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 15 毫秒
1.
Conclusions The temperature dependences of the tangent of the mechanical loss angle, tan , the dynamic shear modulus, G', and the velocity of shear waves, ct, of arimide PM fibre have been studied by the method of free twisting oscillations in the temperature range 20–900°K, at frequencies of 0.2–1 Hz.Relaxation transitions are observed in the tan = f(T) dependence in the temperature regions of 115, 220, 400, and 644°K, with activation energies of 23, 59, 84, and 770 kJ/mole, respectively.Values of the dynamic shear modulus of the investigated fibre have been obtained over a wide temperature range.Translated from Khimicheskie Volokna, No. 4, pp. 41–42, July–August, 1984.  相似文献   

2.
Summary -Methyl styrène polymerization was carried out using t-BuCl/Et2AlCl/CH2Cl2 system in the temperature range of 0° to –63.5°C. The effect of temperature on yield and molecular weights of Poly -Methyl styrene was determined. Based on Arrhenius plots, average activation energies of molecular weights were determined to be –9.2 ± 1.0 K cal/mole. (0° to 40°C) and 0 ± 0.5 K cal/mole (–40° to –63.5°C). These were postulated to reflect the molecular weight governing mechanisms such as transfer to monomer and termination respectively.  相似文献   

3.
Ammonia decomposition on Ir(100) has been studied over the pressure range from ultrahigh vacuum to 1.5 Torr and at temperatures ranging from 200 to 800 K. The kinetics of the ammonia decomposition reaction was monitored by total pressure change. The apparent activation energy obtained in this study (84 kJ/mol) is in excellent agreement with our previous studies using supported Ir catalysts (Ir/Al2O3 82 kJ/mol). Partial pressure dependence studies of the reaction rate yielded a positive order (0.9±0.1) with respect to ammonia and negative order (–0.7 ±0.1) with respect to hydrogen. Temperature-programmed desorption data from clean and hydrogen co-adsorbed Ir(100) surfaces indicate that ammonia undergoes facile decomposition on both these surfaces. Recombinative desorption of N2 is the rate-determining step with a desorption activation energy of 63 kJ/mol. Co-adsorption data also indicate that the observed negative order with respect to hydrogen pressure is due to enhancement of the reverse reaction (NH x + H NH x+1, x=0–2) in the presence of excess H atoms on the surface.  相似文献   

4.
The Relationship between surface structure and reactivity is investigated by means ofn-butane hydrogenolysis, a known structure sensitive reaction, for planar and faceted Pt/ W(111) surfaces. The W(111) surface reconstructs to form pyramidal facets with [211] orientation upon vapor deposition of Pt (>1.3 ML) and annealing above 750 K. The hydrogenolysis kinetics over the planar and the faceted surface are found to be quite different. The planar surface has a higher selectivity towards ethane formation and a higher reaction rate. The apparent activation energies are found to be 33 ± 4 kJ/mol for the planar surface and 76 ± 6 kJ/mol for a surface covered with 20 nm facets. There appears to be a correlation between the concentration of fourfold coordination (C4) sites on the surface and the amount of ethane produced. The C4 concentration is altered by changing the facet size (annealing temperature). The results indicate the presence of a different intermediate on the C4 sites as evidenced by the differences in the apparent activation energy, the reaction rate and the overall selectivity.  相似文献   

5.
The rate equation for the overall reaction of NO and O2 over Pt/Al2O3 was determined to be r=kf[NO] 1.05±0.08[O2]1.03±0.08[NO2]0.92±0.07(1-), with kf as the forward rate constant, =([NO2]/K[NO][O2]1/2), and K as the equilibrium constant for the overall reaction. An apparent activation energy of 82 kJ mol–1 ± 9 kJ mol–1 was observed. The inhibition by the product NO2 makes it imperative to include the influence of NO2 concentration in any analysis of the kinetics of this reaction. The reaction mechanism that fits our observed orders consists of the equilibrated dissociation of NO2 to produce a surface mostly covered by oxygen, thereby inhibiting the equilibrium adsorption of NO, and the non-dissociative adsorption of O2, which is the proposed rate determining step.  相似文献   

