You are on page 1of 350

A Theoretical and Experimental Study on a Stratified Downdraft Biomass Gasifier

San Shwe Hla

Submitted in total fulfilment of the requirements for the degree of Doctor of Philosophy

January 2004

Department of Civil and Environmental Engineering The University of Melbourne

Abstract

A transient model has been developed for a stratified downdraft wood gasifier by applying a two-step pyrolysis mechanism in which primary tars (oxygenates) are initially formed and then cracked into secondary tars (hydrocarbons) and other combustible gases. The model describes the complex physical and chemical processes taking place in the reactor by the use of mass and energy balances, together with information about rates of chemical reactions and physical transport processes. The computer model is capable of predicting the primary and secondary tar profile as well as the gas composition profile and temperature profile along the axis of the gasifier. The model results show that the concentration of carbon monoxide and hydrogen, the two main combustible components in the gas produced, is very low initially but gradually increases to a steady value. The model showed that 20-25 min was required to reach steady conditions. Model results also showed that the average range of gravimetric tar content in the gas is between 20 to 200 mg/Nm for the range of input parameters investigated provided the gasifier bed below the ignition port level is initially charged with charcoal. Reaction zone stability was also investigated and under conditions where reaction zones were steady the specific gasification rate was found to be 290-296 kg/m2hr (dry basis), a value which is close to previously published experimental results for an open core biomass gasifier. In order to obtain a detailed understanding of the models sensitivity to changes in key parameters, a systematic analysis of the effect of varying individual operational and model parameters on predictions of temperature profile, outlet gas tar content, gas composition and reaction zone stability was carried out. The results show that a higher air supply rate increases the combustion zone temperature and reduces outlet tar levels. The final amount of condensable tars in the producer gas was found to be affected by the following model parameters: solid-to-gas heat transfer coefficient, mass transfer coefficient, tar cracking kinetics and effective thermal conductivity. Reaction zone stability was found to be significantly affected by changes in effective thermal conductivity values.

ii

An experimental downdraft gasifier was designed, fabricated and tested, and the experimental data obtained from it was used to validate model predictions. The performance of the experimental gasifier was determined for various supplied airflow rates. Wood blocks in two different sizes were used as reactor fuel. It was found that the gasifier could be operated over a fairly wide range of airflow rates of from 7.5 to 27.6 kg/hr, with corresponding energy outputs ranging from 54 -168 MJ/hr. An average cold gas efficiency of 70% was obtained. Operating the reactor with wood fuel alone produced an outlet gas with a final tar content that was about 20 times higher than the average values of tar content measured in gas produced when using an initial charge of charcoal. Transient temperature profiles, outlet gas compositions, and outlet tar contents predicted by the model were in generally good agreement with experimental values. The predicted rate of increase in the concentrations of two combustible components (H2 and CO) inside the gasifier, however, was much greater than that observed experimentally. There was good agreement between the experimentally observed rates at which the reaction zones moved within the gasifier at different airflow rates and those predicted using the model. By considering two important factors, namely the cold gas efficiency profile and the tar profile along the axis of the gasifier, it was determined that the optimum reaction zone length is 425 mm. Prospects for obviating the need to provide an initial charge of charcoal were investigated using the model. The model was also used to investigate the concept of two-stage gasification. This suggests that provided secondary air can be distributed uniformly across the reactor at a suitable distance downstream of the main oxidation zone, the gas produced will have a lower tar content and a higher cold gas efficiency than those produced in a single-stage gasifier for the same total air supply rate.

iii

Declaration

This is to certify that (i) (ii) (iii) the thesis comprises only my original work towards the PhD; due acknowledgement has been made in the text to all other material used; the thesis is less than 100,000 words in length, exclusive of tables, maps, bibliographies and appendices.

San Shwe Hla January 2004

iv

Acknowledgements
I wish to express my gratitude to those who have contributed guidance, encouragement, and assistance in the completion of this research. First of all I would like to extend my sincere thanks and appreciation to my supervisors, Dr. Lu Aye, Dr. Mike Connor and Associate Professor Don Stewart for their guidance, suggestions, and encouragement throughout the course of this study. I also would like to acknowledge their invaluable comments which have significantly improved my manuscript. In particular, I wish to express my deepest gratefulness to Dr. Mike Connor for his invaluable participation as a co-supervisor during the tough years. I gratefully acknowledge the University of Melbourne for financial support which enabled me to complete my PhD study. The experimental work has been partially funded by the Open Society Institute, and this is gratefully acknowledged. I also would like to express my gratitude to my supervisors for their financial support of my experimental work and for help with payment of my tuition fees for the last three months of my candidature. I am also sincerely grateful to Dr. Paul Fung and Mr. Soo Ng (Energy and Recycling Team, Forestry and Forest Products, CSIRO) for their enthusiastic cooperation and in-kind contributions to the experimental work. I would like to thank to all my co-postgraduate friends who in one way or the other have helped me and encouraged me throughout the study, most especially to Chat, Eddy and Wirachai. My gratitude is expressed to Mrs. Fiorella Chiodo and other staff members in the Department of Civil and Environmental Engineering for their cooperation and assistance. I am deeply indebted to my beloved parents for their spiritual encouragement and moral support in the pursuit of my study at the University of Melbourne. Finally, I express my sincerest thanks to my beloved Latt for her constant encouragement and deep understanding throughout the years of struggle.

Contents

Abstract Declaration

................................................................................................................ ii ............................................................................................................... iv

Acknowledgements ....................................................................................................... v List of tables .............................................................................................................. xii List of figures ............................................................................................................. xiii Nomenclature ............................................................................................................. xix Chapter 1 Introduction................................................................................................. 1 1.1 1.2 1.3 1.4 1.5 Background ...................................................................................................... 1 Problem statement............................................................................................ 2 Aim of the study............................................................................................... 4 Objectives of the study..................................................................................... 4 References........................................................................................................ 6

Chapter 2 Literature review ........................................................................................ 8 2.1 2.2 2.3 Introduction...................................................................................................... 8 Different models of fixed-bed gasification ...................................................... 8 Reaction rates................................................................................................. 10 2.3.1 Drying...................................................................................................... 10 2.3.2 Pyrolysis .................................................................................................. 10 2.3.3 Heterogeneous gas-char reactions ........................................................... 13 2.3.4 Homogeneous gas-phase reactions.......................................................... 16 2.4 Heat and mass transfer processes in packed beds.......................................... 23 2.4.1 Heat transfer ............................................................................................ 24

vi

2.4.2 Mass transfer ........................................................................................... 27 2.4.3 Effective thermal conductivities in packed beds ..................................... 29 2.4.4 Heat loss through the reactor wall ........................................................... 31 2.5 Tar in producer gas ........................................................................................ 33 2.5.1 Tar definition ........................................................................................... 33 2.5.2 Tar quantities as a function of gasifier type ............................................ 34 2.5.3 Tar removal methods............................................................................... 34 2.5.4 Tar removal methods outside the gasifier (Secondary methods) ............ 34 2.5.5 Tar removal methods inside the gasifier (Primary methods) .................. 35 2.5.6 Factors influencing chemical tar conversion in the absence of catalyst..................................................................................................... 36 2.6 Concluding remarks ....................................................................................... 40 2.6.1 Theoretical modelling.............................................................................. 41 2.6.2 Tar reduction ........................................................................................... 41 2.7 References...................................................................................................... 43

Chapter 3 Mathematical model development .......................................................... 50 3.1 3.2 Introduction.................................................................................................... 50 Reaction rates................................................................................................. 51 3.2.1 Drying...................................................................................................... 51 3.2.2 Pyrolysis .................................................................................................. 52 3.2.3 Heterogeneous char combustion and gasification ................................... 56 3.2.4 Homogeneous gas reactions .................................................................... 57 3.3 Gas-particle interchange ................................................................................ 61 3.3.1 Heat and mass transfer coefficients......................................................... 61 3.3.2 Effective thermal conductivity ................................................................ 63 3.4 3.5 Bed-to-wall heat transfer coefficient ............................................................. 64 Thermophysical and transport properties of solid and gas mixture ............... 66 3.5.1 Solid phase............................................................................................... 66 3.5.2 Gas mixture ............................................................................................. 66 3.6 Conservation of mass..................................................................................... 68 3.6.1 Solid phase component............................................................................ 69

vii

3.6.2 Gas phase species .................................................................................... 70 3.7 3.8 3.9 Conservation of energy .................................................................................. 72 The boundary and initial conditions .............................................................. 73 Numerical methods ........................................................................................ 75 3.9.1 Method of discretization.......................................................................... 76 3.9.2 Source-term linearization ........................................................................ 78 3.9.3 Underrelaxation ....................................................................................... 79 3.9.4 Model discretization ................................................................................ 80 3.9.5 Adaptive grid method .............................................................................. 89 3.9.6 Solution procedure................................................................................... 91 3.10 3.11 Conclusions.................................................................................................... 95 References...................................................................................................... 96

Chapter 4 Model results and discussions................................................................ 100 4.1 4.2 Introduction.................................................................................................. 100 Generalised model predictions..................................................................... 101 4.2.1 Transient behaviour of temperatures and gas compositions.................. 101 4.2.2 Temperature profiles ............................................................................. 105 4.2.3 Velocity profiles .................................................................................... 107 4.2.4 Solid densities profiles .......................................................................... 108 4.2.5 Gas composition profiles....................................................................... 110 4.2.6 Reaction rate profiles............................................................................. 114 4.2.7 Stability of reaction zones ..................................................................... 117 4.3 Parametric sensitivity analysis of the model................................................ 119 4.3.1 Airflow rate ........................................................................................... 120 4.3.2 Initial particle size ................................................................................. 122 4.3.3 Initial moisture content.......................................................................... 123 4.3.4 Solid-gas heat transfer coefficient ......................................................... 127 4.3.5 Mass transfer coefficient ....................................................................... 129 4.3.6 Heat and mass exchange area ................................................................ 131 4.3.7 Kinetics of primary pyrolysis ................................................................ 133 4.3.8 Kinetics of secondary pyrolysis............................................................. 133

viii

4.3.9 Kinetic rate of oxidation of secondary tar ............................................. 135 4.3.10 Effective thermal conductivity............................................................ 138 4.3.11 Heat loss through reactor wall ............................................................ 138 4.4 4.5 Conclusions.................................................................................................. 140 References.................................................................................................... 144

Chapter 5 Test rig and experimental procedure.................................................... 146 5.1 5.2 Introduction.................................................................................................. 146 Experimental set-up ..................................................................................... 147 5.2.1 Experimental stratified downdraft gasifier............................................ 147 5.2.2 Thermal insulation................................................................................. 151 5.2.3 Flare and burner..................................................................................... 151 5.2.4 Filter ...................................................................................................... 152 5.2.5 Air blower.............................................................................................. 152 5.2.6 Suction blower....................................................................................... 153 5.3 Measurement systems and instrumentation ................................................. 153 5.3.1 Gas sampling train................................................................................. 153 5.3.2 Measurement of tar content ................................................................... 157 5.3.3 Measurement of gas composition.......................................................... 159 5.3.4 Measurement of temperature................................................................. 159 5.3.5 Measurement of pressure....................................................................... 159 5.3.6 Measurement of supplied airflow rate................................................... 160 5.3.7 Measurement of gas flow rate ............................................................... 160 5.3.8 Measurement of wood consumption rate .............................................. 161 5.3.9 Cold gas efficiency ................................................................................ 161 5.4 Raw materials............................................................................................... 162 5.4.1 Wood blocks.......................................................................................... 162 5.4.2 Red gum charcoal .................................................................................. 164 5.5 5.6 5.7 Experimental procedure ............................................................................... 164 Summary ...................................................................................................... 165 References.................................................................................................... 167

ix

Chapter 6 Experimental results and model validation.......................................... 168 6.1 6.2 Introduction.................................................................................................. 168 Experimental results..................................................................................... 168 6.2.1 Material balance and elemental balance................................................ 169 6.2.2 Temperature........................................................................................... 173 6.2.3 Gas composition .................................................................................... 175 6.2.4 Tar content............................................................................................. 176 6.2.5 Energy output ........................................................................................ 176 6.2.6 Cold gas efficiency ................................................................................ 178 6.3 Comparison of model predictions and experimental results........................ 178 6.3.1 Temperature profiles ............................................................................. 179 6.3.2 Gas composition profiles....................................................................... 182 6.3.3 Tar profile.............................................................................................. 182 6.3.4 Fuel consumption .................................................................................. 185 6.3.5 Air-fuel ratio.......................................................................................... 185 6.3.6 Gas production rate................................................................................ 186 6.3.7 Variation of reaction zone movement with air flow rate and wood block size ............................................................................................... 187 6.4 6.5 Conclusions.................................................................................................. 190 References.................................................................................................... 192

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier................. 193 7.1 7.2 7.3 Introduction.................................................................................................. 193 Optimum length of reaction zones............................................................... 193 Investigation of low tar gas production for the stratified downdraft 7.3.1 Performance of the reactor when using wood fuel only........................ 198 7.3.2 Alternative operating procedures for the stratified downdraft gasifier ................................................................................................... 201 7.3.3 Two different concepts on two-stage gasification................................. 205 7.3.4 Simulation on two-stage gasification .................................................... 207 7.4 Conclusions.................................................................................................. 213

gasifier..................................................................................................................... 197

7.5

References.................................................................................................... 217

Chapter 8 Conclusions and recommendations....................................................... 218 8.1 8.2 8.3 8.4 8.5 8.6 Introduction.................................................................................................. 218 Summary and conclusions for the theoretical study .................................... 218 Summary and conclusions for the experimental study ................................ 222 Summary and conclusions related specifically to tar reduction................... 225 Recommendations for further research works ............................................. 226 Reference ..................................................................................................... 229 Average particle temperature for use in primary pyrolysis process calculations .......................................................................... 230 Appendix B Properties of gas species.................................................................... 234 Appendix C Discretization of model partial differential equations ................... 237 Appendix D Computer code listing ....................................................................... 259 Appendix E Air mass flow measurement using orifice plate with D & D/2 tappings ............................................................................................. 299 Appendix F Experimental test summaries ........................................................... 301 Appendix G Photo documentation ........................................................................ 322 Appendix H Soft copy of the computer code on CD ............................................ 329

Appendix A

xi

List of tables

Table 3.1 List of drying, devolatilization, heterogeneous and homogenous reaction rate constants used in the mathematical model.............................. 62 Table 3.2 Thermophysical data and transport properties of solid fuel ......................... 66 Table 4.1 Baseline operating parameter values used in the model ............................. 100 Table 5.1 Size and specification of main air blower and suction blower ................... 152 Table 5.2 Specifications of common vacuum pump and vacuum pump for gas collection.................................................................................................... 156 Table 5.3 Average properties of red gum wood blocks ............................................. 163 Table 5.4 Average properties of red gum charcoal blocks ......................................... 164 Table 6.1 Performance summary of experimental stratified downdraft gasifier ........ 170 Table 6.2 Material balance summary.......................................................................... 171 Table 6.3 Carbon balance summary............................................................................ 171 Table 6.4 Oxygen balance summary........................................................................... 172 Table 6.5 Hydrogen balance summary ....................................................................... 172 Table 7.1 Estimation of optimum reaction zone length based on tar reduction rate.............................................................................................................. 196 Table 7.2 Measured tar content levels from different throatless downdraft gasifier studies ......................................................................................... 197 Table 7.3 Results from simulation of two-stage gasification concept (at 1000 sec after ignition) ....................................................................................... 214

xii

List of figures

Figure 3.1 Pyrolysis mechanism used in the numerical model and product yields [adapted from Boroson et al. (1989)] .............................................. 55 Figure 3.2 Initial conditions in model reactor............................................................... 75 Figure 3.3 A generic node point i (non-uniform grid) and the control volume in one dimension ........................................................................................ 77 Figure 3.4 Flow chart for the numerical solution of stratified downdraft gasifier........................................................................................................ 94 Figure 4.1 Transient behaviour of solid surface temperature profile.......................... 103 Figure 4.2 Transient behaviour of gas temperature profile......................................... 103 Figure 4.3 Transient behaviour of gas composition profile (a) CO, (b) CO2, (c) H2, (d) H2O............................................................................................... 104 Figure 4.4 Transient behaviour of outlet producer gas properties .............................. 104 Figure 4.5 The predicted solid (Ts) and gas (Tg) temperature profiles along the axis of the gasifier (a) over the total reactor length (b) in the domain where rates of change are greatest .............................................. 106 Figure 4.6 The solid and gas velocity profiles (Vs and Vg respectively) along the axis of the gasifier .............................................................................. 107 Figure 4.7 Solid density profile along the axis of the gasifier (a) over the total reactor length (b) in the domain where rates of change are greatest........ 109 Figure 4.8 Major gas species profiles along the axis of gasifier (a) over the total reactor length (b) in the domain where rates of change are greatest ..................................................................................................... 111 Figure 4.9 Minor gas species profiles along the axis of the gasifier (a) over the total reactor length (b) in the domain where rates of change are greatest ..................................................................................................... 113 Figure 4.10 Primary tar, secondary tar, and oxygen profiles along the reactor axis ........................................................................................................... 114

xiii

Figure 4.11 Drying, pyrolysis and char oxidation rates along the axis of the gasifier...................................................................................................... 115 Figure 4.12 Gas temperature (Tg) and rates of gaseous and solid phase oxidation reaction along the axis of the gasifier ...................................... 116 Figure 4.13 Heterogeneous gasification reaction rates along the axis of the gasifier...................................................................................................... 116 Figure 4.14 Illustration showing downward movement of the reaction zone as a result of an increase in the air supply rate (24 kg/hr)............................ 118 Figure 4.15 Illustration showing upward movement of the reaction zone as a result of a decrease in the air supply rate (8 kg/hr) .................................. 118 Figure 4.16 Effects of air supply rate (a) on solid (Ts) and gaseous (Tg) temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward) .......... 121 Figure 4.17 Effects of initial wood particle size (a) on solid (Ts) and gaseous (Tg) temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward) .......... 124 Figure 4.18 Effects of initial moisture content (a) on solid (Ts) and gaseous (Tg) temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward) .......... 126 Figure 4.19 Effects of solid-to-gas heat transfer coefficient (a) on solid (Ts) and gaseous (Tg) temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward).......................................................................................... 128 Figure 4.20 Effects of mass transfer coefficient (a) on solid (Ts) and gaseous (Tg) temperature profiles (b) on peak temperatures (c) on outlet

xiv

primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward) .......... 130 Figure 4.21 Effects of heat and mass transfer exchange area (a) on solid (Ts) and gaseous (Tg) temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward).......................................................................................... 132 Figure 4.22 Effects of changes in primary pyrolysis rate (a) on solid (Ts) and gaseous (Tg) temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward).......................................................................................... 134 Figure 4.23 Effects of changes in secondary pyrolysis rates (a) on solid (Ts) and gaseous (Tg) temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward).......................................................................................... 136 Figure 4.24 Effects of changes in secondary tar oxidation rate (a) on solid(Ts) and gaseous(Tg) temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward).......................................................................................... 137 Figure 4.25 Effects of changes in effective thermal conductivity (a) on solid (Ts) and gaseous temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward).......................................................................................... 139 Figure 4.26 Effects of changes in the extent of heat loss through the reactor wall (a) on solid (Ts) and gaseous (Tg) temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels

xv

(d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward) ................................................................ 141 Figure 5.1 Experimental set-up................................................................................... 146 Figure 5.2 Picture of experimental stratified downdraft gasifier ................................ 148 Figure 5.3 Schematic diagram of experimental stratified downdraft gasifier ............ 149 Figure 5.4 Schematic diagram of the modified middle section of the reactor ............ 150 Figure 5.5 Schematic diagram of burner set-up.......................................................... 151 Figure 5.6 Gas sampling train..................................................................................... 154 Figure 5.7 The moisture collector and impinger bottle used in experiments.............. 155 Figure 5.8 Modified gas sampling train...................................................................... 156 Figure 5.9 Schematic showing post-sampling procedures for gravimetric tar measurement (Source: Neeft et al. 2002)................................................. 158 Figure 5.10 Constructional arrangement of experimental orifice plate ...................... 160 Figure 5.11 The 30 mm and 20 mm sized wood blocks ............................................. 162 Figure 6.1 Temperature record for experiment 314 (air flow 16.92 kg/hr, with 20 mm blocks).......................................................................................... 173 Figure 6.2 Temperatures measured at three different levels above the ignition port during run 510 (Airflow 7.5 kg/hr, ignition started at 400 mm above the grate)........................................................................................ 174 Figure 6.3 Temperature profiles at three different air flow rates (average temperature values between 30 min and 40 min after ignition)............... 175 Figure 6.4 Outlet gas tar content as a function of air supply rate ............................... 177 Figure 6.5 Energy output versus dry wood consumption rate .................................... 177 Figure 6.6 Cold gas efficiency versus air supply rate................................................. 178 Figure 6.7 Comparison of predicted and experimentally measured temperature profiles along the axis of the gasifier (air flow rate: 15.5 kg/hr, wood block size 20x20x20 mm) .............................................................. 180

xvi

Figure 6.8 Predicted gas outlet temperatures compared with experimental results as a function of air supply rate (experimental points are average values over the 2 minute interval between 29 -31 minutes after ignition; predicted values are those at 30 min after ignition) .......... 181 Figure 6.9 Comparison of model and experimental dry gas composition profiles along the gasifier for two different time intervals (air flow rate 19.6 kg/hr, small wood blocks) (a) 10-15 min after ignition (b) 30-40 min after ignition...................................................................... 183 Figure 6.10 Comparison of predicted and experimentally measured condensable tar profiles along the gasifier for different airflow rates under steady state operating conditions........................................... 184 Figure 6.11 Wood fuel consumption versus air supply rate ....................................... 185 Figure 6.12 Air-fuel ratio (kg/kg d.b) versus air supply rate ...................................... 186 Figure 6.13 Gas production versus air supply rate ..................................................... 187 Figure 6.14 (a) Temperature profiles at two different times (1000 seconds and 3000 seconds respectively after ignition of the charcoal bed) for two different air flow rates (same fuel size) (b) Temperature profiles at the same two times for two different fuel sizes (same air supply rate)............................................................................... 188 Figure 6.15 Variation of reaction zone movement rate with air supply rate (comparison between model and measured values)................................. 189 Figure 7.1 Cold gas efficiency profiles along the axis of the gasifier ........................ 194 Figure 7.2 Predicted tar reduction rates along the gasification zone and inert char zone .................................................................................................. 196 Figure 7.3 Dynamic behaviour of solid temperature at reactor outlet ........................ 199 Figure 7.4 Volatile matter profile along the gasifier axis ........................................... 200 Figure 7.5 Tar profiles along the axis of the gasifier for wood only operation .......... 201 Figure 7.6 Experimental procedure that obviates the need to use an initial charge of charcoal .................................................................................... 203

xvii

Figure 7.7 Changes over time of temperature profiles in the gasifier for the procedure of producing charcoal inside the gasifier ................................ 203 Figure 7.8 Variations over time of outlet gas tar content and cold gas efficiency for the procedure of producing charcoal inside the gasifier...................................................................................................... 204 Figure 7.9 Two-stage gasification systems proposed by TUD (Denmark) and AIT (Thailand) ......................................................................................... 205 Figure 7.10 Variation with time of temperature profiles along the axis of the gasifier for the first 1000 sec after ignition (Ignition port: 500 mm above grate; secondary air supply: 300 mm above grate; primary air flow rate = 15 kg/hr; secondary air flow rate = 5 kg/hr)..................... 208 Figure 7.11 Gas and solid phase oxidation rate profiles for the first and second stage, at 1000 sec after ignition (Ignition port: 500 mm above grate; secondary air supply: 300 mm above grate; primary air = 15 kg/hr; secondary air = 5 kg/hr)................................................................. 209 Figure 7.12 Heterogenous gasification reaction rates for two-stage gasification (Ignition port: 500 mm above grate; secondary air supply: 300 mm above grate; primary air flow rate = 15 kg/hr, secondary air flow rate = 5 kg/hr)........................................................................................... 210 Figure 7.13 Gas composition profiles along the axis of the gasifier for twostage gasification (a) major gas species profile (b) minor gas species profile (Ignition port: 500 mm above grate; secondary air supply point: 300 mm above grate; primary air flow rate = 15 kg/hr, secondary air flow rate = 5 kg/hr) ................................................. 211 Figure 7.14 Tar profile along the gasifier for both single-stage and two-stage units, at 1000 sec after ignition (Ignition port: 500 mm above grate for both cases; single-stage air flow rate: 20 kg/hr; two-stage total air supply rate: 20 kg/hr (primary air flow rate = 15 kg/hr; secondary air flow rate = 5 kg/hr); secondary air supply: at 300 mm above grate)....................................................................................... 213

xviii

Nomenclature

A Av C C Cp dp D D E H h h h hrs hrv H jd jH k k k keff km krg kro KC m" M mp N

pre-exponential factor particle surface area per unit volume, m2/m3 molar concentration of gas species carbon specific heat, J/kg.K particle diameter, m diffusivity, m2/s reactor diameter, m activation energy, J/mol enthalpy of reaction, J/kg specific enthalpy, J/kg heat transfer coefficient, W/m.K absolute enthalpy, J/kg solid radiation coefficient, W/m.K void-to-void radiation coefficient, W/m.K hydrogen dimensionless mass transfer factor dimensionless heat transfer factor intrinsic reaction rate thermal conductivity, W/m.K conductivity ratio, ks/kg effective thermal conductivity, W/m.K mass transfer coefficient, m/s gas effective radial conductivity, W/m.K static effective radial conductivity, W/m.K equilibrium constant for water gas shift reaction mass flux, kg/m2.s molecular weight, g/mol mass of single particle, kg nitrogen

xix

Nu O P Pr R R Re S Sc t T V W x X1i X2i

Nusselt number oxygen gas mixture pressure, kPa Prandtl number (= Cp./k) reaction rate, kg/m.s (or) kmol/m.s universal gas constant, J/kmol.K Reynolds number (= dpm/) Source term Schmidt number (=/Di) time, s temperature, K velocity, m/s mesh monitor function vertical coordinate, m yield of species i in primary pyrolysis process (mass fraction) yield of species i in secondary pyrolysis process (mass fraction)

Greek symbols prefix indicating grid or time spacing (e.g. t) emissivity bed void fraction, void volume/ bed volume ratio of reacting to non-reacting heat transfer coefficient stoichiometric coefficient of char oxidation reaction, mol/mol stoichiometric coefficient, mol oxidant /mol carbon viscosity, kg/m.s density, (kg/m) Stephan-Boltzmann constant (=5.6703E-8, W/m2K4) equivalence ratio packing parameter summation of atomic diffusion volumes

xx

Subscripts avg bw c g gw m m p1 p2 s sg sw tar1 tar2 v w average value fuel bed-to-reactor wall char gas phase gas-to-reactor wall moisture drying process primary pyrolysis process secondary pyrolysis process solid phase solid-to-gas solid-to-wall primary tar (condensable tar, oxygenates) secondary tar (taken as benzene) volatiles reactor wall

xxi

Chapter 1 Introduction

Chapter 1

Introduction
1.1 Background

Research and development in renewable energy has been receiving increased attention due to worldwide sustainable energy concerns and issues of global climate change. A wide range of renewable energy technologies were conceived and developed during the last three decades. Among the renewable energy sources, biomass is widely considered to be a fuel of major potential both for the present time and for the future. According to the World Energy Assessment report, 67% of the worlds renewable energy is contributed by biomass (Rogner & Popescu 2000).

The name Biomass was invented in around 1975 to describe all organic materials that originate from plants (Reed 2003). Biomass includes standing forests, energy crops and organic wastes. Biomass fuel, in contrast to fossil fuels, has a unique potential for making a positive environmental impact. This is because, in any plan for sustainable biomass production and use, the carbon dioxide (CO2) emitted would be absorbed by the growth of new biomass plantations. Biomass can be burned without emitting large amounts of nitrogen oxides (NOX). Moreover, because of the low sulphur content of biomass, emissions of sulphur dioxide (SO2) will also be low, especially compared with the emissions from coal-fired combustion. From ancient times until relatively recently, biomass was the principal chemical fuel of mans activities. It is presently the fourth largest energy resource after coal, oil and natural gas and is estimated to contribute of the order of 10-14% of the worlds primary energy supply (McKendry 2001).

Biomass can be converted for use into a number of energy forms. It can be used directly through combustion for heat, or for conversion into electricity. It can also be transformed into a liquid fuel, such as methanol or ethanol, or converted into a gaseous fuel. Among the various kinds of thermal conversion processes, biomass gasification

Chapter 1 Introduction

has been paid most attention as it offers higher overall conversion efficiencies compared to direct combustion for electricity production (Maniatis 2001).

Gasification is the process of converting a solid fuel to a combustible gas by supplying a restricted amount of oxygen, either as pure oxygen or from air. It is a centuries old technology, which found vital applications during the Second World War when approximately one million gasifiers were used to operate cars, trucks, boats, trains and electric generators in Europe (Foley, Barnard & Timberlake 1983). It quickly lost its competitive edge once petroleum fuels again became widely available. Another wave of development began in the 1970s, following the drastic increase in world oil prices, and a growing awareness of possible climatic effects of the continued use of fossil fuels.

Downdraft gasifiers are attractive for small-scale applications (<1.5 MWth) as there is a very big market not only in developed but also developing economies. According to an extensive review of gasifier manufacturers in Europe, USA and Canada, 75 % of 50 manufacturers offering commercial gasification plants use the downdraft mode (Knoef
2000). However, along with the rebirth of thermal gasification of biomass, came many

still unresolved problems, mainly the efficient and economic removal of tar and stability of reactor operation, that limit its application (Bridgwater 1995; Reed & Agua 1988). These problems still hinder the penetration of biomass gasification technology into world energy markets (Maniatis 2001).

1.2

Problem statement

Tom Reed, who spent more than 20 years working with various gasifier systems, offered the following insight into his experience to date (Reed & Gaur 1998):

The typical project starts with new ideas, announcements at meetings, construction of the new gasifier: then it is found that gas contains 0.1-10% tars. The rest of the time and money is spent trying to solve this problem. Most of the gasifier projects then quietly disappear. In some cases the cost of cleaning up the experimental site exceeds

Chapter 1 Introduction

the cost of the project! Thus tars can be considered the Achilles heel of biomass gasification.

As can be inferred from Tom Reeds remark, the elimination of tar by auxiliary systems such as scrubbers is normally expensive, the equipment required is bulky and not reliable for long-term operation, and in addition, there are serious environmental problems due to the large quantities of condensate produced. Maniatis and Beenackers (2000) also mentioned that the tar problem is conceived of as one of the most important technical barriers to the penetration of biomass gasification technology into the power market. Because of this, the concept of reducing tar content inside the reactor has become more important and several attempts have been made to reduce tar content by modifying the traditional downdraft biomass gasification system (Bhattacharya, Hla & Pham 2001; Brandt, Larsen & Henriksen 2000; Bui, Loof & Bhattacharya 1994; Susanto & Beenackers 1996).

Modelling is a very useful tool for designing a clean and efficient biomass gasification unit, since it increases understanding of the gasification process and reactor optimisation. Therefore modelling of gasification processes has received significant attention over the last few decades. However, among the gasification models developed, those applicable for biomass fuel are less numerous than those for coal. The usefulness of coal-oriented models is limited as there are significant differences between coal and biomass due to their different physical properties, chemical composition and impurities.

There are a limited number of numerical models of fixed bed biomass gasification available in the literature and most of them are formulated based on very simplified assumptions. For example, the equilibrium model (in which no kinetic data is applied) can predict the exit gas composition by assuming solid and gas phase reactions reach equilibrium at a fixed reactor temperature (Buekens & Schoeters 1985). Some of the steady state models (Chen & Gunkel 1988; Groeneveld 1980) neglect the kinetic mechanisms of pyrolysis and cracking of volatile products and the first three zones of downdraft gasifier (drying, pyrolysis and combustion) are lumped together. This assumption is only acceptable for estimating the major gas components of outlet

Chapter 1 Introduction

producer gas and cannot be used to predict mechanisms for pyrolysis product formation inside the gasifier. For an understanding of the tar formation, cracking and oxidation processes inside the downdraft gasifier, a numerical model must include kinetic expressions for those processes. Besides, to understand the behaviour of the reactor under unsteady state conditions, the model has to include the time dependence of the reactor parameters. The development, analysis and validation of such a numerical model are the main concerns of this research.

1.3

Aim of the study

The aim of this research is to analyse, theoretically and experimentally, the processes occurring inside a stratified downdraft wood gasifier. This is in order to understand the behaviour causing both tar formation and tar reduction inside the gasifier under different operating parameters. This information is then to be used as the basis for proposing a concept for a reactor which will minimize the amount of tar in the producer gas.

1.4

Objectives of the study

The specific objectives of the research are listed as follows:

1. To develop a transient model for a downdraft wood gasifier which enables prediction of the tar content of producer gas along the axis of the reactor. 2. To analyse the sensitivity of the model developed to operational and model parameters such as heat and mass transfer coefficients, reaction kinetic rate coefficients and the heat loss coefficient. 3. To identify the main factors influencing the final tar content of producer gas by varying the input parameters and model parameters in the simulation. 4. To design and fabricate a stratified downdraft gasifier for experimental studies.

Chapter 1 Introduction

5. To study the performance of the experimental gasifier under both steady and transient conditions. 6. To validate the model using the experimental results. 7. To determine the best operating conditions for a stratified downdraft gasifier that can operate stably over prolonged periods producing low tar gas suitable for use in engine applications.

