Article pubs.acs.org/IECR Simulation Research on the Fixed-Bed Gasification Process in a TwoStage Combined Gasifier Yifei Wang,* Weilong Jin, Longchu Zhu, Guangsuo Yu, Zhenghua Dai, and Fuchen Wang Key Laboratory of Coal Gasification and Energy Chemical Engineering of Ministry of Education, Shanghai Engineering Research Center of Coal Gasification, East China University of Science and Technology, Shanghai 200237, China ABSTRACT: The novel two-stage gasification process that is combined entrained-flow for first stage and fixed-bed for second stage was shown in the study. The benefit of the second stage was to recycle the sensible heat from first stage syngas and enhance the energy utilization. The fixed-bed gasification reaction of second stage was individually investigated by numerical simulation. The simulated results were validated well with the experimental data in terms of gas concentration, gas flow rate, and carbon conversion. The distributions of velocity, temperature, gas concentration, solid mass fraction, and carbon conversion during the gasification process were analyzed. The influences of coal amount and particle size in the second stage on gasification reactivity were also discussed in the study. Results showed that the gas temperature was reduced and the effective gas concentration increased when first stage syngas flowing through the second stage coal layer. The increase of coal amount of second stage showed a more significant heat recovery, but the heating rate and ultimate carbon conversion would be reduced. The increase of particle size resulted in the decreasing of effective gas concentration and gas−solid reaction rate, which was attributed to the change of specific surface area and the increase heat transfer resistance from the particle surface to inside. 1. INTRODUCTION Coal gasification is an important way to utilize coal resource efficiently and cleanly, which could convert the primary energy to a secondary clean energy. The product gases can be used as the feed stock to treat as fuel, syngas, hydrogen, carbon monoxide, and so on. Nowadays, coal gasification has been widely used in synthetic ammonia, methanol synthesis, sponge iron production, and other fields, and one of the biggest potential markets is integrated coal gasification combined cycle (IGCC) power generation.1 Entrained-flow gasification technology has become an advanced coal gasification technology due to its advantages of high carbon conversion, large singlefurnace production capacity, and good feedstock flexibility. However, the gasification temperature in the gasifier is extremely high, resulting in much waste of syngas sensible heat from the outlet of gasifier. Therefore, a syngas cooling system is essential for the entrained-flow gasification process. As to the cooling methods for the existing gasifiers, most of them still have deficiencies in the technical process. For example, the Shell gasification process2 uses recycle syngas quench to recover the syngas heat. The disadvantage is the increase the circulation compressor and its compression work. The Texaco gasifier3 featured with water quench process ignores the sensible heat recovery from high temperature syngas, while its radiation syngas cooler process increases the huge investment of the equipment. In view of these situations, a novel coal-based two-stage combined gasification process with heat recovery via chemical reaction was proposed by Institute of Clean Coal Technology, East China University of Science and Technology, which is aimed at the insufficient heat recovery of high temperature syngas for the current entrainedflow gasifiers. The novel coal gasification process is an organic combination of the opposed multiburner (OMB) gasifier at the first stage and lump coal moving bed gasifier at the second stage.4,5 Two endothermic reactions of carbon−H2O and © 2014 American Chemical Society carbon−CO2 are utilized to recover the sensible heat of high temperature syngas. The components of CO2 and H2O in the syngas undergo further reaction with the lump coal in the second stage, and the cold gas efficiency can be further improved. The sensible heat of syngas from entrained-flow gasifier provides the heat that the reactions needed and does not need additional oxidant for the second stage gasification reaction. Then, the sensible heat of hot temperature syngas from entrained flow gasifier is shifted to chemical energy and stored in the gas products with a simple and compact structure. It has a dual function for energy recycling efficiently and reduction of carbon dioxide emission.6 According to the patent named “Coal-Based Two-Stage Combined Gasification Process”, Huang et al.7 set up a hotmodel experimental device and investigated the influences of second stage coal amount and gas composition from first stage on integrated gasification efficiency. Jin et al.8 continued to perform the experimental research by focusing on the gasification efficiency influenced by the lump coal type and particle size. In order to further comprehend the changing trend of lump coal layer in the second stage and gas concentration and temperature profile in the furnace, detailed simulation research should be carried out for the hot-model two-stage experimental device. Biasi9 adopted the mathematical method to formulate the biomass gasification process in a downdraft reactor. The simulation research could predict the influences of model parameters, kinetic constants, and operational variables on the produced gas quality well. Yang et al.10 used a mathematical model to predict the main chemical Received: Revised: Accepted: Published: 7611 January 22, 2014 March 