6.
The influence of conditions (e.g., ratios of components, temperature etc.) on the reaction of Cu(OCOCH3)2·2H2O with polyethylene grafted-polyacrylic on the amount of the metal and the composition of the immobilized Cu(II) complexes was studied. The concentration dependence obeys the Langmuir law. Analysis of the data leads to an evaluation of the stability constant for the Cu(II) complexes (K=300 l/mole at 333 K). The constant corresponds to a Cu(II) fixation value, k=0.35 mole/l (22.22 mg Cu(II)/g). The multistage fixation mechanism for Cu(II) complex formation was demonstrated by the marked atom technique. Cu(II) is fixed by one carboxylate group (to 16 mol% of the supported Cu(II), K 1=16×10–2 mole/g) and by two carboxylate groups (K 2=2.54×10–3 mole/g) of the grafted ligands. The PE-gr-PAA–Cu(II) system mimics the situation-insoluble support-soluble functional polymer covering and realizes the advantages of both the soluble and the three-dimensional crosslinked polymer. Steady-state magnetic susceptibility measurements and ESR spectroscopy were used to study the distribution of cupric ion attached to a polyethylene-grafted poly(acrylic acid) support. The existence of three types of cupric ion complexes was demonstrated: (1) isolated complexes, (2) complexes bonded by dipole–dipole interactions, and (3) clusters with strong exchange interactions. The mean distances between the isolated ions (¯r22–15 Å) and between the dipole–bound complexes (¯r agreg7 Å) were estimated. The results obtained were compared to the data for other immobilized catalysts. Preliminary results on the fixation of bimetallic Cu(II) and Pd(II) complexes to the polymers as well as on their distributions were obtained.  相似文献   

7.
This paper deals with a method of estimating single electrode heat balances during the electrolysis of molten NaCl-ZnCl2 in a cell using a-alumina diaphragm. By measuring the thermoelectric power of the thermogalvanic cells: (T) Na/-alumina/NaCl-ZnCl2/-alumina/Na(T+dT) and (T) C,Cl2/NaCl-ZnCl2/Cl2,C(T+dT) the single electrode Peltier heat for sodium deposition and for chlorine evolution at 370° C were estimated to be –0.026±0.001 JC–1 and+0.614±0.096 J C–1, respectively.  相似文献   

8.
Summary The polymerization of -methylstyrene (MeSt) using the H2O/SnCl4 initiating system and ethyl chloride solvent has been investigated over the temperature range from –40° to –122°C in the presence and absence of the proton trap 2,6-di-tert-butylpyridine (DtBP). Arrhenius (In ¯Mn versus 1/T) plots obtained with poly (methylstyrene) (PMeSt) samples prepared in the absence of DtBP reveal the existence of two sharply defined temperature regimes (see Figure 1): at higher temperatures from –40 to –86°C, the slope of the Arrhenius plot yields HM n = –4.90±0.25 kcal/mole whereas at lower temperatures, from –86° to –122°C,HM n = –0.3 kcal/mole. With PMeSt samples prepared in the presence of DtBP the HM n obtained for the higher temperature regime increases to –1.67 ± 0.20 kcal/mole whereas the HM n reflecting the lower temperatures remains the same as that obtained in the absence of DtBP. These observations are readily explained by postulating a change in mechanism at –86°C: Evidently the ¯Mn of PMeSt is determined over the higher temperature regime by chain transfer to monomer which is frozen out at lower temperatures where termination becomes ¯Mn determinant. In the absence of DtBP chain transfer to monomer is operative which leads to the higher Arrhenius slope over the higher temperature regime; however, over the lower temperature regime where chain transfer is absent and termination is ¯Mn controlling, the Arrhenius slope remains unchanged. Evidence obtained from a Mayo plot (negligible intercept in the 1/¯Mn versus 1/[MeSt]o plot, Figure 2) with samples prepared at –92°C, corroborate this postulate. Molecular weight dispersities (¯MW/¯Mn) as a function of temperature have been determined in the presence and absence of DtBP (Figure 3). The proton trap affects ¯MW/¯Mn only over the higher temperature regime which also suggests that chain transfer to monomer is frozen out at –86°C and that the polymerization becomes termination dominated at low temperatures.  相似文献   

9.
Value for the activation energy, U act, and the entropy change, S, for the reaction 2Li + S2O 4 2– Li2S2O4+2e in acetonitrile have been found to be 72 kJ mol–11 and — 0.3 kJ mol–1 K–1, respectively, by a combination of impedance techniques and the use of a temperature-controlled environment on commercially manufactured cells which acted as constant volume containers.  相似文献   