To achieve the objectives of this research, a thorough review of experimental work on tar reduction and a detailed understanding of previous numerical modelling works are required, and these are presented in the next chapter.

Chapter 1 Introduction

1.5

References

Bhattacharya, SC, San Shwe Hla & Pham, H-L 2001, A study on a multi-stage hybrid gasifier-engine system, Biomass & Bioenergy, vol. 21, pp. 445-60. Brandt, P, Larsen, E & Henriksen, U 2000, High tar reduction in a two-stage gasifier Energy & Fuels, vol. 14, pp. 816-9. Bridgwater, AV 1995, The technical and economic feasibility of biomass gasification for power generation, Fuel, vol. 74, no. 5, pp. 631-53. Buekens, AG & Schoeters, JF 1985, Modelling of biomass gasification, Fundamentals of Thermochemical Biomass Conversion, eds. RP Overend, TA Milne & LK Mudge, pp. 619-89. Bui, T, Loof, R & Bhattacharya SC 1994, Multi-stage reactor for thermal gasification of wood, Energy, vol. 19, pp. 397-404. Chen, J & Gunkel, W 1988, Modeling and simulation of Co-current moving bed gasification reactors- Part II A detailed gasifier Model, Biomass, vol. 14, pp. 75-98. Foley, G, Barnard, G & Timberlake, L 1983, Gasifiers: Fuel for siege economies, International Institute for Environmental and Development, London, England. Groeneveld, HJ 1980, The co-current moving bed gasifier, PhD Dissertation, Twente University of Technology, Enschede, The Netherlands. Knoef, HAM 2000, Inventory of biomass gasifier manufacturers & installations, Final report to European commission, Contract DIS/1734/98-NL, Biomass Technology Group B.V, University of Twente, Enschede. Maniatis, K & Beenackers, AACM 2000, Tar Protocols. IEA Bioenergy Gasification Task, Biomass and Bioenergy, vol. 18, pp. 1-4. Maniatis, K 2001, Progress in biomass gasification: An overview, Progress in thermochemical biomass conversion, ed. AV Bridgewater, pp. 1-31. McKendry, P 2001, Energy production from biomass (part 1): overview of biomass, Bioresource Technology, vol. 83, pp. 37-46.

Chapter 1 Introduction

Reed, TB & Agua, D 1988, Handbook of biomass downdraft gasifier engine systems, Biomass Energy Foundation Press, Washington. Reed, TB & Gaur, S 1998, Survey of biomass gasification- 1998, Vol. 1 Gasifier Project and Manufacturers around the World, Golden, CO: The National Renewable Energy Laboratory and The Biomass Energy Foundation, Inc. Reed TB 2003, The Biomass Energy Foundation, viewed 22 August 2003,

<http://www.woodgas.com/> Rogner, HH & Popescu, A 2000, An introduction to energy, World energy assessment: energy and the challenge of sustainability, ed. J Goldemberg (Chair), United Nations Development Programme, New York. Susanto, H & Beenackers, AAC 1996, A moving-bed gasifier with internal recycle of pyrolysis gas, Fuel, vol. 75, pp. 1339-1347.

Chapter 2 Literature review

Chapter 2

Literature review

2.1

Introduction

Since one of the main objectives of this study is to develop a transient model for stratified downdraft gasification, relevant literature on modelling in gasification, and on related chemical and physical phenomena such as reaction rates and heat and mass transfer processes, is reviewed in the first part of this chapter. The second part discusses different experimental investigations focused on tar content in producer gas from biomass gasification and various approaches to producing clean producer gas.

2.2

Different models of fixed-bed gasification

According to Buekens and Schoeters (1985), gasifier models can be classified into five categories, namely equilibrium models, kinetics free models, steady-state models, semi-transient models and transient models. In equilibrium models, the gas composition is calculated assuming a fixed reactor temperature and that gasifier reactions are in equilibrium. In kinetics free models, the reactor is subdivided into different zones. Gas composition is calculated from equilibrium data whereas the reaction temperature is calculated for each zone by a separate heat balance. These two models only allow one to predict exit gas compositions and to estimate a fixed reaction temperature. models. In steady-state models, gas composition and the temperature profile along the gasifier axis can be estimated by using mass balance and energy balance equations, for both the solid and gas phases, in which the time derivative terms are considered to be zero. Analysis of detailed physical and chemical processes inside the gasifier is not possible using these two

Chapter 2 Literature review

These models result in a set of non-linear ordinary differential equations and algebraic equations. In semi-transient models, a set of results from a previous steady-state model is used to calculate the effect on reactor performance of small changes in operating conditions, such as solids feed rate or blast gas flow rate. None of the above models can predict the transient behaviours of the fixed bed gasifier. To analyse the unsteady behaviour of such a reactor, transient models must be applied. Transient models are based on mass and energy balance expressions that include the time derivative terms; these expressions take the form of a set of nonlinear partial differential equations. Except for the very simple model types (equilibrium & kinetics free), the formulating of comprehensive steady state and transient gasifier models requires a series of mass and energy balance equations together with information about rates of drying, pyrolysis and chemical reactions (both heterogeneous and homogeneous), and complex physical transport processes, though simplifications are usually made. However, most of the existing steady state models of downdraft gasification were formulated based on a very simplified approach, namely that the processes occurring in the first three zones of the downdraft gasifier (drying, pyrolysis and combustion) can be lumped together as a single process, and output values of these three zones estimated by applying elemental mass balance and energy balance procedures (Chen & Gunkel 1988;
Groeneveld 1980). This simplified approach neglects an important factor in determining

gasifier performance, namely the degree of completion of carbonization attained at the end of the pyrolysis and oxidation zones. Such information requires consideration of the kinetics both of biomass pyrolysis and of oxidation of volatile gaseous products, but instances where this is done are very limited in wood gasifier models. In the next section, a review of areas related to modelling of biomass gasification is presented.

Chapter 2 Literature review

2.3

Reaction rates

2.3.1

Drying

Drying is the first process undergone by the fuel introduced to the gasifier. Drying of moist biomass involves only physical processes such as heat and mass transport. Although drying seems simple compared to other processes in biomass gasification, formulation of the process for gasifier operating conditions is not straightforward. This is because most drying studies have been carried out on single particles, not particle beds. Fortunately, any errors introduced by neglecting the effect of particle size in the formulation of fixed bed gasifier models can be disregarded because the characteristic times of moisture evaporation are orders of magnitude shorter than those of char combustion and gasification (Di Blasi 2000).

2.3.2

Pyrolysis

Once dry, the fuel particles in the gasifier heat up and undergo pyrolysis. Pyrolysis, by definition, implies decomposition through heating (Hastaoglu & Berruti 1989). During pyrolysis, biomass undergoes a sequence of decomposition reactions that give rise to a wide variety of chemical species. To make the mathematical modelling of the process manageable these are usually grouped; generally the product groups used are char, gas and tar. Chars are carbon-rich non-volatile residues. Tars are high molecular weight products that are volatile at the pyrolysis temperature but are liquid at room temperature. Depending on the degree of cracking, tars may range from light, oxygenated hydrocarbons to heavy, poly-aromatic hydrocarbons. The gas fraction includes all lower molecular weight products (mainly CO and CO2 and also water) which have a measurable vapour pressure at room temperature (Di Blasi 1993). There is no strict fixed temperature of pyrolysis; the decomposition takes place over a temperature range. Pyrolysis of biomass can start as low as 200C and is essentially complete by 500C (Milne & Evans 1998).

10

Chapter 2 Literature review

The reactions involved in wood pyrolysis are complex because wood components have different reactivities and yield different product spectra. Due to the complexity of the reaction processes, involving different reaction paths and also the generation of different products, the detailed kinetics of pyrolysis are still poorly understood (Peters & Bruch 2001). Therefore, in modelling the results of experimental investigations, overall kinetic expressions have to be used. Many studies have been conducted on the pyrolysis both of wood and of biomass materials in general, and a large variety of pyrolysis models have been presented in the literature. These have been extensively reviewed by Di Blasi (1993). The different approaches used by researchers in describing the kinetics of pyrolysis reactions are discussed briefly below. The simplest approach, taken by many researchers, is to consider wood pyrolysis as an overall one-step reaction process in which a single reaction accounts for the different fractions of gases, tars and chars formed: Wood k (a)Gases + (b)Tars + (c)Char

(2.1)

where a, b and c are the yield coefficients, expressed as grams of gas, condensable species and solid per gram of reacted wood. The kinetics rate k is normally described using a first-order Arrhenius-type rate expression. Thus, the rate of devolatilization for one-step global models is expressed as
E k = A exp RT g

(2.2)

where A and E are the pre-exponential factor and apparent activation energy respectively.

11

Chapter 2 Literature review

A better but more complex approach is to assume in the model that virgin solid fuel decomposes directly to each reaction product i, by a single independent reaction.
Wood ki Product i

(2.3)

Such single-step kinetic studies neglect the effects of secondary reactions and residence time within the particle, although both are known to be important considerations affecting the final product yield. Where the experimental work on pyrolysis has been conducted using fine particles, primary products have only a very limited residence time in the heated zone. Therefore, the effect of secondary reactions is not significant and the errors introduced by using a one-step kinetic model are small. However it is not appropriate to apply one-step kinetic expressions derived from fine particle experimental work, to gasifiers where the fuel beds contain a wide range of particle sizes and a variety of secondary pyrolysis reactions can be expected to occur. The above approaches ignore the fact that wood contains a number of different components that follow different decomposition pathways. Models have been developed that account for this by assuming that the overall wood pyrolysis reaction is made up of several parallel one-step reactions each describing the degradation of a different component to char and several gaseous species. The overall rate of reaction is obtained by summing the rates of all parallel reactions (Anthony & Howard 1976). In most cases each parallel reaction is treated as a first-order reaction following Arrhenius kinetics. In most such models, the three main components of wood (cellulose, hemicelluloses and lignin) are assumed to decompose via a series of parallel reactions, as shown below:
Cellulose k1 Products Hemicellulose k 2 Products Lignin k 3 Products

(2.4.a-2.4c)

In an attempt to overcome the deficiencies of the one-step pyrolysis reaction models, a group of researchers (Chan, Kelbon & Krieger 1985) introduced a multi-step pyrolysis kinetics model in which four reaction steps occurred: first virgin wood decomposes

12

Chapter 2 Literature review

into primary gases, char and primary tars, and then the primary tar decomposes into secondary gases and secondary tars by thermal cracking. Wood k1 Gas1 Wood k 2 Tar1 Wood k 3 Char Tar1 k 4 ( )Gas 2 + ( )Tar2 The two-step overall reaction model of Chan, Kelbon and Krieger (1985) has been used in a number of subsequent experimental studies (Boroson et al. 1989; Diebold 1985; Liden, Berruti & Scott 1988). In addition, judging by the series of recent experimental studies using a two-stage pyrolysis reactor (Fagbemi, Khezami & Capart 2001; Morf et al. 2001; Morf, Hasler & Nussbaumer 2002; Rath & Staudinger 2001) there is growing interest in secondary pyrolysis reactions such as the vapour phase cracking of wood pyrolysis tars. Although the experimental processes and methods used in these investigations were not the same, all had the same aim, which was to characterise the thermal conversion of primary pyrolysis products into secondary products during gasification and combustion processes. (2.5.a-2.5.d)

2.3.3

Heterogeneous gas-char reactions

The char remaining after wood devolatilization in the pyrolysis zone of a gasifier consists primarily of carbon, together with ash. The time required for consumption of a single char particle by oxidation and gasification is generally large compared to that needed for devolatilization. In fixed-bed gasifiers, the time for devolatilization to take place is of the order of 1,000 seconds while for char particle burnout the time is of the order of 10,000 seconds (Hobbs, Radulovic & Smoot 1993). Heterogeneous reactions between gases and solid particles have been conventionally modelled using the following assumptions (Skinner & Smoot 1979):

13

Chapter 2 Literature review

(1) Char particles are considered to be carbon spheres in an essentially infinite gas stream. (2) The gas-phase reactant first diffuses to the solid surface or into the pores in the char, and then absorbs and reacts with the surface. (3) The gas-phase products desorb and diffuse away from the particle. There are normally four heterogeneous reactions that can take place during the high temperature gasification of carbonaceous materials. These are listed as follows: (Klm & Joseph 1983)

CH O + ( )O2 k1 (2 2 +

)CO + (2 +

1)CO2 + ( ) H 2 O 2 2

(2.6) (2.7) (2.8) (2.9)

CH O + CO2 k 2 2CO + ( ) H 2 O + (

)H 2

CH O + (1 ) H 2 O k 3 CO + (1 + CH O + (2

)H 2

+ ) H 2 k 4 CH 4 + ( ) H 2 O

The first reaction is known as char combustion and is an exothermic reaction. The other reactions (the so-called char gasification reactions) are endothermic reactions and are relatively slow. In most of the fixed bed gasification models, for simplicity, char entering the combustion and gasification zones is assumed to be pure carbon (ie,

==0). Since both biomass charcoal and coal usually contain some amount of
hydrogen and oxygen in addition to carbon, the validity of this assumption can only be assured by including the concept that the hydrogen and oxygen content of the solid fuel is totally released during pyrolysis. The kinetic rates for gas-char reactions have typically been taken from small particle experiments at high temperatures and heating rates. Use of this small particle kinetic data to describe reactions involving large particles could be expected to produce erroneous results due to size related effects such as intra-particle gradients. However, large particle oxidation and gasification data are scarce, which means that use of small particle kinetic data is usually unavoidable. The consequences of using this data in

14

Chapter 2 Literature review

gasifier models may be less important than it first appears, though, since mass transport tends to be the rate controlling step for fixed-bed combustion and gasification processes. For example, an experimental study with large charcoal particles in a CO2 gas environment has shown that the particle gasification proceeds largely by reaction with the outer shell, with the interior of the particle remaining relatively untouched until very high conversions have been reached (Standish & Tanjung 1988). Thus, shrinking core-reaction control models using effective internal diffusion may often be adequate for comprehensive fixed-bed modelling (Hobbs, Radulovic & Smoot 1993). There are two common char oxidation sub-models, namely the shell progressive model (SP model) and the ash segregation model (AS model). The difference between the two models is in the treatment of the ash. In the SP model the ash is assumed to remain intact. The oxidant is required to diffuse through both the film boundary layer and the ash layer. In the AS model, the ash is assumed to crumble slowly and fall away from the char particle with the oxidant being required to diffuse through the film boundary layer only. The AS model is only appropriate for low ash content carbonaceous solid fuels like hard wood charcoal; the ash content of charcoals is typically 0.5-2 wt% whereas that of coal chars is typically 5-15 wt%, depending on the coal type (Standish & Tanjung 1988). For the formulation of char oxidation reactions, there is still one major uncertainty, namely the composition of the oxidation product. Both CO and CO2 are formed as primary reaction products. The CO/CO2 ratio increases with increasing reaction temperature. The temperature dependence of this ratio has been described by different researchers (Arthur 1951; Evans & Emmons 1977; Laurendeau 1978; Monson et al. 1995; Rossberg 1956) using the following Arrhenius-type equation:

E CO = A exp RT CO2 s

(2.10)

As indicated earlier, the principal gas-phase reactants in heterogeneous carbon reactions of the gasification process are oxygen, carbon dioxide, steam, and hydrogen. Of the gases that react with char, oxygen is the most reactive. The char-carbon dioxide

15

Chapter 2 Literature review

and char-water vapour reactions have reaction rates that are of the same order of magnitude, while the char-hydrogen reaction rate is the slowest (Laurendeau 1978). The intrinsic rates of these reactions have been investigated by many researchers; common experimental systems have included entrained flow reactors, thermogravimetric analysis, and fixed bed gasifiers. Most of the experimental work on combustion and gasification kinetic reaction rates has been conducted using coal char and only limited information has been found for wood char heterogeneous reactions (for instance, Dasappa et al. 1994; DeGroot & Shafizadeh 1984; Evans & Emmons 1977; Groeneveld & Swaaij 1980; Standish & Tanjung 1988). Since the kinetics often depend on the experimental environment, the results obtained from the TGA and fixed bed gasifier analyses seem the more appropriate for the formulation of fixed bed char combustion and gasification sub-models.

2.3.4

Homogeneous gas-phase reactions

a) Water-gas shift reaction A key gas-phase reaction in gasification processes is that known as the homogeneous water-gas shift reaction, where hydrogen and carbon dioxide are produced from carbon monoxide and steam as described in following equation.

CO + H 2 O CO2 + H 2

(2.11)

This reaction is slightly exothermic and is an unfavourable reaction as it reduces the heating value of the product gas in two ways. It exchanges a H2 molecule for a CO molecule with a corresponding reduction in the heat of combustion of 41kJ/gmol. It also exchanges a CO2 molecule for a H2O molecule. Since H2O is easily condensed out of the product gas while CO2 is not so easily removed, the product gas ends up being diluted with an incombustible gas and with its specific heating value further reduced. The water-gas shift reaction is very fast and proceeds to equilibrium within the gasifier. Therefore, the concentrations of hydrogen, carbon dioxide, carbon monoxide

16

Chapter 2 Literature review

and water can be related throughout the gasification zone by the following equilibrium relationship (Gumz 1950):

KC =

[CO][H 2 O] [CO2 ][H 2 ]

(2.12)

where [CO], for example, means the molar concentration (moles per unit volume) of CO, and KC = Equilibrium constant for water gas shift reaction, which is reaction temperature dependent.

b) Oxidation of devolatilization products Other gas phase reactions of importance in gasifier and combustor models are those involving the oxidation of devolatilization products. These reactions have been paid less attention than gas-char reactions. This is probably due to the complicated reaction mechanisms involved, and also to the previous lack of need for such information (Thurgood & Smoot 1979). In updraft mode reactors, on which most early work was conducted, the location of the pyrolysis zone means that volatile product combustion is of little importance in determining the overall combustion rate and therefore did not need characterising. In downdraft reactors, however, gaseous pyrolysis products flow through the hot char combustion zone and may be ignited as long as oxygen is available. When a combustible gas reacts with an oxidising agent, many sequential processes can occur involving many intermediate species. To convert the reactants to final products may involve only a few steps or as many as several hundred (Turns 2000). To include all of the elementary reactions in combustion sub-models is very difficult. Therefore, global reaction mechanisms are formulated by using a minimum number of elementary steps.

17

Chapter 2 Literature review

In the global reaction mechanism, the overall reaction of a mole of fuel with X moles of an oxidizer (say O2) to form Y moles of combustion products can be expressed as Fuel + XO2 k Yproduct (2.13)

From experimental measurements, the rate at which the fuel is consumed can be expressed as
d [Fuel ] n m = kTP a [Fuel ] [O2 ] dt

(2.14)

where the parameters in brackets are in molar concentrations (moles per unit volume). This rate equation states that the rate of disappearance of the fuel is proportional to each of the reactants raised to a power. The constant of proportionality, k, is called the global rate coefficient, and in general is not constant, but rather a strong function of temperature and is usually described by an Arrhenius-type rate expression. For global reactions, the reaction orders m and n are not necessarily integers and arise from curve fitting of experimental data. In general, a particular global expression in the form of the above equation holds only over a limited range of temperatures and pressures whereas elementary reaction rates can apply over a wide range of temperatures and pressures (Turns 2000). Simplified reaction mechanisms for the oxidation of devolatilization products (including hydrocarbons) are reviewed below.

c) Oxidation of hydrocarbon As indicated earlier, one of the main pyrolysis products is tar. Tar is not a single substance, but a mixture of oxygenates and heavy hydrocarbons. Some of these tar forming molecules combust while passing through the char bed in a downdraft gasifier.

18

Chapter 2 Literature review

However, dealing with the combustion of these components individually is not practical. Therefore oxygenates and other hydrocarbons are often treated as a single substance, represented as a pseudo-molecule, CxHyOz (Adams 1980). For the processes of modelling, the hydrocarbon oxidation mechanism has been greatly simplified and a global reaction proposed to allow for the partial oxidation of gaseous hydrocarbons. This reaction, shown below, was assumed to have an infinitely fast forward rate (Hammond & Mellor 1970).
x y y C x H y + + O2 ( x )CO + H 2 O 2 4 2

(2.15)

Siminski et al. (1972) proposed an alternative global reaction where the combustion products are CO and H2 rather than CO and H2O, and have correlated kinetic rates for the combustion of heavy hydrocarbons by means of this reaction:
x y C x H y + O2 ( x )CO + H 2 2 2

(2.16)

In the two previous equations, only partial oxidation was involved. Complete oxidation has been investigated by Westbrook and Dryer (1981), who presented and evaluated one-step and two-step global kinetics for a wide variety of hydrocarbons. A one-step reaction mechanism is the simplest approach, representing the oxidation process of a conventional hydrocarbon fuel going directly to CO2 and H2O:
y y C x H y + x + O2 ( x )CO2 + H 2 O 4 2

(2.17)

The advantages of a single-step process are obvious in that only four chemical species are involved in the formulation, and the heat release calculation is also quite simple. But it has several drawbacks, which can be important in certain applications. For instance, by assuming that the reaction products are CO2 and H2O the total heat of reaction is overpredicted and this results in too high a value of the adiabatic flame

19

Chapter 2 Literature review

temperature. Since the formation of intermediate hydrocarbons and CO is not taken into account, oxidation to CO2 is predicted to occur much faster than is observed experimentally. More realistic is a two-step mechanism, which separates the highly exothermic oxidation of CO to CO2 from the less exothermic oxidation of the hydrocarbon to CO. One such mechanism, adapted from the work of Dryer and Glassman (1973) is:
x y y C x H y + + O2 ( x )CO + H 2 O 2 4 2 1 CO + O2 CO2 2

(2.18) (2.19)

d) Oxidation of Methane Methane yields from both primary and secondary pyrolysis are reported to be about 56% of the total dry weight of virgin wood (Boroson et al. 1989). Therefore, methane combustion is a process that needs to be accounted for in the formulation of downdraft mode reactor models. Methane oxidation at combustion temperatures has been studied in more detail than the oxidation of any other hydrocarbon (Thurgood & Smoot 1979). A review of its combustion kinetics indicated that the methane combustion mechanism involves 149 elementary steps, with 144 reverse reactions, involving 33 species (Frenklach, Wang and Rabinowitz 1992). To make some sense of this complex system, by minimizing the number of elementary steps, a two-step global model for methane oxidation has been constructed using experimental results from a turbulent flow reactor (Dryer & Glassman 1973).

3 CH 4 + O2 CO + 2 H 2 O 2 1 CO + O2 CO2 2

(2.20) (2.21)

20

Chapter 2 Literature review

e) Oxidation of carbon monoxide Oxidation of carbon monoxide is another very important gas phase reaction that has to be allowed for in the formulation of combustor and gasifier models. As discussed in an earlier subsection, the reaction sequence in hydrocarbon flames has been recognized for some time to be the rapid oxidation of fuel to carbon monoxide followed by the much slower oxidation of the latter to carbon dioxide. Besides, at higher combustion temperatures, solid char is initially oxidized to carbon monoxide rather than carbon dioxide. So, carbon monoxide oxidation is an essential component of mathematical models of fixed bed gasifiers operating in the downdraft mode. The rate of carbon monoxide oxidation is available from an extensive experimental analysis in which oxidation of aliphatic hydrocarbons was studied in a high temperature, turbulent flow reactor to develop kinetic rates (Hautman et al. 1981). The rate of the CO oxidation, developed primarily from propane oxidation results, was reported as d [CO ] 20138 1.0 0.25 0.5 = 1014.6 exp [CO ] [O2 ] [H 2 O ] 7.93 exp( 2.48 ) dt T where the rate is expressed in (mole/cms) and is the initial equivalence ratio.

(2.22)

Another expression describing the rates of CO oxidation was developed by Howard, Williams and Fine (1973), who determined the rates of carbon monoxide oxidation in post-flame gases at 1 atm by concentration and temperature measurements, both made in a differential flow reactor comprising the second stage of a two-stage combustion system. The carbon monoxide-oxygen global reaction rate which was derived from the experimental results and which applies over a temperature range of 840-2360 K is as follows: d [CO ] 15104 1.0 0.5 0.5 = 1.3 1014 exp [CO ] [O2 ] [H 2 O ] dt T where the kinetic rate is in (mole/cms)

(2.23)

21

Chapter 2 Literature review

The rate expression reported by Hautman et al. (1981) resulted from experimental investigations which encompassed a temperature range of 960 to 1540 K. In the downdraft gasifier, a potential flame temperature from gas oxidation can be higher than 1750 K (Reed & Markson 1983). Therefore, the rate expression of Williams and Fine (1973) is more reasonable to apply in gasifier models since their correlation covers a wider temperature range.

f) Oxidation of Hydrogen The hydrogen-oxygen system is also important as a subsystem in the oxidation of hydrocarbons if a multi-step mechanism is applied. Detailed reviews of H2-O2 kinetics and their elementary reactions can be found in standard combustion books (Borman & Ragland 1998; Keating 1993; Turns 2000). A difficulty in predicting rates of hydrogen oxidation is that these depend on what other oxidisable components are present. For example, one of the early experimental analyses of the simultaneous combustion of hydrogen and carbon monoxide showed that the ratio of the velocity constants for these two reactions is kH2/kCO = 2.86, showing that hydrogen burns 2.86 times faster than carbon monoxide (Haslam 1923). In addition, Hautman et al. (1981) investigated the effect on the rate of hydrogen oxidation of the presence of hydrocarbons. They found, for their system, that in the presence of hydrocarbons, the combustion rate of hydrogen could be calculated from: d [H 2 ] 20642 0.85 1.42 0.56 = 1013.52 exp [H 2 ] [O2 ] [C 2 H 4 ] dt T

(2.24)

where kinetic rate is in (mole/cms). But the inclusion of the [C2H4] term implies that this rate expression is dependent on the nature of the hydrocarbons present, so its applicability is limited.

22

Chapter 2 Literature review

The global kinetic rates of hydrogen oxidation have also been estimated, using elementary kinetic reactions, from experimental analysis of the combustion of premixed laminar hydrogen-air mixtures of equivalence ratios ranging from 0.5 to 5.0 (Varma, Chatwani & Bracco 1986). From their analyses they developed the following expression:
d [H 2 ] E 1 .1 1 .1 = A exp [H 2 ] [O2 ] dt RT

(2.25)

where A and E are the pre-exponential factor and activation energy respectively, and A and E vary as the equivalence ratio alters.

2.4

Heat and mass transfer processes in packed beds

The pioneering study of heat and mass transfer in the flow of fluids through packed beds was that of Gamson, Thodos and Hougen (1943). This has been followed by a number of other studies during the last six decades. All of the published results have been based on experimental work in which no chemical reactions occurred. The heat and mass transfer coefficients computed from correlations developed for these nonreactive systems appear to exceed the experimentally observed values in reacting gasifiers (Hobbs, Radulovic & Smoot 1993). This may be attributable to the fact that fixed-bed gasifiers usually operate outside of the range in which heat and mass transfer coefficients have been adequately measured and correlated. To overcome this problem and enable the experimentally determined correlations of heat and mass transfer coefficients to be used in the formulation of gasification models, a factor called reacting to nonreacting heat transfer ratio was introduced in a coal gasification model (Hobbs, Radulovic & Smoot 1992).

23

Chapter 2 Literature review

2.4.1

Heat transfer

Experimental determinations of heat transfer coefficients have been made for a wide variety of systems involving heat transfer to or from particles. Various experimental techniques have been used under both steady-state and unsteady-state conditions. Most of the relevant heat transfer studies reported in the literature have been summarized and comprehensively reviewed by Wakao and Kaguei (1982). In most of the studies, the results have been presented in terms of the jH factor, which was introduced by Chilton and Colburn (1934).
hsg Cp g m g

jH =

(Pr ) 3

(2.26)

where, jH hsg mg Pr = Dimensionless heat transfer factor = (Solid-to-gas) heat transfer coefficient, = Mass velocity of the fluid based on the total (or) superficial bed cross-section measured normal to the mean flow direction = Prandtl number (Cpg/k)

Cpg = Specific heat of the gas mixture

Gamson, Thodos and Hougen (1943), who experimentally determined rates of evaporation of water into air streams from spherical and cylindrical porous pellets, developed jH factor correlations for modified Reynolds numbers in the range of 200 to 3750. The surface temperature of the wet porous pellets was assumed the same as the wet-bulb temperature of the inlet air and experiments were carried out by varying inlet dry bulb temperatures from 26 to 71C and sphere particle sizes from 2.3. to 11.6 mm. The relationship between jH factor and Reynolds number (Re=Dpmg/) was reported as follows:

24

Chapter 2 Literature review

j H = 1.064(Re ) j H = 18.1(Re )

0.41

for Re > 350

1.0

for Re < 40

(2.27)

The validity of the assumption that the surface temperature of the wet porous pellets is equal to the wet bulb temperature of inlet air is questionable at low air velocities. Therefore, Acetis and Thodos (1960) conducted a series of experiments in which the temperature of the water evaporating from spherical catalyst carriers was not assumed but directly measured and it was found that the temperature of the evaporating surface was the same as the wet-bulb temperature of the inlet air only at high air velocities. They derived the following relationship between the jH factor and Reynolds number:
1 .1

jH =

(Re )

0.41

0.15

(2.28)

for the modified Reynolds number region of 13 to 2136 covered in their study. In the experimental works of Gamson, Thodos & Hougen (1943) and Acetis and Thodos (1960), no attempt was made to account for the temperature difference between the wetted particles and surroundings which caused radiation heat transfer between the surroundings walls and the particles. Therefore, further experimental improvements were made by Gupta and Thodos (1963), who modified their experimental system to keep the temperature of the surrounding walls at the same temperature as that of the wetted particles so that radiation effects could be safely neglected. In their experiments, particle surface temperatures were measured directly and the relationship they derived between the jH factor and Reynolds number was as follows:
2.06(Re )
0.575

jH =

(2.29)

where is the void fraction in the particle bed. Since the source of errors caused by radiation effects was eliminated in their experimental work, the heat transfer coefficient equation proposed by Gupta and Thodos (1963) seems more appropriate for use in modelling of packed bed reactors.

25

Chapter 2 Literature review

Subsequently, Waitaker (1977) developed the Nusselt number correlation shown below. This was derived from heat transfer data obtained from various sources for beds packed with a variety of packings.
1 2 0 .4 Nu = 0.4(Re ) 2 + 0.2(Re ) 3 (Pr )

(2.30)

This correlation applies over a Reynolds number range from 3.7 to 8000 and a void fraction range from 0.34 to 0.74. For the above equation, Nusselt number and Reynolds number are defined as follows:

D p mg hD p Nu = , Re = (1 ) k 1

(2.31.a, 2.31.b)

In an attempt to come up with a more comprehensive correlation than those then available, Wakao and Kaguei (1982) revised and correlated selected reliable data (from both steady-state and unsteady-state measurements) and developed an empirical relationship, which they believed would enable more accurate prediction of particle-tofluid heat transfer coefficients. The Nu-Re relationship applicable over a Reynolds number range from 3 to 3000 proposed by Wakao and Kaguei (1982) is as follows:
1 0 .6 Nu = 2 + 1.1(Pr ) 3 (Re )

(2.32)

hD p where Nu = k

D m , Re = p g

Very recently (and since the inception of this project) a new forced convection heat transfer correlation covering the flow of gases and liquids through shallow packed beds has been proposed (Bird, Stewart & Lightfoot 2002):
2 3

j H = 2.19(Re )

+ 0.78(Re )

0.381

(2.33)

26

Chapter 2 Literature review

Equations 2.30, 2.32 and 2.33 are empirical relationships derived from heat transfer data obtained from various experimental works for packed beds. They all predict heat transfer coefficient values of the same order of magnitude. So the predictions of these equations were compared with the average values of heat transfer coefficients from two experimental correlations (Gupta & Thodos 1963; Acetis & Thodos 1960). Predictions made using the heat transfer correlation of Bird, Stewart and Lightfoot (2002) were close to the experimental values while the other two correlations gave values lower than those observed experimentally.

2.4.2

Mass transfer

As in the case of heat transfer, relevant solid-to-gas mass transfer studies were first carried out by Gamson, Thodos and Hougen (1943). They obtained a mass transfer coefficient from measurements of the rates of evaporation of water from wet porous particles. Since their pioneering work, a large number of other experimental studies have been carried out on mass transfer coefficients in packed bed systems (Wakao & Kaguei 1982). These mass transfer coefficients are frequently expressed in terms of the jd factor given by Chilton and Colburn (1934).
2 k m PM (Sc ) 3 mg

jd = where, jd km, P M Sc

(2.34)

= Dimensionless mass transfer factor = (Solid-to-gas) mass transfer coefficient, = Log mean partial pressure of the non-transferred gases in the gas film = Mean molecular weight of gas stream = Schmidt number (/D) where = gas mixture density and D = diffusivity

A number of empirical relationships between the jd factor and Re have been developed by different researchers for a range of experimental conditions.

