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 22 毫秒
1.
Taro Sonobe 《Fuel》2008,87(3):414-421
The pyrolysis behaviors of several agricultural residues have been investigated by using thermogravimetric analysis. The evolving rates of the gaseous products during the pyrolysis such as H2, CH4, H2O, CO and CO2 were also measured by the TG-MS techniques. Without any assumption and mathematical fitting, we could obtain the very proper kinetic parameters (the distribution curve of activation energy, f(E), and the activation energy dependent frequency factor, k0(E)) of biomass pyrolysis by utilizing the distributed activation energy model (DAEM) proposed by Miura and Maki [Miura K, Maki T. Energ Fuel 1998;12:864]. The peaks of f(E) curve for rice straw, rice husk, corncob and cellulose were found to be 170, 174, 183, and 185 kJ/mol, respectively. The k0 value increased from an order of 1011 to an order of 1018 s−1, while E increased from 120 to 250 kJ/mol. The catalytic effects of alkali and alkaline earth metals during the pyrolysis play a major role in the variation of f(E) curve among the different biomass species.  相似文献   

2.
The propagation and termination rate coefficients for bulk polymerization of the butyl acrylate dimer (BA dimer) are determined by pulsed laser techniques. The rate coefficient for propagation, kp, is deduced for temperatures from 20 to 90 °C via the pulsed laser polymerization-size exclusion chromatography (PLP-SEC) method at pulse repetition rates between 1 and 10 Hz. The Arrhenius parameters were found to be: EA(kp) = (34.2 ± 1.0) kJ mol−1 and A(kp)/L mol−1 s−1 = (1.08 ± 0.49) × 107 L mol−1 s−1. The termination rate coefficient, kt, has been measured via SP-PLP-ESR, single pulse-pulsed laser polymerization in conjunction with time-resolved electron spin resonance detection of radical concentration. The resulting Arrhenius parameters as deduced from the temperature range −15 to +30 °C are: EA(〈kt〉) = (22.8 ± 3.7) kJ mol−1 and log(A/L mol−1 s−1) = 10.6 ± 1. The chain-length dependence of kt was studied at 30 °C. For short chains a significant dependence was found which may be represented by an exponent α = 0.79 in the power-law expression kt(i) = kt0i−α.  相似文献   

3.
M. Sreekanth  Ajit Kumar Kolar 《Fuel》2010,89(5):1050-1055
This work presents the results of experiments conducted to determine the mass loss characteristics of a cylindrical wood particle undergoing devolatilization under oxidation conditions in a bubbling fluidized bed combustor. Cylindrical wood particles having five different sizes ranging from 10 to 30 mm and aspect ratio (l/d = 1) have been used for the study. Experiments were conducted in a lab scale bubbling fluidized bed combustor having silica sand as the inert bed material and air as the fluidizing medium. Total devolatilization time and mass of wood/char at different stages of devolatilization have been measured. Studies have been carried out at three different bed temperatures (Tbed = 750, 850 and 950 °C), two inert bed material sizes (mean size dp = 375 and 550 μm) and two fluidizing velocities (u = 5umf and u = 10umf). Devolatilization time is most influenced by the initial wood size and bed temperature. Most of the mass is lost during the first half of the devolatilization process. There was no clear influence of the fluidization velocity and bed particle size on the various parameters studied. The apparent kinetics estimated from the measured mass history show that the activation energy varied narrowly between 15 and 27 kJ/mol and the pre-exponential factor from 0.11 and 0.45 s−1 for the wood sizes considered.  相似文献   

4.
Devolatilization of five coals having volatile matter in the range of 31 to 41% was studied in argon and in air under fluidized bed conditions. The diameter of the coal particles varied between 4 and 9.5 mm. The variation of devolatilization time with particle diameter was expressed by the correlation, tv = Advn. The superficial gas velocity was found to have a significant effect on the rate of devolatilization. The devolatilization rate increased with the increase in the oxygen concentration in the fluidizing gas. The correlations developed in this study fitted the mass versus time profiles of the coal particles satisfactorily. The same correlations were found to be appropriate for predicting devolatilization of a batch of coal particles. The correlations developed in the present study will be useful for the design of fluidized bed combustors.  相似文献   