10.
When the impedance is measured on a battery, an inductive impedance is often observed in a high frequency range. This inductance is frequently related to the cell geometry and electrical leads. However, certain authors claimed that this inductance is due to the concentration distribution of reacting species through the pores of battery electrodes. Their argument is based on a paper in which a fundamental error was committed. Hence, the impedance is re-calculated on the basis of the same principle. The model shows that though the diffusion process plays an outstanding role, the overall reaction rate is never completely limited by this process. The faradaic impedance due to the concentration distribution is capacitive. Therefore, the inductive impedance observed on battery systems cannot be, by any means, attributed to the concentration distribution inside the pores. Little frequency distribution is found and the impedance is close to a semi-circle. Therefore depressed impedance diagrams in porous electrodes without forced convection cannot be ascribed to either a Warburg nor a Warburg-de Levie behaviour.Nomenclature A D¦C¦ (mole cm s–1) - B j+K¦C¦ (mole cm s–1) - b Tafel coefficient (V–1) - C(x) Concentration ofS in a pore at depthx (mole cm–3) - C 0 Concentration ofS in the solution bulk (mole cm–3) - C C(x) change under a voltage perturbation (mole cm–3) - ¦C¦ Amplitude of C (mole cm–3) - D Diffusion coefficient (cm2 s–1) - E Electrode potential (V) - E Small perturbation inE namely a sine-wave signal (V) - ¦E¦ Amplitude of E(V) - F Faraday constant (96500 A s mol–1) - F(x) Space separate variable forC - f Frequency in Hz (s–1) - g(x) KC(x)¦E¦(mole cm s–1) - I Apparent current density (A cm–2) - I st Steady-state current per unit surface of pore aperture (A cm–2) - j Imaginary unit [(–1)1/2] - K Pseudo-homogeneous rate constant (s–1) - K Potential derivative ofK, dK/dE (s–1 V–1) - K * Heterogeneous reaction rate constant (cm s–1) - L Pore depth (cm) - n Reaction order - P Reaction product - p Parameter forF(x), see Equation 13 - q Parameter forF(x), see Equation 13 - R e Electrolyte resistance (ohm cm) - R p Polarization resistance per unit surface of pore aperture (ohm cm2) - R t Charge transfer resistance per unit surface of pore aperture (ohm cm2) - S Reacting species - S a Total surface of pore apertures (cm2) - S 0 Geometrical surface area - S p Developed surface area of porous electrode per unit volume (cm2 cm–3) - s Concentration gradient (mole cm–3 cm–1) - t Time - U Ohmic drop - x Distance from pore aperture (cm) - Z Faradaic impedance per unit surface of pore aperture (ohm cm2) - Z x Local impedance per unit pore length (ohm cm3) - z Charge transfer number - Porosity - Thickness of Nernst diffusion layer - Penetration depth of reacting species (cm) - Penetration depth of a.c. signal determined by the potential distribution (cm) - Electrolyte (solution) resistivity (ohm cm) - 0 Flow of S at the pore aperture (mole cm2 s–1) - Angular freqeuncy of a.c. signal, 2f(s–1) - Integration constant  相似文献   

11.
The relaxation parameter K sthat is equal to the ratio of the viscosity to the Kohlrausch volume relaxation time s is analyzed. It is shown that this parameter can be evaluated from the temperature T 13(corresponding to a viscosity of 1013P) and the glass transition temperature T 8 +determined from the dilatometric heating curve. The maximum error of the estimate with due regard for experimental errors is equal to ±(0.4–0.5)logK sfor strong glasses and ±(0.6–0.8)logK sfor fragile glasses, which, in both cases, corresponds to a change in the relaxation times with a change in the temperature by ±(8–10) K. It is revealed that the viscosity, the Kohlrausch volume relaxation time s , and the shear modulus Gof glass-forming materials in silicate, borate, and germanate systems satisfy the relationship log( s G/) 1. The procedure for calculating the temperature dependences of the viscosity and the relaxation times in the glass transition range from the chemical composition and the T 8 +temperature for glass-forming melts in the above systems is proposed. The root-mean-square deviations between the calculated and experimental temperatures T 11and T 13are equal to ±(6–8) K for all the studied (silicate, borate, germanate, and mixed) oxide glass-forming systems. The proposed relationships can be useful for evaluating the boundaries of the annealing range and changes in the properties and their temperature coefficients upon cooling of glass-forming melts.  相似文献   