14, 2014 April 3, 2014 April 3, 2014 dx.doi.org/10.1021/ie500309a | Ind. Eng. Chem. Res. 2014, 53, 7611−7621 Industrial & Engineering Chemistry Research Article and physical processes and comparing with the experiment results in fixed-bed gasifier, assessing the subsequent effect of model parameters varying on the char gasification characteristics. In recent years, the rapid development of computer hardware promotes the use of computational fluid dynamics (CFD) software to simulate and analysis fixed-bed gasification phenomena. Murgia et al.11 used the Euler−Euler approach to simulate an air-blown updraft fixed-bed coal gasifier based on CFD software. The solid phase was considered as continua according to the kinetic and plastic theory of granular flows. The operation of the gasifier was investigated, and the characteristics of space- and time-dependent behavior were also addressed. Wu et al.12 used the Euler−Euler approach to comprehend the biomass gasification in a downdraft fixed-bed gasifier. The model results exhibited a reasonable agreement with the experimental data, and the parametric studies of varying preheating temperature and steam/air ratio were performed on the basis of the developed model. Their work is based on a common fixed-bed gasifier with the steady condition, while no simulation research is about the heat recovery fixed-bed gasification process in the combined gasifier. As to this simulation, the high temperature syngas reacting with the lump coal under the unsteady condition, and a thorough endothermic reaction is carried out in the second stage fixedbed. A two-dimensional fixed-bed gasification model is established to simulate the second stage fixed-bed gasification process in the proposed two-stage gasifier, which concerned the heat utilization and gasification efficiency in the syngas atmosphere. The Euler−Euler approach is adopted to solve gas and solid phase equations in this model. The accuracy of the model is verified by the comparison of the simulated results and experimental data, and the gasification characteristics of fixedbed for the second stage are also analyzed, including the distributions of velocity, temperature, gas concentration, gas− solid reaction rate, and carbon conversion. Furthermore, the influences of coal amount in the second stage and lump coal particle size on fixed-bed gasification reactivity are also discussed. Figure 1. Structure of the combined two-stage gasifier. Inner Mongolia lignite is used as the feedstock of the second stage fixed-bed. The proximate and ultimate analyses are shown in Table 1. The produced gas from second stage outlet is scrubbed and quenched in a chilling chamber, and the syngas volume is measured by a V flow-meter. The syngas composition is analyzed by Aglient7890a GC with the sampling frequency of 10 min after being dried and purified. After the experiment is finished, the mass of residue in the second stage and its ash are weighed to calculate the ultimate carbon conversion. The stable parameters for the first stage syngas are tabulated in Table 2 on the basis of practical operation conditions. 3. MATHEMATICAL EQUATIONS The multiphase flow model with Euler−Euler approach is adopted to simulate the fixed-bed gasification process. In this model, the high temperature syngas is treated as gas phase and the lump coal in the fixed-bed as solid phase, in which both of the phases are dealt with interpenetrating. The conservation equations of mass, momentum and energy are solved for gas and solid phases simultaneously. The standard k−ε turbulence model13 is used to describe the gas phase flowing. 3.1. Continuity Equation. The continuity equations for the gas and solid phases are defined as14 2. EXPERIMENTAL DEVICE AND FEEDSTOCK The combined two-stage gasifier is sketched in Figure 1. The size of alumina tube in the furnace is ϕ160 × 20 mm with the height of 1010 mm. The first stage is entrained-flow gasification with diesel and oxygen feed, and the fixed-bed pattern is adopted as the second stage with coal feed. The purpose of the study is to address the transient gasification phenomena in the second stage, so the first stage is adopted to provide the proposed syngas condition for the second stage inlet. The experiment process can be divided into two periods: initial heating process and subsequent gasification reaction process. The syngas produced from the first stage will be exhausted from the first stage outlet in order to avoid reacting with coal during the heating process. When the syngas temperature and composition reaches the stable in the first stage, the valve of second stage outlet is turned to open and meanwhile the valve of first stage outlet is turned to closed. Subsequently, the high temperature syngas from first stage flows to the lump coal in the second stage and the produced gas flows out from the second stage outlet. The latter period of fixed-bed gasification process in the combined gasifier is just the research process in this simulation. ∂ (αgρg ) + ∇(αgρg Ug) = Sgs ∂t (1) ∂ (αsρs ) + ∇(αsρs Us) = Ssg ∂t (2) In which α, ρ, U are the volume fraction, density and instantaneous velocity, respectively. S is the mass source term, resulting from the heterogeneous reaction between gas and solid phases, which meets the equation of Ssg = wc∑γcRc = −Sgs, where wc is the molecular weight, γc is the stoichiometric coefficient, and Rc is the reaction rate. 3.2. Momentum Equation. 3.2.1. Gas Phase. 