5.
Pyrolysis of sawdust was studied using a thermogravimetric analyser (TGA) to understand the devolatilisation process and to obtain its global kinetic parameters. The influences of particle size, initial weight of the sample and heating rate on the devolatilisation of sawdust particles have been studied. Results from proximate analysis show that smaller particle size has more ash content compared to larger particle size. The TG and derivative TG curve for variation in particle size and initial weight of the sample showed significant difference in the third stage of the pyrolysis. In addition, the pyrolysis of sawdust differed significantly for variation in heating rate. As the heating rates increased, the char yield also increased. The devolatilisation kinetics was studied considering different stages of pyrolysis. The kinetic parameters for thermal devolatilisation of the sawdust were determined through a nonlinear optimisation method of two independent parallel nth‐order reaction models. The kinetic parameters such as activation energy, frequency factor and order of the reaction for the two stages considered in the model were: E2 = 79.53 (kJ/mol), E3 = 60.71 (kJ/mol); k02 = 1.90 × 106 (1/min), k03 = 1.01 × 103 (1/min); n2 = 0.91, n3 = 1.78, respectively. The results show good agreement between the proposed model and the experimental data of the sawdust pyrolysis.  相似文献   

6.
A 1-D mathematical model describing the thermal decomposition, or calcination, of a single gibbsite particle to alumina has been developed and validated against literature data. A dynamic, spatially distributed, mass and energy balance model enables the prediction of the evolution of chemical composition and temperature as a function of radial position inside a particle. In the thermal decomposition of gibbsite, water vapour is formed and the internal water vapour pressure plays a significant role in determining the rate of gibbsite dehydration. A thermal decomposition rate equation, developed by closely matching experimental data reported previously in the literature, assumes a reaction order of 1 with respect to gibbsite concentration, and an order of −1 with respect to water vapour pressure. Estimated values of the transformation kinetic parameters were k0 = 2.5 × 1013 mol/(m3 s) for the pre-exponential factor, and Ea = 131 kJ/mol for the activation energy. Using these kinetic parameters, the gibbsite particle model is solved numerically to predict the evolution of the internal water vapour pressure, temperature and gibbsite concentration. The model prediction was shown to be very sensitive to the values of heat transfer coefficient, effective diffusivity, particle size and external pressure, but relatively less sensitive to the mass transfer coefficient and particle thermal conductivity. The predicted profile of the water vapour pressure inside the particle helps explain some phenomena observed in practice, including particle breakage and formation of a boehmite phase.  相似文献   

7.
The standard rate constant of a simple electrode reaction Ox + ne ↔ Red, in which both Ox and Red are solution soluble, can be determined by the variation of frequency in the square-wave voltammetry with inverted scan direction: log ks = log f01/2 + log D1/2 − 0.60 ± 0.01. In this equation log f0 is the abscissa of the intersection of straight lines Ep,2 = a log f + b and Ep,2 = E0, where Ep,2 is the potential of the second peak of the net response, E0 is the standard potential, a = 2.3RT/2(1 − α)nF, b = E0 − 2a log ks + a log D − 0.0353/(1 − α)n and D is a common diffusion coefficient.  相似文献   

8.
Elanio A. Medeiros 《Fuel》2011,90(4):1696-1699
The rate constants for the quenching of biacetyl phosphorescence by a series of conjugated dienes were measured. 1,3-cyclohexadiene (kqP = 2.94 × 109 s−1 mol−1 L), 2,5-dimethyl-2,4-hexadiene (kqP = 1.91 × 109 s−1 mol−1 L), 2,4-dimethyl-1,3-pentadiene (kqP = 1.78 × 108 s−1 mol−1 L), 3-methyl-1,3-pentadiene (kqP = 1.22 × 108 s−1 mol−1 L), 2,4-hexadiene (kqP = 1.35 × 108 s−1 mol−1 L) and trans-2-methyl-1,3-pentadiene (kqP = 3.84 × 108 s−1 mol−1 L). Cyclooctene also quenched biacetyl phosphorescence but with a lower rate (kqP = 1.97 × 107 s−1 mol−1 L). Quenching was not observed with 1-methylnaphthalene. Since conjugated dienes quench biacetyl phosphorescence preferentially, this method was studied using gasoline samples with known diene composition. A good correlation was found between the rate of quenching of biacetyl by the gasoline samples and the quantity of conjugated dienes present.  相似文献   