27

Chapter 2 Literature review

An early study was that of Acetis and Thodos (1960) who published a jd-Re relationship which applied to both packed and expanded fixed beds and covered the Reynolds number region of 13 to 2136:
0.725

jd =

(Re )0.41 0.15

(2.35)

Subsequently Gupta and Thodos (1963) reported that the ratio jH /jd was approximately unity, implying that separate measurement of mass transfer coefficients is unnecessary if a jH versus Re relationship is available. The mass transfer coefficient associated with the vaporization of water and heavy hydrocarbons from the surface of porous spheres, 1.8 to 9.4 mm in diameter, was determined by Petrovic and Thodos (1968). Air was used as a carrier gas which flowed through packed and dispersed beds of these spheres. Porous spheres saturated with a liquid were placed in a matrix of dry glass spheres to produce a dispersed bed. The following jd-Re relationship was obtained:
0.357(Re )
0.359

jd =

(2.36)

where is the void fraction, and (3<Re<230) Mass transfer experimental data for a Reynolds number range of from 3 to 3000 was collected from different literature sources by Wakao & Kaguei (1982). These were used to develop an empirical Sherwood number correlation for both gas and liquid phase mass transfer in packed beds:
1 0 .6 Sh = 2 + 1.1(Sc ) 3 (Re )

(2.37)

km Dp where Sh = D im

D m , Re = p g

, Sc = D im

28

Chapter 2 Literature review

Very recently, an empirical mass transfer correlation for packed beds, based on the Chilton-Colburn analogy, (jH = jD = a function of Reynolds number) has been proposed by Bird, Stewart and Lightfoot (2002):
2 3

j D = 2.19(Re )

+ 0.78(Re )

0.381

(2.38)

As discussed in the heat transfer section, using the mass transfer coefficient correlation of Gupta and Thodos (1963) in the fixed bed model seems most appropriate because of the modifications made in their experimental work to minimise possible error sources. Besides, the mass transfer coefficient correlation produced by Bird, Stewart and Lightfoot (2002), who re-evaluated and correlated experimental data from a number of sources, gives similar results to those calculated using Gupta and Thodoss correlation.

2.4.3

Effective thermal conductivities in packed beds

Heat transfer between the particles in a high temperature packed bed takes place by a mix of conduction and other heat transfer mechanisms, particularly radiation. It is often convenient to combine these heat transfer processes by introducing the concept of an effective thermal conductivity. The heat transfer mechanisms involved in the determination of an effective thermal conductivity for a bed of particles can be listed as follows (Kunii & Smith 1960): 1. Heat transfer through the solid phase a. Thermal conduction through the solid particles b. Thermal conduction through the contact surfaces of solid particles c. Thermal conduction through the stagnant fluid near the contact surface d. Radiation heat transfer between the surfaces of solid particles 2. Heat transfer through the gas phase e. Thermal conduction through the fluid in the void space f. Radiation heat transfer through fluid in the void space g. Heat transfer by lateral mixing of fluid

29

Chapter 2 Literature review

The effective thermal conductivity in a fuel bed is determined less and less by pure conduction as temperatures increase. At temperatures normally found in gasifiers, radiation becomes the most significant process, especially for low conductivity solid materials such as wood. In a bed of particles heat transfer mechanisms 1 and 2 (see above) occur in parallel with each other. Mechanism a occurs in series with the combined result of parallel mechanisms b, c and d while mechanisms e, f and g occur in parallel with each other. The effective thermal conductivity (keff) can be written as follows (Kunii & Smith 1960; Yagi and Kunii 1957): Total heat flux [Keff t / x] Heat flux through solid phase (Mechanism 1) Heat flux through fluid in void space (Mechanism 2) (2.39)

In their approach to formulation of an effective thermal conductivity, Yagi and Kunii (1957) did not include mechanism e. They also assumed that mechanism b, direct conduction between solid particles in contact, was less important than the other mechanisms and this was neglected too. The validity of this assumption was confirmed by Kunii and Smith (1960) whose theoretical study of effective thermal conductivities of porous media showed heat transfer mechanism b to be negligible except in a high vacuum. Kunii and Smith (1960) did, however, take account of the thermal conduction through the fluid in the void space, mechanism e, in their formulation. Since both of the research groups (Kunii & Smith 1960; Yagi & Kunii 1957) concentrated in their work on beds in which the fluid was stagnant, mechanism g (heat transfer by lateral mixing of fluid) was not considered. However, in the case where fluid flows through a particle bed and the Reynolds number is large, the effect of mechanism g becomes significant. This contribution by mechanism g has been allowed for by Froment and Bischoff (1979) in their formulation of an effective thermal conductivity.

30

Chapter 2 Literature review

2.4.4

Heat loss through the reactor wall

The mechanisms that contribute to heat loss through the reactor wall can be listed as follows: 1. Heat transfer from reactor core to inner reactor wall 2. Thermal conduction through reactor wall 3. Thermal conduction through insulation (if reactor is insulated) 4. Convective heat loss to the surroundings from outer reactor wall (or) outer surface of insulation 5. Radiation heat loss to the surroundings from outer reactor wall (or) outer surface of insulation where mechanisms 1, 2 and 3 are in series with the combined result of parallel mechanisms 4 and 5 (Wakao & Kaguei 1982). Mechanisms 2 and 3 are controlled by the thicknesses and thermal conductivities of the reactor wall and the surrounding insulation. To estimate the heat loss via mechanism 4, the free convective heat transfer coefficient from the outer surface of the insulation must be found. To estimate the heat loss via mechanism 5, the emissivity of this outer surface must be known. Methods for estimating the heat loss via mechanisms 2-5 can be found in standard heat transfer texts. Factors affecting the determination of heat flows via mechanism 1 are discussed below:

The gas-wall & solid-wall heat transfer coefficient (hwf) The direct determination of the gas-wall heat transfer coefficient from heat transfer experiments is made difficult by the heat transfer occurring simultaneously between the wall and contacting particles. However, it is possible to estimate hwf from mass transfer experiments using a heat and mass transfer analogy. Yagi and Wakao (1959) estimated the wall mass transfer coefficient by measuring the dissolution rate of a coated material on the inner wall of a packed tube through which flowed a water stream. The corresponding fluid-wall heat transfer coefficient was obtained by

31

Chapter 2 Literature review

analogy. Kunii and Suzuki (1968) conducted a similar experiment but in determining their wall heat transfer coefficient made allowance for the average void fraction of the wall region. Other relevant experiments have been carried out by Olbrich and Potter (1972), who determined mass transfer coefficients by measuring the vaporization of mercury from the wall into a nitrogen stream. They considered simultaneous axial and radial gas mixing in the bed voids as well as accounting for the pressure drop across the bed, side effects which were not accounted for in the work of Yagi and Wakao (1959). Olbrich (1970) also developed an approach for estimating the solid-wall heat transfer coefficient by considering the ideal case of heat transfer between a plane wall and a contacting hexagonal close packed array of spheres.

The overall wall heat transfer coefficient (hw) More data is available on overall rates of heat transfer between the inside reactor wall and reactor contents (both gas and solid) than on individual gas-wall and solid-wall heat transfer processes. Calderbank and Pogorski (1957) proposed the following correlation to estimate the overall wall heat transfer coefficient
0.365

hw d p kg

mg d p = 3.6

(2.40)

while DeWasch and Froment (1971) put forward a rather different correlation:
o hw d p

hw d p kg

kg

+ 0.033 Re Pr

(2.41)

In the latter correlation, the influence of the tube diameter and of properties of catalytic
o reactors is accounted for in the correlation through hw .

32

Chapter 2 Literature review

Subsequently, Li and Finlayson (1977) re-examined experimental data from a number of authors and developed correlations for the asymptotic wall heat transfer coefficient for spherical and cylindrical packings, for constant wall temperature conditions. The best fit correlation for spherical packing was as follows:
hw d p kg dp dt

= 0.17 Re

0.79

(20 Re 7600 ), (0.05

0.3)

(2.42)

The average deviation was 14%. All the experiments were done with air, so there was no Prandtl number dependence. However, it was suggested that if working with fluids other than air it would be reasonable to replace the constant 0.17 by 0.17(Pr/0.7)0.33 (Li & Finlayson 1977).

2.5

Tar in producer gas

Tar is the most awkward and problematic parameter in any gasification commercialization effort. In this section, detailed discussions about tar in producer gas and different methods and approaches for tar removal are presented.

2.5.1

Tar definition

Tar has been defined in different ways by different researchers. In a recent guideline, methods for tar sampling and analysis are standardized (Neeft et al. 2002). According to that guideline, Tar is defined as follows: Generic (unspecific) term for entity of all organic compounds present in the producer gas excluding gaseous hydrocarbons (C1 through C6).

33

Chapter 2 Literature review

2.5.2

Tar quantities as a function of gasifier type

Gasifiers can be divided into three principal types from the point of view of the types of tar each produce: namely, updraft, downdraft and fluidized bed. There is general agreement about the relative order of magnitude of tar production, with updraft gasifier being the dirtiest, downdraft the cleanest and fluid beds intermediate. A very crude generalization would place updraft at 100g/Nm, fluidized beds at 10g/Nm and downdraft at 1g/Nm (Milne & Evans 1998).

2.5.3

Tar removal methods

Different approaches for tar reduction (or) elimination have been reported in the literature. All the methods available can be categorized into two groups depending on the location where tar reduction is carried out; either inside the gasifier itself (known as primary methods) or outside the gasifier (secondary methods) (Devi, Ptasinski & Janssen 2003). The following subsections describe both methods with emphasis on the primary method.

2.5.4

Tar removal methods outside the gasifier (Secondary methods)

Secondary methods (also known as downstream cleaning methods) are the conventional treatments applied to the hot producer gas. Tar reduction can be achieved by both physical (mechanical methods such as use of demisters, granular filters, electrostatic precipitators and scrubbers) and chemical (catalytic tar cracking) treatments. Although downstream gas cleaning methods are reported to be very effective in tar reduction, in some cases they are not economically viable. In addition, physical tar removal, which is mainly done through wet or wet-dry scrubbing, causes some undesirable consequences. The main problem arising from tar scrubbing is that condensed tar components are merely transferred into another phase (water or solids

34

Chapter 2 Literature review

such as scrubbing lime), which then has to be disposed of in an environmentally acceptable manner. The problems associated with the management of these wastewater or solid residues are described by Milne & Evans (1998). Secondary methods have been widely investigated and are well understood. They are not discussed further here since it is primary methods of tar reduction that are the main concern of this study.

2.5.5

Tar removal methods inside the gasifier (Primary methods)

The term primary methods encompasses all the processes that either prevent formation of tar or reduce tar levels while the producer gas is still inside the gasifier. Primary methods consist largely of chemical tar conversions. The chemical tar conversion processes can be divided into four generic categories: thermal, steam, partially oxidative, and catalytic processes (Milne & Evans 1998). Temperature is the most important parameter influencing chemical tar conversion processes. Tar can be reduced by thermal cracking at temperatures higher than 800C (Milne & Evans 1998). Studies aimed at producing a relatively clean gas by increasing its temperature have been carried out at various temperatures. Generally, the higher the temperature and the longer the residence time, the more complete will be the tar removal. Converting tar completely to lighter combustible gas requires temperatures greater than 1,100C if a catalyst is not used (Donnot, Magne & Deglise 1985). The addition of steam, over and above that formed from the water and oxygen in the feedstock, has been reported to produce fewer refractory tars, enhance phenol formation (Dayton & Evans 1997), reduce the concentration of other oxygenates (Evans et al. 1988), and has only a small effect on the conversion of aromatics (Jess 1996). Steam gasification is endothermic and hence sometimes requires a complex design for heat supply to the process.

35

Chapter 2 Literature review

When oxygen (or) air is added downstream of the pyrolysis zone, tar can be preferentially oxidized. Partial oxidation could effectively reduce tars, but contact between oxygen and tar is limited in gas producers (Kaupp, Creamer & Goss 1983). Many types of catalysts have been investigated to reduce tars to lower levels, at lower temperatures. Non-metallic catalysts such as dolomites, and metallic catalysts such as nickel (Ni), have been extensively studied. However, the duration of most reported catalyst tests has been quite short, especially considering the long activity requirements for expensive catalysts such as Ni to be economical (Milne & Evans 1998). Catalytic gas cleaning is being considered mostly for large-scale applications of biomass gasification, such as biomass gasifier- gas turbine systems. The technique appears to be rather too complicated for small power gasifier systems. There are several reports available in the literature on experimental studies of tar reduction by modifying the reactor design. All the modifications introduced have been done to enhance one or more of the above chemical tar conversion processes. These experimental investigations are discussed in the following section although, since catalytic tar conversion is beyond the scope of this study, only experimental investigations of tar reduction not involving catalysts are considered here.

2.5.6

Factors influencing chemical tar conversion in the absence of catalyst

In this subsection, experimental results showing how the tar content of producer gas is affected by various factors are discussed. These factors include the primary pyrolysis temperature, the cracking temperature, split level air supply, preheating of supplied air, partial oxidation, recycling of the pyrolysis product, superficial velocity, air supply technique and reactor type. a) Effect of primary pyrolysis temperature on tar content of raw gas In the pyrolysis reactor, thermal cracking of tar takes place to some extent but to accomplish a more efficient decomposition, higher temperatures and greater residence

36

Chapter 2 Literature review

times are desirable. It can be generally concluded, by analysing the results from experimental work using a pyrolysis reactor surrounded by a tube furnace, that a higher pyrolysis temperature produces a lower tar content in the final raw gas (Jnsson 1985). Yu et al. (1997) also performed pyrolysis experiments on birch wood in a free-fall reactor to observe the temperature effect on the process and found that increasing temperature promotes the formation of gaseous products at the expense of total tar. A more than 40% reduction in tar yield was reported when the temperature was raised from 700C to 900C.

b) Effect of cracking temperature on tar content of raw gas It is reported that thermal cracking does not occur if the cracking temperature is lower than the primary pyrolysis temperature (Black, Bircher & Chisholm 1979). A very high temperature will most probably drive the cracking towards very simple components such as CO and H2. As a consequence of this, the content of hydrocarbons will decrease at high temperatures. However, there are several other factors that limit the operating temperature. Some of the drawbacks that come with increased temperature are a lowering of the gas heating value, higher char conversion and increased risk of sintering (Milne & Evans 1998). After taking account of these and other critical factors Hallgren (1997) has recommended typical operating temperature ranges for various feed materials.

c) Two-stage gasification The two-stage gasifier is of a simple cylindrical shape, with air intakes arranged at two levels in order to have separate reaction zones. The first zone is dedicated to the flaming pyrolysis process, while the second zone is designed to be the reduction zone. Higher temperatures can be achieved by introducing a secondary air supply to the second stage. The tar vapour is mainly generated in the first zone and subsequently passes through the second zone where it can either be burned or cracked to simpler molecules due to the higher temperature in this zone.

37

Chapter 2 Literature review

Use of the second air intake has been shown by Bui, Loof & Bhattacharya (1994) to help in burning of the products from the first stage, as well as raising the reactor temperature to a maximum value that is about 100C higher than the maximum temperature for a conventional single stage reactor. The high temperature obtained in the second stage of the two-stage reactor made this stage more favourable for tar cracking as well as for gasification reactions. The lowest value of the tar content experimentally obtained was 50mg/Nm, that is 40 times less than that obtained with a single stage reactor under similar operating conditions (Bui, Loof & Bhattacharya 1994). The two-stage gasification system, however, largely depends on the stability of the two different reaction zones with their different supplied air levels. If the first stage air supply is lower than the flame sustaining limit, the flame in the first stage extinguishes and the system works as a conventional one-stage gasifier at the secondary air supply level. On the other hand, if the first-stage air supply is too high, then it could drive the combustion zone down to the second stage and again, the system will not work with two separate reaction zones as planned. Therefore, adjustment of first-stage air flow rate is a key parameter in maintaining the stability of a two-stage gasification system.

d) Effect of supplied air preheating on tar content By heating the supplied air using six electrical heaters in the form of copper tubes (2000W, 2m long) having internal insulated heating wire, Bhattacharya and Dutta (1999) investigated the effect of inlet air temperature on the final tar content of raw gas. Their gasifier was tested with different combinations of primary and secondary air flow rates, with and without the preheating option. It was reported that tar in the producer gas was significantly lower in the case where a preheated air supply was used. The low tar content obtained was a result of increasing the general temperature profile inside the gasifier due to air preheating.

38

Chapter 2 Literature review

e) Effect of partial oxidation on tar reduction The partial oxidation of biomass tar has been investigated by constructing a two-step experimental oven where biomass pyrolysis and tar destruction could be studied under well-defined conditions (Jensen, Larsen & Jorgensen 1996). It was found that adding oxygen above 700C resulted in a considerable reduction in the tar content.

f) Effect of recycling of the pyrolysis product on tar reduction Susanto (1984) conducted a study of a co-current moving bed gasifier with internal recycle and separate combustion of pyrolysis gas. The air was introduced through a pipe from the top of the gasifier. This air also acted as the motive gas in the injector for suction of the recycle gas. The combustor was installed in the discharge of the injector and was mounted on the ash grate, which could be rotated. Recycle ratios (ratio of recycled gas rate to air flow rate) in the range of 0.5 to 1.5 resulted in stable combustion of the recycled gas. A significant tar reduction was reported, from 1.4 g/Nm without recycle down to less than 0.1 g/Nm with recycle.

g) Effect of superficial velocity on low tar gas production It is reported that the superficial velocity of a gasifier, which is controlled by the supplied air rate, is the most important determinant of its performance, controlling the tar production rate (Reed et al. 1999). This conclusion was reached from an experimental investigation in which the gas velocity in an inverted downdraft gasifier with a 7.5 cm diameter was controlled. As the superficial velocity was varied from 0.05m/s to 0.26 m/s, it was found that tar in the gas decreased from 8.33 to 0.3 g/kg. It was also reported that a higher superficial velocity causes very fast pyrolysis, producing less than 10% char-ash at 1050C and hot gas temperatures of 1200-1400C in the flaming pyrolysis zone.

39

Chapter 2 Literature review

h) Effect of air supply technique and reactor type In the traditional downdraft gasifier, air is introduced to the system through nozzles attached to the side of the reactor and located just above a throat, also called a hearth. Since air is supplied to the reactor from the sides, there are regions towards the middle of the reactor to which little air penetrates and where biomass can pass through unpyrolyzed and also some pyrolysis products can pass through without cracking. If tarry gas is produced from this type of gasifier, a common practice is to reduce the hearth constriction area until a low-tar gas is produced (Reed & Agua 1988). This limits the size of gasifier and means that a throated gasifier cannot be scaled-up to a larger diameter. There have been some attempts to scale up the traditional throated gasifier and a disastrous increase in tar production has been observed as a result (Goss 1979; Graham & Huffman 1984). Attempts have been made to get round this problem. For example, Groeneveld (1980) modified the air supply system of a throated gasifier and found that a central air nozzle promotes the combustion of volatiles produced in pyrolysis. Also a new type of gasifier, known as the stratified downdraft gasifier has been developed during the last two decades (Reed & Agua 1988). The stratified downdraft gasifier consists of a cylindrical vessel in which all reaction zones exist and air is introduced to the system from the top of the reactor. This type of gasifier operates as a plug-flow reactor, and air and fuel can be assumed to be mixed uniformly. A 600 mm internal diameter stratified downdraft gasifier has been operated successfully by the Buck Rogers Co. of Kansas (Walawender, Chern & Fan 1985). The tar content reported from this 600 mm diameter reactor was found to be of the same order of magnitude as that from much smaller stratified downdraft gasifiers (75 mm ID) (Milligan 1994).

2.6

Concluding remarks

From the above review, a number of conclusions can be reached:

40

Chapter 2 Literature review

2.6.1

Theoretical modelling

1. The available literature on modelling of downdraft gasification is limited and existing steady state models mostly neglect the kinetics of pyrolysis processes. 2. To understand the behaviour of a gasifier under transient conditions, any model has to include the time dependence of the reactor parameters. 3. Two or multi-step pyrolysis mechanisms should be incorporated into any model since these can account for the thermal conversion of primary pyrolysis products into lighter combustibles, a significant phenomenon in downdraft gasification. 4. The shrinking core-reaction control model, using effective internal diffusion, is suitable for estimation of heterogeneous gas-char reaction rates inside fixed-bed gasifiers and combustors. 5. For the gasification of biomass material, the gas species involved in the water-gas shift reaction can be related by equilibrium relationships throughout the gasification zone. 6. Oxidation of pyrolysis products is important in downdraft gasification and the relevant chemical reaction processes should be included in the numerical model formulation. For pyrolysis gas (including hydrocarbons) oxidation, single or twostep global models constructed by using experimental results are available in the literature. 7. Correlations giving heat and mass transfer coefficients inside packed beds are available. However, since these are based on experimental work using non-reactive systems, a factor known as the reacting to non-reacting heat transfer ratio should be included in the gasification model. 8. Both thermal conduction and radiation processes need to be considered for formulation of the effective thermal conductivity in fuel beds, and radiation becomes the dominant process at high temperatures.

2.6.2

Tar reduction

1. Tar removal by means of external physical processes causes waste treatment problems which are usually costly to manage.

41

Chapter 2 Literature review

2. Steam has only a small influence on the conversion of aromatics in the tar and requires a complex design for steam injection. 3. Two-stage gasification and partial oxidation are considered to be attractive methods to produce low tar gas for engine applications; however, a careful adjustment of air supply rates is required to ensure the stability of the two different oxidation zones. 4. Pre-heating the supplied air is another option for tar reduction. Heat from engine exhaust gas can be utilized to provide the required external heat source but it means that additional heat exchange systems are needed. 5. Superficial velocity (air supply rate) has to be considered as an important factor in achieving low tar generation. 6. The stratified downdraft gasifier has advantages over throated gasifiers; stratified gasifiers have no problems associated with restriction of throat size, which controls the rate of thermal cracking of volatiles from the pyrolysis zone.

42

Chapter 2 Literature review

2.7

References

Acetis, JD & Thodos, G 1960, Flow of gases through spherical packings Industrial and Engineering Chemistry, vol. 52, pp. 1003-6. Adams, TN 1980, A simple fuel bed model for predicting particle emissions from a woodwaste boiler, Combustion and Flame, vol. 39, pp. 225-39. Anthony, DB and Howard, JB 1976, Coal devolatilization and hydrogasification, AIChE J, vol. 22, no. 4, pp. 625-56. Arthur, JA 1951, Reactions between carbon and oxygen, Trans Faraday Soc, vol. 47, pp. 164-78. Bhattacharya, SC & Dutta, A 1999, A two-stage gasification of wood with pre-heated air supply: A promising technique for producing gas of low tar content, paper presented to ISES 99 Solar World Congress, Jerusalem, Israel, 4-9 July. Bird, RB, Stewart, WE & Lightfoot, EN 2002, Transport phenomena, 2nd edn, J Wiley, New York. Black, JW, Bircher, KG & Chisholm, KA 1979, Fluidized bed gasification of solid wastes and biomass, paper presented to the symposium on thermal conversion of solid waste and biomass, The C.I.L program Amer Chem Soc Meeting, Washionton D.C., September 10-14. Borman, GL & Ragland, KW 1998, Combustion Engineering, Mc Graw-Hill. Boroson, ML, Howard, JB, Longwell, JP & Peters, WA 1989, Product yields and kinetics from the vapor phase cracking of wood pyrolysis tars, AIChE Journal, vol. 35, no. 1, pp. 1208. Buekens, AG & Schoeters, JF 1985, Modelling of biomass gasification, Fundamentals of Thermochemical Biomass Conversion, eds. RP Overend, TA Milne & LK Mudge, pp. 619-89. Bui, T, Loof, R & Bhattacharya, SC 1994, Multi Reactor for Biomass Gasification for Power Generation, Fuel, vol. 19, no. 4, pp. 397-404. Calderbank, PH & Pogorski, LA 1957, Heat transfer in packed beds, Transactions of Institution of Chemical Engineers, vol. 35, pp. 195-207.

43

Chapter 2 Literature review

Chan, W-CR, Kelbon, M & Krieger, BB 1985, Modeling and experimental verification of physical and chemical processes during pyrolysis of a large biomass particle, Fuel, vol. 64, pp. 1505-13. Chen, J & Gunkel, W 1988, Modeling and simulation of Co-current moving bed gasification reactors- Part II A detailed gasifier Model, Biomass, vol. 14, pp. 75-98. Chilton, TH, & Colburn, AP 1934, Mass transfer (absorption) coefficients: Prediction from data on heat transfer and fluid friction, Ind Eng Chem, vol. 26, pp. 1183-7. Dasappa, S, Paul, PJ, Mukunda, HS & Shrinivasa, U 1994, The gasification of wood-char spheres in CO2-N2 mixture: Analysis and experiments, Chemical Engineering Science, vol. 49, pp. 223-32. Dayton, DC & Evans, RJ 1997, Laboratory gasification studies via partial oxidation of biomass pyrolysis vapours, Proceedings of the 3rd biomass conference of the Americas, vol. 1, eds. Overend, RP & Chornet, E, Canada, August, pp. 673-82. DeGroot, WF & Shafizadeh, F 1984, Kinetics of gasification of Douglas fir and cotton wood char by carbon dioxide Fuel, vol. 63, pp. 210-6. Devi, L, Ptasinski KJ & Janssen FJJG 2003, A review of the primary measures for tar elimination in biomass gasification processes, Biomass and Bioenergy, vol. 24, pp. 125-40. DeWasch, AP & Froment, GF 1971, A two-dimensional heterogeneous model for fixed bed catalytic reactors, Chemical Engineering Science, vol. 26, pp. 629-34. Di Blasi, C 1993, Modelling and simulation of combustion processes of charring and noncharring solid fuels, Progress in Energy and Combustion Science, vol. 19, pp. 71-104. Di Blasi, C 2000, Dynamic behaviour of stratified downdraft gasifiers, Chemical Engineering Science, vol. 55, pp. 2931-44. Diebold, JP 1985, The cracking kinetics of depolymerized biomass vapours in a continuous, Tubular reactor, MS Thesis, Dept Chem Petroleum-Refining Eng, Colorado Sch Mines, Golden. Donnot, A, Magne, P & Deglise, X 1985, Flash pyrolysis of tar from the pyrolysis of pine bark, J Analy Appl Pyrolysis, vol. 8, pp. 401-14.

44

Chapter 2 Literature review

Dryer, FL & Glassman, I 1973, High-temperature oxidation of CO and CH4, 14th Symposium (International) on Combustion, pp. 987-1003. Evans, DD & Emmons, HW 1977, Combustion of wood charcoal, Fire Research, vol. 1, pp. 57-66. Evans, RJ, Knight, RA, Onischak, M & Babu, SP 1988, Development of biomass gasification to produce substitute fuels, Richland, WA, Pacific Northwest Laboratory, PNL-6518. Fagbemi, L, Khezami, L & Capart, R 2001. Pyrolysis products from different biomasses: application to the thermal cracking of tar, Applied Energy, vol. 69, pp. 293-306. Frenklach, M, Wang, H & Rabinowitz, MJ 1992, Optimization and analysis of large chemical kinetic mechanisms using the solution mapping method- Combustion of Methane, Progress in Energy and Combustion Science, vol. 18, pp. 47-73. Froment, GF & Bischoff, KB 1979, Chemical reactor analysis and design, J Wiley, New York. Gamson, BW, Thodos, G & Hougen, OA 1943, Heat, mass and momentum transfer in the flow of gases through granular solids, American Institute of Chemical Engineers, vol. 39, pp. 1-35. Goss, JR 1979, An investigation of the downdraft gasification characteristics of agricultural and forestry residues, Interim report, California Energy Commission P 500-79-0017. Graham, RG & Huffman DR 1984, Gasification of wood in commercial-scale downdraft gasifiers, Research paper from Forintek Canada Corp., Canada. Groeneveld, HJ 1980, The co-current moving bed gasifier, PhD Dissertation, Twente University of Technology, Enschede, The Netherlands. Groeneveld, MJ & Swaaij, WPMV 1980, Gasification of char particles with CO2 and H2O, Chemical Engineering Science, vol. 35, pp. 307-13. Gumz, W 1950, Gas Producers and Blast Furnaces, Wiley, New York. Gupta, AS & Thodos, G 1963, Direct Analogy between mass and heat transfer to beds of spheres, AIChE Journal, vol. 9, no. 6, pp. 751-4.

45

Chapter 2 Literature review

Hallgren, A 1997, Improved technologies for the gasification of energy crops Publishable final report (TPS AB), European Commission JOULE III Programme, Project no. JOR3-CT970125. Hammond, DC & Mellor, AM 1970, A preliminary investigation of gas turbine combustor modelling, Combustion Science and Technology, vol. 4, pp. 67-80. Hastaoglu, MA & Berruti, F 1989, A gas-solid reaction model for flash wood pyrolysis, Fuel, vol. 68, pp. 1408-15. Haslam, RT 1923, The simultaneous combustion of hydrogen and carbon monoxide, Industrial and Engineering Chemistry, vol. 15, pp. 679-81. Hautman, DJ, Dryer, FL, Schug & Glassman, I, 1981, A Multiple-step overall kinetic mechanism for the oxidation of hydrocarbons, Combustion Science and Technology, vol. 25, pp. 219-35. Hobbs, ML, Radulovic, FT & Smoot, LD 1992, Modelling of fixed-bed coal gasifiers, AIChE Journal, vol. 38, no. 5, pp. 681-702. Hobbs, ML, Radulovic, FT & Smoot, LD 1993, Combustion and gasification of coals in fixed-beds, Progress in Energy and Combustion Science, vol. 19, pp. 505-86. Howard, JB, Williams, GC & Fine, DH 1973, Kinetics of carbon monoxide oxidation in postflame gases, 14th Symposium (International) on Combustion, pp. 975-86. Jensen, PA, Larsen, E & Jorgensen, KH 1996, Tar reduction by partial oxidation, Biomass for Energy and the Environment, Proceeding of the 9th European Bioenergy Conference, vol. 2, Copenhagen, Denmark 24-27 June. Jess, A 1996, Mechanisms and kinetics of thermal reactions of aromatic hydrocarbons from pyrolysis of solid fuels, Fuel, vol. 75, no. 12, pp. 1441-8. Jnsson, O 1985, Thermal cracking of tars and hydrocarbons by addition of steam and oxygen in the cracking zone, Fundamentals of Thermal Conversion of Biomass, eds. RP Overend, TA Milne & LK Mudge, pp. 733-45. Kaupp, A, Creamer, K & Goss, JR 1983, The characteristics of rice hulls for the generation of electricity and shaft power on a small (5-30 Hp) scale, Energy Res., vol. 3, pp. 103-17.

46

Chapter 2 Literature review

Keating, EL 1993, Applied Combustion, Marcel Dekker, Inc. Klm, M & Joseph, B 1983, Dynamic behaviour of moving-bed coal gasifiers, Ind Eng Chem Process Des Dev, vol. 22, pp. 212-7. Kunii D & Smith JM 1960, Heat transfer characteristics of porous rocks, AIChE Journal, vol. 6, no.1, pp. 71-8. Kunii, D & Suzuki, M 1968, paper presented to Symposium on Heat and Mass transfer, Minsk. Laurendeau, NM 1978, Heterogeneous kinetics of coal char gasification and combustion, Progress in Energy and Combustion Science, vol. 4, pp. 221-70. Li, C-H & Finlayson, BA 1977, Heat transfer in packed beds-A reevaluation, Chemical Engineering Science, vol. 32, pp. 1055-66. Liden, AG, Berruti, F & Scott, DS 1988, A kinetics model for the production of liquids from the flash pyrolysis of biomass, Chemical Engineering Communication, vol. 65, pp. 207-21. Milligan, JB 1994, Downdraft gasification of biomass, PhD Dissertation, The University of Aston in Birmingham. Milne, TA & Evans, RJ 1998, Biomass gasifier tars: Their nature, formation and conversion, Golden, CO, NREL, USA, Report no. NREL/TP-570-25357. Monson, RC, Germane, GJ, Blackham, AU & Smoot, LD 1995, Char oxidation at elevated pressures, Combustion and Flame, vol. 100, pp. 669-83. Morf, P, Hasler, P & Nussbaumer, T 2002, Mechanisms and kinetics of homogenous secondary reactions of tar from continuous pyrolysis of wood chips, Fuel, vol. 81, pp. 843-53. Morf, P, Hasler, P, Hugener, M & Nussbaumer, T 2001, Characterization of products from biomass tar conversion. Progress in Thermochemical Biomass Conversion, ed. AV Bridgwater, pp. 150-61. Neeft, JPA, Knoef, HAM, Zielke, U, Sjstrm, Hasler, P, Simell, PA, Dorrington, MA, Abatzoglou, N, Deutch, S, Greil, C, Buffinga, G.J, Brage, C & Suomalainen, M 2002, Guideline for sampling and analysis of tar and particles in biomass producer gases (Version 3.3), Energy project EEN5-1999-00507 (Tar protocol).

47

Chapter 2 Literature review

Olbrich, WE & Potter, OE 1972, Mass transfer from the wall in small diameter packed beds, Chemical Engineering Science, vol. 27, pp. 1733-43. Olbrich, WE 1970, A two-phase diffusional model to describe heat transfer processes in a non-adiabatic packed tubular bed, Chemeca70: Proceedings of a conference, Melbourne and Sydney, August, 19-26, Butterworth, London, England, pp. 101. Peters, B & Bruch, C 2001, A flexible and stable numerical method for simulating the thermal decomposition of wood particles, Chemosphere, vol. 42, pp. 481-90. Petrovic, LJ & Thodos, G 1968, Mass transfer in the flow of gases through packed beds, Industrial & Engineering Chemistry Fundamentals, vol. 7, pp. 275-80. Rath, J & Staudinger, G 2001, Cracking reactions of tar from pyrolysis of spruce wood, Fuel, vol. 80, pp. 1379-89. Reed, TB & Agua, D 1988, Handbook of biomass downdraft gasifier engine systems, Biomass Energy Foundation Press, Washington. Reed, TB & Markson, M 1983, A predictive model for stratified downdraft gasification of biomass, Progress in biomass conversion, vol. 4, eds. DA Tillman & EC John, pp. 217-54. Reed, TB, Walt, R, Ellis, S, Das, A & Deutch, S 1999, Superficial velocity- The key to downdraft gasification, paper presented to 4th Biomass of the Americas Conference, September. Rossberg, M 1956, Experimental results concerning the primary reactions in the combustion of carbon, Z Elecktrochem, vol. 60, pp. 952-56. Siminski, VJ, Wright, FJ, Edelman, RB, Economos, C & Fortune, OF 1972. Research on methods of improving the combustion characteristics of liquid hydrocarbon fuels, AFAPL TR 72-74, vols. 1 & 2, Air Force Aero propulsion Lab, Wright Patterson air Force Base, Ohio. Skinner, FD & Smoot, LD 1979, Heterogeneous reactions of char and carbon, Pulverizedcoal combustion and gasification, eds. LD Smoot & DT Pratt, pp. 149-67. Standish N & Tanjung, AFA 1988, Gasification of single wood charcoal particles in CO2, Fuel, vol. 64, pp. 666-72.