7612 dx.doi.org/10.1021/ie500309a | Ind. Eng. Chem. Res. 2014, 53, 7611−7621 Industrial & Engineering Chemistry Research Article Table 1. Proximate and Ultimate Analyses of Inner Mongolia Lignite proximate analysis/dry basis ultimate analysis/dry ash-free basis FC VM ash C H O N S 0.5913 0.3556 0.0531 0.8355 0.0480 0.0995 0.0111 0.0059 Table 2. Stable parameters of the first stage syngas item temp. (K) pressure (Pa) inlet gas velocity (m·s−1) Gas Components (%) CO H2 CO2 H2O 3.3. Energy Equation. 3.3.1. Gas Phase. ∂ (αgρg hg ) + ∇(αgρg Ughg ) ∂t ∂p = −αg + tg : ∇Ug − ∇qg⃗ + Sg + Q sg + Ssghsg ∂t value 1573.15 101 325 0.5153 where hg is the specific enthalpy for gas phase; q⃗g is the heat flux for the gas phase; Sg is the source term; Qsg is the heat exchange between gas and solid phases; hsg is the enthalpy between phases. 3.3.2. Solid Phase. 34.07 27.94 16.52 21.47 ∂ (αgρg Ug) + ∇(αgρg UgUg) = −αg∇p + K sg(Us − Ug) ∂t + (∇τg) + αgρg g + SgsUs ∂p ∂ (αsρs hs) + ∇(αsρs Uh + ts: ∇Us − ∇qs⃗ + Ss s s) = − αs ∂t ∂t (3) + Q gs + Sgshgs where Ksg is the momentum exchange coefficient between gas and solid phases. The Syamlal−O’Brien approach15 is adopted for drag force model, and the momentum exchange coefficient is given by the formula of K sg = 3αsαlρl 4vr2,sds ⎛ Re ⎞ C D⎜⎜ s ⎟⎟|vs⃗ − v l⃗ | ⎝ vr ,s ⎠ ⎞2 4.8 ⎟ Res/vr ,s ⎟⎠ ∂ (αgρg Yi ) + ∇(αgρg UgYi ) = −∇αgJi + αR g, i + R s, i ∂t (4) (5) vr ,s = 0.5(A − 0.006Res (0.006Res)2 + 0.12Res(2B − A) + A2 ) ps = αsp* (6) The relative Reynolds number is shown as Res = where p* is expressed by an empirical power function ρg ds|vs⃗ − vg⃗ | μg p* = 1025(αs − αg*)10 (7) (8) 3.2.2. Solid Phase. μs = ∂ (αsρs Us) + ∇(αsρs UU s s) = − αs∇p − ∇ps + ∇· τs + αsρs g ∂t + SsgUs p sin ϕ 2 I2D (15) where ϕ is the angle of internal friction and I2D is the second invariant of the strain rate tensor. 3.6. Heat Transfer. The interphase heat transfer is considered as a function of temperature difference between phases. (9) where SsgUs is the momentum transfer from solid particle to gas phase. ⎛ 2 ⎞ τs = αsμs (∇Us − ∇UsT ) + αs⎜λs − μs ⎟∇UI s ⎝ 3 ⎠ (14) The solid volume fraction for the solid phase is close to the packing limit due to the fact that the solid phase flow is a dense flow in the fixed-bed. The generation of shear stress is mainly due to friction between particles. Therefore, only frictional viscosity is considered for the shear stress. The shear stress can be showed as τ g is the viscous stress tensor and is expressed as follows: ⎛ 2 ⎞ τg = αgμg (∇Ug − ∇UgT ) + αg ⎜λg − μg ⎟∇UgI ⎝ 3 ⎠ (13) where Rg,i and Rs,i represent the net rate of production of component i by homogeneous reaction, and the net rate by heterogeneous reaction, respectively. Ji is the diffusion flux of species i, resulting from concentration gradients. 3.5. Solid-Phase Stress. The solid-phase stress is composed by the solid pressure and shear stress. As for the fixed-bed gasification process, most of solid phase is in packed condition. The solid stress arises because of Coulomb friction between particles in enduring contact. This kind of granular materials can be treated as plastic flow. The Schaeffer model17 is used for the solid pressure, in which where vr,s is the correlation for the solid phase terminal velocity and expressed as follows: + (12) The heat exchange between phases must comply with the local balance condition Qgs = −Qsg. 3.4. Species Transport Equation. The species transport equation for the mass fraction Yi of gas phase is given by The drag coefficient is used by the formula proposed by Dalla Vallue.16 ⎛ C D = ⎜⎜0.63 − ⎝ (11) Q gs = hgs(Tg − Ts) (10) 7613 (16) dx.doi.org/10.1021/ie500309a | Ind. Eng. Chem. Res. 2014, 53, 7611−7621 Industrial & Engineering Chemistry Research Article Table 3. Coefficients of Each Devolatilization Gas α1 α2 α3 α4 α5 α6 α7 α8 avg. mol wt. 0.2040 0.0589 0.0176 0.0687 0.3919 0.2287 0.0096 0.0206 16.852 where hgs = hsg is the heat transfer coefficient between gas and solid phases; Tg is the gas temperature; Ts is the solid temperature. The heat transfer coefficient is related to the Nusselt number for the solid phase and calculated by hsg = (17) where κg is the heat conductivity coefficient of gas phase and Nus is the solid phase Nusselt number and expressed by the correlation proposed by Guun18 Nus = (7 − 10αg + 5αg2)(1 + 0.7Res0.2Prg0.33) (18) (R5) 4. REACTION MODELS When the first stage syngas temperature and composition reaches stable, the second stage fixed-bed gasification reaction could be started. According to practical experiment condition, the initial value of lump coal temperature in the second stage is assumed at 100 °C before the syngas flows to the second stage. Therefore, the lump coal is regarded as dried lignite, and its composition is considered to be a mixture of volatile matter, carbon, and ash. The reaction model can be subdivided into devolatilization, heterogeneous reaction, and homogeneous reaction. 4.1. Devolatilization. When the high temperature syngas from first stage flows to the coal particle in the fixed-bed, the devolatilization reaction will be carried out immediately. The nitrogen element in coal is assumed to be converted to N2 entirely during the devolatilization. The sulfur element is assumed to H2S, and tar is simplified as the benzene. The composition of the devolatilization gas is calculated by a mathematical model based on the element balance.19 The coefficients of each devolatilization gas are listed in Table 3. The secondary devolatilization of tar is ignored in this study due to the small amount produced from the devolatilization process. (R1) Coal → Char + VM Ac = (1/hM , i) + (1/kj) (20) 6 y d p char (21) where Ychar is the carbon volume fraction and dp is the particle diameter. The chemical reaction rate can be expressed by the Arrhenius equation kj = kj ,0T b e−Ea / RT (22) The heterogeneous reaction kinetic parameters are listed in Table 4.23 Table 4. Gas−Solid Reaction Kinetic Parameters kj (m·s−1·K−b) Ea (J·kmol−1) b 19.4 208 2083 2.36 × 10 2.40 × 108 2.30274 × 108 1 1 0 CO2 H2O CH4 Di,mix is the diffusion coefficient of component i; Sc is the Schmidt number. 