9.
A kinetic analysis of the pyrolysis of various types of biomass (trunk, bark, leaf, shell, herbage, food dregs, and polysaccharide) as well as synthetic biomass consisting of cellulose and lignin was performed using thermogravimetric analysis data. The reaction rates of biomass pyrolysis were found to be expressed simply by a single nth-order reaction model. The kinetic parameters (frequency factor k0, activation energy E, and reaction order n) were estimated first by differentiating the thermogravimetric curves and then by the nonlinear estimation method. The rate parameters of the pyrolysis of both 38 biomass samples and 9 synthetic biomass samples were successfully correlated in terms of the solid residue yield ω; charts are presented showing the correlations. Furthermore, a linear correlation was found between ω and the lignin content L in the woody biomass. This allows the kinetic parameters of biomass pyrolysis to be estimated using the value of ω, which is obtained from thermogravimetric measurements or estimated from the value of L for the biomass feedstock.  相似文献   

10.
Study of pyrolysis kinetics of oil shale   总被引:2,自引:0,他引:2  
Shuyuan Li  Changtao Yue 《Fuel》2003,82(3):337-342
The pyrolysis experiments on oil shale samples from Fushun, Maoming, Huangxian, China, and Colorado, USA, were carried out with the aid of thermogravimetric analyzer (TGA) at a constant heating rate of 5 °C/min. A kinetic model was developed which assumes several parallel first-order reactions with changed activation energies and frequency factors to describe the oil shale pyrolysis. The kinetic parameters of oil shale pyrolysis were determined on the basis of TGA data. The relationship between the kinetic parameters was further investigated and the correlation equations of x-E and ln A-E were obtained. These equations show that the final fractional conversion of each parallel reaction, x(j), can be expressed as an exponential function of the corresponding activation energy. The plot of ln A-E for different reactions becomes a straight line. These relationship equations can provide important information to understand the pyrolysis mechanism and to investigate the chemical structure of oil shale kerogen.  相似文献   

11.
Liping Wang 《Electrochimica acta》2006,51(26):5961-5965
The electrochemical behaviour of the anticancer herbal drug emodin was investigated by cyclic voltammetry (CV) at glassy carbon electrode. In 0.05 M NH3-NH4Cl (50% ethanol, pH 7.2) buffer solution, a pair of quasi-reversible redox peaks at potentials of Ep1 = −0.688 V and Ep2 = −0.628 V and one irreversible anodic peak, which was a typical anodic peak of emodin, at Ep3 = −0.235 V appeared at a scan rate of 100 mV/s. The irreversible anodic peak currents are linearly related to the emodin concentrations in a range from 8.9 × 10−8 M to 7.8 × 10−6 M with a pre-concentration time of 80 s under −0.620 V. Using the established method without pretreatment and pre-separation, emodin in herbal drug was determined with satisfactory results. Moreover, the electrode process dynamics parameters were also investigated by electrochemical techniques.  相似文献   

12.
The bulk phase of nonionic surfactant C10E4 solution was monitored by a dynamic light scattering (DLS) system at 20 °C in a narrow range of concentration near the cmc. Two particle aggregations were observed. The DLS data show (i) there exist premicellar multimers (or called sub-micelles) and (ii) micelles coexist with multimers. The C10E4 sub-micelles have a narrow size distribution with an averaged hydrodynamic diameter (Dh) of 1.35 nm. The Dh of the micelles is around 10.5 nm at 1.0 × 10−6 mol/mL and increases slightly with C10E4 concentration. It is illustrated from the DLS data that (i) at C = 0.78-0.82 μmol/mL, monomers and premicellar multimers coexist and (ii) at C = 0.84-0.92 μmol/mL, monomers + submicellar multimers + micelles coexist. At more elevated concentrations, only the signals from the micelles are detected by DLS.  相似文献   