48

Chapter 2 Literature review

Susanto, H 1984, Moving-bed gasifier with internal recycle and separate combustion of pyrolysis gas, PhD Dissertation, Institute of Technology Bandung. Thurgood, JR & Smoot, LD 1979, Volatiles combustion, Pulverized-coal combustion and gasification, eds. LD Smoot & DT Pratt, pp. 169-82. Turns, SR 2000, An introduction to combustion: Concepts and application, 2nd edn, McGrawHill, New York. Varma, AK, Chatwani, AU & Bracco, FV 1986, Studies of premixed laminar hydrogen-air flames using elementary and global kinetics models, Combustion and Flame, vol. 64, pp. 2336. Wakao, N & Kaguei, S 1982, Heat and mass transfer in packed beds, Gordon and Breach: New York. Walawender, WP, Chern, SM & Fan, LT 1985, Wood chip gasification in a commercial downdraft gasifier, Fundamentals of Therochemical Biomass Conversion, eds. RP Overend, TA Milne & LK Mudge, pp. 911-21. Westbrook, CK & Dryer, FL 1981, Simplified reaction mechanisms for the oxidation of hydrocarbon fuels in flames, Combustion Science and Technology, vol. 27, pp. 31-43. Whitaker, S 1977, Fundamental principles of heat transfer, Pergamon Press, New York. Yagi, S & Kunii, D 1957, Studies on effective thermal conductivities in packed beds, AIChE Journal, vol. 3, pp. 373-81. Yagi, S & Wakao, N 1959, Heat and mass transfer from wall to fluid in packed beds, AIChE Journal, vol. 5, pp. 79-85. Yu Q, Brage C, Chen G & Sjstrm K 1997, Temperature impact on the formation of tar from biomass pyrolysis in a free-fall reactor, Journal of Analytical and Applied Pyrolysis, vols. 4041, pp. 481-9.

49

Chapter 3 Mathematical model development

Chapter 3

Mathematical model development


3.1 Introduction

In this chapter, the formulation and development of a one-dimensional transient model for a stratified downdraft gasifier is presented. As discussed in the previous chapter, the stratified downdraft gasifier has a number of advantages over the traditional throated gasifier, namely: simple fabrication, uniform air and fuel flow along the axis of the gasifier, no scale up problems associated with a throat. Also, it is easier to comprehend both conceptually and mathematically. In addition, it has more facilities for measuring gas composition and temperatures within the bed, so that it is easier to compare modelling results with empirical observations. In this study, therefore, the stratified type of downdraft gasifier was selected for both theoretical and experimental investigation. In stratified downdraft gasification, solid fuel is introduced, together with air, at the top of the gasifier. It flows downwards through the gasifier, undergoing drying, pyrolysis, char oxidation, volatiles oxidation, and char gasification. Although the bed can be divided up into reaction zones, each dominated by different types of reaction, it is hard to determine where along the reactor the boundaries between adjacent zones are located. This is because many reactions occur simultaneously. In the drying and pyrolysis processes, moisture and volatiles are released from the fuel and the density of the fuel dramatically decreases. In the numerical model developed, the fuel particle size was assumed to remain constant in these zones even though in practice some shrinkage occurs during pyrolysis. It was also assumed that, in the heterogeneous combustion and gasification zone, the size of the particle decreases as predicted by the shrinking core reaction control model. The model developed incorporates the phenomena of biomass drying, pyrolysis, cracking of tars and oxidation of pyrolysis products. The model describes the complex physical and chemical processes taking place in the reactor by the use of mass and energy balances, together with information

50

Chapter 3 Mathematical model development

about rates of chemical reactions and physical transport processes. Details are presented in the following sections.

3.2

Reaction rates

3.2.1 Drying

Drying of the solid fuel is normally formulated in reactor models by one of three different approaches: it is assumed that moisture from the moist biomass or charcoal (in the case of charcoal gasification) is instantaneously evaporated before it enters the combustion and gasification zone (Basak & Bhattacharya 1987) the process of drying is considered as a diffusion limited process (Di Blasi 2000; Hobbs, Radulovic & Smoot 1992) the drying rate is assumed to be kinetically controlled (Bryden 1998; Purnomo, Aerts & Ragland 1990). In this model, drying was formulated by assuming it follows the Arrhenius type kinetics equation developed by Chan, Kelbon and Krieger (1985a):

Em Rm = (1 )Am exp RT m s
where Rm Am Em R Ts m = moisture release rate (kg/ms) = bed void fraction (void volume/bed volume) = pre-exponential factor for drying (s-1) = activation energy for drying (J/mol) = universal gas constant (J/mol/K) = solid surface temperature (K) = density of moisture (kg/m)

(3.1)

51

Chapter 3 Mathematical model development

3.2.2 Pyrolysis

A two-step pyrolysis process, as shown in Figure 3.1, was used in this model. For the primary wood pyrolysis step, a series of experiments have been conducted by various researchers and the kinetics rates obtained are available in the literature. Almost all of the experimental studies and kinetics rates, however, apply only for finely divided particles or for sample sizes less than 1.5 mm, so kinetic data which can be applied directly to the gasification process of large wood particles is scarce. In this model, the kinetic data obtained for the rapid pyrolysis of large wood particles by Chan, Kelbon and Krieger (1985b) was used:

E p1 R p1 = (1 )A p1 exp RT v avg
where

(3.2)

Rp1 Ap1 Ep1 Tavg v

= primary pyrolysis rate (kg/ms) = pre-exponential factor for primary pyrolysis (s-1) = activation energy for primary pyrolysis (J/mol) = average wood particle temperature (K) = density of volatile matter (kg/m)

When wood is decomposed by application of an external heat source, a thermal wave moves from the wood particle surface to its centre. This involves a complex combination of heat and mass transfer processes that is very hard to model accurately. There are many works on modelling of single wood particles during the pyrolysis process but no attempt to model what occurs in a particle bed was found. This meant that a single particle pyrolysis rate expression would have to be used in the present study. The expression selected was that of Chan, Kelbon and Krieger (1985b). This necessitated calculating an average particle temperature, which is a function of wood surface temperature, primary pyrolysis kinetic coefficient and particle size and which was determined using equation 3.3. This average temperature was used in equation 3.2 for estimating the primary pyrolysis rate of particle beds.

52

Chapter 3 Mathematical model development

Tavg =

1 1 2 + R ln 953.138d p + 46861.984d p T E p1 s

(3.3)

where, dp = particle diameter (m)

Equation 3.3 is derived from the correlation of Reed and Markson (1983) who determined the time required for completion of the flaming pyrolysis process and correlated this with particle size, temperature, moisture content and oxygen fraction. Details of the derivation of equation 3.3 are presented in Appendix A. Secondary pyrolysis is the process of cracking of primary tar into lower molecular weight gases and secondary tar. A limited number of experimental studies of vapour phase cracking of wood pyrolysis tar are available in the literature including several that are comparatively recent (Boroson, et al. 1989; Diebold 1985; Fagbemi, Khezami & Capart 2001; Kosstrin, 1980; Liden, Berruti & Scott 1988; Morf et al. 2001; Rath and Staudinger 2001). Although all of the kinetic parameters reported relate to unprocessed wood, the nature of the wood used as well as the experimental processes were not the same for all the investigations. Diebold (1985), Kosstrin (1980) and Liden, Berruti and Scott (1988) used data obtained from a fluidised-bed technique where both the solid and tar decomposition reactions took place simultaneously in the same reactor. Boroson et al. (1989), Morf et al. (2001) and Rath and Staudinger (2001) have studied the tar decomposition reaction separately by using a set-up consisting of two distinct reactors, the first one devoted to solid pyrolysis, the second to tar cracking under a flow of carrier gas with no packing present. Fagbemi, Khezami and Capart (2001) obtained the kinetic rates for thermal cracking of tar (in the range of 400-900C) formed by the pyrolysis of three different biomass materials (wood, coconut shell and straw). Arrhenius constants determined by Boroson et al. (1989) (see above) show very good agreement with those in a recent report (Morf et al. 2001). Moreover, the secondary pyrolysis kinetic data obtained when using two separate experimental reactors seem more appropriate than those obtained when using one reactor. Therefore, the Arrhenius

53

Chapter 3 Mathematical model development

kinetic data provided by Boroson et al. (1989) were used in this model for the secondary pyrolysis (primary tar cracking) process:

E p2 R p 2 = ( ) M tar1 A p 2 exp RT Ctar1 C


where Rp2 Ap2 Ep2 TC Ctar1 Mtar1 = primary tar cracking rate (kg/ms) = pre-exponential factor for secondary pyrolysis (s-1) = activation energy for secondary pyrolysis (J/mol) = primary tar cracking temperature (K) = molar concentration of primary tar (kmol/m) = molecular weight of primary tar (kg/kmol)

(3.4)

Tar reduction, however, can occur by both homogeneous and heterogenous thermal cracking. Since there is no data available on heterogenous tar cracking mechanisms, a means of modifying the activation energy of Borosons data was introduced in this study in order to avoid the overpredicting of tar output by considering only the homogenous tar cracking process. It was established, by a trial and error method, that multiplying Borosons activation energy values by 0.85 yielded final gas tar contents that were in a similar range to observed values. In addition, the tar cracking temperature in eq. 3.4 was taken as:

TC =

Ts + Tg 2

for for

(T

> Tg ) Tg )

TC = Tg

(T

Product yields for primary and secondary pyrolysis were chosen on the basis of literature data (Boroson et al. 1989). By carrying out an elemental mass balance using the data provided by Boroson et al. (1989), and lumping the results, a formula for the tars (primary tar) which are the product of primary pyrolysis (also known as oxygenates) was derived: C6H10.71O3.264. In addition the composition of secondary tar (hydrocarbon) was taken to be 100% benzene (C6H6), since benzene was the major

54

Chapter 3 Mathematical model development

component found in tar from downdraft gasifiers when this was analysed by direct mass spectrometry (Milne & Evans 1998). The assumption that secondary tar can be represented by benzene (C6H6) has been previously used in a wood particle decomposition model (Peters & Bruch 2001) and a coal particle combustion model (Veras et al. 1999).

Dry wood (CH1.4O0.6 ) = 100kg

Primary Pyrolysis

Char (C) = 20.5kg

Primary Tar (C6H10.71O3.264)= 52.8kg

Secondary Pyrolysis

Primary Gases* = 26.7kg CO = 3.2kg CO2 = 6.8kg CH4 = 0.4kg H2O = 16.3kg

Secondary Tar (C6H6) = 6.65kg

Secondary Gases* = 46.15kg CO = 27.62kg CO2 = 6.4kg CH4 = 5.1kg H2 = 1.63kg C2H4 = 5.4kg

(Note: the gaseous products include Acetylene (C2H2) and Ethane (C2H6) but these have been omitted

due to the very small quantities involved.)

Figure 3.1 Pyrolysis mechanism used in the numerical model and product yields [adapted from Boroson et al. (1989)]

Some uncertainty exists concerning the energetics of biomass pyrolysis as over the temperature range concerned both endothermic as well as exothermic processes have been observed to occur (Robert 1970). In this model, the heat of the primary pyrolysis process was assumed to be zero (Bryden & Ragland 1996) and a value of the heat of reaction of 50 kJ/kg was assumed for the secondary pyrolysis process (tar cracking) (Gronli & Melaaen 2000).

55

Chapter 3 Mathematical model development

3.2.3 Heterogeneous char combustion and gasification

The unreacted shrinking core model, in which reaction rate depends on film diffusion and intrinsic chemical reaction rate, was used for the heterogeneous char oxidation and gasification reactions, eq. 3.5-3.8.

C + ( )O2 k1 (2 2 )CO + (2 1)CO2 C + CO2 k 2 2CO C + H 2 O k 3 CO + H 2 C + 2 H 2 k 4 CH 4

(3.5) (3.6) (3.7) (3.8)

The heterogeneous reaction rates were estimated using eq. 3.9: this is adapted from Hobbs, Radulovic and Smoot (1993).
MC Ri = j M j Av j 1 1 + k m k i

(3.9)

where

Ei k i = Ai Ts exp RT s
Ri Ai Ei MC Mj
j

, i = 1-4, j = O2, CO2, H2O, H2,

= heterogeneous reaction rate for reaction i (kg/ms) = pre-exponential factor for heterogeneous reactions i (m/sK) = activation energy for heterogeneous reactions i (J/mol) = molecular weight of carbon (kg/kmol) = molecular weight of gaseous specie j (kg/kmol) = the stoichiometric coefficient defining the moles of oxidant required per mole of carbon, e.g. 1 = , 2 = 1, 3 = 1, 4 = 2

Av j km

= the specific surface area or the reaction surface area (m-1) = density of gaseous species j (kg/m) = mass transfer coefficient (m/s)

56

Chapter 3 Mathematical model development

Av, the specific surface area or the reaction surface area, is one of the important fuel characteristics where heat and mass transfer to the particles in the bed are concerned. This reaction surface area is related to the bed void fraction and the particle diameter as follows:
6 Av = dp (1 )

(3.10)

CO and CO2 are generally the main products of char oxidation (eq. 3.5) and the ratio of CO/CO2 increases with the reaction temperature. The ratio has been correlated by a first order Arrhenius-type relationship by different researchers for different types of char. In this study, the ratio of CO/CO2 was calculated from one of the correlations described by Evans and Emmons (1977) who correlated their experimental data for wood char by the following expression:

EC CO = AC exp RT CO2 s

(3.11)

From eq. 3.11, the stoichiometric coefficient can be estimated using


EC 2 + AC exp RT s = EC 2 + 2 AC exp RT s

(3.12)

where AC EC = pre-exponential factor for ratio of CO/CO2 (-) = activation energy for ratio of CO/CO2 (J/mol)

3.2.4 Homogeneous gas reactions

Six homogeneous reactions were considered when developing the model.

57

Chapter 3 Mathematical model development

The first of these is the well-known water-gas shift reaction while the rest involve the oxidation of volatile gases, including the secondary tar (Hydrocarbon). These are significant reactions in the downdraft gasification process. The water-gas shift reaction (eq. 3.13) is an important reaction between major gaseous species and the process is slightly exothermic. The reaction is fairly rapid over carbon surfaces at gasification temperatures and is assumed to be in equilibrium at all locations in the reaction zone. An expression for the forward water-gas shift reaction rate was proposed by Grebenshchikova (1957).

CO + H 2 O R5 CO2 + H 2
E5 F R5 F = A5 F exp RT g where R5F A5F E5F CCO CH2O = forward water-gas shift reaction rate (kmol/ms) C CO C H O 2

(3.13)

(3.14)

= pre-exponential factor for forward water-gas shift reaction (m/kmol.s) = activation energy for forward water-gas shift reaction (J/mol) = molar concentration of CO (kmol/m) = molar concentration of H2O (kmol/m)

The reverse water-gas shift reaction rate was determined using the equilibrium expression as

R5 R where R5R

E5 F A5 F exp RT g = KC

CO2

CH2

(3.15)

= reverse water-gas shift reaction rate (kmol/ms)

KC is the equilibrium constant, a function of gas phase temperature and can be found from the following correlation (Benson 1981):

58

Chapter 3 Mathematical model development

1 K C = exp RTg

11321 31.08Tg + 3Tg ln Tg 2.8 E 4Tg 2 91500 Tg

(3.16)

The next homogeneous reaction considered here is the partial oxidation of hydrocarbon (secondary tar), which is a significant reaction in downdraft gasification. Even though tar is not a single substance, treating the combustion of individual pyrolysis products (from both primary and secondary pyrolysis) separately is not practical. As mentioned in subsection 3.2.2, the mixture of products making up secondary tar was grouped and assumed to behave like benzene (C6H6). For the partial combustion of this hydrocarbon group the reaction products were assumed to be CO and H2 (Smoot & Smith 1979):

C 6 H 6 + 3O2 k 6 3H 2 + 6CO

(3.17)
0.3

E6 C R6 = A6 exp RT C6 H 6 g

) (C )T P
0.5 O2 g

(3.18)

where R6 A6 E6 CO2 P = hydrocarbon oxidation rate (kmol/ms) = pre-exponential factor for hydrocarbon oxidation [(m1.5)/(kPa0.3kmol0.5Ks)] = activation energy for hydrocarbon oxidation (J/mol) = molar concentration of O2 (kmol/m) = gas mixture pressure (kPa)

CC6H6 = molar concentration of C6H6 (kmol/m)

The other homogeneous reactions considered in this model were oxidation of gaseous products (CO, CH4, C2H4 and H2) released from primary and secondary pyrolysis processes. CO is a major gas component released during the secondary pyrolysis process. The mechanism for oxidation of carbon monoxide, proposed by Howard, Williams and Fine (1973), was used in this model.

59

Chapter 3 Mathematical model development

CO + 0.5O2 k 7 CO2
E7 (CCO )(CO 2 )0.5 C H O R7 = A7 exp 2 RT g

(3.19)

0.5

(3.20)

where R7 A7 E7 = carbon monoxide oxidation rate (kmol/ms) = pre-exponential factor for carbon monoxide oxidation (m3/kmol.s) = activation energy for carbon monoxide oxidation (J/mol)

The mechanism for oxidation of methane, in which the reaction products are CO and H2O (Dryer & Glassman 1973), was used in this model:

CH 4 + 1.5O2 k 8 CO + 2 H 2 O
E8 C R8 = A8 exp RT CH 4 g

(3.21)
0.8

) (C )
0.7 O2

(3.22)

where R8 A8 E8 CCH4 = methane oxidation rate (kmol/ms) = pre-exponential factor for methane oxidation (m1.5/kmol0.5s) = activation energy for methane oxidation (J/mol) = molar concentration of CH4 (kmol/m)

The oxidation of ethylene was considered in the study to occur by a mechanism similar to that followed by other hydrocarbons (Smoot & Smith 1979):

C 2 H 4 + O2 k 9 2 H 2 + 2CO
E9 C R9 = A9 exp RT C2 H 4 g

(3.23)
0.3

) (C )T P
0.5 O2 g

(3.24)

where R9 A9 = ethylene oxidation rate (kmol/ms) = pre-exponential factor for ethylene oxidation [(m1.5)/(kPa0.3kmol0.5Ks)]

60

Chapter 3 Mathematical model development

E9

= activation energy for ethylene oxidation (J/mol)

CC2H4 = molar concentration of C2H4 (kmol/m) The last homogenous gas reaction considered in this model is oxidation of hydrogen which is formed during partial oxidation of hydrocarbons (C6H6 and C2H4). The global kinetic expression proposed by Varma, Chatwani and Bracco (1986) was used in this study.
k 10 H 2 + 0.5O2 H 2 O

(3.25)
1.1

E10 C R10 = A10 exp RT H 2 g where R10 A10 E10 CH2 = hydrogen oxidation rate (kmol/ms)

( ) (C )
1.1 O2

(3.26)

= pre-exponential factor for hydrogen oxidation (m3.6/kmol1.2s) = activation energy for hydrogen oxidation (J/mol) = molar concentration of H2 (kmol/m)

All of the reaction rate constants used in this model are summarized in Table 3.1.

3.3

Gas-particle interchange

3.3.1 Heat and mass transfer coefficients

The solid-gas heat transfer process is characterized by a convective heat transfer coefficient, which has been experimentally determined using various experimental techniques, under either steady state or unsteady-state conditions. Recently, the quite extensive collection of data on forced convection for the flow of gases through packed beds has been critically analysed and correlated. This correlation (see below) was used in the model (Bird, Stewart & Lightfoot 2002):

61

Chapter 3 Mathematical model development

Table 3.1 List of drying, devolatilization, heterogeneous and homogenous reaction rate constants used in the mathematical model

Reaction

Mechanism

Kinetic rate constants

Reference

Drying Primary pyrolysis Secondary pyrolysis Char oxidation (Equation 3.4) Boudouard reaction (Equation 3.6) Water Gas reaction (Equation 3.7) Methane formation reaction (Equation 3.8) CO/CO2 ratio Water Gas shift reaction (Equation 3.13) Hydrocarbon oxidation (Equation 3.17) Carbon monoxide oxidation (Equation 3.19) Methane oxidation (Equation 3.21) Ethylene oxidation (Equation 3.23) Hydrogen oxidation (Equation 3.25)

Equation 3.1 Equation 3.2 Equation 3.4

Am = 5.13 106 (s-1) Em = 87,900 (J/mol) Ap1 = 3.20 10 (s ) Ep1 = 113,000 (J/mol) Ap2 = 9.55 10 (s ) Ep2 = 93,300 (J/mol) A1 = 0.685 (m/sK) E1 = 74,830 (J/mol) A2 = 589 (m/sK)
4 -1 5 -1

Chan, Kelbon & Krieger (1985a) Chan, Kelbon & Krieger (1985b) Boroson et al. (1989) Evans & Emmons (1977) Hobb, Radulovic & Smoot (1992) Yoon, Wei & Denn (1978) Hobb, Radulovic & Smoot (1992) Evans & Emmons (1977) Grebenshchikova (1957) Smoot & Smith (1979) Howard, Williams & Fine (1973) Dryer & Glassman (1973) Smoot & Smith (1979) Varma, Chatwani & Bracco (1986)

Equation 3.9 Equation 3.10

E2 = 222,825 (J/mol) A3 = 5.7144 (m/sK) E3 = 129,700 (J/mol) A4 = 3.42 10-3 (m/sK) E4 = 129,700 (J/mol) AC = 4.3 EC = 28,185 (J/mol) A5 = 0.03 (m/kmol.s) E5 = 60,279 (J/mol) A6 = 2.07 10
4.9

Equation 3.11 Equation 3.14

Equation 3.18

[(m1.5)/(kPa0.3kmol0.5Ks)] E6 = 80,230 (J/mol) A7 = 1.3 1011 (m3/kmol.s) E7 = 125,580 (J/mol) A8= 1.0 1011.7 (m1.5/kmol0.5s) E8 = 202,600 (J/mol) A9 = A9 E6 = E6 A10 =3.53108.4(m3.6/kmol1.2s) E10 = 30,514 (J/mol)

Equation 3.20 Equation 3.22 Equation 3.24 Equation 3.26

62

Chapter 3 Mathematical model development

2 2 hsg = Cp g m 2.19(Re) 3 + 0.78(Re) 0.381 (Pr) 3 g

(3.27)

where hsg Cpg mg" Re g Pr kg = solid to gas heat transfer coefficient (W/mK) = specific heat of gas mixture (J/kgK) = gas mixture mass flux (kg/ms) = Reynolds number (-) = dpmg"/g = gas mixture viscosity (kg/m.s) = Prandtl number (-) = Cpg g/kg = gas mixture conductivity (W/mK)

Mass flux from a fluid to a solid surface is given by the product of the mass transfer coefficient and the concentration driving force. By applying an empirical analogy, known as the Chilton-Colburn analogy, a correlation for mass transfer in packed beds can be derived (Bird, Stewart & Lightfoot 2002). This correlation was used in this study:
2 m 2 g 2.19(Re) 3 + 0.78(Re) 0.381 ( Sc) 3 km = g

(3.28)

where km g Sc Dg = mass heat transfer coefficient (m/s) = gas mixture density (kg/m) = Schmidt number (-) = g/ g Dg = gas mixture diffusivity (m/s)

3.3.2 Effective thermal conductivity

The two modes of heat transfer, conduction and radiation, were considered in developing the effective thermal conductivity term, since at the temperatures normally found in gasifiers, radiation becomes the most significant process, especially for a low

63

Chapter 3 Mathematical model development

conductivity solid material such as wood. The correlation proposed by Kunii and Smith (1960) was used for this model.

k eff = k g

(1 )
kg k s 1 + 1 d p hrs + kg

+ (k g + d p hrv ( )) + Re Pr k g

(3.29)

where keff ks = effective thermal conductivity (W/mK) = thermal conductivity of solid phase (W/mK)

=2/3, =0.9~1 and = 0.1 ,

1 3 hrv =Void-to-void radiation coefficient = 4Tg / 1 + 2(1 )


3 hrs = Solid radiation coefficient = 4Ts 2
= solid phase emissivity (-)

(3.30)

(3.31)

depends on the ratio of particle diameter to effective reactor diameter (DeWasch & Froment 1971).

0.14 dp 1 + 46 D
2

, where D is reactor effective diameter in metre. (3.32)

3.4

Bed-to-wall heat transfer coefficient

The effective bed-to-wall heat transfer coefficient used in this model is calculated using a correlation developed by Froment and Bischoff (1979). Heat losses to the gasifier wall from the solid and gas phases were determined using equations initially proposed by DeWasch and Froment (1971) and modified later by Hobbs, Radulovic

64

Chapter 3 Mathematical model development

and Smoot (1992). The equations for estimating the bed-to-wall heat transfer coefficient, together with their auxiliary equations, are listed below:

Bed -to-wall heat transfer coefficient1 hbw = 2.44k ro D Gas-to-wall heat transfer coefficient2 Solid-to-wall heat transfer coefficient2 Static effective radial conductivity1
d p hrv k ro = k g 1 + kg k g (1 ) + 1 1 hrs d p + + 2 kg 3k

+ 0.033k g Pr Re d p

(3.33) (3.34) (3.35)

h gw = hbw k rg / k rg + k rs hsw = hbw k rs / k rg + k rs

) )

(3.36)

Gas effective radial conductivity1


k rg d p hrv = k g 1 + kg d + 0.14 Pr Re/ 1 + 46 p D
1

(3.37)

Gas effective radial conductivity

1 h d rs p k rs = k s (1 ) / + ks
2

2 3k

( 3.38)

Packing parameter3

k 1 0.3525 2 k = 0.4569(k 1) 3k ln[k 0.5431(k 1)] k

(3.39)

where k = Conductivity ratio =

ks kg

1 2 3

Froment and Bischoff (1979) DeWasch and Froment (1971) Kunii and Smith (1960)

65

Chapter 3 Mathematical model development

3.5

Thermophysical and transport properties of solid and gas mixture

3.5.1 Solid phase

The components of the solid phase specified in this study are biomass, fixed carbon and unbound water. The physical properties of the solid feed can readily be established but, from a modelling perspective, one needs to know how the physical properties of the solid alter as it changes from its initial state to its final state. The physical and transport properties of biomass fuel were extensively reviewed by Ragland, Aerts and Baker (1991) for the purpose of combustion analysis. The relationships they developed for predicting specific heat values have been adopted here but information on other properties was obtained from different sources. The expressions used in this study for predicting temperature dependent thermophysical and transport properties of solid fuel are listed in Table 3.2, together with the references from which they were obtained.

Table 3.2 Thermophysical data and transport properties of solid fuel

Quantity

Expression (or) value

Source

Thermal conductivity (solid particle) (W/mK) Specific heat (solid) (kJ/kgK) Char emissivity (-)

= 0.13 + (3.0 x 10-4) TS = 0.1031 + 0.003867 TS = 0.85

Koufopanos and Panayannakos (1991) Ragland, Aerts and Baker (1991) Gronli and Mellaaen (2000)

3.5.2 Gas mixture

Expressions for predicting gas-phase physical properties, including heat capacity, thermal conductivity, and diffusivity are required for model formulations. These properties are all functions of temperature and composition.

66

Chapter 3 Mathematical model development

For estimation of the thermal conductivity of the gas mixture, the correlation proposed by Wayne (1983) was adopted here:

Thermal conductivity (gas mixture) [W/mK] = (4.77 x 10-4) Tg 0.717

(3.40)

The temperature dependent gas mixture viscosity used in this model is (White 1979):

Viscosity (gas mixture) [kg/m.s] = (1.98 x 10-5) x (Tg/T0)

(3.41)

where T0 is a reference temperature, taken as 300 K. Polynomial expressions for gas mixture specific heat capacities as functions of gas temperature were adapted from Rogers and Mayhew (1995) by using the values for air since the average molecular weight of the gas mixture is similar to that of air. The general form of these expressions is shown below:

Cpg = 9.41079 x 102 + 1.66918 x 10-1 Tg + 6.50858 x 10-5 Tg2 3.63930 x 10-8Tg3 (3.42)

Reid, Prausnitz and Poling (1987) proposed the following equation for estimating the diffusivity of a dilute component i in a homogenous gas mixture:

Dim = (1 X i )

(X
j i

Dij )

(3.43)

where Dij is the binary diffusion coefficient for component i diffusing into component j. For the diffusion coefficient for binary gas systems at low pressures, Reid, Prausnitz and Poling (1987) recommended the use of the equation proposed by Fuller, Schettler and Giddings (1966):

67

Chapter 3 Mathematical model development

D AB = PM

0.00143T 1.75
2 AB 1

( v )A 13 + ( v )B 13

(3.44)

where DAB T MAB P = binary diffusion coefficient (cm/s) = temperature (K) = 2[(1/MA)+(1/MB)]-1 = pressure (bar) = summation of atomic diffusion volumes (provided in Appendix B)

MA,MB = molecular weights of A and B (g/mol)

Heats of reaction for both homogenous and heterogeneous reactions were estimated from the absolute enthalpy of reactants and products as shown in the following equation:

h f (T ) =

reactants

0 i

(T )

products

0 j

(T )

(3.45)

Absolute enthalpies of gas species were calculated from the empirical coefficients provided by Turns (2000) (also given in Appendix B).

3.6

Conservation of mass

Mass balance equations were developed for several components, as described below. These describe how the mass of particular components changes with time and distance along the reactor.