4.3. Homogeneous Reaction. The main gas phase species includes CO, H2, CO2, and H2O, so the water−gas shift reaction is only considered for homogeneous reaction. (R2) The devolatilization rate is expressed by a single-step global reaction equation, and its kinetic parameters can be consulted from ref 20. k wg CO + H 2O ← → CO2 + H 2 −(E / RT ) ρV (19) (R6) The reaction rate can be expressed by the following equation:24 −1 where the pre-exponential factor A1 = 1.1 × 10 s , and the activation energy E = 8.86 × 107 kJ·kmol−1. 4.2. Heterogeneous Reaction. In the gasification reaction region of second stage, three gas−solid heterogeneous reactions are considered: carbon−CO2, carbon−H2O, and carbon−CH4. 5 8 The gas−solid mass transfer is determined by the formula of Di ,mix hM, i = (2 + 1.1Sc1/3Re 0.6) dp (23) VM → α1CH4 + α2CO + α3CO2 + α4 TAR + α5H 2 + α6 H 2O + α7 H 2S + α8 N2 Ac ci ,bulk where Ac is the specific surface area and shown as follows; ci,bulk is the gas concentration in the bulk; hM,i is the mass transfer coefficient for solid to gas phase; and kj is the chemical reaction rate. where Res is the relative Reynolds number for the solid particle and Prg is the Prandtl number for gas phase. C + CO2 → 2CO C + 2H 2 → CH4 rj = + (1.33 − 2.4αg + 1.2αg2)Res0.7Prg0.33 k = A1 e (R4) The reaction rate of R5 is significantly slower than R3 and R4. The overall reaction rate should consider the chemical reaction rate and mass transfer between gas and solid phases. It is supposed that the contribution of pores for the gasification reaction is small,21 and then, it can be considered that the gas− solid reaction is conducted on the particle surface. The overall reaction rate can be expressed as follows:22 6κgαsαgNus ds2 C + H 2O → CO + H 2 (R3) ⎛ CCO2C H2 ⎞ R wg = εk wg ⎜CCOC H2O − ⎟ KE ⎠ ⎝ (24) k wg = 2.78 exp( −1513/Tg) (25) KE is the equilibrium constant for water−gas shift reaction. 7614 dx.doi.org/10.1021/ie500309a | Ind. Eng. Chem. Res. 2014, 53, 7611−7621 Industrial & Engineering Chemistry Research KE = 0.0265exp(3966/Tg) Article The concentration of CO shows a significant decrease during the initial stage, followed by a maximum value, and then turns to be decreased gradually. The concentration of H2 first increases and then decreases throughout the gasification process, while the concentration of CO2 shows an opposed variation trend of H2. The concentration of CH4 increases dramatically during the initial stage, and becomes to close to nil after 40 min, showing that in the first 40 min the main reaction is dominated by devolatilization. At the end of the gasification reaction, the concentrations of CO, H2, and CO2 are close to the initial values. The comparison of gas flow rate between simulation and experiment is showed in Figure 2b. It also turns to be a reasonable agreement between two values, which can further validate the accuracy of the gasification model. The flow rates of H2 and CH4 rapidly increase in the initial stage, while the flow rate of CO increases smoothly. This is mainly due to the devolatilization process in the initial stage, in which a large amount of H2 and CH4 is emitted and dilutes the concentration of CO. Therefore, this can reasonably explain the drop of CO concentration in the initial stage in Figure 2a. The curve of H2O flow rate shows the amount of H2O increases due to devolatilization in the beginning and reaches the maximum value after 10 min. Then, it decreases gradually and reaches the minimum value after 50 min, showing that the gasification reaction starts to play the main role in the time period. The slight decreasing of CO2 flow rate also shows the CO2 is reacted with coal char simultaneously. Finally, the flow rates of H2O and CO2 increase to the near initial values. 5.2. Analysis of Simulation Results. 5.2.1. Velocity and Temperature Distributions. Figure 3a shows the distributions of gas velocity at the moment of 60, 180, and 300 min. It can be found that the gas velocity rapidly increases, when hot temperature syngas flows through the fixed-bed layer. The superficial gas velocity increases because of the porous region of second stage, and also, the total gas volume flow rate increases due to gas generated from gasification reactions. As the gasification reaction proceeds, the influence of gas velocity from the fixed-bed layer becomes weaker as a result of the mass decreasing of lump coal in the second stage. Furthermore, the syngas is more inclined to flow from the center during the later reaction process, which is attributes to that the velocity distribution in the inlet syngas is decreasing from the center to the tube wall,25 resulting a smaller pressure drop resistance at center in the coal layer ascribed to a more enhanced gasification reaction process than that nearby wall surface. Figure 3b shows the distributions of gas temperature at the moment of 60, 180, and 300 min. After the syngas flowing through the coal layer, a significant temperature drop has taken on comparing with the initial gas temperature. The outlet gas temperatures are 1372.1, 1422.3, and 1507.4 K at 60, 180, and 300 min, respectively. This shows that the heat energy from the inlet syngas used in the gasification is decreased due to the decreasing of coal mass during the gasification process. It also can be found that the interface of the temperature drop between the coal layer and inlet syngas also shows the