13.
An alumina precursor was prepared by the aluminium sulphate (0.20 M) and excess urea reaction in boiling aqueous solution. The precursor was calcined at 900 °C for 2 h and then δ-Al2O3 powder having volumetric agglomeration degree of 80% was obtained. Cylindrical compacts having diameter of 14 mm were prepared under 32 MPa by axial pressing using oleic acid as binder. Each compact was fired isothermally at various temperatures between 950 and 1400 °C. The firing time was changed from zero to 2 h. The fired compacts were examined by scanning electron microscopy (SEM) and nitrogen adsorption techniques. The specific surface areas (S/m2 g− 1) of the samples were calculated using the Brunauer, Emmett, and Teller (BET) procedure. The rate constant (k) and mechanism-characteristic parameter (n) were obtained for different temperatures between 950 °C and 1150 °C from the application of the neck-growth sintering rate (NGRM) model on the surface area reduction data. An Arrhenius equation and the parameter n for the sintering were found in the forms of k = (7.648 × 106 h− 1) exp (− 186,234 J mol− 1 / RT) and n = 4.0 × 10− 7 T3-1.7 × 10− 3 T2 + 2.3 T − 1030.8 respectively. The parameter n changes in the interval 0.61 <  n < 1.34 with rising temperature having maximum at about 1025 °C. Based on the SEM images and NGRM data, the intra-particle sintering was discussed.  相似文献   

14.
Thermal hydrocracking and catalytic hydrocracking over NiMo/γ-Al2O3 of a pentane-insoluble asphaltene were conducted in a microbatch reactor at 430 °C. The experimental data of asphaltene conversion fit second-order kinetics adequately, to give the apparent rate constants of 2.435 × 10−2 and 9.360 × 10−2 wt frac−1 min−1 for the two processes respectively. A three-lump kinetic model is proposed to evaluate rate constants of parallel reactions from asphaltenes to liquid oil (k1) and to gas + coke (k3), and consecutive reaction from liquid to gas + coke (k2). The evaluated k1 is 2.430 × 10−2 and 9.355 × 10−2 wt frac−1 min−1, k2 is 2.426 × 10−2 and 6.347 × 10−3 min−1, and k3 is 5.416 × 10−5 and 4.803 × 10−5 wt frac−1 min−1 for asphaltenes hydrocracking in the presence or absence of the catalyst, respectively. Analysis of selectivity shows that the catalytic hydrocracking process promotes liquid production and inhibits coke formation effectively.  相似文献   

15.
Cytosine plays an important role in many biological processes since it constitutes the buildings blocks of DNA and RNA. A two-step reduction of Zn2+ ions at the dropping mercury electrode in acetic buffers at pH 4 and 5 in the presence of cytosine was examined. The measurements were performed using an impedance method in a wide potential and frequency ranges.The values of the standard rate constants ks in the both studied system decrease from 3.8 × 10−3 to 2 × 10−3 cm s−1 at pH 4 and from 5.1 × 10−3 to 2.5 × 10−3 cm s−1 at pH 5. The values of the standard rate constants ks1 characterizing the stage of the first electron transfer decrease similarly. However, the values of the standard rate constants ks2 characterizing the stage of the second electron exchange decrease more markedly in the buffer at pH 4 than in the buffer at pH 5.  相似文献   

16.
A pentane-insoluble asphaltene was processed by thermal cracking and catalytic hydrocracking over NiMo/γ-Al2O3 in a microbatch reactor at 430 °C. Kinetic analysis shows that the first-order kinetics fits the data of conversion in reaction times ≤ 30 min approximately, but deviates from the data of times over 30 min significantly; whereas the second-order kinetics fits the data of the reaction times up to 60 min adequately, to give the apparent rate constants of 1.704 × 10−2 and 9.360 × 10−2 wt frac−1min−1 for the two cracking processes. Furthermore, a three-lump kinetic model is proposed to include parallel reactions of asphaltenes to produce liquid oil (k1) and gas + coke (k3), and consecutive reaction from liquid to gas + coke (k2). The evaluated value of k1 is 1.697 × 10−2 and 9.355 × 10−2 wt frac−1min−1, k2 is 3.605 × 10−2 and 6.347 × 10−3 min−1 , and k3 is 6.934 × 10−5 and 4.803 × 10−5 wt frac−1min−1 for asphaltenes thermal cracking and catalytic hydrocracking, respectively. Selectivity analysis shows that the catalytic hydrocracking process promotes liquid production and inhibits coke formation effectively.  相似文献   