68

Chapter 3 Mathematical model development

3.6.1 Solid phase component

Char is consumed by several heterogeneous reactions (eq.3.5-3.8) and its rate of disappearance is given by

(1 )

4 V C + (1 ) C S = Ri t x i =1

(3.46)

Char density along the reactor was assumed constant and it can be equated to 0.205 times the biomass density (this is taken from pyrolysis mechanism and product yield data described in Figure 3.1). This assumption eliminates the time dependent char density term and the solid velocity along the gasifier can then be estimated using eq 3.46 excluding the first term on the right hand side of the equation. The volatile matter content of wood diminishes during the primary pyrolysis process. This can be described by:
(1 ) V V + (1 ) V S = R p1 x t

(3.47)

Changes in the moisture content of the fuel also needed to be characterised. In this model, movement of liquid water within the pores was not considered and evaporation of water from the particle was assumed not to be influenced by internal diffusion processes. This assumption is reasonable for moisture levels less than the fibre saturation point in which no free water is available (Bryden 1998). The relevant mass balance equation is:
(1 ) V m + (1 ) m S = Rm x t

(3.48)

The first terms on the left hand side of equations (3.46-3.48) are the transient terms and the second terms are convective terms. The terms on the right hand side of equations (3.46-3.48) account respectively for the conversion of char to gaseous matter by

69

Chapter 3 Mathematical model development

heterogenous reactions, for loss of volatile matter to the gas phase, and for loss of moisture to the gas phase by evaporation. The rates of mass loss from particles are important when it comes to determining local heat and mass transfer coefficients. These are a function of char particle diameter, which diminishes along the combustion and gasification zone of the reactor as char is consumed. Assuming that no fragmentation, attrition or agglomeration of particles takes place, the reduction in the particle diameter can be related to the solid velocity, VS, (Cooper & Hallett 2000) and can be calculated as follows:

(1 ) C t m P

VS C + x m P

=0

(3.49)

where, m P = C

dp

3.6.2 Gas phase species

Ten gas phase species, N2, O2, CO, CO2, H2, H2O, CH4, C2H4, primary tar and secondary tar were considered in this model. The general form of the mass balances for gas species can be written as:

10 i mi 4 Mi + = i R j + i M i R j + X 1i R p1 + X 2 i R p 2 + C i R m t x MC j =1 j =5

(3.50)

where, i = O2, CO, CO2, H2, H2O, CH4, C2H4, C6H10.71O3.264, C6H6, CH2O =1 and for i H2O balance equation) X1i = (Yield of species i in primary pyrolysis) / (Total yield of volatiles) X2i = (Yield of species i in secondary pyrolysis) / (Total yield of primary tar) Ci=0 (Rate of drying, Rm, is only considered for H2O

70

Chapter 3 Mathematical model development

The first term on the left hand side of equation (3.50) is the transient term and the second term accounts for convective transport of gas. On the right hand side, the terms represent respectively the loss and formation of gaseous species involved in gas-char heterogenous reactions, gaseous homogenous reactions, primary pyrolysis reactions, secondary pyrolysis reactions, and moisture evaporation. From equation (3.50) the differential mass balance equations for the mass balance of each species can be derived: Oxygen:

O2 t + mO2 x = M O2 MC R1 3M O2 R6 0.5M O2 R7 1.5M O2 R8

(3.51)

M O2 R9 (0.5) M O2 R10

Carbon monoxide:

m M M M CO = 2(1 ) CO R + 2 CO R + CO R + 6 M R 1 2 3 CO 6 x M M M (3.52) C C C 3.20 27.62 +M M M R + 2M R +M R R R + R + R CO 8 CO 9 CO 5 R CO 5 F CO 7 79.50 p1 52.80 p 2 CO + t

Carbon dioxide:

CO2 t M CO2 MC

mCO2 x

= (2 1)

M CO2 MC

R1 + M CO2 R5 F + M CO2 R7 M CO2 R5 R

6.80 6.40 R2 + R p1 + R p2 79.50 52.80

(3.53)

Hydrogen:

H 2 t

m 2 H x

M H2 MC

R3 + M H 2 R5 F + 3 R 6 M H 2 + 2 R9 M H 2 1.63 + R p2 52.80

M H 2 R5 R 2

M H2 MC

(3.54)

R4 R10 M H 2

71

Chapter 3 Mathematical model development

Water vapour:

H 2O t M H 2O MC

m 2O H x

= 2 M H 2O R8 + M H 2O R10 + M H 2O R5 R 16.30 R p1 + Rm + 79.50

(3.55)

R3 M H 2O R5 F

Methane:

CH 4 t

mCH 4 x
C2 H 4 t

M CH 4 MC
mC2 H 4 x

R4 M CH 4 R8 +

0.40 5.10 R p1 + R p2 79.50 52.80


5.40 R p2 52.80

(3.56)

Ethylene:

= R9 M C2 H 4 +

(3.57)

Primary Tar:

tar1 mtar1 52.80 + = R p2 + R p1 x 79.50 t

(3.58)

Secondary Tar:

tar 2 mtar 2 6.65 = R6 M tar 2 + R p2 + x 52.80 t

(3.59)

In addition to the mass balances for individual species, the overall gas mass balance is required. The relevant differential equation can be written as:

g t + m g x = R p1 + Rm + Ri
i =1 4

(3.60)

3.7

Conservation of energy

The energy balance relationships include individual equations for the solid phase and for the gas phase. They are explicitly coupled by the heat exchange that occurs between the gas and the solid. Each energy balance equation consists of 6 terms, namely: transient heat transfer rate, heat transfer due to solid/gas flow, conduction through the medium (including radiative effects), convective heat exchange between the solid and gaseous phases, heat loss through the reactor wall, and heat production

72

Chapter 3 Mathematical model development

(or extraction) due to all chemical reactions (including drying). Energy balance differential equations for the solid phase and the gaseous phase are given below: Solid phase:
(1 ) T ( s hs ) + (m S hs ) = k eff s hsg Av(Ts Tg ) t x x x

(3.61)

4 4h sw (Ts Tw ) + Rm H m + R p1 H p1 + R j H j D j =1

Gaseous phase:
( ) T ( g hg ) + (mg hg ) = k g g + hsg Av(Ts Tg ) x x t x 10 4h gw (Tg Tw ) + R p 2 H p 2 + R j H j D j =5

(3.62)

3.8

The boundary and initial conditions

Operator determined boundary conditions include the inlet air composition, flow rate and temperature, and the fuel (biomass) composition and temperature at the inlet to the reactor. Both the fuel and air inlet temperature were set to be the same as the ambient temperature since the fuel bed was relatively thick and since the rate of radiative heat transfer towards the unreacted biomass at the top was found to be negligible. The gas flow rate at the top of the fuel bed was set equal to the air supply rate and the gas composition was taken to be that of air. The composition of the raw biomass at the reactor inlet was assumed to be the same as that of the incoming fuel. At the bottom of the char bed, temperatures are still high and radiation to the surroundings has to be allowed for since ash pit temperatures are likely to be lower than that of the base of the char bed due to heat losses. In this model, for simplicity, the ash pit temperature was defined as a function of the temperature at the base of the reactor bed. Since no significant reaction was observed at the bottom of the reactor, the

73

Chapter 3 Mathematical model development

temperature gradients at the bottom boundary can be defined by the heat lost to the ash pit. For the solid phase,
k eff dTs = hrs (Ts Tashpit ) dx

(3.63)

For the gas phase,


kg dT g dx = hrv (T g Tashpit )

(3.64)

Zero char loss through the grate is assumed so that Vs, the solid fuel velocity, can be set to zero at the bottom of the reactor. Before the model could be run, the initial conditions in the system needed to be specified. Gas temperatures were set equal to the inlet air temperature. It was assumed that a 500 mm depth of char was present in the reactor. The top 50 mm of char was assumed to be the combustion initiating fuel. This was necessary, since the chemical reactions are temperature dependent and some fraction of the fuel must be assumed to be at the ignition temperature. Suitable ignition temperatures range from 900 K to 1200 K. In this model an intermediate ignition temperature value of 1000 K was specified. It was assumed that wood is only added to the reactor once ignition has occurred and that, as suggested by Bhattacharya, Siddique and Pham (1999), it is supplied continuously together with air, so that the release of large amounts of tar and smoke formed as the bed is warming up can be avoided. Other parameters, such as mole fractions and particle diameter, were assumed to be uniform along the reactor at the start of the run. The initial conditions in the reactor are illustrated in Figure 3.2.

74

Chapter 3 Mathematical model development

Air

Wood

Biomass (wood), 3000K Ignited char, 10000K

20 cm 5 cm 70 cm (Effective length of reactor)

Unignited char, 3000K

45 cm

Figure 3.2 Initial conditions in model reactor

Other assumptions used in the model are: 1. The gas mixture behaves as an ideal gas. 2. Gas and solid phases move through the reactor in plug flow (i.e. temperatures can be assumed uniform in the radial direction). 3. Void fraction is assumed to remain constant at its initial value. 4. No pressure difference exists along the reactor. 5. Particles are of uniform diameter and spherical in shape. With the above assumptions, it was possible to solve the model equations numerically, as described in the next section.

3.9

Numerical methods

The sixteen partial differential equations (3 solid mass balance equations, 10 gas mass balance equations, 2 energy balance equations and the equation which relates char particle size to solid velocity) describing the transient one-dimensional downdraft

75

Chapter 3 Mathematical model development

gasifier cannot be solved analytically, and must be solved using a numerical method. In this section, step-by-step techniques used for solving the model equations are presented.

3.9.1 Method of discretization

To derive the algebraic equations from the model partial differential equations, it is necessary to select a discretization method. The control volume method proposed by Patankar (1980), was used to solve the model equations. In the control volume method, the differential equations were examined as they related to the underlying conservation principle. The calculation domain was divided into a number of nonoverlapping control volumes such that there was one control volume surrounding each grid point. The distances between successive nodes were not necessarily equal. To write a discrete form of the second order differential term, the governing equation was integrated over the control volume placed on either side of the point of interest i (Figure 3.3). Using linear interpolation functions between the grid points, the discretization procedures can be described as follows:
T T k k x i + 12 x i 12 xi

T x k x = i

1 k i + 12 (Ti +1 Ti ) k i 12 (Ti Ti 1 ) = xi xi + 1 xi 1 2 2

It is convenient to assume that i+ is located midway between i and i+1, and i- midway between i-1 and i. Thus, the discrete form becomes:

(k i +1 k i )(Ti +1 Ti ) (k i k i 1 )(Ti Ti 1 ) T x k x = i x x i + 1 + xi 1 xi 1 xi + 1 + xi 1 i+ 1 2 2 2 2 2 2

(3.65)

This form is also known as the central-difference scheme and is the natural outcome of a Taylor-series formulation (Patankar 1980).

76

Chapter 3 Mathematical model development

xi

Control volume

Ti-1 i-1 i-

Ti i i+

Ti+1 i+1

xi-

xi+

Figure 3.3 A generic node point i (non-uniform grid) and the control volume in one dimension

Using the central-difference scheme to discretize both first and second order differential terms in the diffusion-convection problem is limited to a particular case (i.e. low Reynolds numbers). That limitation can be overcome by using another scheme for discretization of the convective term while a central difference term is used for discretization of diffusion terms. In this numerical solution, an upwind-difference scheme was used for discretization. The upwind-difference scheme improves the stability when both diffusion and convective terms exist in an equation and the cell Peclet number is larger than 2. In an upwind scheme, the value of unknown T at an interface is taken to be equal to the value of T at the grid point on the upwind side of the face. In other words, a central differencing was used for the second order terms and backward differencing was used for the first order terms. The first order term in energy balance equations (eq. 3.61 & 3.62) represents convective terms, which can be discretized as follows:

(mS hS )i (mS hS )i 1 S x (m hS ) = x i 1 i
2

(3.66)

Since the model was transient, it also has to be discretized in time. For discretization of the time derivative, a backward difference approximation was used as follows:

77

Chapter 3 Mathematical model development

( g )i ( g )i g t = t i

(3.67)

The superscript, 0, denotes the previous time level. A fully implicit scheme was used for discretization of space and time, that is, the unknown variables in the spatial discretization were evaluated at the current time level of solution.

3.9.2 Source-term linearization

In the model equations, chemical reaction rates are highly nonlinear functions of temperature and also nonlinear functions of the mass fraction of gaseous species in some cases. Such nonlinear source terms cause a divergence problem during iterations and must be linearized. A linearization method in which a nonlinear term is linearized with respect to the variable being solved was recommended by Patankar (1980) and applied in this model. The method can be expressed as:
dS S = S + ( * ) d
* *

(3.68)

where is the unknown variable being solved for and the symbol * is used to denote the previous-iteration value of . The final form of the linearized equation can also be written as:

S = S C + S P

(3.69)

To apply this linearization method, there are some basic rules which must be obeyed. The most important rule is that the coefficient SP must always be less than or equal to zero since the physical situation could become unstable if SP were positive. Another requirement is that SC must always be positive. These rules must be followed while linearization is being applied to source terms in both energy balance equations and mass balance equations.

78

Chapter 3 Mathematical model development

For the energy balance equations, it should be noted that only reaction source terms with negative (endothermic) relations gave a negative SP. Thus, in the solid phase energy balance equation (eq. 3.61), heat source terms concerned with drying, primary pyrolysis and all heterogeneous reactions were linearized while no linearization is required for source terms of the gaseous phase energy balance equation (eq. 3.62) since all related reactions for the gas phase are exothermic. Similarly, in the case of mass balance equations, linearization was done only for the nonlinear reactions in which the unknown gas species being solved for was a reactant since only the terms associated with consumption rates produce a negative SP. The step-by-step linearization of reaction source terms for solving the mass and energy balance equations is presented in Appendix C.

3.9.3 Underrelaxation

Because of the strongly nonlinear nature of the equations, rapid changes can occur in the magnitude of source terms between successive iterations. This can cause divergence problems, making it desirable to reduce the extent of the changes in the values of dependent variable between successive iterations. This process is known as underrelaxation, which is a very useful device for strongly nonlinear problems to achieve stability of the solution. There are many ways of performing underrelaxation. In the underrelaxation method used in this study, the source term can be underrelaxed using the relationship

S C = ( )S C + (1 )S * C

(3.70)

where is the underrelaxation factor, which is less than unity, and the superscript, *, denotes the value of the source term from the previous iteration.

79

Chapter 3 Mathematical model development

In this gasifier model the exponential nature of the kinetic rate equations means that rapid changes in heat release rates can occur, leading to instabilities in the computation. To obviate this problem, underrelaxation was carried out for all source terms related to gas oxidation reactions. There are no general rules for choosing the best value of . The optimum value depends upon a number of factors, such as the nature of the problem, the number of grid points, the grid spacing, and the iterative procedure used. Usually, a suitable value of can be found by experience and from exploratory computations for the particular problem (Patankar 1980). A small value of improves the stability of the iteration but increases the computing time. A large value of enables the program to be run faster but it can result in a failure to converge. Exploratory tests showed that an value of 0.3 is the most suitable for use in the computer program developed.

3.9.4 Model discretization

This subsection presents the algebraic equations in their final form, after discretization, which was carried out together with linearization and underrelaxation, of the model partial differential equations (eq. 3.46-3.62). The step-by-step formulations of algebraic equations are presented in Appendix C. For mass balance equations, as discussed in subsection 3.9.1, both the space and time derivatives were discretized by a backward-difference scheme since all mass balance equations are in a first order differential form. Char: By eliminating the transient char density terms of equation 3.46, the solid velocity was estimated as follows:
VS i 1 = BVS i + C A

(3.71)

80

Chapter 3 Mathematical model development

where

A = (1 ) c B = (1 ) c
C = (R1 + R2 + R3 + R4 ) xi 1
i 2

It should be noted that solid velocity was calculated working from the bottom to the top of the reactor whereas other parameters were estimated working from the top to the bottom.
Volatile matter:

After discretization was carried out on eq. 3.47, the volatile matter density of the solid fuel can be estimated by

V i =
where
A= x i 1 t
2

B V i 1 + C A

(3.72)

+ (V S )i + (k p1 )i xi 1

B = (VS )i 1
C=

( V )i0 xi 1
t

Moisture content:

Moisture density of biomass fuel was calculated via a discrete form of equation 3.48.

mi =

B m i 1 + C A

(3.73)

81

Chapter 3 Mathematical model development

where
A= x i 1 t
2

+ (VS )i + (k m )i xi 1

B = (VS )i 1
C=

( m )i0 xi 1
t

Particle diameter:

The particle diameter of the solid fuel in each horizontal slice through the reactor can be determined by the following equation discretized from eq. 3.49.
3 xi 1 + (Vs )i t 2 = xi 1 (V ) t 2 + s i 1 d 3 0 d 3 i 1 i
1

(d )i

(3.74)

( )

( )

By following a similar approach to that used for the solid mass balance equation, gaseous mass balance equations can be written in the general form

i =

B i 1 + C A

(3.75)

where is the molar fraction of the gaseous species concerned. Since all of the reaction rate equations were originally derived in terms of mass concentrations, these equations needed to be rewritten in terms of molar fractions. This was done using the ideal gas law:

j =

pjM j RTg

(3.76)

where, p j = PY j and Y j = mole fraction of species j

82

Chapter 3 Mathematical model development

Reaction rate equations rewritten with molar fraction terms are presented in Appendix C. Given below are the equations used in the model for each gaseous species and the total gas mixture. These were obtained by discretization of the model partial differential equations (eq. 3.51- 3.60).

Oxygen:

(YO 2 )i
where
( ) PM O2 xi 1 2 A= RTg t

B(YO 2 )i 1 + C A

(3.77)

+ m M O2 g Mg i

M + ( ) O2 MC i

Av M C RT g

Pxi 1 2 1 1 + k m k1
* (1.8) M O2 R8 xi 1 * YO2 2

+ (3) M O2 ( )k 6Tg0.5 P 1.8YC06.5 6 xi 1 + H


2

* (0.6) M O2 R7 xi 1 * YO2

+ M O2 ( )k 9T

0.5 g

P Y

1.8

0.5 C6 H 6

xi 1 +
2

* (0.55) M O2 R10 xi 1

* O2

M O2 B = m g M g

i 1

( )YO2 PM O2 xi 1 2 C = RTg t

* * + (0.1) M R * x + (0.3) M O2 R8 xi 1 + (0.05) M O2 R10 xi 1 O2 7 i 1 2 2 2 i

Carbon monoxide:

(YCO )i
where
( ) PM CO xi 1 2 A= RTg t

B (YCO )i 1 + C A

(3.78)

+ m M CO g Mg i
2

+ M CO ( )k 5 P 2Tg 2YH O x 1 i 2 2 i

0 0 + M CO ( )k 7 P 2Tg 2YO2.5YH .25 xi 1 O

83

Chapter 3 Mathematical model development

M B = m CO g Mg

i 1
0

( )YCO PM CO x i 1 M 3.20 27.62 2 + R p1 x i 1 + R p 2 x i 1 + 2(1 ) CO R1 x i 1 C = 2 2 2 29.50 52.80 MC RTg t i M M + (2) CO R2 x i 1 + CO R3 x i 1 + M CO R5 R x i 1 + 6 M CO R6 x i 1 2 2 2 2 MC MC + M CO R8 x i 1 + 2 M CO R9 x i 1


2 2

Carbon dioxide:

(YCO 2 )i
where

B(YCO 2 )i 1 + C A

(3.79)

( ) PM CO2 xi 1 M + m CO2 A= g M RTg t g i E5 F A5 F exp RT g + M CO2 KC


M CO2 B = m g Mg i 1
0

M CO2 + MC i

Av M C RT g

Pxi 1 2 1 1 + k m k 2

P 2T 2Y x g H2 i 1

( )YCO2 PM CO2 xi 1 M CO2 6.80 6.40 2 C = R p1 xi 1 + R p 2 xi 1 + (2 1) + R1 xi 1 2 2 2 RTg t 79.50 52.80 MC i + M CO2 R5 F xi 1 + M CO2 R7 xi 1
2 2

Hydrogen:
B(YH 2 )i 1 + C A

(YH 2 )i

(3.80)

84

Chapter 3 Mathematical model development

where
( ) PM H 2 xi 1 2 A= RTg t + m M H 2 g Mg i P 2T 2Y
g

M + 2 H2 MC i

Av M C RT g

Pxi 1 2 1 1 + k m k 4

E5 F A5 F exp RT g + M H2 KC
M H2 B = m g Mg i 1

CO2

xi 1 +
2

* 1.1M H 2 R10 xi 1 * YH 2

( )YH 2 PM H 2 xi 1 2 C = RTg t
2

M H2 + 1.63 R x + R3 xi 1 + M H 2 R5 F xi 1 + 3R6 M H 2 xi 1 p2 i 1 2 2 2 2 52.80 MC i


2 2

* * + 2 R9 M H 2 xi 1 M H 2 R10 xi 1 + (1.1) M H 2 R10 xi 1

Water vapour:

(YH 2O )i
where
( ) PM H 2O x i 1 2 A= RTg t

B(YH 2O )i 1 + C A

(3.81)

+ m M H 2O g Mg i

M H 2O + MC i

Av M C RT g

Px i 1 2 1 1 + k m k 3

+ M H 2O ( )k 5 F P 2Tg 2YCO x i 1

M H 2O B = m g M g i 1 ( )YH 2O PM H 2O xi 1 16.30 2 C = + R m x i 1 + R p1 xi 1 + M H 2O R5 R xi 1 2 2 2 RTg t 79.50 i + 2 M H 2O R8 xi 1 + M H 2O R10 xi 1


2 2

85

Chapter 3 Mathematical model development

Methane:

(YCH 4 )i
where
( ) PM CH 4 xi 1 2 A= RTg t
M CH 4 B = m g Mg i

B(YCH 4 )i 1 + C A

(3.82)

+ m M CH 4 g Mg i

* (1.1) M CH 4 R8 xi 12 + * YCH 4 i

( )YCH 4 PM CH 4 xi 1 2 C = RTg t
* + (0.1) M CH 4 R8 xi 1 2

M CH 4 5.10 + 0.40 R x R p 2 xi 1 + R4 xi 1 p1 1 + i 79.50 2 2 2 52.80 MC i

Ethylene:

(YC 2 H 4 )i =
where
( ) PM C2 H 4 xi 1 2 A= RTg t

B(YC 2 H 4 )i 1 + C A

(3.83)

+ m M C2 H 4 g Mg i

M R* + 1.2 C2 H 4 9 x 1 i 2 YC*2 H 4 i

M C2 H 4 B = m g M g

i1
* + 5.40 R x + 0.2M C2 H 4 R9 xi 1 p2 i 1 2 2 52.80 i
0

( )YC2 H 4 PM C2 H 4 xi 1 2 C = RTg t

Primary tar:
B(Ytar1 )i 1 + C A

(Ytar1 )i

(3.84)

86

Chapter 3 Mathematical model development

where
( ) PM tar1 xi 1 2 A= RTg t + m M tar1 + ( ) PT 1 k x g p2 i 1 g Mg 2 i i

M B = m tar1 g M g i 1
( )Ytar1 PM tar1 xi 1 2 C = RTg t + 52.80 R x p1 i 1 2 79.50 i
0

Secondary tar:
B(Ytar 2 )i 1 + C A

(Ytar 2 )i
where
( ) PM tar 2 xi 1 2 A= RTg t

(3.85)

+ m M tar 2 g Mg i

* + (1.2) M tar 2 R6 x 1 * i 2 Ytar 2 i

M B = m tar 2 g M g

i 1
* + 6.65 R x + 0.2 M tar 2 R6 xi 1 p2 i 1 2 2 52.80 i
0

( )Ytar 2 PM tar 2 xi 1 2 C = RTg t

Total gas:

mg i =
where
A =1 B =1

Bmg i 1 + C A

(3.86)

( ) PM g xi 1 ( ) PM g xi 1 2 2 C = (R p1 + Rm + R1 + R2 + R3 + R4 )xi 1 + 2 RTg t RTg t i i

87

Chapter 3 Mathematical model development

For energy balance equations, discretization was carried out by using a central difference scheme for the second order terms and a backward difference scheme for the first order terms as discussed in subsection 3.9.1. After linearization and underrelaxation were carried out, the solid and gas phase energy balance differential equations (eq. 3.61 & 3.62) became:

Solid phase:

Ai (Ts )i = Bi (Ts )i +1 + Ci (Ts )i 1 + Di


where

(3.87)

(K ) + (K ) (K ) + (K ) (1 ) eff i +1 eff i eff i eff i 1 ( mCpm + bioCpbio + char Cpchar )i + 2x x + 2x x t i i i+ 1 i 1 2 2 (1 )( mCpm + bioCpbio + char Cpchar )i (vs )i 4(hsw )i + + (hsg Av )i + D xi 1 Ai =
2

(K ) + (K eff i +1 eff Bi = 2x x i i+ 1 2

)
i

(K ) + (K ) (1 )( Cp + Cp + Cp ) (v ) eff i eff i 1 m m bio bio char char i 1 s i 1 Ci = 2x x + xi 1 i i 1 2 2


(1 ) (( mCpm + bio Cpbio + char Cpchar )Ta )i + (1 ) (( mCpm + bio Cpbio + char Cpchar )Ts )i0 t t (1 )(( m Cp m + bio Cpbio + char Cp char )vsTa )i (1 ) (( mCpm + bio Cpbio + char Cpchar )Ta )i0 + t xi 1
2

Di =

(1 )(( m Cp m + bio Cpbio + char Cp char )vsTa )i 1 4 + (hsg AvTg ) + (TW hsw )i xi 1 D
2

+ Rm H m + R p1H p1 + (R j H j )i
4 j =1

88

Chapter 3 Mathematical model development

Gas phase: Ai (Tg )i = Bi (Tg )i +1 + Ci (Tg )i 1 + Di where (K ) + (K ) (K ) + (K ) (m Cp ) 4(hgw )i g i g i 1 g i g i +1 g g i + + + + (hsg Av ) + 2x x 2x x D xi 1 i i i 1 i+ 1 2 2 2 (3.88)

Ai =

( Cp )
g g

(K ) + (K ) g i g i +1 Bi = 2x x i i+ 1 2 (K ) + (K ) (m Cp ) g i g i 1 g g i 1 Ci = + 2x x xi 1 i i 1 2 2
Di =

Cp g Ta )i +

Cp g Tg )i
0

Cp g Ta )i +
0

(m Cp T ) (m Cp T )
g g a
i

+ (hsg Av Ts )i +

4(hgw )i Tw D

xi 1

i 1

xi 1

+ (R p 2 H p 2 )i + (R5 F H 5 F )i + (R5 R H 5 R )i
*

10 10 + ( ) R j H j + (1 ) R j H j j =6 i j =6 i

Discretization equations for the energy balance equations (eq. 3.87 & 3.88) were solved by applying the Thomas algorithm, which is also known as TDMA (TriDiagonal-Matrix Algorithm). The tridiagonal-matrix algorithm is a very powerful and convenient equation solver whenever the algebraic equations can be represented in the form of three diagonals of the matrix along which all the nonzero coefficients line up. The step-by-step solution procedures of TDMA are available in Patankar (1980).

3.9.5 Adaptive grid method

The reaction rates of homogenous gas phase oxidation processes included in this numerical model were controlled by the chemical kinetics expressions, which are

89

Chapter 3 Mathematical model development

exponential functions. This resulted in a very narrow zone of sharp gradients the length of which was of the order of millimetres while the effective reactor thickness was of the order of metres. In addition, the reaction front cannot be located at the same position in the transient condition, since it moves upward or downward according to the input operational parameters. Because of this, if an equal grid size method were used, then to retain the accuracy and stability of the solution, it would be necessary to use very small grids along the whole length of the reactor and a very lengthy computation time would be required. This problem can be overcome by using computational grids whose nodes are highly nonuniformly distributed in space: a small mesh spacing was imposed within the oxidation zone, whereas a coarse grid was used outside the combustion zone, where the gradients of variables are not as steep. This procedure is known as adaptive gridding, which is an efficient method to resolve the sharp gradients accurately while avoiding wasting computer time by over-resolving the smooth-solution regions. Varieties of adaptive mesh methods are available in the literature. In the numerical solution used in this study, a global static rezone method based on the equidistribution principle was used. There are two main parts of the global static rezone: the first part is to find a new grid using the old (current) grid and the old solution, and the second part is to interpolate the solution from the old grid to the new one. In the static rezone method, by applying the equidistribution principle, a new grid can be computed satisfying

xi +1 W = constant xi
where W is the mesh monitor function.

(3.89)

The mesh monitor function can be defined by different approaches (i.e. based on the local truncation errors, the arc-length, the curvature and the combined arc-length and curvature). In this study, an arc-length monitor was used; this can be defined as

90

Chapter 3 Mathematical model development

1 Wi = 1 + N

j =1

(U

2 U i, j ) (xi +1 xi )2 i +1, j

(3.90)

where N is the number of partial differential equations and U is the value of the variable concerned. It was not necessary to use all of the partial differential equations of the gasifier model to estimate the mesh monitor. In this numerical solution, only gas temperature was used to find the mesh monitor since the steep gas temperature gradients were the most important parameters causing the thin combustion zone. The new mesh satisfying eq. 3.89 and 3.90 was determined using the inverse interpolation method, which is comprehensively discussed by Hyman and Naughton (1983). Once the new grid is known, it remains to define the values of the variables at the new points. There exist several ways of projecting the old solution onto the new grid (i.e. linear interpolation, conservative interpolation, cubic Hermite interpolation, monotone cubic Hermite interpolation, and quintic Hermite interpolation.) The simplest one consists in linearly interpolating the variables on each mesh of the old grid. This rather rough procedure naturally introduces a certain loss of information (Larrouturou 1986). The monotone cubic Hermite interpolation, which was introduced by Fritsch and Carlson (1980), is recommended for problems with very steep wave fronts. So, this later interpolation method was used in this study for estimating the solution at new grid points.

3.9.6 Solution procedure

The sixteen unknowns (densities of moisture and volatile matter; mass flux of total gas mixture; solid velocity; particle diameter; molar fractions of O2, CO, CO2, H2O, H2, CH4, C2H4, primary tar and secondary tar; temperature of solid phase and gas phase)

91

Chapter 3 Mathematical model development

were solved for by using the 16 algebraic linear equations (eq. 3.71-3.74, 3.77-3.88) derived from the governing mass and energy balance partial differential equations. Even after linearization has been carried out, the solution of the algebraic linear equations is not straightforward because of the presence of several parameters which are dependent on each other. For example, the reaction rate is a function of the solid and gas temperatures but these temperatures themselves are dependent on the amount of heat released in the reactions. Another such parameter is the effective thermal conductivity, Keff, which can be determined only if the solid temperature at each grid point is known, but for calculating the solid temperature the value of Keff is required. Therefore, the solution had to be obtained by a trial and error method with iterative steps. The boundary and initial conditions were set as discussed in section 3.8. Before computation started, the other operational parameters needed to be defined. These were air supply rate, biomass fuel properties (initial particle diameter, density, moisture content, void fraction) and reactor dimensions (effective length and diameter of the reactor). The computation started at the top of the reactor since all boundary values except VS were defined at the top of the reactor. For the solid velocity, the calculation started from the grate where VS was equal to zero. The detailed iteration procedures were as follows: 1. After setting the initial conditions, calculations started using the defined initial temperature profiles for the temperature dependent properties, such as reaction rates and mass transfer coefficients; these have to be known to solve the mass balance equations. 2. The local velocities of the solid were estimated starting from the bottom of the reactor by using heterogeneous reaction rates. 3. The local particle diameters of the solid from the top of the reactor were determined by using the estimated local velocities.

92

Chapter 3 Mathematical model development

4. The mass balance equations (solid phase) were solved and the moisture density and volatile matter density in each control volume calculated. 5. The molar fraction of the 9 different gas species was estimated starting with the computation of the total mass flux since it appears in every mass balance equation. 6. The other temperature dependent properties, such as wall heat loss coefficient and effective conductivity were calculated, as these were required for solving energy balance equations. 7. The solid and gas temperatures were calculated by solving energy balance equations with TDMA. 8. The new solid and gas temperature profiles were compared with previous ones to check how close they were to one another; the process then returned to step 1 using the new temperature values and repeated the process. 9. After a considerable number of iterations, the difference between the old temperature profile and new temperature profile diminished to less than 0.01%, the preset convergence criterion. The computations were then stopped for that current time step. Since a static rezone procedure was employed, before starting the calculation for next time step, a new mesh was computed using the old mesh and old solutions, and variables were also redefined on the new mesh. Then, calculation for the next time step commenced. The new grid and new temperature profiles were used as the starting values for this new time step. The size of the time step used in this model was determined by trial and error. A small time step improves the accuracy of the transient solution but increases the computing time. A large time step enables the program to be run faster but it could result in a failure to converge. The range of suitable time steps determine for this model was between 1 and 4 seconds. Tests show that a 3 second time step was the most suitable for use in the computer program developed. The computational algorithm used to obtain the solution is shown in Figure 3.4 as a flow chart. A copy of the computer code in FORTRAN 90 that is used in implementing the discretized model equations is listed in Appendix D.

93

Chapter 3 Mathematical model development

Start

Enter input data

Solve for energy balance equations by TDMA: Calculate the solid and gas temperatures

Start calculation for next time step Calculate the values of heat transfer coefficient and effective conductivity Calculate reaction rates, mass transfer coefficients no Calculate solid velocities (from bottom to top) (Tnew-Told)/Tnew < 10-4 yes Calculate particle diameters (from top to bottom) Find the new mesh and define the solutions on new grid points

Solve for mass balance equation (solid phase): Find moisture & volatile matter densities

no

time = output time yes

Solve for mass balance equations (gas phase): Estimate the molar fraction of gas species

Write up of results

End

Figure 3.4 Flow chart for the numerical solution of stratified downdraft gasifier

94

Chapter 3 Mathematical model development

3.10 Conclusions

A transient model has been developed for a stratified downdraft wood gasifier by applying a two-step pyrolysis mechanism in which primary tars (oxygenates) are initially formed and then cracked into secondary tars (hydrocarbons) and other combustible gases. The model is, therefore, capable of predicting the primary and secondary tar profile along the gasifier as well as the gas composition profile and temperature profile along the axis of the gasifier. The model describes the complex physical and chemical processes taking place in the reactor by the use of mass and energy balances, together with information about rates of chemical reactions and physical transport processes. The factors taken into account in this model were temperature differences between the gas and solid surface, the particle shrinkage during heterogeneous reaction, the effect of reaction temperature on the ratio of (CO/CO2) for heterogeneous char combustion, and the effective thermal conductivity, which included the importance of radiation at a high reactor temperature. It also included the oxidation of volatile gases and secondary tars (hydrocarbon). Temperature dependent thermophysical and transport properties of the solid and gas mixture were applied in this model. The control volume method proposed by Patankar (1980), was used to solve the model governing equations. The discretization equations were derived from the partial differential equations by applying the fully implicit scheme for time and the upwind scheme for space. Since the reaction source terms were highly nonlinear, these were linearized to achieve a converged solution. As well as this, to avoid the rapid temperature changes in the iterative solution of strongly nonlinear equations, underrelaxation, which slows down the changes from iteration to iteration, was applied. To solve the very steep temperature and mole fraction gradients at the oxidation zone, the adaptive method (static rezone method) was used to produce computational grids whose nodes are highly non-uniformly distributed in space. The results and sensitivity analysis of the model developed are described in the following chapter.

95

Chapter 3 Mathematical model development

3.11 References
Basak, AK & Bhattacharya, SC 1987, Performance of a down-draft charcoal gasifier, Applied
Energy, vol. 26, no. 3, pp. 193-216.

Benson, HE 1981, Chemistry of coal utilisation, 2nd supplementary volume, ed. MA Elliott, John Wiley. Bhattacharya SC, Siddique AHM, & Pham H-L 1999, A study on wood gasification for lowtar gas production, Energy- The International Journal, vol. 24, pp. 285-96. Bird, RB, Stewart, WE & Lightfoot, EN 2002, Transport phenomena, 2nd edn, J Wiley, New York. Boroson, ML, Howard, JB, Longwell, JP & Peters, WA 1989, Product yields and kinetics from the vapor phase cracking of wood pyrolysis tars, AIChE Journal, vol. 35, no. 1, pp. 1208. Bryden, KM 1998, Computational modeling of wood combustion, PhD Dissertation, University of Wisconsin-Madison. Bryden, KM & Ragland, KW 1996, Numerical modeling of a deep, fixed bed combustor,
Energy and Fuel, vol. 10, pp. 269-75.

Chan, W-CR, Kelbon, M & Krieger, BB 1985a, Modeling and experimental verification of physical and chemical processes during pyrolysis of a large biomass particle, Fuel, vol. 64, pp. 1505-13. Chan, W-CR, Kelbon, M & Krieger, BB 1985b, Product formation in the pyrolysis of large wood particles, Fundamentals of thermochemical biomass conversion, eds. RP Overend et al., pp. 219-36. Cooper, J & Hallett, WLH 2000, A numerical model for packed-bed combustion of char particles, Chemical Engineering Science, vol.55, pp. 4451-60. DeWasch, AP & Froment, GF 1971, A two-dimensional heterogeneous model for fixed bed catalytic reactors, Chemical Engineering Science, vol. 26, pp. 629-34.