12.
The effects of thiourea (TU), benzotriazole (BTA) and 4,5-dithiaoctane-1,8-disulphonic acid (DTODSA) on the deposition of copper from dilute acid sulphate solutions have been studied using potential sweep techniques. Tafel slopes and exchange current densities were determined in the presence and absence of these organic additives. TU and BTA were found to inhibit the copper deposition reaction; increases in the BTA concentration gave a systematic lowering of the exchange current density, whilst TU behaved in a less predictable manner. For BTA and TU concentrations of 10–5 mol dm–3,j 0 values of 0.0027 ± 0.0001 and 0.0028 ± 0.0002 mA cm–2 were obtained compared to a value of 0.0083 ± 0.0003 mA cm–2 for the additive free acid sulphate solution. In contrast, in the presence of DTODSA, an increased exchange current of 0.043 ± 0.0003 mA cm–2 was observed. The presence of additives gave rise to measured Tafel slopes of –164, –180 and –190 mV for TU, BTA and DTODSA, respectively, compared to that of –120 mV for copper sulphate alone.List of symbols A electrode area (cm2) - b C cathodic Tafel slope (mV) - c B bulk concentration (mol cm–3) - D Diffusion coefficient (cm2 s–1) - F Faraday constant (A s mol–1) - I L Limiting current (A) - j Current density (A cm–2) - j CT Charge transfer current density (A cm–2) - j 0 Exchange current density (A cm–2) - k L Mass transport coefficient (cm s–1) - R Molar gas constant (J K–1 mol–1) - T Temperature (K) - z Number of electrons (dimensionless) Greek symbols C Cathodic transfer coefficient (dimensionless) - Overpotential (V) - v Kinematic viscosity (cm2 s–1) - Rotation rate (rad s–1)  相似文献   

13.
A procedure is described for computer-assisted optimization of an electrolytic process flowsheet. Material, energy, and economic balances for all process units were incorporated in a nonlinear optimization routine for predicting the minimum selling price based on a discounted cash flow rate of return on investment. The optimization utilized a simultaneous-modular approach which was incorporated into the public version of the Aspen flowsheeting package, and used an infeasible path convergence method based on successive quadratic programming procedures. Electrolyte vapour-liquid equilibrium data were estimated by the non-random two-liquid model. The Lagrangian multipliers of the constraint equations were used to determine the sensitivity of the optimum to key process variables. The method was illustrated by evaluation of two process flowsheets for electrosynthesis of methyl ethyl ketone (MEK) from 1-butene based on pilot-plant performance reported in the patent literature.List of symbols A c cell cost factor ($ cell–1) - A H heat exchanger cost factor ($ m–2) - A p pump cost factor ($ sl–1) - A R rectifier cost factor ($ kVA–1) - A T tank cost factor ($l –0.5) - A cm cell maintenance factor ($ A–1 y–1) - A cl cell labour ($ cell–1 y–1) - A cw cooling water cost ($ m–3) - A e electricity cost ($ kWh) - A m membrane cost ($ cell–1 y–1) - A om other maintenance factor, fraction of plant capital less cell cost - C p cooling water heat capacity (kJ kg–1 °C–1) - H operating hours per year - I C current to each cell (A) - I TOT total current to all cells (A) - L A Lang factor for auxiliaries - L C Lang factor for cells - L R Lang factor for rectifiers - N number of cells in plant - Q heat removal load (kJ h–1) - R production rate (kgh–1) - T cw cooling water temperature rise (°C) - T LM cooler log mean temperature difference (°C) - U heat transfer coefficient for cooler (kW m–2 °C–1) - v c electrolyte flow to each cell (l -1) - v C cell voltage (V) - R rectifier efficiency - cooling water density (kg m–3) - T surge tank residence time (s)  相似文献   

