (kg m–3)  相似文献   

16.
Xiong  Qingfeng  Ni  Peihong  Zhang  Feng  Yu  Zhangqing 《Polymer Bulletin》2004,53(1):1-8
Summary Homopolymers of 2-(dimethylamino)ethyl methacrylate (DMAEMA) have been synthesized directly in aqueous media by reversible addition-fragmentation chain transfer polymerization (RAFT) using 4,4-azobis(4-cyanopentanoic acid) (V501) as a water-soluble azo initiator and 4-cyanopentanoic acid dithiobenzoate (CPADB) as a chain transfer agent. The resulting polymers were controlled in the range of narrow molecular weight distributions, with lower than 1.3. Using the produced dithioester-capped DMAEMA homopolymer as a macro chain transfer agent, miniemulsion RAFT polymerization of methyl methacrylate and styrene were carried out, respectively. 1H NMR analysis showed that the diblock copolymer PDMAEMA-b-PMMA in the form of stable latices was obtained as expected. This revised version was published online at the end of November 2004. Unfortunately, the received date was incorrect due to a technical problem.  相似文献   

17.
Intrinsic viscosity-number average molecular weight relationships have been measured, at 30C in benzene, for poly (n-octadecyl acrylate) as [η]=2.72×10−4 Mn0.638 and for poly (N-n-octadecylacrylamide) as [η]=0.82×10−4 Mn0.676. Whole polymers of various molecular weights were prepared in benzene solution at 65C with dodecyl mercaptan as primary regulator. By the use of these parameters, the molecular weight of such polymers and their homologs may now be measured by simple solution-viscosity determinations. In the expression { } (relating degrees of polymerization { } to the mercaptan/monomer ratio), intercept { } and apparent transfer constant Cs for n-octadecyl acrylate were 6.28×10−3 and 0.68; for N-n-octadecylacrylamide 1.10×10−3 and 0.62 respectively. These parameters permit preparation of homopolymers of chosen molecular weight. Presented at the AOCS Meeting, Philadelphia, October 1966 E. Utiliz. Res. Devel. Div., ARS, USDA.  相似文献   

18.
An efficient approach for more selective synthesis of higher linear α-olefins was achieved by utilizing suitable reaction media in FTS reaction. About 41.4% average α-olefins content in C5–C25 fractions was obtained at = 5 g-cat h mol−1 in n-C10 solvent, which is markedly higher than the value (2.5%) obtained in 85% N2 + 15% n-C6.  相似文献   

19.
A mathematical model for an absorption of nitrogen oxides into water in packed column was developed based on the mass-transfer coefficient in packed column and the chemical reaction accompanying NO x absorption produces HNO3 and HNO2 in the liquid phase. The subsequent dissociation of HNO2 in the liquid-phase results in the formation of HNO3 and NO gas, and then this NO gas follows to be oxidized by O2 in air. The important factors influenced on the absorption of NO x are the oxidation state of NO gas and the partial pressure of nitrogen oxides in the gas. The efficiency of NO x absorption increases with the increase of the partial pressure of NO x . The most critical value for using the model is the constant of .The selection of 11×10−4kg·mol/atm·m2·sec for resonable for this model.  相似文献   

20.
A parallel-plate constant-stress rheometer is used to measure the yield stress τy, and the post-yield flow curve T(\documentclass{article}\pagestyle{empty}\begin{document}$ \dot \gamma $\end{document}), where τ is shear stress and \documentclass{article}\pagestyle{empty}\begin{document}$ \dot \gamma $\end{document} is shear rate, for microphase-separated triblock copolymer melts. Five polymer samples, all styrene-butadiene-styrene but with differing composition ratios and molecular weights, are tested at 125°C. Specimens are prepared by casting sheets from solutions made with different solvents. The τ(\documentclass{article}\pagestyle{empty}\begin{document}$ \dot \gamma $\end{document}) is found usually to be sigmoidal, for the range 10?5 < \documentclass{article}\pagestyle{empty}\begin{document}$ \dot \gamma $\end{document} < 10?3 s?1, representing different stages of microstructural degradation in flow. Measurements indicate that a true τy exists, with values in the range 100 < τy < 500 Pa for these melts. A general trend is detected for τy to decrease as the casting solvent solubility parameter increases. A scheme for correlating the dependence of τy, on composition and molecular weight is proposed for the various polymers. For selected samples, the effect of mechanical history (sequence of stress application) and a temperature variation that crosses Ts (110 to 150°C) are also explored.  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