surface of coal char in lump coal, and the depth area for the high temperature in the fixed-bed layer increases with gasification reaction as a result of the absence of carbon in the upper layer. In the later stage of reaction, the gas temperature distribution in the coal layer shows a more obvious characteristic of deep at center and shallow nearby wall surface. (26) 4.4. Solution Method and Boundary Condition. The unsteady two-dimensional model is used to simulate the second stage fixed-bed gasification process, and the total number of the grids is chosen as 24 800 after the grid independence test. The convective terms in transport equations are discretized with second order upwind scheme. A phase coupled SIMPLE scheme is adopted for the velocity−pressure coupling calculation. The adiabatic wall is chosen for the wall heating condition. The gas conditions from first stage based on experiment results is used as the inlet condition, and pressure outlet is set as outlet. The calculated step time is controlled as 1 s, and the total physical time is 300 min. 5. RESULTS AND DISCUSSION 5.1. Comparison between Simulated and Experimental Results. The settings in the simulation model are based on experimental operating conditions. The mass of coal in the second stage is 1400 g, and the particle size range is 17.5 mm instead of the 15−20 mm in experiment for the near-spherical shape of raw coal particle. The true density of Inner Mongolia lignite is 1200 kg·m−3 and 707 kg·m−1 for the bulk density. The comparison of gas concentration at outlet between simulation and experiment is showed in Figure 2a based on H2O free, which exhibits a good agreement with the experimental values7 and an identical variation trend during the gasification process. Figure 2. Comparisons of outlet gas concentration and flow rate between simulation and experiment. 7615 dx.doi.org/10.1021/ie500309a | Ind. Eng. Chem. Res. 2014, 53, 7611−7621 Industrial & Engineering Chemistry Research Article Figure 3. Distributions of gas velocity and temperature at the moment of 60, 180, and 300 min. 5.2.2. Gas Concentration Distribution. Figure 4 shows the variation of average gas concentrations along the axial position at the moment of 60, 180, and 300 min. It can be seen that the gas concentration outside the fixed-bed layer basically remains unchanged. The concentrations of CO and H2 increase, while the concentrations of CO2 and H2O show downward trend when the syngas flows through the fixed-bed layer. Furthermore, the declining rate of H2O is much faster than CO2, revealing the conversion rate of H2O is higher than that of CO2 in the fixed-bed gasification reaction. This is because that the chemical reaction rate of C + H2O is faster than that of C + CO2, and the H2O concentration in syngas is higher than that of CO2, resulting a higher partial pressure of H2O than that of CO2. As the gasification reaction proceeds, the changes of gas concentration above and below the fixed-bed layer become smaller, coupling with a smaller temperature drop, which indicates a weaker endothermic gasification reaction. At the reaction time of 300 min, there is no significant change of gas concentration in the upper lump coal layer due to much consuming of carbon and remaining by ash layer. 5.2.3. Solid Components Distribution. Figure 5 shows the solid mass and mass fraction of each component in coal layer Figure 4. Average gas concentrations along the axial position at the moment of 60, 180, and 300 min. with reaction time. It can be found that volatile gas has been released out within 40 min, which reveals that the devolatilization process has already been performed in this time period. The mass of char is almost unchanged within 20 min, indicating a weak gasification reaction during the initial stage, and then is attenuated significantly and eventually approaches nil. As inert matter, the ash mass amount remains unchanged in the coal layer during the whole gasification process, while its mass fraction gradually increases as a result of the consumption of the other two matters. Figure 6 shows the mass fractions of each component in coal layer along the axial position at the moment of 60, 180, and 300 min. It can be seen that the mass fraction of char in the upper layer is lower than that at bottom, which indicates that char in the coal layer is consumed from the top toward downward during the fixed-bed gasification process. As a result, ash will become the main material in the upper layer, when the char is reacted gradually. As the gasification reaction proceeds, the 7616 dx.doi.org/10.1021/ie500309a | Ind. Eng. Chem. Res. 2014, 53, 7611−7621 Industrial & Engineering Chemistry Research Article Figure 5. Solid mass and mass fraction of each component in coal layer. content of char in the upper layer decreases significantly, resulting a moving-down process of gasification reaction zone, which can also be observed from the declining rate of solid temperature. 5.2.5. Carbon Conversion and Conversion Rate. Figure 7 shows the variation of carbon conversion and conversion rate in the fixed-bed during the gasification process. The carbon conversion gradually increases as the reaction proceeds, while the conversion rate shows a bimodal distribution. According to the variation of conversion rate curve, the gasification reaction process can be divided into two zones, namely Zone 1 and Zone 2. In Zone 1, the first peak of conversion rate is formed in 50 min with a faster conversion rate, indicating that the carbonaceous matter in volatiles is released to the gas block rapidly. In Zone 2, a secondary peak for conversion rate is formed after 50 min. The conversion rate increases gradually due to a more obvious gasification reaction between coal char and gasifying agent. As the gasification reaction proceeds, the conversion rate of carbon decreases, resulting from the mass decrease of effective carbon in coal layer during gasification reaction.26 It can also be found that the ultimate carbon conversion of simulation is close to the experimental data, which is 97.9%. 