96

Chapter 3 Mathematical model development

Diebold, JP 1985, The cracking kinetics of depolymerized biomass vapours in a continuous, Tubular reactor, M.S Thesis, Dept. Chem. Petroleum-Refining Eng., Colorado Sch. Mines, Golden. Di Blasi, C 2000, Dynamic behaviour of stratified downdraft gasifiers, Chemical Engineering
Science, vol. 55, pp. 2931-44.

Dryer, FL & Glassman, I 1973, High-temperature oxidation of CO and CH4, 14th Symposium
(International) on Combustion, pp. 987-1003.

Evans, DD & Emmons, HW 1977, Combustion of wood charcoal, Fire Research, vol. 1, pp. 57-66. Fagbemi, L, Khezami, L & Capart, R 2001, Pyrolysis products from different biomasses: application to the thermal cracking of tar, Applied Energy, vol. 69, pp. 293-306. Fritsch, FN & Carlson, RE 1980, Monotone piecewise cubic interpolation, Siam J Numer
Anal, vol. 17, no. 2, pp. 238-46.

Froment, GF & Bischoff, KB 1979, Chemical reactor analysis and design, Wiley, New York. Fuller, EN, Schettler, PD & Giddings, JC 1966, A new method for prediction of binary gasphase diffusion coefficients, Industrial & Engineering Chemistry, vol. 58, no. 5, pp. 18-27. Grebenshchikova, GB 1957, Study of the kinetics of the conversion of carbon monoxide by steam in the pressure of ash from Lisichansk Coal, Podzemnaya Gazifikatsiya Uglei: no. 2, pp. 54-57, (Lawrence Livermore Laboratory Translation UCRL-TRANS-10900). Gronli, MG & Mellaaen, MC 2000, Mathematical model for wood pyrolysis- Comparison of experimental measurements with model predictions, Energy & Fuels, vol. 14, pp. 791-800. Hobbs, ML, Radulovic, FT & Smoot, LD 1992, Modelling of fixed-bed coal gasifiers, AIChE
Journal, vol. 38, no. 5, pp. 681-702.

Hobbs, ML, Radulovic, FT & Smoot, LD 1993, Combustion and gasification of coals in fixed-beds, Progress in Energy and Combustion Science, vol. 19, pp. 505-86. Howard, JB, Williams, GC & Fine, DH 1973, Kinetics of carbon monoxide oxidation in postflame gases, 14th Symposium (International) on Combustion, pp. 975-86.

97

Chapter 3 Mathematical model development

Hyman, JM & Naughton, MJ 1983, Static rezone methods for tensor-product grids, Report LA-UR-83-3245, Los Alamos National Laboratory, Los Alamos, NM. Kosstrin, H 1980, Direct formation of pyrolysis oil from biomass, Proceedings of specialists
workshop on fast pyrolysis of biomass, Copper Mountain, Colorado, 19-22 October, pp. 105-

21. Koufopanos, CA & Panayannakos, N 1991, Modelling of the pyrolysis of biomass particles: Studies on kinetics, thermal and heat transfer effects, The Canadian Journal of Chemical
Engineering, vol. 69, pp. 907-15.

Kunii, D & Smith, JM 1960, Heat transfer characteristics of porous rocks, AIChE Journal, vol. 6, no.1, pp. 71-78. Larrouturou, B 1986, Adaptive numerical methods for unsteady flame propagation, Lectures
in Applied Mathematics, vol. 24, pp. 415-35.

Liden, AG, Berruti, F & Scott, DS 1988, A kinetics model for the production of liquids from the flash pyrolysis of biomass, Chemical Engineering Communication, vol. 65, pp. 207-21. Milne, TA & Evans, RJ 1998, Biomass gasifier tars: Their nature, formation and
conversion, Golden, CO: National Renewable energy Laboratory, the U.S Department of

Energy. Morf, P, Hasler, P, Hugener, M & Nussbaumer, T 2001, Characterization of products from biomass tar conversion. Progress in thermochemical biomass conversion, ed. AV Bridgwater, pp. 150-61. Pantankar, SV 1980, Numerical heat transfer and fluid flow, New York: McGraw-Hill. Peters, B & Bruch, C 2001, A flexible and stable numerical method for simulating the thermal decomposition of wood particles, Chemosphere, vol. 42, pp. 481-90. Purnomo, Aerts, DJ & Ragland, KW 1990. Pressurized downdraft combustion of wood chips,
Twenty-third International Symposium on Combustion, The combustion Institute, pp. 1025-32.

Ragland, KW, Aerts, DJ & Baker, AJ 1991, Properties of wood for combustion analysis,
Bioresources Technology, vol. 37, pp. 161-8.

98

Chapter 3 Mathematical model development

Rath, J & Staudinger, G 2001, Cracking reactions of tar from pyrolysis of spruce wood, Fuel, vol. 80, pp. 1379-89. Reed, TB & Markson, M 1983, A predictive model for stratified downdraft gasification of biomass, Progress in biomass conversion, vol. 4, eds. DA Tillman & EC John, pp. 217-54. Reid, RC, Prausnitz, JM & Poling, RE 1987, The properties of gases and liquids, 4th edn, McGraw-Hill. Roberts, AF 1970, A review of kinetics data for the pyrolysis of wood and related substances,
Combustion and Flame, vol. 14, pp. 261-72.

Rogers, GFC & Mayhew, YR 1995, Thermodynamic and transport properties of fluids: SI
Units, 5th edn, Black Well, Oxford, UK.

Smoot, LD & Smith, PJ 1979, Volatiles combustion, Pulverized-coal combustion and


gasification, eds. LD Smoot & DT Pratt, pp. 169-82.

Turns, SR 2000, An introduction to combustion: Concepts and application, 2nd edn, McGrawHill, New York. Varma, AK, Chatwani, AU & Bracco, FV 1986, Studies of premixed laminar hydrogen-air flames using elementary and global kinetics models, Combustion and Flame, vol. 64, pp. 2336. Veras, CAG, Saastamoinen, J, Carvalho JR, JA & Aho, M 1999, Overlapping of the devolatilization and char combustion stages in the burning of coal particles, Combustion and
Flame, vol. 116, pp. 567-79.

Wayne, WS 1983, Analysis of single particle wood combustion in convective flow, PhD Thesis, Department of mechanical engineering, University of Wisconsin- Madison. White, FM 1979, Fluid Mechanics, McGraw-Hill, New York. Yoon, H, Wei, J & Denn, MM 1978, A model for moving-bed coal gasification reactors,
AIChE Journal, vol. 24, no. 5, pp. 885-903.

99

Chapter 4 Model results and discussions

Chapter 4 Model results and discussions

4.1

Introduction

Chapter Three described the development of a model that is capable of predicting the transient performance of a stratified downdraft gasifier over different ranges of operating conditions. The model can predict the gas composition, primary and secondary tar contents, and gas and solid surface temperature profiles along the gasifier axis under transient conditions. In this chapter, the performance of the transient model developed is presented and discussed in detail. There are two main sections in this chapter; firstly, the general characteristics of the model output are described, and then the sensitivity of the model to small changes in parameter values is analysed. As discussed in the previous chapter, to run the computer program the operating parameters must be defined. The selected operating parameters, which also formed the base case for the sensitivity analysis, were chosen to be as close as possible to the experimental conditions. These parameter values are shown in Table 4.1.

Table 4.1 Baseline operating parameter values used in the model

Quantity Particle diameter (mm) Apparent density of fuel (kg/m) Moisture content of the fuel (%db) Inner diameter of reactor (mm) Effective length of the reactor (mm) Reactor pressure (kPa) Inlet air temperature (K) Ignition temperature (K) Ignition depth (mm) Air supply rate (kg/hr) Fuel bed initial void fraction (-)

Value 20 950 10 206 700 101.325 300 1000 50 15 0.46

100

Chapter 4 Model results and discussions

4.2

Generalised model predictions

Using the input operating parameter values shown in Table 4.1, computer simulations were employed to develop plots showing how a variety of parameters varied along the reactor as a function of time (Figure 4.1 to Figure 4.15). Some figures are given as paired plots, with each showing the same quantity, first over the computational effective reactor length (the 700 mm stretch above the level of the grate) and then over a more limited domain in the area where rates of change with distance are highest. In this section, the transient behaviour of temperature and gas composition under warm up conditions is discussed first, then the predicted solid and gas temperature profiles, velocity profiles, profiles of solid phase components (unbound water and volatile matter), gas composition profile, and reaction rates profile under steady state conditions are presented. The stability of the reaction zone is also evaluated by varying the specified rate at which inlet air is supplied.

4.2.1 Transient behaviour of temperatures and gas compositions At the beginning of a simulation, a 50mm deep section of the char bed was assumed to be ignited and char oxidation was therefore the dominant heat source. Most of the heat released from this combustion zone is initially utilized to heat up the lower part of the char bed but some is used to devolatilize the wood particles above the combustion zone. While the char below the combustion zone is heating up, gasification reaction rates in this zone are very low. However, above the ignition zone, pyrolysis of the lower part of the wood bed is by this time taking off quite rapidly. The volatiles released in this process burn readily and, as a result, the combustion zone moves upwards. This phenomenon is discussed in detail in subsection 4.2.7. After sufficient time has elapsed for the entire char bed to heat up, temperature profiles stabilise and the gasifier starts producing combustible gas. This latter condition is known as the steady operating mode of the reactor. To obtain comparable results for the different ranges of model and operating parameter inputs, all of the results presented were generated by the model for the same operating period, namely 30 minutes of operation. As is shown later, in all cases this was sufficient to achieve steady state operation.

101

Chapter 4 Model results and discussions

The predicted changes with time (from ignition to steady operating condition) of temperatures in the solid phase are shown in Figure 4.1. The gas temperature was initially assumed to be the same as the inlet air temperature. It then increases as the gas absorbs heat released during the char oxidation reaction. Since there are no gas phase oxidation reactions at the beginning of the start up period, the gas temperature is initially lower than the solid temperature. Once gas phase oxidation reactions of pyrolytic decomposition products (volatile gases and secondary tar) begin, the gas temperature increases very rapidly to a peak level. Heat released from oxidation of volatiles is utilized to produce more volatile matter from the pyrolysis zone and this continues till pyrolysis is complete. This phenomenon is also known as flaming combustion in the case of a combustor and flaming pyrolysis in the case of a downdraft gasifier. An important aspect of flaming pyrolysis is the rapid evolution of large volumes of a highly combustible gas having a potential flame temperature > 1500C for air and > 2500C for oxygen (Reed & Markson 1983). Before the reactor reaches a steady operating condition, the heat released during gas phase oxidation also contributes to heating up the lower part of the char bed by convective gas-to-solid heat transfer. Figure 4.2 shows the predicted changes of gas phase temperature with time. Figure 4.3 presents the predicted changes in gas composition along the axis of the gasifier during the period from start up to achievement of a steady state. Initially, while the char bed below the ignition level is being heated up by heat released in the oxidation zone, the rate of the heterogeneous gasification reaction is not high enough to produce self-burning producer gas; this is because the concentrations of the two main combustible components, carbon monoxide and hydrogen, are very low. As temperatures in the gasification zone increase over time, then, as would be expected, the amounts of CO and H2 along the gasification zone increase while the amounts of CO2 and H2O decrease correspondingly.

102

Chapter 4 Model results and discussions

1600

Solid surface temperature (K)

1400 1200 1000 800 600 400 200 600 400 200 0

30 sec 90 sec 300 sec 600 sec 900 sec 1200 sec 1500 sec 1800 sec

Distance from grate (mm)

Figure 4.1 Transient behaviour of solid surface temperature profile

1800 1600
30 sec 90 sec 300 sec 600 sec 900 sec 1200 sec 1500 sec

Gas Temperature (K)

1400 1200 1000 800 600 400 200 600 400 200 0

1800 sec

Distance from grate (mm)

Figure 4.2 Transient behaviour of gas temperature profile

In Figure 4.4, outlet gas composition changes over time are presented together with the changes in the corresponding higher heating value of the producer gas (dry basis). It can be seen that, not unexpectedly, the heating value of the produced gas is very poor at the beginning but increases over time till it reaches a steady value. Model results show that it takes about 20-25 min to reach a steady condition.

103

Chapter 4 Model results and discussions

18 16 14 CO (vol %) 12 10 8 6 4 2 0 600 400

18 16 14 12 CO2 (vol %) 10 8 6 4 2

t=0

t = 30 min

t=0

t = 30 min
200 0

0 600 400 200 0

Distance from grate (mm) Fig 4.3 (a)

Distance from grate (mm) Fig 4.3 (b)

16 14 12 H2O (vol %) H2 (vol %) 10 8 6 4 2 0 600 400

30 25 20 15 10 5 0

t=0

t = 30 min

t=0

t = 30 min
200 0 600 400 200 0

Distance from grate (mm) Fig 4.3 (c)

Distance from grate (mm) Fig 4.3 (d)

Figure 4.3 Transient behaviour of gas composition profile (a) CO, (b) CO2, (c) H2, (d) H2O

25 Gas composition (vol %) 20 15 10 5 0 0 5 10 15 Time (min) 20 25 30

6 5 4 3 2 1 0 Higher heating value (MJ/Nm)


CO CO2 H2 H2O HHV

Figure 4.4 Transient behaviour of outlet producer gas properties

104

Chapter 4 Model results and discussions

4.2.2 Temperature profiles Figure 4.5 shows the predicted steady state temperature profiles for the set of baseline operating parameters. These results show that as one moves down the gasifier, the solid surface temperature increases very sharply at the point where the highly exothermic char oxidation reaction starts. The increase in solid temperature causes an increase of gas temperature through convective solid-to-gas heat transfer. When the gas temperature is hot enough to initiate oxidation of volatile gases and secondary tars, which are the products of primary and secondary pyrolysis, the gas temperature increases very rapidly. Then, once all the volatiles have been oxidised, the gas temperature ceases to rise any further. This peak in the gas temperature occurs slightly further down the gasifier than the peak in the solid temperature. Beyond its peak, the solid temperature starts to decrease. This happens mainly because exothermic surface oxidation reactions cease, due to utilization of the remaining oxygen in gas phase oxidation reactions. Also, the predominantly endothermic reactions of char with steam and carbon dioxide that are characteristic of the gasification zone start to increase in importance. Gas temperatures, having peaked at the end of the oxidation zone, now decrease, mainly due to heat loss to the solid phase and reactor walls. The increase in solid temperature from ambient levels to a peak of around 1450 K was predicted to occur in a zone only about 20 mm long (Figure 4.5 (b)). The length of the region over which the gas phase temperature increased from ambient to its maximum (about 1700 K) was found to be similar to the value computed for the solid phase. Solid temperatures lead gas temperatures over the first few millimetres of this zone of rapid change and then gas temperatures lead over the rest of the reactor length. Differences between the solid and gas temperatures were substantial only in the combustion zone and tended to become smaller in the reduction zone.

105

Chapter 4 Model results and discussions

1800 1600 1400

Temperature (K)

1200 1000 800 600 400 200 700 Ts Tg

600

500

400

300

200

100

Distance from grate (mm)


Fig 4.5 (a)

1800 1600 1400

Temperature (K)

1200 1000 800 600 400 200 650 Ts Tg

600

550

500

Distance from grate (mm)


Fig 4.5 (b)

Figure 4.5 The predicted solid (Ts) and gas (Tg) temperature profiles along the axis of the gasifier (a) over the total reactor length (b) in the domain where rates of change are greatest

106

Chapter 4 Model results and discussions

4.2.3 Velocity profiles The predicted solid and gas velocity profiles along the gasifier are shown in Figure 4.6. The solid velocity at the top of the reactor is constant, because one of the assumptions made in developing the model was that particle shrinkage due to moisture evaporation and primary pyrolysis processes could be neglected. This was not expected to have any significant impact on what was predicted to happen lower down in the gasifier. In the char oxidation and gasification zones, however, the particle size reduces and causes the fuel velocity to decrease. This velocity drops to zero at the grate since there was assumed to be no char loss through the grate. The predicted superficial gas velocity profile along the gasifier shows a very different pattern. Along the unreacted biomass particle bed, the gas velocity is the same as the air inlet velocity. It then increases rapidly due to the large amount of gas released by drying, volatilisation, the heterogeneous gas-char reactions, and expansion due to the increase of the gas phase temperature. Its peak occurs close to the maximum gas temperature point and thereafter the gas velocity gradually decreases as gas phase temperatures fall along the rest of the gasifier.

0.14 0.12 0.10 0.08 0.06 0.04 0.02 0.00

1.0

0.6

Velocity of gases mixture (m/s)

0.8

Velocity of solid (mm/s)

Vs Vg

0.4

0.2

0.0 700 600 500 400 300 200 100 0

Distance from grate (mm)

Figure 4.6 The solid and gas velocity profiles (Vs and Vg respectively) along the axis of the gasifier

107

Chapter 4 Model results and discussions

4.2.4 Solid densities profiles In this study, wood was treated as a mixture of three solid species, namely unbound water, volatile matter and char (which was assumed to be pure carbon). The density of char was assumed constant along the gasifier as discussed in chapter three. (In this subsection, the term density is used to refer to the mass of the component under consideration that is present in unit mass of reactor volume; it has units of kg/m.) Figure 4.7 shows the predicted changes in density of two solid phase components, unbound water and volatiles, in the drying and pyrolysis zones along the axis of the gasifier. Also shown is the solid surface temperature profile. From Figure 4.7(b) it can be observed that moisture is predicted to start vaporizing and evaporate completely within a very narrow region only 3-4 mm in length. This is because solid surface temperatures in the drying zone increase steeply and the model for drying was formulated by simply applying an Arrhenius type equation without considering diffusion effects. This result may or may not be physically accurate. For the gasifier model, however, it was acceptable because times required to complete the drying process are orders of magnitude shorter than those for char combustion and gasification (Di Blasi 2000). From Figure 4.7, it can be seen that pyrolysis of wood particles is predicted to occur over the solid surface temperature range from 850 to 1300 K, which is about 350 K higher than the temperature required to complete the devolatilization of finely divided virgin wood. This is not unreasonable because, in particles 20-30 mm in size, heat takes a while to move from the surface of wood particles to the particle interior. The model allows for this by estimating the rate of the primary pyrolysis process of the whole wood block using an equivalent temperature which is considerably less than the outer surface temperature of the wood blocks (see subsection 3.2.2).

108

Chapter 4 Model results and discussions

800

1600

600

Solid density (kg/m)

1200 1000 800

400

200

600 400

0 200 700 600 500 400 300 200 100 0

Distance from grate (mm)


Fig 4.7 (a)

800

1600 1400

600

Solid density (kg/m)

400

200

unbound water density volatile matter density solid temperature

1200 1000 800 600 400

0 200 640 630 620 610 600

Distance from grate (mm)


Fig 4.7 (b)

Figure 4.7 Solid density profile along the axis of the gasifier (a) over the total reactor length (b) in the domain where rates of change are greatest

109

Solid surface temperature (K)

Solid surface temperature (K)

unbound water density volatile matter density solid temperature

1400

Chapter 4 Model results and discussions

4.2.5 Gas composition profiles Figure 4.8 and Figure 4.9 show the predicted gas composition profile along the reactor under steady state conditions. Figure 4.8 shows that the volume percent of oxygen drops rapidly as the air enters the drying and volatilization zone. This does not mean that oxygen is being consumed in these zones. In fact the amount of oxygen remains almost the same, but due to dilution by water vapour and primary pyrolysis volatiles, the concentration of oxygen in the total gas is reduced. When the solid surface temperature reaches a temperature of around 900 K, oxygen begins to be consumed in heterogenous char oxidation reactions, causing the solid temperature to increase further. At this stage, heat produced during solid phase combustion reactions is transferred to the gas phase by convection, increasing the gas temperature. Once the gas temperature is high enough for homogenous oxidation of combustible volatiles to begin, the rest of the oxygen is consumed very rapidly. Figure 4.8 (b) illustrates how rapidly the oxygen volume percent decreases from 21 to zero. This plot can be divided into 3 parts, each with a different gradient: the first drop is largely attributable to mass transfer from the solid phase to the gas phase as a result of drying and primary pyrolysis processes; the second is largely attributable to oxygen consumption during char oxidation reactions; and the last is associated with gas oxidation reactions. Since gas phase oxidations occur very rapidly and are completed over just a few millimetres of the oxidation zone, the last gradient of the oxygen content plot is very steep. Carbon monoxide is the main combustible component in the producer gas. The model predicts that the volume percentage of CO initially increases along the reactor because of CO evolution during pyrolysis processes, especially secondary pyrolysis. Whilst free oxygen is present most of the CO produced in the pyrolysis zone is subsequently lost by oxidation to CO2. This accounts for the small peak of CO within the combustion zone shown in Figure 4.8 (b). A similar result has been reported for a char combustor model (Cooper & Hallett 2000). Once the char gasification zone is reached, CO levels build up again as a result of the reaction of char with H2O and CO2.

110

Chapter 4 Model results and discussions

25

20

Gas composition (vol %)

15

O2 CO

CO2
10

H2O H2

0 700 600 500 400 300 200 100 0

Distance from grate (mm)


Fig 4.8 (a) 25

20

Gas composition (vol %)

15

O2 CO

CO2
10

H2O H2

0 650 600 550 500 450 400

Distance from grate (mm)


Fig 4.8 (b)

Figure 4.8 Major gas species profiles along the axis of gasifier (a) over the total reactor length (b) in the domain where rates of change are greatest

As would be expected, hydrogen concentrations are predicted to be very low in the oxidation zone. Although hydrogen is formed during oxidation of C6H6 and C2H4, temperatures are so high that in the presence of free oxygen it is instantaneously oxidized to water. The content of hydrogen in the gas mixture, therefore, only increases

111

Chapter 4 Model results and discussions

once oxygen levels are depleted. The final hydrogen content in the producer gas is determined by the extent of char reactions with H2O in the char gasification zone. Initially CO2 is formed by combustion of char. Further amounts of CO2 are then produced by combustion of CO and, as shown in Figure 4.8, CO2 levels reach a maximum quite close to the point where the gas temperature is at its maximum. Then, in the gasification zone, CO2 concentrations decrease monotonically due to its reaction with char to produce CO, finally levelling out at a more-or-less constant value. Initial H2O concentrations in the gas phase depend upon the moisture content of the biomass fuel. Figure 4.8 shows H2O levels building up due to the evaporation of moisture and being augmented by the water formed when hydrogen formed in primary pyrolysis processes is oxidised. Concentrations of H2O, like those of CO2, decrease in the gasification zone; this occurs as a result of the heterogeneous reaction of carbon with water vapour. Other, minor, gaseous components include methane and ethylene, products of tar cracking; Figure 4.9 shows that these are partially oxidized in the combustion zone with only small amounts being found in the final producer gas. Also shown in Figure 4.9 are the predicted levels of primary tars (oxygenates), the major products of the primary pyrolysis reaction. The initial mass of these tars is more than half that of the initial biomass fuel but most primary tars are rapidly cracked into lower molecular weight gases and secondary tars (C6H6). These products are mostly oxidized in the combustion zone. However, as shown, not all primary tars are cracked before the gases leave the combustion zone. Much of this residual tar decomposes below the combustion zone, producing secondary tars that cannot be oxidized since no oxygen is available. This behaviour is more clearly shown in Figure 4.10, which shows that cracking of primary tars still continues in the oxygen depleted zone beyond the point where the peak gas temperature occurs. This helps explain why small amounts of tar are still present in the gas produced from the conventional stratified downdraft gasifier.

112

Chapter 4 Model results and discussions

6 5

Gas composition (vol %)

4 3 2 1 0 CH4 C2H4

primary tar C6H6

700

600

500

400

300

200

100

Distance from grate (mm)


Fig 4.9 (a)

6 5

Gas composition (vol %)

4 3 2 1 0 CH4 C2H4

primary tar C6H6

650

600

550

500

Distance from grate (mm)


Fig 4.9 (b)

Figure 4.9 Minor gas species profiles along the axis of the gasifier (a) over the total reactor length (b) in the domain where rates of change are greatest

113

Chapter 4 Model results and discussions

25

20

Gas composition (vol %)

15

O2

primary tar
10

C6H6

640

620

600

580

560

540

Distance from grate (mm)

Figure 4.10 Primary tar, secondary tar, and oxygen profiles along the reactor axis

4.2.6 Reaction rate profiles Figure 4.11 to Figure 4.13 show how the rates of the main process reactions are predicted to change along the gasifier. As mentioned in subsection 4.2.2, the predicted drying front is steep, most probably as a consequence of the highly simplified description of the drying process in the model (intrinsic kinetic rate only), which does not take account of intra-particle temperature and moisture gradients. As Figure 4.11 shows, once drying is complete the primary pyrolysis reactions begin, followed by the secondary pyrolysis reactions in which primary tar is cracked to combustible gases including secondary tar. Figure 4.11 shows that secondary pyrolysis occurs over a greater distance and extends further downstream than the other reaction processes illustrated. This reflects the fact that, as discussed earlier, the cracking of the primary tar is not completed by the end of the oxidation zone and small amounts of secondary tar are still produced beyond this point. In Figure 4.12, the gas and solid phase oxidation reaction rate profiles are presented together with the gas temperature profile. Figure 4.12 shows that char oxidation reaches its peak quite quickly and then decreases as one moves down to the point where gas
114

Chapter 4 Model results and discussions

phase oxidation reactions commence. Rates of oxidation of volatile products show a major peak at the end of the oxidation zone where gas temperatures increase sharply. However, hydrogen and secondary tar oxidation reactions are shown to start at a lower gas temperature than CO oxidation and take place over a greater distance due to their greater reactivity. The total rate of hydrogen oxidation, however, is lower than other oxidation rates because of the smaller amounts released by the primary and secondary pyrolysis processes compared with CO and secondary tar. CO reaction rates are markedly higher since the amount of CO released by pyrolysis processes is significantly higher than the amounts of other gaseous components. As would be expected, predicted oxidation reaction rates decline to zero once oxygen is totally consumed.

1.8 1.6 Volumetric rate (kg/m/s) 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 630 620 610 600 590 580 Drying x 0.1 Primary pyrolysis x 0.05 Secondary pyrolysis x 0.1 Char oxidation

Distance from grate (mm)

Figure 4.11 Drying, pyrolysis and char oxidation rates along the axis of the gasifier

Figure 4.13 shows how the rates of heterogeneous char gasification reactions are predicted to vary along the gasification zone. Among the heterogeneous gasification processes, steam gasification was found to have the highest rate; this is because large amounts of water vapour are released during drying and the primary pyrolysis processes. The methane formation reaction (Carbon-H2 reaction) rate was found to be very low since the reaction is much slower than the other heterogeneous reactions (Hobbs, Radulovic & Smoot 1992). It is evident that the char gasification reactions occur over longer distances in the reactor than other reactions.

115

Chapter 4 Model results and discussions

1.6 1.4 Volumetric rate (kg/m/s) 1.2 1.0 0.8 0.6 0.4 0.2 0.0 625 620 615 610 605 600 C6H6 oxidation CO oxidation x 0.1 H2 oxidation Char oxidation

1800 1600 1400 1200 1000 800 600 400 200 Distance from grate (mm) Gas temperature (K)

Tg

Figure 4.12 Gas temperature (Tg) and rates of gaseous and solid phase oxidation reaction along the axis of the gasifier

0.14 0.12 0.10 0.08 0.06 0.04 0.02 0.00 600 525 Distance from grate (mm) 450 Carbon-H2O reaction Carbon-CO2 reaction x 10 Carbon-H2 reaction x 500

Figure 4.13 Heterogeneous gasification reaction rates along the axis of the gasifier

Volumetric rate (kg/m/s)

116

Chapter 4 Model results and discussions

4.2.7 Stability of reaction zones In the downdraft stratified biomass gasifier, the reaction front moves slightly upwards while the reactor is heating up due to heat losses to the reactor wall and a low gasification rate. After it reaches a steady condition, three operating modes are possible (Earp et al. 1990): 1. Gasification dominant, where the rate of char formation by pyrolysis is slower than the rate of char consumption and as a result the reaction zone moves downwards. 2. Pyrolysis dominant, where the rate of char formation by pyrolysis is faster than the rate of char consumption and as a result the reaction zone rises up. 3. Stable operation, where the rate of char deposition by pyrolysis is equal to the rate of char depletion and, as a result, the reaction zone is stable. This is the optimum mode of operation since the gasifier can be run for long periods with a relatively constant output compared to operation in other modes. There is only one particular air to fuel ratio that will permit a stable reaction zone (Earp 1988) and a small change in any operating parameter may cause the zone to move in one direction or the other (Reed & Markson 1985). The upward movement of the front is similar to the movement of a flame front in a flame speed experiment involving a premixed gas mixture. In both cases the front moves into an unburnt zone; in the latter case this is a premixed fuel-air gaseous mixture, while in the gasifier it is a zone containing unburnt char and air (Dassappa & Paul 2001). The air supply rate to the reactor was found to be one of the more sensitive parameters affecting the stability of the reaction zone. An increase in the airflow rate to the gasifier increased the oxygen supply and hence the bed temperature and char consumption rates. As a result, reaction zones moved downwards towards the grate. When the air supply rate to the gasifier was reduced, the oxidation and gasification zone temperatures decreased, the rate of char consumption was observed to be lower than the char production rate by pyrolysis, and, as a result, the reaction zone rose up. This is illustrated in the model outputs shown in Figure 4.14 and Figure 4.15.

117

Chapter 4 Model results and discussions

2000

20 min

25 min 30 min Ts

Temperature (K)

Tg 1500

1000

500

600

580

560

540

520

500

Distance from grate (mm)

Figure 4.14 Illustration showing downward movement of the reaction zone as a result of an increase in the air supply rate (24 kg/hr)

1600 1400 1200 1000 800 600 400 200 640 620 600 580 560 540 520 500 Distance from grate (mm) 30 min 25 min 20 min Ts Temperature (K) Tg

Figure 4.15 Illustration showing upward movement of the reaction zone as a result of a decrease in the air supply rate (8 kg/hr)

118

Chapter 4 Model results and discussions

The range of air supply rates for which a stable mode of operation can be achieved was found to be between 19 and 20 kg/hr for the base case (parameters defined in Table 4.1) giving a specific gasification rate of 290-296 kg/m2hr (dry basis). This result is close to the value found in experimental work on an open core gasifier in which the specific capacity for stable operation was found to be 271 kg/m2hr (DAF basis) (Milligan 1994).

4.3

Parametric sensitivity analysis of the model

Values of parameters used in the numerical model were determined using empirical correlations taken from various independent experimental reports in the open literature. Therefore it was felt important to investigate the effect on model predictions of varying the values of these parameters so that the sensitivity of the model to particular parameters could be established. In this study, a parametric sensitivity analysis was carried out to explore the effects of variations in operating and model parameters on the temperatures profile, gas composition, final tar content and stability of the reaction zones. The parametric sensitivity analysis was broken down into two parts dealing respectively with operating parameters and model parameters. The effects of three operating parameters were investigated, namely airflow rate, initial particle diameter and moisture content. The effects of eight model parameters were also examined, namely solid-gas heat transfer coefficient, mass transfer coefficient, heat and mass exchange area for heterogeneous reaction, kinetic rate coefficients of primary pyrolysis, kinetic rate coefficients of secondary pyrolysis (tar cracking), kinetic rate coefficients of secondary tar oxidation, effective thermal conductivity and heat loss to the reactor wall. The parametric studies were conducted by changing the value of one specific parameter while all the other model parameters were kept constant at the base level given in Table 4.1. An airflow rate of 15 kg/hr was selected as the base value for the sensitivity analysis. It should be noted, in the graphical presentation of sensitivity analysis results, that primary tar content is shown in units of mg/Nm while secondary tar content is shown as a volume percentage of the total gas mixture. The reason for this is that only

119

Chapter 4 Model results and discussions

in the case of primary tar can the quantities of tar present be determined with any accuracy (by collection and weighing); only some of the secondary tars are present in the condensed tar due to their volatility (Reed, Levie & Graboski 1987). As discussed in a previous subsection (4.2.7), reaction zone movement is not constant throughout the reactor warming up stage. Therefore, to obtain comparable results for the different ranges of model and operating parameter inputs used in the sensitivity analysis, all of the values of reaction front movement presented are those calculated for the case where steady state conditions have been achieved (i.e. for the period between 25 minutes and 30 minutes after ignition).

4.3.1 Airflow rate

Air supply rate to the gasifier is one of the most important operating parameters, strongly influencing the temperature profiles, tar content, and stability of the reaction zones. Figure 4.16 shows the effects of variations in airflow rate on gasifier behaviour. Increasing the air supply rate means a greater oxygen supply to the system and a higher heat release from both the heterogeneous and homogenous oxidation reactions and an increased peak temperature for both the solid and gas phases in the combustion zone. Figure 4.16 (b) shows that peak gas phase temperatures increased faster than peak solid phase temperatures as the air supply rate was increased. This is attributed to a more rapid increase in gas phase oxidation rates than in solid phase oxidation rates as temperatures rose. As a result, both primary and secondary tar outlet levels drop at higher air supply rates (see Figure 4.16 (c)); this is attributed to higher thermal cracking rates at higher gas phase temperatures.