14.
The effect of benzotriazole, BTA, on mass transfer in dissolution-corrosion of the copper rotating disk electrode in 0.02 M Fe(III)–0.5 M H2SO4 has been studied by means of atomic absorption spectrometry. The mass transfer coefficient, K, was determined from the slope of ln(C 0/C)Fe(III) vs. time plots. In the absence of BTA the corrosion process can be described by the correlation Sh = KR/D = 4.47Re 0.5. The difference in values between Sh and Sh Levich, and the change in slope in the Arrenhius plot points to mixed control for the cathodic process Fe3+ + 1e Fe 2+ and charge transfer control for the anodic process, Cu Cu2+ + 2e. The average activation energies were 7.7 kJ mol–1 and 19.5 kJ mol–1 at (500–1500) and (2000–3000) rpm, respectively. At low concentration of BTA the inhibiting action of BTA increases with concentration and with rotation speed. For [BTA] 5 × 10–3 M, the K value, 10–4 cm s–1, remains constant and is independent of rotation rate. The morphology of the copper rotating disk after corrosion in the absence and presence of BTA was examined using scanning electron microscopy (SEM).  相似文献   

15.
An electrochemical ozone generation process was studied wherein glassy carbon anodes and air depolarized cathodes were used to produce ozone at concentrations much higher than those obtainable by conventional oxygen-fed corona discharge generators. A mathematical model of the build up of ozone concentration with time is presented and compared to experimental data. Products based on this technology show promise of decreased initial costs compared with corona discharge ozone generation; however, energy consumption per kg ozone is greater. Recent developments in the literature are reviewed.Nomenclature A electrode area (m2) - Ar * modified Archimedes number, d b 3 gG/2 (1 — G) - C O 3 (aq) concentration of dissolved ozone (mol m–3) - C O 3 i concentration at interface (mol m–3) - C O 3 1 concentration in bulk liquid (mol m–3) - D diffusion coefficient (m2 s–1) - E electrode potential against reference (V) - F charge of one mole of electrons (96 485 C mol–1) - g gravitational acceleration (9.806 65 m s–2) - i current density (A m–2) - i 1 limiting current density (A m–2) - I current (A) - j material flux per unit area (mol m–2 s–1) - k obs observed rate constant (mol–1 s–1) - k t thermal conductivity (J s–1 K–1) - L reactor/anode height (m) - N O 3 average rate of mass transfer (mol m–2 s–1) - Q heat flux (J s–1) - r i radius of anode interior (m) - r a radius of anode exterior (m) - r c radius of cathode (m) - R gas constant (8.314 J K–1 mol–1) - S c Schmidt number, v/D - Sh Sherwood number, k m d b/D = i L d b/zFD[O3] - t time (s) - T i temperature of inner surface (K) - T o temperature of outer surface (K) - U reactor terminal voltage (V) - electrolyte linear velocity (m s–1) - V volume (m3) - V O 3 volume of ozone evolved (10–6 m3 h–1) - z i number of Faradays per mole of reactant in the electrochemical reaction Greek symbols G gas phase fraction in the electrolyte - (mean) Nernst diffusion layer thickness (m) - fractional current efficiency - overpotential (V) - electrolyte kinematic viscosity (m2 s–1) - electrolyte resistivity (V A–1 m)  相似文献   

16.
Conclusions In the range 1200–1500°C the mechanism of oxidation of industrial heating elements containing 97–99% silicon carbide is approximately identical with, and is controlled in the same way, as the oxidation of pure silicon carbide [1, 2], in the main by oxygen diffusion through a film of SiO2.The relationship between the oxidation of various silicon carbide heating elements and time is described the equation q=Kb, where b comes within the range 0.3–0.5. The relationship between the oxidation rate of the heating elements and the temperature obeys the Arrhenius equation: K=K0exp(–E/RT).The calculated activation energy for the process for various heating elements alters from 23 to 43 kcal/mole.The porosity of the heaters (up to values not exceeding 18%) substantially affects the oxidation capacity.Translated from Ogneupory, No.6, pp. 52–57, June, 1970.  相似文献   

17.
A method for studying the heat resistance of composite aluminum nitride-based ceramic materials in air at 1073 – 1273 K is developed that allows the change in mass to be measured with an accuracy of 0.15 – 0.17 mg. The interaction between AlN-based composite materials and a phosphate binder (H3PO4) is studied and compared with hot-pressed specimens. A mechanism for the effect of the binder on the kinetics of oxidation is proposed. The relatively low activation energies (152 and 205 kJ/mole) suggest that the oxidation process is mainly determined by the diffusion of aluminum ions through the -Al2O3 film.  相似文献   