5.3. Effects of Feedstock Parameters. 5.3.1. Various Coal Amounts. The effect of various coal amounts for the second stage on gasification efficiency is also investigated in the proposed model. The added coal amount is 1000 g, 1200 g, 1400 g, and 1600 g, respectively. Figure 8a shows the gas Figure 6. Mass fractions of each component in coal layer along the axial position at the moment of 60, 180, and 300 min. Figure 7. Carbon conversion and conversion rate in the fixed-bed. temperature at the outlet as the reaction with various coal amounts proceeds. The outlet gas temperature shows a temperature fluctuation in the initial stage, which is caused by a distinct heat recovery through devolatilization reaction. A 7617 dx.doi.org/10.1021/ie500309a | Ind. Eng. Chem. Res. 2014, 53, 7611−7621 Industrial & Engineering Chemistry Research Article Figure 9. Effective gas concentration at outlet with various coal amounts. be the main factor for gasification reaction in the condition of less temperature difference. In the case of 1600 g, the height of the coal layer and the carbon content are higher than the other cases, which will increase the surface area for gasification reaction and lengthen the flow path in coal layer. This results in the highest effective gas concentration in Zone 2 of the 1600 g coal amount. Figure 10 shows the gas−solid reaction rates of C + CO2 and C + H2O in the reaction with various coal amounts. It can be Figure 8. Average temperatures of outlet gas and coal layer with various coal amounts. larger coal amount shows a more obvious heat recovery from high-temperature syngas. As the gasification reaction proceeds, the gas temperature at the outlet approaches to the initial value (1573.15 K), indicating a gradual weaker effect of endothermic reaction at the later gasification process. Figure 8b shows the average temperature of coal layer with various coal amounts. It can be found that the temperature of 1000 g case is always higher than other three cases, especially in the first 50 min, indicating that a smaller coal amount results in a faster heating rate. After the reaction time of 50 min, the increase of temperature becomes slowly and the temperature difference among the four cases becomes smaller. The effective gas concentration is defined as the concentration of CO + H2 on dry basis. Figure 9 shows the effective gas concentration at the outlet with various coal amounts. It can be seen that for all the cases the effective gas concentration shows a significant decrease in the initial stage due to the dilution by a large amount of CH4 released from devolatilization process. After about 60 min, the effective gas concentration gradually decreases as the gasification reaction proceeds, due to the reducing of carbon content in the coal layer. With various coal amounts, the effective gas concentration of 1000 g case is a little higher than that of larger coal amounts in the first 50 min, while effective gas concentration for 1600 g case takes on the highest value after 50 min. This is because a faster heating rate of the 1000 g case in the initial stage will take the lead in the gasification reaction process. Subsequently, the contact surface area between high temperature syngas and lump coal turns to Figure 10. Gas−solid reaction rates of C + CO2 and C + H2O with various coal amounts. seen that the gas−solid reaction rates exist an obvious trend of increasing first and then decreasing. The gas−solid reaction rate of the 1000 g case is much higher than other three cases in the initial gasification process. As the reaction proceeds, the smaller coal amount shows a much more rapid decreasing trend on the gas−solid reaction rate due to the consuming of carbon content. Furthermore, it is also confirmed that the reaction rate of C + H2O is much higher than that of C + CO2, revealing that the main gasification reaction in the gasifier is C + H2O. Figure 11 shows the carbon conversion and conversion rate with various coal amounts. The carbon conversion of the 1000 g case is always the highest during the gasification process, and the ultimate value is 99.6%. The carbon conversion of 1600 g case turns to be lowest (96.9%) after 300 min reaction time. Therefore, the carbon conversion will be reduced within the limited reaction time for a large coal amount in the second stage. In the viewpoint of conversion rate, a smaller coal 7618 dx.doi.org/10.1021/ie500309a | Ind. Eng. Chem. Res. 2014, 53, 7611−7621 Industrial & Engineering Chemistry Research Article Figure 11. Carbon conversion and conversion rate with various coal amounts. amount shows a more rapid devolatilization and gasification reacting rate during the reaction process. 