120

Chapter 4 Model results and discussions

1800 Temperature (K) 1500 1200 900 600 300 640

15 kg/hr 13 kg/hr

17 kg/hr

Ts Tg

620

600

580

560

540

520

500

Distance from grate (mm) Fig 4.16 (a)


Primary tar (Condensable) 1900 1800 Primary tar (mg/Nm) 1700 1600 Peak Ts 1500 1400 1300 1200 12 13 14 15 16 17 18 Air supply rate (kg/hr) Peak Tg 200 180 160 Temperature (K) 140 120 100 80 60 40 20 0 12 13 14 15 16 17 18 Air supply rate (kg/hr) 0.3 0.2 0.1 0 Secondary tar (Non-condensable) 0.7 0.6 0.5 0.4 Secondary tar (vol%)

4.16 (b)

4.16 (c)

18
Reaction zone velocity (mm/sec)

0.050 0.040 0.030 12 kg/hr 0.020 0.010 0.000 -0.010 14 kg/hr 16 kg/hr 18 kg/hr 20 kg/hr

16
Compostion (vol%)

14 12 10 8 6 4 2 0
CO CO2 H2O Gas species H2 CH4 12 kg/hr 14 kg/hr 16 kg/hr 18 kg/hr

4.16 (d)

4.16 (e)

Figure 4.16 Effects of air supply rate (a) on solid (Ts) and gaseous (Tg) temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward)

121

Chapter 4 Model results and discussions

The composition of the outlet producer gas as a function of air supply rate is presented in Figure 4.16(d). Of particular interest is that changes in the concentration of CO in the outlet gas at different air flow rates are not significant. Increases in the solid temperature (Ts) would be expected to increase the rate of heterogeneous gasification reactions and therefore the CO level. However, this is offset by the higher gas temperature, which increases the rate of oxidation of CO released in the pyrolysis processes. Higher solid phase temperatures also increase the heterogeneous water-gas reaction rates; as a result the quantity of water vapour found in the final gas outlet would be expected to decrease and the concentration of H2 to increase. However, as Figure 4.16 (d) shows, the predicted drop in H2O level appears more significant than the increase in H2 level as air flow rates rise. The main reason for this is that increasing the airflow rate also increases nitrogen concentrations in the gas; this has the effect of emphasising the decrease in the volume percentage of H2O and retarding the increase in the volume percentage of hydrogen. The other gaseous component that changes significantly is methane. At higher air supply rates, lower CH4 levels are observed because of the higher CH4 oxidation rates at higher temperatures. As discussed in section 4.2.7, air supply rate controls the stability of the reaction zones. At higher air supply rates, the char consumption rate increases, which leads to a higher solid velocity and causes the reaction zone to move downwards towards the grate (see Figure 4.16 (e)).

4.3.2 Initial particle size The predicted sensitivity of system output parameters to changes in initial wood particle diameter is shown in Figure 4.17. Varying the particle diameter affects heat and mass transfer coefficients and therefore has an influence both on primary pyrolysis processes and on the heterogeneous gas char reactions. In the pyrolysis zone, the rate at which heat from the particle surface penetrates to the particle core is faster in the case of a smaller particle. Thus, the rate of volatiles production is greater for a small particle than for a larger particle. In the

122

Chapter 4 Model results and discussions

combustion and gasification zones, a smaller char particle diameter increases the total surface area per unit volume and decreases mass transfer resistances, thereby promoting both char oxidation and gasification. As a result, smaller particles enhance char combustion rates and lead to higher solid peak temperatures. At larger particle sizes, char oxidation rates fall and more oxygen is utilized for gas phase combustion, leading to increases in gas phase temperatures. As a result, lower primary and secondary tar outputs are observed in the case of large wood particles. These effects are illustrated in Figure 4.17 (a), (b) and (c). Predicted exit gas compositions as a function of initial wood particle size are presented in Figure 4.17 (d). Because of the lower specific surface areas in beds of larger particles, higher solid temperatures are needed to attain particular levels of carbon conversion in the gasification zone than are required in the case of smaller particles. This results in more combustible gas species being consumed within the oxidation zone when larger char particles are present. As a result, lower exit CO and H2 concentrations, and higher exit CO2 concentration are predicted for larger solid fuel particles (see Figure 4.17 (d)). CH4 is also present in decreasing amounts as particle sizes increase. As particle sizes decrease, reaction zones show a tendency to move upwards; this can be attributed to the faster pyrolysis rate (Figure 4.17 (e)).

4.3.3 Initial moisture content The effect of increasing fuel moisture content on the phenomena occurring within the reaction region is shown in Figure 4.18 for five fuel moisture contents, 5, 8, 10, 12 and 15% (dry basis).

123

Chapter 4 Model results and discussions

1800 Temperature (K) 1500 1200 900 600 300 700 650 15 mm

20 mm

25 mm

Ts Tg

600

550

500

450

400

Distance from grate (mm) Fig 4.17 (a)


Primary tar (Condensable) 2000 1900 1800 Primary tar (mg/Nm) Temperature (K) 1700 1600 1500 1400 1300 1200 15 20 25 30 Peak Ts Peak Tg 250 200 150 0.3 100 50 0 15 20 25 30 Initial particle diam eter (m m ) 0.2 0.1 0 0.6 0.5 0.4 Secondary tar (vol%) 300 Secondary tar (Non-condensable) 0.7

Initial particle diam eter (m m )

4.17 (b)

4.17 (c)

20 16
Compostion (vol%) Reaction zone velocity (mm/sec)

0.070 0.060 0.050 0.040 0.030 0.020 0.010 0.000 15mm 20mm 25mm 30mm

18 14 12 10 8 6 4 2 0
CO CO2 H2O Gas species H2 CH4 15 mm 20 mm 25mm 30 mm

4.17 (d)

4.17 (e)

Figure 4.17 Effects of initial wood particle size (a) on solid (Ts) and gaseous (Tg) temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward)

124

Chapter 4 Model results and discussions

As discussed earlier, drying of the wood is expected to be complete before temperatures high enough for char combustion to begin are reached. Therefore, as shown in Figure 4.18 (a), increases in wood moisture content favour the char gasification processes involving water vapour, and the reaction zones move downwards. Figure 4.18 (b) shows that changing initial moisture content does not affect the peak solid temperature but that a slight increase in peak gas temperature occurs at higher moisture content levels. This rather unexpected result seems to be linked to CO oxidation processes. The CO oxidation reaction rate depends in part on the water vapour concentration (eq. 3.19) and a higher water vapour concentration increases the CO combustion rate, the associated heat release rate, and the local gas temperature. It also increases the hydrocarbon (secondary tar) combustion rate and, as a result, a slight decrease in both the primary and secondary tar content of the gas is found at higher initial moisture contents (Figure 4.18 (c)). Figure 4.18 (d) shows that changing the moisture content also causes slight changes in the levels of all gas phase components. As discussed above, higher moisture contents lead to an increase in CO oxidation rates, and a slight decrease in gas phase CO level results. Figure 4.18 (d) also shows a slight decrease of hydrogen at higher moisture content levels, due to the higher gas temperatures and consequent higher hydrogen oxidation rate. This presumably contributes to the observed increase in gas phase H2O content as the wood moisture content increases. The CO2 concentration appears not to be affected by changes in fuel moisture content even though the CO oxidation rate increases. This is likely to be a result of the increased amount of H2O in the gas phase, which affects volume percentage calculations. At higher wood moisture contents, a lower CH4 level is observed; this can be attributed to the higher peak gas temperatures and the higher associated CH4 oxidation rates. The effect of fuel moisture content increases is to slow the upward movement of the reaction zones in the gasifier, as shown in Figure 4.18 (e).

125

Chapter 4 Model results and discussions

1800 Temperature (K) 1500

5%

10 %

15 %

Ts 1200 900 600 300 700 650 600 550 500 450 400 Tg

Distance from grate (mm) Fig 4.18 (a)


Primary tar (Condensable) 1800 1700 Temperature (K) 1600 1500 1400 1300 1200 5 8 10 12 15 Peak Ts Peak Tg Primary tar (mg/Nm) 70 60 50 40 30 20 10 0 5 8 10 12 15 Secondary tar (Non-condensable) 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 Secondary tar (vol%)

Initial m oisture content (%d.b)

Initial m oisture content (%d.b)

4.18 (b)

4.18 (c)

18
Reaction zone velocity (mm/sec)

0.045 0.040 0.035 0.030 0.025 0.020 0.015 0.010 0.005 0.000 8% 10% 12% 15%

16
Compostion (vol%)

14 12 10 8 6 4 2 0
CO CO2 H2O Gas species H2 CH4 8% 10% 12% 15%

4.18 (d)

4.18 (e)

Figure 4.18 Effects of initial moisture content (a) on solid (Ts) and gaseous (Tg) temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward)

126

Chapter 4 Model results and discussions

4.3.4 Solid-gas heat transfer coefficient Heat transfer coefficients derived for a reacting surface are scarce. So, the basic solidgas heat transfer coefficient used in this model was one formulated for a non-reacting surface. As reported by Dzhapbyev, Miropolskii and Malkovskii (1986), however, the solid-gas heat transfer coefficient for a nonreacting system exceeds that for a reacting system. Therefore the initially calculated heat transfer coefficient needs to be modified for use in the gasifier model. The ratio of reacting to non-reacting heattransfer coefficient () has been introduced as a means of doing this (Cho & Joseph 1981; Ghani, Radulovic & Smoot 1996). According to Dzhapbyev, Miropolskii and Malkovskii (1986), the factor () that accounts for the reactive nature of the fixed bed can have a value ranging from 0.02 to 1.0. For the present study a base case value of 0.6 was selected by comparison with results from experimental work (see Chapters 5 and 6 for details). The effects of changing on the model outputs are shown in Figure 4.19. A reduction in the value of leads to a decrease in the heat transfer rate from the solid phase to the gas phase. Since oxidation of volatile gases depends on gas phase temperature, a smaller results in a more prolonged gas combustion process and consequently a lower peak gas phase temperature (Figure 4.19 (a) & Figure 4.19 (b)). More oxygen is consumed in char oxidation reactions and so less is available for hydrocarbon oxidation reactions. Because of this, a higher primary and secondary tar content is observed at smaller values of (Figure 4.19 (c)). By increasing , a higher heat transfer rate and a higher peak gas phase temperature are achieved and increases occur in the solid temperature along the gasification zone (not the peak solid temperature) (Figure 4.19 (a)) and hence in heterogeneous gas-char reaction rates. As a result, solid velocities increase, and reaction zones move towards the bottom of the bed (Figure 4.19 (e)). A higher value of leads to higher gas temperatures and more rapid gas phase combustion reactions in the oxidation zone as a result. The trend of increasing H2O and CO2 and decreasing H2, CO and CH4 with increasing is a direct consequence of this increase in gas phase reaction activity (Figure 4.19 (d)).

127

Chapter 4 Model results and discussions

1800 Temperature (K) 1500 1200 900 600 300 700 650

= 0. 6

= 0.7

= 0.5
Ts Tg

600

550

500

450

400

Distance from grate (mm) Fig 4.19 (a)


Primary tar (Condensable) 2000 1900 1800 Primary tar (mg/Nm) Temperature (K) 1700 1600 1500 1400 1300 1200 0.5 0.6 0.7 0.8 0.9 Peak Ts Peak Tg 160 140 120 100 80 60 40 20 0 0.5 0.6 0.7 0.8 0.9 0.3 0.2 0.1 0 Secondary tar (Non-condensable) 0.7 0.6 0.5 0.4 Secondary tar (vol%)

Reacting to non-reacting heat transfer coefficient ratio (-)

Reacting to non-reacting heat transfer coefficient ratio (-)

4.19 (b)

4.19(c)

18
Reaction zone velocity (mm/sec)

0.035 0.030 0.025 0.020 0.015 0.010 0.005 0.000 = 0.6 = 0.7 = 0.8 = 0.9

16
Compostion (vol%)

14 12 10 8 6 4 2 0
CO CO2 H2O Gas species H2 CH4 = 0.6 = 0.7 = 0.8 = 0.9

4.19 (d)

4.19 (e)

Figure 4.19 Effects of solid-to-gas heat transfer coefficient (a) on solid (Ts) and gaseous (Tg) temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward)

128

Chapter 4 Model results and discussions

4.3.5 Mass transfer coefficient There is always some uncertainty associated with the use of empirical mass transfer correlations derived from experimental data obtained in equipment different from ones own. This is especially true at low values of the Reynolds number (as occur in the case of packed beds of small particles at high temperature). Kato, Kubota and Wen (1970) describe the wide divergence between different correlations at low Re and also showed that the actual mass-transfer coefficient in solid-gas fixed beds becomes considerably smaller than the theoretical coefficient at low Re values. In this study, direct application of existing empirical correlations to determine the rate of the diffusion-controlled combustion reactions in the gasifier resulted in unrealistic temperature values. To overcome this problem, use of a limiting maximum value of the mass-transfer coefficient has been recommended (Bhattacharya et al. 1986; Di Blasi 2000). A value of 0.045 m/s, which was found to give results closely matching those obtained experimentally, was therefore selected as a maximum mass transfer coefficient value to use in the simulation for the base case. The effects of varying the mass transfer coefficient value on the system outputs are presented in Figure 4.20. The mass transfer coefficient strongly affects the rates at which gaseous reactants diffuse to the fuel surface and react with the fuel. Thus a lower mass transfer coefficient leads to decreased heterogeneous gas-char reaction rates. A reduced mass transfer rate results in heterogeneous char oxidation reactions slowing down in comparison with gas phase combustion reactions. This leads to a greater proportion of the oxygen being consumed in reactions with volatile products, with the result that the gas temperature increases (Figure 4.20 (b)), and both primary and secondary tar levels decrease (Figure 4.20 (c)). In contrast, if mass transfer rates increase (due to an increase in the mass transfer coefficient value), gas temperatures drop, the amounts of combustible components consumed in the oxidation zone fall, and larger quantities of CO and CH4 are present in the exit gas. A slight increase in H2 levels is also observed (Figure 4.20 (d)).

129

Chapter 4 Model results and discussions

1800 Temperature (K) 1500 1200 900 600 300 700

0.045 m/s 0.04 m/s 0.05 m/s Ts Tg

650

600

550

500

450

400

Distance from grate (mm) Fig 4.20 (a)


Primary tar (Condensable) 1900 1800 Primary tar (mg/Nm) 1700 1600 Peak Ts 1500 1400 1300 1200 0.04 0.045 0.05 0.055 Peak Tg 200 250 Secondary tar (Non-condensable) 0.7 0.6 0.5 150 100 0.4 0.3 0.2 50 0.1 0 0.04 0.045 0.05 0.055 Secondary tar (vol%)

Temperature (K)

Maxim um m ass transfer coefficient (m /s)

Maxim um m ass transfer coefficient (m /s)

4.20 (b)

4.20 (c)

18
Reaction zone velocity (mm/sec)

0.050 0.045 0.040 0.035 0.030 0.025 0.020 0.015 0.010 0.005 0.000 0.040 m/s 0.045 m/s 0.050 m/s 0.055 m/s

16
Compostion (vol%)

14 12 10 8 6 4 2 0
CO CO2 H2O Gas species H2 CH4 0.040 m/s 0.045 m/s 0.050 m/s 0.055 m/s

4.20 (d)

4.20 (e)

Figure 4.20 Effects of mass transfer coefficient (a) on solid (Ts) and gaseous (Tg) temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward)

130

Chapter 4 Model results and discussions

Since gas temperatures are lower in the case where mass transfer rates increase, solid temperatures along the gasification zone (not peak temperatures) also decrease. As a result, despite the higher value of the mass transfer coefficient, there is nevertheless a decrease in overall gas-char reaction rates and reaction zones move upwards (Figure 4.20 (e)).

4.3.6 Heat and mass exchange area In this model, the biomass particles were taken to be spherical in shape and the effect of char particles breaking up was not accounted for. In reality, however, particles are not spherical and char particles do break up into smaller pieces. The effect of this is to increase the particle surface area per volume above the value assumed in the model. The model also neglects the loss of effective surface area when particle surfaces are in contact. Therefore a sensitivity analysis was carried out to assess the effect of changes in heat and mass exchange area. This was done by multiplying the base case area by factors ranging from 0.8 to 1.2. The effects of varying the heat and mass transfer exchange area (Av) are presented in Figure 4.21. A reduction in heat and mass transfer exchange area reduces the extent of heterogeneous reactions. This results in more oxygen being consumed in volatiles combustion processes, an increase in gas temperature (Figure 4.21 (b)), and a decrease in primary and secondary tar levels in the outlet gas (Figure 4.21 (c)). As also happens when the mass transfer coefficient is reduced, decreasing the heat and mass exchange area decreases the proportions of CO, H2, and CH4 in the outlet gas due to the higher combustion rates at higher gas phase temperatures (Figure 4.21 (d)). Reducing the heat and mass transfer exchange area also decreases char consumption rates and, as a result, reaction zones tend to move upwards (Figure 4.21 (e)).

131

Chapter 4 Model results and discussions

1800 Temperature (K) 1500 1200 900 600 300 640 620 600 580 560 Av x 1.2 Av x 1.0 Ts Tg Av x 0.8

Distance from grate (mm) Fig 4.21 (a)


Primary tar (Condensable) 1800 1700 Primary tar (mg/Nm) Temperature (K) 1600 1500 1400 1300 1200 0.8 0.9 1.0 1.1 1.2 Heat and m ass transfer exchange area m ultiplication factor (-) Peak Ts Peak Tg 120 100 80 60 40 20 0 0.8 0.9 1.0 1.1 1.2 Secondary tar (Non-condensable) 0.6 0.5 0.4 0.3 0.2 0.1 0 Secondary tar (vol%)

Heat and m ass transfer exchange area m ultiplication factor (-)

4.21 (c)

4.21 (b)

18
Reaction zone velocity (mm/sec)

0.040 0.035 0.030 0.025 0.020 0.015 0.010 0.005 0.000 Av x 0.9 Av x 1.0 Av x 1.1 Av x 1.2

16
Compostion (vol%)

14 12 10 8 6 4 2 0
CO CO2 H2O Gas species H2 CH4 Av x 0.9 Av x 1.0 Av x 1.1 Av x 1.2

4.21 (d)

4.21 (e)

Figure 4.21 Effects of heat and mass transfer exchange area (a) on solid (Ts) and gaseous (Tg) temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward)

132

Chapter 4 Model results and discussions

4.3.7 Kinetics of primary pyrolysis As described in chapter 3, in this model rate coefficients for the primary pyrolysis reactions were derived from experimental results obtained during the rapid pyrolysis of large wood particles. The effect of varying primary pyrolysis reaction rates was explored in a similar way to that in which the sensitivity analysis for the heat and mass exchange area was carried out, except that in this case, the factor ranged in value from 0.5 to 2.0. Primary pyrolysis is the process in which primary tars (oxygenates) are released from virgin wood. An increase in primary pyrolysis rates produces more primary tar within a given time and completion of tar cracking is hindered. Thus a higher tar content in the gas results at higher primary pyrolysis rates (Figure 4.22 (c)). A slight decrease in both gas and solid temperatures is observed at higher primary pyrolysis rates, probably because of the lower secondary tar oxidation rate reducing the peak gas temperature (Figure 4.22 (b)). No significant changes in gas composition result from changing the value of the primary pyrolysis reaction rate (Figure 4.22 (d)). As discussed in subsection 4.2.7, the primary pyrolysis process is one of the main factors affecting the stability of the reaction zone; a faster pyrolysis rate leads to the reactor running in a pyrolysis dominant mode and, as a consequence, reaction zones move upwards (Figure 4.22 (e)).

4.3.8 Kinetics of secondary pyrolysis The effect of changes in the rate coefficients for the secondary pyrolysis process on the model output parameters are shown in Figure 4.23. Changes in the rate coefficients were made by multiplying the pre-exponential factor in the rate equation by values ranging between 0.5 and 2.0. Secondary pyrolysis involves the cracking of primary tar into secondary tar and other combustible gases. Thus, increasing the secondary pyrolysis reaction rate directly reduces the amount of primary tar, as Figure 4.23 (c) shows.

133

Chapter 4 Model results and discussions

Ap1 = pre-exponential factor in primary pyrolysis rate equation

1800 Temperature (K) 1500 1200 900 600 300 640 620

Ap1 x 1.0 Ap1 x 1.5

Ap1 x 0.5

Ts Tg

600

580

560

Distance from grate (mm) Fig 4.22 (a)


Primary tar (Condensable) 1800 80 1700 Temperature (K) 1600 1500 1400 1300 1200 0.5 1.0 1.5 2.0 Peak Ts Peak Tg Primary tar (mg/Nm) 70 60 50 40 30 20 10 0 0.5 1.0 1.5 2.0 0.1 0 0.4 0.3 0.2 0.6 0.5 Secondary tar (vol%) Secondary tar (Non-condensable)

Prim ary pyrolysis rate m ultiplication factor (-)

Prim ary pyrolysis rate m ultiplicaton factor (-)

4.22 (b)

4.22 (c)

18
Reaction zone velocity (mm/sec)

0.045 0.040 0.035 0.030 0.025 0.020 0.015 0.010 0.005 0.000 Ap1 x 0.5 Ap1 x 1.0 Ap1 x 1.5 Ap1x 2.0

16
Compostion (vol%)

14 12 10 8 6 4 2 0
CO CO2 H2O Gas species H2 CH4 Ap1 x 0.5 Ap1 x 1.0 Ap1 x 1.5 Ap1x 2.0

4.22 (d)

4.22 (e)

Figure 4.22 Effects of changes in primary pyrolysis rate (a) on solid (Ts) and gaseous (Tg) temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward)

134

Chapter 4 Model results and discussions

Increasing the secondary pyrolysis reaction rate also leads to the formation of more combustible gases, the consumption of more oxygen in gas combustion reactions and increases in gas phase temperatures (Figure 4.23 (b)). The effect of changes in secondary pyrolysis rates on gas composition was insignificant (Figure 4.23 (d)). Increases in secondary pyrolysis rates lead to higher gas phase oxidation rates, reduced char combustion rates and a tendency for reaction zones to move upwards (Figure 4.23 (e)).

4.3.9 Kinetic rate of oxidation of secondary tar In the model, the amount of secondary tar in the outlet gas is determined mainly by the rate of secondary tar oxidation. An analysis of the sensitivity of the model to changes in this rate was carried out by multiplying the pre-exponential factor in the rate equation by values ranging from 0.8 to 1.2. The results of the analysis are presented in Figure 4.24. As Figure 4.24 shows, changing secondary tar oxidation rates seems to have no significant effect on model operation. This suggests that secondary tar oxidation is not a rate limiting step in the process.

135

Chapter 4 Model results and discussions Ap2 = pre-exponential factor in secondary pyrolysis rate equation

1800

Ap2 x 1.0 Ap2 x 1.2 Ts

Temperature (K)

1500 1200 900 600 300 640

Tg Ap2 x 0.8

620

600

580

560

Distance from grate (mm)

Fig 4.23 (a)


Primary tar (Condensable) 1800 1700 Temperature (K) 1600 1500 1400 1300 1200 0.8 0.9 1.0 1.1 1.2 Peak Ts Peak Tg Primary tar (mg/Nm) 300 250 200 150 100 50 0 0.8 0.9 1.0 1.1 1.2 Secondary tar (Non-condensable) 0.6 0.5 0.4 0.3 0.2 0.1 0 Secondary tar (vol%)

Secondary pyrolysis rate m ultiplication factor (-)

Secondary pyrolysis rate m ultiplication factor (-)

4.23 (b)

4.23 (c)

18
Reaction zone velocity (mm/sec)

0.035 0.030 0.025 0.020 0.015 0.010 0.005 0.000 Ap2 x 0.8 Ap2 x 1.0 Ap2 x 1.1 Ap2 x 1.2

16
Compostion (vol%)

14 12 10 8 6 4 2 0
CO CO2 H2O Gas species H2 CH4 Ap2 x 0.8 Ap2 x 1.0 Ap2 x 1.1 Ap2 x 1.2

4.23 (d)

4.23 (e)

Figure 4.23 Effects of changes in secondary pyrolysis rates (a) on solid (Ts) and gaseous (Tg) temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward)

136

Chapter 4 Model results and discussions A6 = pre-exponential factor in secondary tar oxidation rate equation

1800

A6 x 1.0 A6 x 0.5 A6 x 1.5 Ts

Temperature (K)

1500 1200 900 600 300 640

Tg

620

600

580

560

Distance from grate (mm)

Fig 4.24 (a)


Primary tar (Condensable) 1800 50 1700 Temperature (K) 1600 1500 1400 1300 1200 0.5 1.0 1.5 2.0 Peak Ts Peak Tg Primary tar (mg/Nm) 45 40 35 30 25 20 15 10 5 0 0.5 1.0 1.5 2.0 0 0.2 0.1 0.3 0.5 0.4 Secondary tar (vol%) 0.6 Secondary tar (Non-condensable)

Tar oxidation ratem ultiplication factor (-)

Tar oxidation rate m ultiplication factor (-)

4.24 (b)

4.24 (c)

18
Reaction zone velocity (mm/sec)

0.035 0.030 0.025 0.020 0.015 0.010 0.005 0.000 A6 x 0.5 A6 x 1.0 A6x 1.5 A6 x 2.0

16
Compostion (vol%)

14 12 10 8 6 4 2 0
CO CO2 H2O Gas species H2 CH4 A6 x 0.5 A6 x 1.0 A6x 1.5 A6 x 2.0

4.24 (d)

4.24 (e)

Figure 4.24 Effects of changes in secondary tar oxidation rate (a) on solid(Ts) and gaseous(Tg) temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward)

137

Chapter 4 Model results and discussions

4.3.10 Effective thermal conductivity The effective thermal conductivity for the gasifier fuel bed was derived from a complex model as explained in subsection 2.4.3 (Kunii & Smith 1960). An analysis of the sensitivity of the model outputs to changes in effective thermal conductivity (Keff) is presented in Figure 4.25. All model outputs except for gas composition were found to be sensitive to variations in the effective thermal conductivity. At a higher effective thermal conductivity, more heat from the oxidation zone is transferred to the upper part of the reactor bed while the solid and gas phase temperatures in the combustion zone fall (Figure 4.25 (a) & Figure 4.25 (b)). As a result, rates of tar cracking decrease and the gas has a higher tar content (Figure 4.25 (c)). Since the solid temperature falls in the case of a higher effective thermal conductivity, the rates of the heterogeneous gasification reactions are reduced and, as a result, a lower amount of CO and H2, and a higher amount of H2O is found in the outlet gas. Also, a larger amount of CH4 is present due to less CH4 combustion in the oxidation zone (Figure 4.25 (d)). At higher effective thermal conductivity levels, conduction of heat upwards in the bed increases and pyrolysis occurs at a faster rate (Figure 4.25 (e)). Under such conditions the reaction zones rise up the gasifier. To restore stable operation under such conditions the input air flow rate would have to be increased.

4.3.11 Heat loss through reactor wall In this study, an empirical correlation suggested in the literature (Froment & Bischoff 1979) was used to find the bed-to-wall heat loss coefficient. Total heat loss through the reactor wall is directly controlled by the bed-to-wall heat transfer coefficient and by the reactor inner wall temperature, which depends on the thickness and type of materials in the reactor wall and insulation. In this model, the inner wall temperature was assumed to vary linearly with solid and gas temperatures over time.

138

Chapter 4 Model results and discussions

1800

Keff x 1.5

Keff x 1.0

Keff x 0.5 Ts Tg

Temperature (K)

1500 1200 900 600 300 640 620 600 580 560 540 520 500

Distance from grate (mm)

Fig 4.25 (a)


Primary tar (Condensable) 1900 180 1800 Primary tar (mg/Nm) Temperature (K) 1700 1600 1500 1400 1300 1200 0.5 1.0 1.5 2.0 Peak Ts Peak Tg 160 140 120 100 80 60 40 20 0 0.5 1.0 1.5 2.0 0.1 0 0.5 0.4 0.3 0.2 0.7 0.6 Secondary tar (vol%) Secondary tar (Non-condensable)

Effective conductivity of solid pahse m ultiplication factor (-)

Effective conductivity of solid pahse m ultiplication factor (-)

4.25 (b)

4.25 (c)
0.060 Reaction zone velocity (mm/sec) 0.050 0.040 0.030 0.020 0.010 0.000 keff x 0.5 keff x 1.0 keff x 1.5 keff x 2.0

18 16
Compostion (vol%)

14 12 10 8 6 4 2 0
CO CO2 H2O Gas species H2 CH4 keff x 0.5 keff x 1.0 keff x 1.5 keff x 2.0

4.25 (d)

4.25 (e)

Figure 4.25 Effects of changes in effective thermal conductivity (a) on solid (Ts) and gaseous temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward)

139

Chapter 4 Model results and discussions

The effect of varying the heat loss through the reactor wall was explored by running the model using bed-to-wall heat transfer coefficient values (hbw) ranging from 0.5 to 2.0 times the base case value. The results obtained are presented in Figure 4.26. This shows that, as might be expected, a lower rate of heat loss from the fuel bed to the wall increases the gas phase temperature while a higher heat loss rate has the opposite effect (Figure 4.26 (a) & Figure 4.26 (b)). When the heat loss is small, a low primary tar content is predicted because of the higher gas temperature and the almost complete cracking of the primary tar that this causes. Under these conditions a very clean gas is produced (Figure 4.26 (c)). In the case where heat losses are reduced, elevated H2 and CO levels and lower H2O and CO2 levels are predicted due to the higher steam and CO2 gasification rates which accompany higher solid temperatures along the gasification zone (Figure 4.26 (d)). As shown in Figure 4.26 (e) the effect of a reduction in heat loss is to move the reaction zones downwards.

4.4

Conclusions

The predictions of the transient model developed for a stratified downdraft biomass gasifier are given in this chapter. These results show that solid surface temperatures increase very sharply once the initial char oxidation reaction begins. At the start of the combustion zone, solid surface temperatures are higher than gas temperatures but gas temperatures rise rapidly and exceed solid surface temperatures through the rest of the gasifier. Therefore the peaks of all of the gas phase oxidation reaction rates are located at the end of the oxidation zone where the gas temperature increases sharply. The concentrations of carbon monoxide and hydrogen, the two main combustible components, were very low at the beginning and gradually increased to a stable value. The model showed that 20-25 min was required to reach a steady operating condition. At steady state, some CO was present in the gas leaving the oxidation zone and its concentration increased in the gasification zone. The hydrogen content of the producer gas was mainly determined by steam-carbon reactions in the char gasification zone.

140

Chapter 4 Model results and discussions

1800

hbw x 1.0 hbw x 1.5

hbw x 0.5

Temperature (K)

1500 1200 900 600 300 700

Ts Tg

650

600

550

500

450

Distance from grate (mm)

Fig 4.26 (a)


Primary tar (Condensable) 1900 350 1800 300 Primary tar (mg/Nm) 1700 Temperature (K) 1600 1500 1400 1300 1200 0.5 0.8 1.0 1.5 2.0 250 200 150 100 50 0 0.5 0.8 1.0 1.5 2.0 0.58 0.57 Secondary tar (vol%) 0.56 0.55 0.54 0.53 0.52 0.51 0.5 0.49 Secondary tar (Non-condensable)

Peak Ts Peak Tg

Bed to w all heat transfer coefficient m ultiplication factor (-)

Bed to w all heat transfer coefficient m ultiplication factor (-)

4.26 (b)

4.26 (c)

18
Reaction zone velocity (mm/sec)

0.045 0.040 0.035 0.030 0.025 0.020 0.015 0.010 0.005 0.000 hbw x 0.5 hbw x 0.8 hbw x 1.0 hbw x 1.2

16
Compostion (vol%)

14 12 10 8 6 4 2 0
CO CO2 H2O Gas species H2 CH4 hbw x 0.8 hbw x 1.0 hbw x 1.2 hbw x 1.5

4.26 (d)

4.26 (e)

Figure 4.26 Effects of changes in the extent of heat loss through the reactor wall (a) on solid (Ts) and gaseous (Tg) temperature profiles (b) on peak temperatures (c) on outlet primary and secondary tar levels (d) on outlet gas composition (e) on reaction zone movement (+ve for upward, -ve for downward)

141

Chapter 4 Model results and discussions

Drying and the primary pyrolysis process were observed to start and be completed within a distance of a few millimetres just above the oxidation zone. Within this zone the solid surface temperature rises very steeply. Since most of the primary tars are cracked into secondary tars at the high temperatures present in the oxidation zone, more secondary tars (hydrocarbons) than primary tars (oxygenates) are found in the outlet gas from the gasifier. Model results also predict an average range of tar content between 20 to 200 mg/Nm for the range of input parameters used in the model simulation, assuming that the fuel bed below the ignition ports was initially full of charcoal. The stability of the reaction zones was also investigated and for stable operation it was found that the specific gasification rate needed to be 290-296 kg/m2hr (dry basis) which is in agreement with experimental results obtained on an open core biomass gasifier (Milligan 1994). In order to obtain a detailed understanding of model sensitivity, a systematic analysis of the effect of varying individual operating and model parameters on predictions of temperature profile, outlet gas tar content, gas composition and reaction zone stability has been carried out. A higher air supply rate increased both solid and gas phase temperatures and reduced the outlet gas tar content. Over the wide range of air inlet rates investigated, an increase in the inlet air flow rate was found to lead to a decreased upward movement of the reaction zone, and vice versa. At a certain stage, a further increase in the inlet air rate leads to the reaction zone changing direction and moving downwards. This makes it important for the long term operability of the reactor for the airflow rate to be fixed at a value ensuring stable operation. Using a larger particle size was found to significantly increase the gas phase temperature and to reduce both primary and secondary tar output. However, for a smaller char particle size, due to the higher rates of mass transfer achievable, gasification was possible at lower temperatures than those necessary for a bed of larger particles to achieve the same extent of carbon conversion in the gasification zone. Increasing the initial moisture content of the fuel has a negative effect on the quality of the outlet producer gas.