17.
Peihua Ren 《Polymer》2009,50(20):4801-5711
We report the synthesis and optoelectronic properties of highly soluble poly(9,10-bis(3′,4′-di(2″-ethylhexyloxy))phenyl)-2,6-anthracenevinylene) (HSM-PAV). The key intermediate for the synthesis of HSM-PAV is 2,6-dimethyl-9,10-dibromoanthracene, and the high solubility of HSM-PAV is from the incorporation of lateral 3,4-di(2-ethylhexyloxy)phenyl moieties into the 9,10-positions of anthracene units. The increase of side alkyloxy groups endows HSM-PAV with higher molecular weight (Mn = 3.2 × 104) and better electroluminescence performances (Lmax = 590 cd/m2, LEmax = 0.27 cd/A) compared with the poly(2,6-anthracenevinylene) with lateral monoalkyoxy moieties (Mn = 1.9 × 104, Lmax = 340 cd/m2, LEmax = 0.17 cd/A). The electrical conductivity of doped HSM-PAV film with iodine is 5 × 10−2 S cm−1 that is several order higher than that of doped 9,10-anthracene-based polymers, further demonstrating that linkage position has a dramatic effect on the optoelectronic properties of anthracene-based conjugated polymers.  相似文献   

18.
LaFeO3 were synthesized via a sol-gel route based on polyvinyl alcohol (PVA). Differential scanning calorimetry (DSC), Thermogravimetric (TG), Fourier transform infrared spectroscopy (FT-IR), X-ray diffraction (XRD), Raman spectroscopy and field emission scanning electron microscopy (FESEM) techniques were used to characterize precursors and derived oxide powders. The effect of the ratios of positively charged valences to hydroxyl groups of PVA (Mn+/-OH) on the formation of LaFeO3 was investigated. XRD analysis showed that single-phase and well-crystallized LaFeO3 was obtained from the Mn+/-OH = 4:1 molar ratio precursor at 700 °C. For the precursor with Mn+/-OH = 2:1, nanocrystalline LaFeO3 with average particle size of ∼50 nm was formed directly in the charring procedure. With increase of PVA content to Mn+/-OH = 1:1, phase pure LaFeO3 was obtained at 500 °C.  相似文献   

19.
A study on the pyrolysis of asphalt   总被引:1,自引:0,他引:1  
The pyrolysis of asphalt has been studied using thermogravimetric analysis at atmospheric pressure and with nitrogen as the ambient gas. The heating rates ranged from 50 or 80 °C min−1 to a final temperature of 650 °C. A two-stage first-order model is established to describe the pyrolysis of asphalt. In the model, the activation energy, E, is different for each stage, but is independent of the type of asphalt and its heating rate. The frequency factor depends on the heating rate and is independent of the asphalt. The final yield of volatiles depends on the type of asphalt. The modeled results agree with the experimental measurements, so the model is reasonable.  相似文献   

20.
The thermal conditions for obtaining the glass-ceramic material of Al0.107B0.374Mg0.043Zn0.282Ca0.100Si0.927O3 with a crystalline phase in the form of gahnite (ZnAl2O4) were specified. The activation energy Ea and the Avrami parameter n for the crystallisation process were determined with the non-isothermal DTA procedure. The maximum temperatures of crystallisation of phases, depending on the rate of heating, ranged between 800-840 °C for willemite and 870-915 °C for gahnite. The homogeneous crystalline spinel phase was obtained by heat treatment above 1000 °C. Precipitation solely of a ghanite phase from glass-ceramic causes a relative increase in its fracture toughness and wear resistance compared to the two-phase materials, i.e., KIC = 2.12-1.65 MPam1/2 and ws = 0.21 × 10−4 mm3/Nm to ws = 1.43 × 10−4 mm3/Nm.  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