5.3.2. Various Coal Particle Sizes. The effect of various coal particle sizes on gasification efficiency is studied with the coal mass keeping at 1400 g. The coal particle size for the second stage is ranged from 5−10 mm, 10−15 mm, 15−20 mm, and 20−25 mm, replaced by the intermediate uniform diameter in the simulation. Figure 12 shows the temperature of the outlet gas and the coal layer as the with various coal particle sizes proceeds. It can be seen that the change of particle sizes has slight effect on temperature comparing with that caused by coal amounts. The outlet gas temperature of 5−10 mm case is a little lower than other three cases in the first half period and shows a distinct increase and exceeds other three cases in the later period, which reveals that a more significant heat recovery efficiency for smaller particle size in initial stage and slight endothermic effect at later stage. In Figure 12b the temperature difference of coal layer with various coal particle sizes is small, so it can be obtained that the reaction rate difference is not consisted in the chemical reaction rate. Figure 13 shows the effective gas concentration at outlet with various coal particle sizes. It can be found that the effective gas concentration of 5−10 mm case increases immediately after the initial drop, which shows that the gasification reaction of 5−10 mm case has already been conducted significantly during the devolatilization process. The effective gas concentration of 5− 10 mm case is the highest one among these four cases during the gasification process, in which the maximum value increases by 3.9 percentage point to 82.9% compared with the initial gas concentration. Figure 14 shows the gas−solid reaction rates of C + CO2 and C + H2O with various coal particle sizes. It can be found that the gas−solid reaction rate for 5−10 mm case increases immediately after the reaction time of 5 min, while after 25 min for 20−25 mm case. Therefore, with the increase of coal particle size the time for gasification reaction starting will be prolonged and the gas−solid reaction rate will also be decreased. This is because larger particle size contains a larger heat transfer resistance from the particle surface to inside. Furthermore, the specific surface area of bed layer decreases with the increase of coal particle diameter,27 resulting in a lower reaction rate between high temperature syngas and lump coal as the increasing of coal particle size. Figure 15 shows the carbon conversion and conversion rate with various coal particle sizes. It can be seen that the changing of particle size has a significant Figure 12. Average temperatures of outlet gas and coal layer with various coal particle sizes. Figure 13. Effective gas concentration at outlet with various coal particle sizes. effect on carbon conversion and conversion rate. The carbon conversion declines obviously with the increase of coal particle size. The carbon conversion of 5−10 mm case has already been 87.7% after 120 min reaction time, while 58.7% for 20−25 mm case. As the aspect of conversion rate, the lump coal of 5−10 mm diameter shows a much more rapid devolatilization rate than others, followed by taking the lead of the gasification 7619 dx.doi.org/10.1021/ie500309a | Ind. Eng. Chem. Res. 2014, 53, 7611−7621 Industrial & Engineering Chemistry Research Article a higher coal layer and more carbon content. The largest coal amount (1600 g) leads to the lowest carbon conversion, which is 96.9% within 300 min gasification reaction time. (3) A significant heat recovery is observed for 5−10 mm diameter from the outlet gas temperature during the first half period. An increase of particle size results in the reducing of effective gas concentration and gas−solid reaction rate during the gasification process, which is caused by a smaller specific surface area and relative lower heat transfer from particle surface to inside. The response time for gasification reaction starting is proposed to be prolonged with the increase of particle size, and the carbon conversion is also reduced. A smaller coal particle size is propitious to the gasification reaction in the fixed-bed gasification process. Figure 14. Gas−solid reaction rates of C + CO2 and C + H2O with various coal particle sizes. ■ AUTHOR INFORMATION Corresponding Author *Telephone: +86-021-64252522. Fax: +86-021-64251312 Email: [email protected]. Notes The authors declare no competing financial interest. ■ ■ ACKNOWLEDGMENTS This work is supported by foundation of Shanghai outstanding academic leader (No. 08XD1401306). Figure 15. Carbon conversion and conversion rate with various coal particle sizes. reaction process. Thus, a smaller coal particle size is propitious to the gasification reaction in the second stage. 6. CONCLUSIONS The second stage fixed-bed gasification reaction in the combined two-stage gasifier, which recycles the sensible heat from first stage syngas to improve the energy utilization, is investigated by numerical simulation in this study. Results can be summarized as follows: (1) The simulated results show a reasonable agreement with the experimental data in terms of gas concentration, gas flow rate, and carbon conversion. The gas velocity increases and gas temperature descends when flowing through the fixed-bed layer, and the concentrations of CO and H2 in syngas increase by gasification reaction. As the gasification reaction proceeds, the temperature difference and changes of gas concentration become weaker comparing with initial values. The volatile matter in coal has been released out within 40 min, and the coal char in coal layer is consumed from the top toward the bottom gradually. (2) An increase of coal amount leads to the decline of outlet gas temperature and coal layer temperature in the second stage. A smaller coal amount turns to be a faster heating rate during the initial stage, resulting a little higher gas− solid reaction rate and effective gas concentration. During the later gasification stage, a larger coal amount shows a relative higher effective gas concentration due to NOMENCLATURE A1 = pre-exponential factor Ac = specific surface area (m−1) ci,bulk = gas concentration in the bulk CD = drag coefficient dp = particle