18.
Long service life IrO2/Ta2O5 electrodes for electroflotation   总被引:1,自引:0,他引:1  
Ti/IrO2-Ta2O5 electrodes prepared by thermal decomposition of the respective chlorides were successfully employed as oxygen evolving electrodes for electroflotation of waste water contaminated with dispersed peptides and oils. Service lives and rates of dissolution of the Ti/IrO2-Ta2O5 electrodes were measured by means of accelerated life tests, e.g. electrolysis in 0.5M H2SO4 at 25°C and j = 2 A cm–2. The steady-state rate of dissolution of the IrO2 active layer was reached after 600–700 h (0.095 g Ir h–1 cm–2) which is 200–300 times lower than the initial dissolution rate. The steady-state rate of dissolution of iridium was found to be proportional to the applied current density (in the range 0.5–3 A cm–2 ). The oxygen overpotential increased slightly during electrolysis (59–82 mV for j = 0.1 A cm–2 ) and the increase was higher for the lower content of iridium in an active surface layer. The service life of Ti/IrO2 (65 mol%)-Ta2O5 (35 mol%) in industrial conditions of electrochemical devices was estimated to be greater than five years.List of symbols a constant in Tafel equation (mV) - b slope in Tafel equation (mV dec–1) - E potential (V) - f mole fraction of iridium in the active layer - j current density (A cm–2) - l number of layers - m Ir content of iridium in the active layer (mg cm–2) - r dissolution rate of the IrO2 active layer (g Ir h–1 cm–2) - T c calcination temperature (°C) - O 2 oxygen overpotential (mV) - O 2 difference in oxygen overpotential (mV) - A service life in accelerated service life tests (h) - S service life in accelerated service life tests related to 0.1 mg Ir cm–2 (h) - p polarization time in accelerated service life tests (h)  相似文献   

19.
The thermodynamics of aqueous sulphur-water systems are summarized in the form of potential-pH diagrams, calculated from recently reported critically assessed standard Gibbs energies of formation of the species considered. However, there is convincing evidence from the literature that a value of pK a(HS) = 17–19 is appropriate, whereas a value of 13 is widely accepted; hence, the higher value of 19, corresponding to G f 0 (S2–) = 120.5 kJ mol–1 , was used in these calculations, rather than G f 0 (S2-) = 86.31 kJ mol–1 quoted in the main data source.Under ambient conditions, only – 2 (sulphide), 0 (elemental sulphur) and + 6 (sulphate) oxidation states are thermodynamically stable in water, which is predicted to be oxidized by peroxodisulphate (H2 S2 O8/SO 8 2– and peroxomonosulphate (HSO 5 /SO 5 2– ). However, when sulphate is excluded from the calculations to allow for the large energy of activation/slow kinetics of its formation from sulphide, then other sulphoxy species appear on the diagram for what is then a metastable system. Similarly, if all sulphoxy species (i.e. any species with oxidation states > 0) are excluded, then polysulphide ions (S n 2– , 2 n 5) have areas of predominance at high pH, each with a narrow potential window of predominance. Hence, this information is complemented with S n 2– /HS activity-potential diagrams at pH 9 and 14.Some species have no area of stability even on the metastable diagrams. Hence, potential-pH diagrams are also presented for the sulphite-dithionite system (excluding elemental sulphur), and that involving peroxomonosulphate (HSO 5 /SO 5 2– ) in place of peroxodisulphate (H2S2O8/SO 8 ¨– ).  相似文献   

20.
The polymerization kinetics of methyl methacrylate with K2S2O8/L-serine redox system has been investigated volumetrically at 35±0.1°C under nitrogen atmosphere acidic aqueous medium in DMF/H2O mixture (50% v/v). The rates of polymerization were measured varying concentrations of the monomer, initiator, L-serine as well as temperature; and it was found to increase with increasing of both temperature and concentrations of monomer, initiator, and L-serine. The overall energy of activation (E a ) has been calculated to be 29.48 kJ/mol from the Arrhenius plot in temperature range 25–50°C. The molecular weight of the polymer was determined by gel permeation chromatography (GPC). Based on kinetic studies and depending on the results obtained, a suitable reaction mechanism has been suggested and the rates of polymerization found to obey the following equation: V p [methyl methacrylate]1.09[L-serine]1.03[K2S2O8]0.96.  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