diameter (m) Di,m x = diffusion coefficient of component i (m2·s−1) E = activation energy (kJ·kmol−1) hg = specific enthalpy for gas phase (J·kg−1) hM,i = mass transfer coefficient (m·s−1) hsg = heat transfer coefficient between gas and solid phases (W m−2·K−1) I2D = second invariant of the strain rate tensor k = turbulent kinetic energy kj = chemical reaction rate KE = equilibrium constant of water shift reaction Ksg = momentum exchange coefficient between gas and solid phases (kg m−3·s−1) Nus = Nusselt number of the solid phase ps = solid pressure Prg = Prandtl number of gas phase qg⃗ = heat flux of the gas phase (W·m−2) Qsg = heat exchange between gas and solid phases (W·m−3) Res = Relative Reynolds number S = mass source term Sc = Schmidt number Tg = gas temperature (K) Ts = solid temperature (K) U = instantaneous velocity (m·s−1) vr,s = terminal velocity of solid phase (m·s−1) Ychar = carbon volume fraction Greek Symbols α = volume fraction αi = coefficients of each devolatilization gas 7620 dx.doi.org/10.1021/ie500309a | Ind. Eng. Chem. Res. 2014, 53, 7611−7621 Industrial & Engineering Chemistry Research Article ρ = density (kg·m−3) τ g = viscous stress tensor (Pa) μs = shear stress (Pa·s) ϕ = angle of internal friction κ = thermal conductivity (W·m−1·K−1) ε = turbulent dissipation rate (m2·s−3) ■ (22) Maria, S.; Manuel, P.; Fabrizio, B.; Alberto, B. Simulated Moving Bed Technology Applied to Coal Gasification. Chem. Eng. Res. Des. 2010, 88, 465−475. (23) Brown, B. W.; Smoot, L. D.; Smith, P. J.; Hedman, P. O. Measurement and Prediction of Entrained Flow Gasification Processes. AIChE J. 1988, 34, 435−446. (24) Blasi, C. D. Modeling Wood Gasification in a Countercurrent Fixed-Bed Reactor. AIChE J. 2004, 50, 2306−2319. (25) Chua, L. P.; Lua, A. C. Measurements of a Confined Jet. Phys. Fluids 1998, 10, 3137−3144. (26) Wang, Y. F.; Jin, W. L.; Huang, T. H.; Zhu, L. C.; Wu, C. Q.; Yu, G. S. Characteristics of Alkali and Alkaline-Earth Metals for the Catalytic Gasification of Coal Char in a Fixed-Bed Reactor. Energy Technol. 2013, 1, 544−550. (27) Luo, S. Y.; Xiao, B.; Hu, Z. Q.; Liu, S. M.; Guan, Y. W.; Cai, L. Influence of Particle Size on Pyrolysis and Gasification Performance of Municipal Solid Waste in a Fixed Bed Reactor. Bioresour. Technol. 2010, 101, 6517−6520. REFERENCES (1) Christopher, H.; Maarten, J. B. Gasification, 2nd ed.; Gulf Professional Publishing: San Francisco, 2008. (2) Emanuele, M.; Thomas, K.; Michiel, C.; Stefano, C.; Daniel, J. Shell Coal IGCCS with Carbon Capture: Conventional Gas Quench vs Innovative Configurations. Appl. Energy 2011, 88, 3978−3989. (3) Preston, W. E. The Texaco Gasification Process in 2000 Startups and Objectives. Presented at the Gasification Technologies Conference, San Francisco, CA, Oct. 2000. (4) Yu, Z. H; Liu, H. F; Wang, F. C.; Gong, X.; Yu, G. S.; Shi, J. M. Equipment of Combined Two Stage Gasification Technology. Chinese Patent No. 00111437.9, 2000. (5) Yu, Z. H.; Wang, Y. F.; Yu, G. S.; Wang, F. C.; Chen, X. L.; Dai, Z. H.; Guo, X. L.; Liang, Q. F.; Li, W. F.; Gong, X.; Liu, H. F. Gasification Equipment and Application of Combined Two Stage Gasification Technology Based on Heat Recovery and Washer. Chinese Patent No. 200610119514.4, 2007. (6) Doctor, R. D.; Molburg, J. C.; Thimmapuram, P.; Berry, G. F.; Livengood, C. D.; Johnson, R. A. Gasification Combined Cycle: Carbon Dioxide Recovery, Transport, and Disposal. Energy Convers. Manage. 1993, 34, 1113−1120. (7) Huang, T. H.; Wang, Y. F.; Jiao, Y. T.; Jin, W. L.; Su, P. Hot-State Experiment of High-Performance Two-Stage Combined Coal Gasification. CIESC J. 2010, 61, 2924−2930 In Chinese. (8) Jin, W. L; Wang, Y. F.; Ren, Y. Z. Hot-State Experimental Research on a Novel Coal-Based Two-Stage Gasification Process. Presented at the 30th International Pittsburgh Coal Conference, Beijing, China, Sept. 2013. (9) Blasi, C. D. Dynamic Behavior of Stratified Downdraft Gasifiers. Chem. Eng. Sci. 2000, 55, 2931−2944. (10) Yang, W. H.; Ponzio, A.; Lucas, C.; Blasiak, W. Performance Analysis of a Fixed-Bed Biomass Gasifier Using High-Temperature Air. Fuel Process. Technol. 2006, 87, 235−245. (11) Murgia, S.; Vascellari, M.; Cau, G. Comprehensive CFD model of an air-blown coal-fired updraft gasifier. Fuel 2012, 101, 129−138. (12) Wu, Y. S.; Zhang, Q. L.; Yang, W. H.; Blasiak, W. TwoDimensional Computational Fluid Dynamics Simulation of Biomass Gasification in a Downdraft Fixed-Bed Gasifier with Highly Preheated Air and Steam. Energy Fuel 2013, 27, 3274−3282. (13) Jones, W. P.; Launder, B. E. The Prediction of Laminarization with a Two-Equation Model of Turbulence. Int. J. Heat Mass Transfer 1972, 15, 301−314. (14) ANSYS, Inc. ANSYS FLUENT 12.0 Theory Guide; ANSYS, Inc.: Canonsburg, PA, 2009. (15) Syamlal, M.; O’Brien, T. J. Computer simulation of bubbles in a fluidized bed. AIChE Symp. 1989, 85, 22−31. (16) Dalla valle, J. M. Micromeritics; Pitman: London, 1948. (17) Schaeffer, D. G. Instability in the Evolution Equations Describing Incompressible Granular Flow. J. Differ. Equations 1987, 66, 19−50. (18) Gunn, D. J. Transfer of Heat or Mass to Particles in Fixed and Fluidized Beds. Int. J. Heat Mass Transfer 1978, 21, 467−476. (19) David, M. Mathematical Models of the Thermal Decomposition of Coal. Fuel 1983, 62, 534−539. (20) Wen, C. Y.; Chaung, T. Z. Entrainment Coal Gasification Modeling. Ind. Eng. Chem. Process Des. Dev. 1979, 18, 684−695. (21) Nagpal, S.; Sarkar, T. K.; Sen, P. K. Simulation of Petcoke Gasification in Slagging Moving Bed Reactors. Fuel Process. Technol. 2005, 86, 617−640. 7621 dx.doi.org/10.1021/ie500309a | Ind. Eng. Chem. Res. 2014, 53, 7611−7621
© Copyright 2025 Paperzz