142

Chapter 4 Model results and discussions

As the solid-to-gas heat-transfer coefficient is increased, gas temperatures increase faster, a higher proportion of the oxygen present is used to oxidize the volatile gases, and the gas temperature increases. Both primary tar and secondary tar content were found to be sensitive to the ratio of the reacting to non-reacting solid-to-gas heattransfer coefficient. Changing the mass transfer coefficient affected the location and maximum temperature of both the solid and gas phases. Decreasing either the mass transfer coefficient or the exchange area favoured gas oxidation and reduced the quantity of the outlet tar. An increase in the primary pyrolysis rate leads to the reactor operating in a pyrolysis dominant mode and the reaction zone moves upwards. Increasing the secondary pyrolysis rate has the effect of reducing the primary tar content of the outlet gas. The model was found to be insensitive to changes in the rate of tar oxidation. Temperatures, tar content and reaction zone stability were all significantly affected by variations in effective thermal conductivity. By reducing the heat loss through the wall, the tar content of the outlet gas can be reduced; however, the reaction zone moves downwards due to increases in heterogeneous reaction rates. The sensitivity analysis provided a good understanding of the effects of operating and model parameters on the performance of a stratified downdraft gasifier. However, the model developed still needed to be validated against experimental data. The validation of the developed model could not be done using experimental data in the literature since those focus on temperature profiles and outlet gas composition and there is no experimental data available on tar profiles along the downdraft gasifier. It was therefore necessary to conduct an experimental investigation on a stratified downdraft gasifier for the purpose of model validation. This investigation is described in the following chapters, where details of the experimental set up, experimental procedures, experimental results and a comparison between model predictions and experimental results are presented.

143

Chapter 4 Model results and discussions

4.5

References

Bhattacharya, A, Salam, L, Dudukovi, B & Joseph, B 1986, Experimental and modelling studies in fixed-bed char gasification, Ind. Eng. Chem. Process Des. Dev., vol. 25, pp. 988-96. Cho, YS & Joseph, B 1981, Heterogeneous model for moving-bed coal gasification reactors, Ind. Eng. Chem. Process Des. Dev., vol. 20, pp. 314-8. Cooper, J & Hallett, WLH 2000, A numerical model for packed-bed combustion of char particles, Chemical Engineering Science, vol. 55, pp. 4451-60. Dzhapbyev, K, Miropolskii, AL & Malkovskii, VJ 1986, Investigation of unsteady heat transfer in a packed bed of spheres swept by gas, Thermal Engng., vol. 33, no. 3, p. 159. Dasappa, S & Paul, PJ 2001, Gasification of char particles in packed beds: analysis and results, International Journal of Energy Research, vol. 25, pp. 1053-72. Di Blasi, C 2000, Dynamic behaviour of stratified downdraft gasifiers, Chemical Engineering Science, vol. 55, pp. 2931-44. Earp, DM 1988, Gasification of biomass in a downdraft reactor, PhD Thesis, Aston University, Birmingham, UK. Earp, DM, Reyes-Nunez, LR, Evans, GD & Bridgwater, AV 1990, Mass and energy balances over an open-core downdraft gasifier, Biomass for Energy and Industry, vol. 2, EC Conference 9-13 Oct 1989, eds. Lisbon, G Grassi, G Gosse and G dos Santos, Elsevier Applied Science, London. Ghani, MU, Radulovic, PT & Smoot, LD 1996, An improved model for fixed-bed coal combustion and gasification: sensitivity analysis and applications, Fuel, vol. 75, no. 10, pp. 1213-26. Hobbs, ML, Radulovic, FT & Smoot, LD 1992, Modelling of fixed-bed coal gasifiers, AIChE Journal, vol. 38, no. 5, pp. 681-702. Kato, K, Kubota, H & Wen CY 1970, CEP Symp. Ser., vol. 66, no. 105, p.87. Kunii, D & Smith, JM 1960, Heat transfer characteristics of porous rocks, AIChE Journal, vol. 6, no.1, pp. 71-8.

144

Chapter 4 Model results and discussions

Milligan, JB 1994, Downdraft gasification of biomass, PhD Dissertation, The University of Aston in Birmingham. Reed, TB & Markson, M 1983, A predictive model for stratified downdraft gasification of biomass, Progress in biomass conversion, vol. 4, eds. DA Tillman & EC John, pp. 217-54. Reed, TB & Markson, M 1985, Biomass gasification reaction velocities, Fundamentals of Thermochemical Biomass Conversion, eds. RP Overend, TA Milne & LK Mudge, pp. 951-65. Reed, TB, Levie, B & Graboski, MS 1987, Fundamentals development and scale up of the air oxygen stratified downdraft gasifier, Solar Energy Research Institute, SERI/PR-234-2571.

145

Chapter 5 Test rig and experimental procedure

Chapter 5

Test rig and experimental procedure

5.1

Introduction

Experiments were carried out in a stratified downdraft biomass gasifier using wood blocks as a biomass fuel. The overall schematic of the gasifier system is shown in Figure 5.1. The main components of the experimental set-up consisted of an air supply unit, a stratified downdraft gasifier, a wood fuel ignition system, a producer gas burner and a gas sampling train. The experimental set-up, measurement systems, raw materials and experimental procedure are described in this chapter.
P12 flue gas

Orifice meter

Stratified Downdraft Gasifier Rotameter Thermocouple connections To gas sampling train

Control Valve

Air from main blower

T1 Suction blower

Figure 5.1 Experimental set-up

146

L.P.G

Chapter 5 Test rig and experimental procedure

5.2

Experimental set-up

5.2.1 Experimental stratified downdraft gasifier A small scale stratified downdraft gasifier (40- 75 kW thermal input), pictured in Figure 5.2 and shown schematically in Figure 5.3, was designed and fabricated for the purpose of the experimental work. A range of photographs of the assembled gasifier, gasifier components and auxiliary equipment is shown in Appendix G. The gasifier, which was cylindrical in shape, was fabricated from 6 mm thick mild steel tube with a 206 mm inner diameter. The reactor can be divided into three sections, described in Figure 5.3 and also pictured in Figure G.2, Appendix G. The top section, through which air was supplied to the reactor, consisted of an inlet section shaped like an inverted funnel mounted above the reactor main cylindrical body. No reaction was expected for 700 mm below the inlet so that the lower section acted in practice as a fuel storage bunker. In the middle section, groups of four 12 mm OD stainless steel tubes were inserted into the reactor at 50 mm intervals along the reactor down to the level of the grate. These tubes were used for the purposes of temperature measurement and gas sampling and were located symmetrically around the reactor circumference. Twelve 25 mm OD stainless steel tubes were attached at three different levels in the middle part of the reactor for the purpose of ignition. The bottom section contained a grate located at a distance of 1500 mm below the top of the reactor tube. The grate was designed so that it could be rotated for the purpose of cleaning. The part immediately below the grate served as a gas solid separator, and below that an ashbin was fitted. A horizontal 50 mm (ID) diameter pipe attached to the gasifier below the grate served as the producer gas outlet. After three experimental runs had been carried out, some damage to the reactor wall of the middle section was observed. Several small holes (about 10mm diameter) were found to have developed just below the ignition ports, presumably due to the very high temperatures in this zone. Before the experimental work could be continued, major

147

Chapter 5 Test rig and experimental procedure

repairs had to be carried out to the reactor. A new middle section was fabricated using mild steel tube (O.D 356mm X I.D 340mm). A 67 mm thick layer of bonded castable (Tufcast 1700QT) insulation was applied to the inside surface of the gasifier so that the effective internal diameter of the new middle section was the same as that of the top and bottom sections. A detailed schematic diagram of the modified middle section is shown in Figure 5.4 and the step-by-step fabrication process of the new middle section is illustrated in Figure G.3 in Appendix G. The temperature measuring ports and ignition ports were located such that the original spacing was maintained.

Figure 5.2 Picture of experimental stratified downdraft gasifier

148

Chapter 5 Test rig and experimental procedure

80 mm

cover

300 mm 160 mm

Top Section
6 mm thick 12 mm ports located 50mm apart- for temperature measurement Ignition ports (25 mm stainless steel; 100 mm apart) 1500 mm and gas sampling C C Reactor wall

DETAIL A 206 mm

Middle Section
700 mm 600 mm

12 mm

12 mm

65 mm

120 mm 20 mm bar SECTION CC

Inner Dia 206mm Grate DETAIL A

Gas Outlet Port 50 mm pipe 200 mm 100 mm 100 mm Ash pit 100 mm 170 mm

Bottom Section

Supports

100 mm

Figure 5.3 Schematic diagram of experimental stratified downdraft gasifier

149

Chapter 5 Test rig and experimental procedure

Castable insulation (Tufcast 1700QT)

45

100 mm

80 mm 6 mm thick flange 422 PCD, Mild steel 25 mm 50 mm 50 mm 225 mm

12 mm stainless steel tubes (every 50 mm along the reactor) Castable (Tufcast 1700QT) 206 mm 25 mm stainless 350 mm steel tube 100 mm 100 mm 800 mm

25 mm

Figure 5.4 Schematic diagram of the modified middle section of the reactor

150

Chapter 5 Test rig and experimental procedure

5.2.2 Thermal insulation To prevent excessive heat loss from the reactor wall, two layers of a ceramic fibre blanket of thickness 25 mm were wrapped around the reactor. To prevent the condensation of tar and moisture along the gas-sampling pipes, two layers of ceramic fibre blanket of thickness 25 mm were also wrapped around them.

5.2.3 Flare and burner A L.P.G torch was used both for ignition of fuel inside the gasifier and for ignition of producer gas downstream of the reactor. A stainless steel burner, more details of which are shown in Figure 5.5, was used to flare the producer gas. A pilot flame from the L.P.G torch was injected through one side of the burner, air was introduced through the other sides of the burner, while the producer gas entered through the bottom of the burner. The pilot flame from the L.P.G torch was turned on throughout the reactor warm up process to ensure the complete combustion of the very low calorific value producer gas formed during the reactor start-up stage.
Flue gas

100 mm stainless steel tube 30 mm holes Pilot flame Air

L.P.G Smoke from suction blower Producer gas

Figure 5.5 Schematic diagram of burner set-up

151

Chapter 5 Test rig and experimental procedure

5.2.4 Filter The particulates (mainly fine char particles and some ash) in the producer gas sample had to be removed before moisture and tar collection. Particle removal from the sample gas was undertaken with a cotton wool filter installed in a 35mm steel housing, which was connected at the entrance of the gas sampling line. The filter housing was heated and also insulated with a ceramic fibre blanket to avoid condensation of tar and moisture inside it. The filter was replaced with a new one after each experimental run to avoid blockage and no trace of condensate was observed in the used filters.

5.2.5 Air blower In the gasification process, external power is required for moving the air and producer gas through the gasifier, which may be done by sucking or blowing. In the experimental set-up, a blowing system was employed and a rotary piston compact blower was used for that purpose. The capacity of the air blower was much higher than the required air supply rate for some experimental runs. Therefore a by-pass valve was attached to the blower. This enabled some air to be discharged to the atmosphere preventing pressure build up and overloading of the blower motor. The detailed specifications of the main air blower are presented in Table 5.1.

Table 5.1 Size and specification of main air blower and suction blower

Main air blower (Rotary Piston Compact Blower Unit) Roots Compressor Type Shaft power (kW) Motor power (kW) Max speed (rpm) Air vol. flow rate (m/min) Inlet pressure, Outlet pressure and p (kPa) Suction blower (Rotary Blower Unit) Motor power (kW) Speed (rpm) Amps (A) Voltage (V) 0.375 1430 1.2 400/415 15.6 18.65 2500 25.6 101, 128.6, 27.6

152

Chapter 5 Test rig and experimental procedure

5.2.6 Suction blower In this experiment, ignition of solid fuel inside the gasifier was achieved by sucking in the flame from a torch, which was held in front of the air suction ports, using a small blower. A rotary blower, for which motor specifications are presented in Table 5.1, was used for that purpose. After ignition of the solid fuel, the suction blower was switched off and only the main blower was used.

5.3

Measurement systems and instrumentation

In this section, the gas-sampling set-up and instrumentation are discussed in detail. The measured parameters for the experiments included: tar content of the producer gas at five different levels in the reactor; gas composition at five different levels in the reactor; temperatures at each of 16 different levels in the reactor; inlet air pressure; inlet airflow rate; producer gas flow rate; and wood consumption rate.

5.3.1 Gas sampling train In this experiment, producer gas from five different levels in the gasifier was sampled using a recently proposed standard gas-sampling procedure guideline (Neeft et al. 2002). Figure 5.6 shows the modular sampling train, which was set up according to the guideline, and which consisted of 3 modules; namely, gas preconditioning and particle removal, tar collection and volume measurement. In the preconditioning module, the temperature of the producer gas was maintained within a temperature range of 300-350C using electric heating tape. Fine particles in the sample gas were removed by a heated filter which is described in detail in section 5.2.4. This module also included a stainless steel airtight (sealed) lock. The tar collection module consisted of a moisture collector, five impinger bottles and a water bath with two chambers. The moisture collector and impinger bottles were connected in series so that the outlet of one bottle was joined to the inlet of the next.

153

Chapter 5 Test rig and experimental procedure

The moisture collector was empty and each of the five impinger bottles contained approximately 50 ml of absorbing solvent (isopropanol). The moisture collector and three of the impinger bottles were placed in one chamber of a water bath at 20C. Two other impinger bottles were placed in the other chamber which contained a cooling mixture of salt, ice and water. The main purpose of this module was to cool down the sample gas gradually from 20C to 15/-20C. The detailed dimensions of the moisture collector and impinger bottles used in the experiment are shown in Figure 5.7. The glass tube of the moisture collector was made the same size as the heated stainless steel sampling tube (12 mm OD) so that these two could be joined without any leakage by using a stainless steel tube union. The volume measurement module consisted of a gas suction pump, a flow indicator, pressure and temperature sensors and a volume flow meter. A control valve was also attached at the inlet to the flow indicator to adjust the sample flow rate.

P T
MODULE 1 Electric Heating Filter Rotameter Airtight lock Flow Control Valve Condensate Collection Cooling Water (T= 20C) MODULE 2 Salt/ice/water mixture (T = -15/-20C) Dry Gas Meter To gas sampling container

Vacuum Pump

MODULE 3

Figure 5.6 Gas sampling train

154

Chapter 5 Test rig and experimental procedure

in OD tube Conical ground joints 24/29

Din = 5mm s =1.5 mm Rubber seal 24/29

Din=32mm s = 3 mm

285 mm

(a) Moisture collector

(b) Impinger bottle

Figure 5.7 The moisture collector and impinger bottle used in experiments

The sampling line was kept as short as possible for all sampling points. A sampling train such as that described above is designed to process gas only from one designated point in the reactor at any given time. In this study, gas samples needed to be collected from 5 different levels in the reactor. Therefore 5 sets of modules 1, 2 & 3 were set up. A common water bath and a common vacuum pump were used for these 5 trains. A direct drive high vacuum pump was used as the common vacuum pump and its specifications are presented in Table 5.2. A small vacuum pump, for which details are described in Table 5.2, was also used to transfer the gas from each sampling line to the sampling containers. A schematic of the modified gas sampling train set-up is shown in Figure 5.8 and photographs of the set-up are also presented in Appendix G (Figure G.5).

155

Chapter 5 Test rig and experimental procedure

Electric heating Producer gas Airtight lock Producer gas

Moisture collector

Impinger bottles Flow Control Valve Rotameter Dry Gas Meter


P T

To gas sampling container via small vacuum pump

Filter
P T

Producer gas

Common vacuum pump

Producer gas

Producer gas

Cooling Water (T= 20C)

Salt/ice/water mixture (T = -15/-20C)

Figure 5.8 Modified gas sampling train

Table 5.2 Specifications of common vacuum pump and vacuum pump for gas collection

Common vacuum pump (Direct drive high vacuum pump) Motor Power (kW) Speed (rpm) Amps (A) Voltage (V) Vacuum pump for gas collecting Maximum vacuum (kPa) Amps (A) Voltage (V) Frequency (Hz) 101.325 1.9/2.2 220/240 50 0.375 1430 4.3 230/240

156

Chapter 5 Test rig and experimental procedure

5.3.2 Measurement of tar content Gravimetric tar measurement enables quantitative measurement of the mass of the portion of a tar sample that is not amenable to GC analysis (Brage & Sjstrm 2002). The procedure, described below, is that laid down in a recently proposed standard (Neeft et al. 2002). Figure G.7 in Appendix G illustrates some of the equipment used. After tar collection was complete, the solvent, water and tar mixture collected from the impingers and condenser were mixed in one flask(A). The specified amount (50-100 ml) of solvent was used during rinsing of the moisture collector, impingers and connectors. Due to the solubility of tar in the solvent, the tars can be assumed to be equally distributed through the solvent (S Paasen 2003, pers. comm., 20 January). About 100 ml of the tar solution was poured into a 250 ml flask pre-weighed to an accuracy of 1 mg. The weight of the flask plus tar solution was measured to determine the weight of the solution. The flask was connected to a rotary evaporator and evaporation was started with the water bath at 55C and the vacuum pressure controller set at 100 mbar. The rate at which drops were falling from the cold finger was checked to ensure the rate was around 1-2 drops per second. Once almost all the solvent had evaporated, the rate of drop fall was decreased to 1 drop per 4 seconds. The evaporation was continued for a further 15 minutes. When no further traces of water were observed, the vacuum was released by letting in air and the rotation of the flask was stopped. The flask was removed from the heated water bath and dried. The flask and tar were then left for 5 minutes before the flask was weighed. The amount of gravimetric tar was determined as follows: The concentration of gravimetric tar in producer gas (mg/Nm) where W0 = Weight of empty flask(A) (g) W1 = Weight of empty flask(A) + tar solution (g)
=

(M

M0) V

(W1 W0 ) 1000 (M 1 M 0 )

(5.1)

157

Chapter 5 Test rig and experimental procedure

M0 = Weight of empty 250 ml flask (g) M1 = Weight of 250 ml flask + 100 ml of tar solution (g) Mf = Weight of tar containing 250 ml flask after evaporation (g) V is the sample gas volume in m at normal conditions, which is the difference between the initial and final readings of the gas meter during the sampling procedure, corrected for the actual pressure and temperature of the sampled gas at the dry gas meter. Figure 5.9 summarizes the post sampling procedure of gravimetric tar measurement.

Sampling from gasifier Empty and rinse with solvent Solvent + water + tar Solvent + water + tar Solvent + water + tar

MOISTURE COLLECTOR

IMPINGERS

solvent added

Empty and rinse with solvent

SAMPLING LINES AND CONNECTORS

Rinse with solvent

Weigh residue

Rotary evaporate bulked sample (100ml)

Bulked solvent sample

Calculate concentration of gravimetric tar

Recover solvent

Figure 5.9 Schematic showing post-sampling procedures for gravimetric tar measurement (Source: Neeft et al. 2002)

158

Chapter 5 Test rig and experimental procedure

5.3.3 Measurement of gas composition The gas composition was measured by (Varian 3700) gas chromatograph (GC) using discrete samples taken periodically from the gas-sampling stream. 400ml PVC gas sample containers (see Figure G.6) were used for batch sampling. The sample containers were first thoroughly flushed with a small positive pressure of sampling gas and later filled with it so that air could not leak in before analysis. Next, 5-10 ml of gas sample was extracted through the rubber seal of the gas sample container with a hypodermic syringe for injection into the gas chromatograph. A thermal conductivity detector was used to analyse the gas composition and was able to measure the volume fraction of H2, CO, CH4 and CO2 by integrating the area under the peak in the curve and comparing it with that for a calibration gas of known composition.

5.3.4 Measurement of temperature Chromel-Alumel (K-Type) thermocouples [these are capable of measuring to within 8C accuracy over the temperature range of the combustion zone (900-1200 C)] were used to measure temperatures at 16 different levels along the experimental reactor and also the outlet gas temperature. Copper-Constantan (T-Type) thermocouples were used to measure the sampling gas temperature and inlet air temperature. These thermocouples were connected to a data logger (DataTaker 500) which recorded the temperature values every 3 seconds.

5.3.5 Measurement of pressure Inlet air pressure and the pressure drop across the orifice plate meter were measured by using pressure transducers which were also connected to the data logger; this recorded the sensor outputs every 3 seconds. For the measurement of inlet air pressure, a Setra pressure transducer [Model 206. 25 PSIG (0.15% accuracy)], which can measure a range of gauge pressure between 0 and 25 psi (0 and 172 kPa) was used. Pressure drop through the orifice plate was measured by a Setra pressure transducer [Model 264. 25 WG (1.0% accuracy)], which can measure differential pressure in the range 0-25 in H2O (0-6.2 kPa).

159

Chapter 5 Test rig and experimental procedure

5.3.6 Measurement of supplied airflow rate Air supply rate to the gasifier is one of the most important input operating parameters and it was necessary to measure this accurately and continuously. Airflow rate was measured by an orifice meter consisting of a washer-shaped plate with a central hole of diameter d that was placed in the centre of the air pipeline. The positions of the upstream and downstream pressure tappings, together with detailed dimensions of the orifice plate used in the experiment, are shown in Figure 5.10. By knowing the upstream pressure and temperature, and the differential pressure across the orifice plate (as mentioned in sections 5.3.4 & 5.3.5), the airflow rate can be calculated using a standard method (British Standards Institution 1965), details of which are presented in Appendix E.

D D

D/2

d
D = 28.65 mm t d = 12.27 mm t = 1.6 mm

Figure 5.10 Constructional arrangement of experimental orifice plate

5.3.7 Measurement of gas flow rate When a gas is cool and clean, its flow rate can be directly measured by flow measurement devices such as rotameters, Pitot tubes, orifice meters, venturi meters, etc. However, in this experiment, hot producer gas was flared without prior cooling. Therefore, the total gas flow rate was not measured directly but indirectly estimated by means of a nitrogen balance across the gasifier, assuming the nitrogen content of wood to be negligible.

160

Chapter 5 Test rig and experimental procedure

Producer gas flow rate (Nm/hr) =

Air supply rate (Nm/hr) x N2 Vol fraction in air N2 Vol fraction in producer gas

(5.2)

By knowing the volume fraction of major gas species (H2, CO, CH4 and CO2) in the final producer gas, the balance could be assumed to be N2 since levels of O2 and other gases such as hydrocarbons would be negligible at the reactor outlet.

5.3.8 Measurement of wood consumption rate Average wood consumption rate was estimated by measuring the fuel volume consumed during the experiments. Wood was normally loaded up to the top of the reactor at the start of an experiment and the height of the fuel layer from the reactor top after the experimental run was measured to calculate the wood consumption rate as follows: Wood bulk Wood consumption rate (kg/hr) = density (kg/m) Reactor crossx section area (m) x Length of wood column consumed (m) (5.3)

Duration of experiment (hr)

5.3.9 Cold gas efficiency If the producer gas is used as a fuel in an internal combustion engine, the sensible heat of the gas is not utilized. The conversion efficiency estimated based on the energy in the cold gas is called the cold gas efficiency and it is defined as follows: Cold gas efficiency (%) where Energy carried by producer gas = Energy supplied to gasifier VFgas CRwood VFgas x NHVgas CRwood x NHVwood (5.4)

= volumetric flow rate of producer gas (Nm/hr) = consumption rate of the wood blocks (kg d.b /hr)

NHVgas = net heating value of producer gas (MJ/Nm) NHVwood = net heating value of the wood (MJ /kg d.b)

161

Chapter 5 Test rig and experimental procedure

5.4

Raw materials

The raw materials used for the stratified downdraft gasification study were mainly wood blocks, with some charcoal used for the reactor warming up process.

5.4.1 Wood blocks Red gum (Eucalyptus camaldulensis) wood sticks of square cross-section were purchased from a local red gum wood supplier in two different sizes: as 20mm x 20 mm x 1800 mm long and 30 mm x 30 mm x 1800 mm long. Since the wood sticks were only available in green wood condition, the original average moisture content was found to be as high as 45% db. Therefore, purchased wood sticks were dried in a kiln for about 10 days and the average moisture content was decreased to 13% db. The dried wood sticks were cut into two different cubic block sizes (20 x 20 x 20 mm and 30 x 30 x 30 mm) using a band saw. The wood blocks were stored in drums for experimental use. Pictures of samples of the two different sized wood blocks are shown in Figure 5.11. The measured properties of the red gum wood blocks are presented in Table 5.3.

Figure 5.11 The 30 mm and 20 mm sized wood blocks

162

Chapter 5 Test rig and experimental procedure

Table 5.3 Average properties of red gum wood blocks

Characteristic Gross Heating Value (MJ/kg) Apparent density (kg/m) Small Blocks Size (mm) Bulk Density (kg/m) (20 mm cube) Void fraction, (-) (20 mm cube) Moisture content (% w.b) Larger Blocks Size (mm) Bulk Density (kg/m) (30 mm cube) Void fraction, (-) (30 mm cube) Moisture content (% w.b) Proximate Analysis (wt % d.b) Volatile Matter Fixed Carbon Ash Ultimate Analysis (wt % d.b) C H O N S Ash

Value 20.04 951

Test Method (*) ASTM - D 2395-83

20 x 20 x 20 527 0.45 9.7 ASTM - E 871-82 ASTM - E 873-82

30 x 30 x 30 485 0.49 12.2 ASTM - E 871-82 ASTM - E 873-82

75.95 23.92 0.13

ASTM - E 872-82 ASTM - E 870-82 ASTM - D 1102-82

52.90 5.30 41.63 0.01 0.03 0.13

**HRL Method 1.4 HRL Method 1.4 HRL Method 1.4 AS 1038.6.3.3 HRL Method 1.6

* Calculated from values of ultimate analysis. (Reed & Agua 1988).


**

HRL is the NATA registered laboratory that analysed the wood sample.

163

Chapter 5 Test rig and experimental procedure

5.4.2 Red gum charcoal Suitably sized charcoal is required for initial start-up of the gasifier to avoid the formation of large amounts of tar within the reactor under warm-up conditions. Bags of red gum wood charcoal were purchased from a local charcoal supplier. Since the size of the charcoal blocks available from the market was bigger than the size required in this study, these were manually broken into smaller pieces to facilitate the feeding and burning. To achieve a uniform size, broken charcoal blocks were sieved to obtain the 2030 mm fraction. The selected charcoal pieces were stored in drums for experimental use. The measured properties of the red gum charcoal blocks are presented in Table 5.4.

Table 5.4 Average properties of red gum charcoal blocks

Characteristic Size (mm) Bulk Density (kg/m) (20 mm cube) Moisture content (% w.b) Proximate Analysis (wt % d.b) Volatile Matter Fixed Carbon Ash

Value 20-30 244 5.06

ASTM Standards Test Method No.

E 873-82 E 871-82

13.58 86.90 0.52

E 872-82 E 870-82 D 1102-82

5.5

Experimental procedure

At the beginning of the first run, the gasifier was filled with sized charcoal up to the ignition level (the air suction port) to avoid tar formation during start-up. The rest of the reactor was then filled up with wood blocks. After loading the fuel, the top cover was closed and a suction blower was started up in order to suck in a flame from a torch held in front of the four air suction ports.

164

Chapter 5 Test rig and experimental procedure

After 1-2 minutes of ignition time, combustion was well developed inside the gasifier and the suction ports were closed. Then air at the desired rate was supplied to the top of the reactor. It required about 10-15 minutes to achieve steady operation, by which time the heating value of the producer gas was high enough for self-burning without a pilot flame. For each of a series of predetermined airflow rates, the outlet gas was sampled and the composition and tar content of the producer gas were measured as discussed in section 5.3. The experimental run continued until the wood blocks had burned down to the level of the ignition ports; this took 1.0 to 2.0 hours depending on the rate of air supplied. Experimental runs were carried out for various rates of air supply for both the two different sizes of wood blocks. Details of the experimental conditions are given in the next chapter (Chapter 6). The setting up of the experimental rig and the carrying out of experimental runs were conducted in strict accordance with CSIRO occupational health and safety regulations. Tar collection and gravimetric measurements were carried out in accordance with the safety precautions laid down in the sampling guideline (Neeft et al. 2002).

5.6

Summary

A stratified downdraft gasifier of a small scale (206 mm ID) was designed and fabricated for the purpose of this experimental study. Air was supplied to the gasifier by a rotary piston compact blower. Ignition of solid fuel inside the gasifier was achieved by sucking in a flame from a torch, which was held in front of air suction ports, using a small blower. A stainless steel burner was attached downstream of the system to flare the producer gas. Temperatures at 16 different levels along the reactor were measured by K-type thermocouples. The inlet airflow rate was measured using an orifice meter. Producer gas from five different levels of the reactor was sampled through a gas sampling train which was slightly modified from a standard gas sampling system. The tar content of sampled gas from the five different levels was determined by a standardized gravimetric tar measurement method. The volume fractions of H2, CO, CH4 and CO2

165

Chapter 5 Test rig and experimental procedure

in the gas samples were measured by gas chromatography. Total gas flow rate was calculated by carrying out a nitrogen balance across the gasifier and the average wood consumption rate was estimated by measuring the fuel volume consumed during the experiments. Two different sizes of red-gum wood blocks were used as fuel. Suitably sized red gum charcoal was also used as ignition initiation fuel. The performance of the gasifier was determined for various rates of inlet air supply. In the following chapter the results of the experimental work on the stratified downdraft gasifier are described and the results obtained are discussed.

166

Chapter 5 Test rig and experimental procedure

5.7

References

Brage, C & Sjstrm, K 2002, An outline of R&D work supporting the tar guideline, report related to the project Tar Guideline (project no EEN5-00507). British Standards Institution 1965, Methods for the measurement of fluid flow in pipes: Part1. Orifice Plates, Nozzles and Venturi Tubes, B.S. 1042 : Part 1 : 1964. Neeft, JPA, Knoef, HAM, Zielke, U, Sjstrm, Hasler, P, Simell, PA, Dorrington, MA, Abatzoglou, N, Deutch, S, Greil, C, Buffinga, GJ, Brage, C & Suomalainen, M 2002, Guideline for sampling and analysis of tar and particles in biomass producer gases (Version 3.3), Energy project EEN5-1999-00507 (Tar protocol). Paasen, SVB van 2003, email, 20 January, <vanpaasen@ecn.nl>, Energy Research Centre of the Netherlands (ECN), The Netherlands. Reed, TB & Agua, D 1988, Handbook of biomass downdraft gasifier engine systems, Biomass Energy Foundation Press, Washington.

167

Chapter 6 Experimental results and model validation

Chapter 6

Experimental results and model validation

6.1

Introduction

As discussed earlier, the objective of this project was to improve our understanding of how stratified downdraft gasifiers work. A numerical model was constructed, the detailed development of which is discussed in Chapter 3. The predictions of this model, together with details of a sensitivity analysis carried out to test the effect of changes in different operating and model parameters on model predictions, were described in Chapter 4. Unlike earlier stratified downdraft gasifier studies, that have largely focused on temperature profiles and gas compositions, the model developed in this project was designed also to provide information on tar formation and reduction within the body of a gasifier using wood as a fuel, under a range of operating conditions. Validation of this model could not be done using previously published experimental data since there are no experimental data available on tar profiles along a downdraft gasifier either for a traditional (with throat) or a stratified (without throat) reactor. In addition, there is very limited experimental data on gas composition profiles along the reactor axis. It was therefore necessary to conduct an experimental investigation on a stratified downdraft gasifier constructed specifically for the purpose of model validation. Details of the experimental set-up and experimental procedures are presented in Chapter 5 while this chapter presents the experimental results and model validation procedures.

6.2

Experimental results

A total of 19 experimental runs was conducted. Conditions varied between runs were: the air supply rate (7.5- 27.6 kg/hr) and the size of the wood blocks (20 mm or 30 mm cubes). Normally, as described in Chapter 5, at the start of a run the bed was loaded up

168

Chapter 6 Experimental results and model validation

to the level of the ignition port with charcoal. To determine the effect of this procedure (which was introduced to avoid excessive tar formation during the reactor warm up) an additional experimental run (Run no. 601) was carried out using small wood blocks only and no initial charge of charcoal. During experiments, the main parameters measured or calculated were: temperature profiles along the gasifier, the gas composition profile, the tar profile, fuel consumption, gas production rate, air to fuel ratio, reactor cold gas efficiency and reaction zone velocity. These results are summarised in Table 6.1. More detailed information for individual runs is given in Appendix F.

6.2.1 Material balance and elemental balance The reliance that can be placed on experimental data is determined by the accuracy and completeness of the measurements obtained. Material and elemental balance closures are important indicators of the quality of experimental results since mass has to be conserved across the gasifier. Material balance closure over the entire period of experimental run was calculated by dividing the total mass output by the total mass input (and multiplying by 100 if it was to be expressed as a percentage). The inputs comprised the inlet air and moist wood while the outputs comprised the dry gas and condensate. Elemental balances for oxygen, carbon and hydrogen were conducted in a similar way, i.e. by determining the ratio of the total outputs of each individual element to the total inputs of that element. Details of the overall and elemental mass balances for all experimental runs are presented in Table 6.2 to Table 6.5. Table 6.2 shows that for 14 out of the total 20 experimental runs, material balances closed to within 3%. Another 5 runs had closures within 5% and the last run closed to within 6%. The average value of mass balance closures for all experimental runs was found to be 99.3% and this indicates that considerable reliance can be placed on the accuracy of the experimental measurements. Table 6.3 to Table 6.5 show that good elemental balance closures were also achieved over the full range of input rates and operating conditions.

169

Chapter 6 Experimental results and model validation

Table 6.1 Performance summary of experimental stratified downdraft gasifier


wood block Run no* 311 314 318 320 502 504 506 508 510 513 515 517 522 524 525 527 528 531 604 601
* a b c

Inlet air T (C) 21.3 21.7 32.6 17.5 22.4 20.7 23.2 27.1 20.9 19.7 21.3 23.9 16.2 23.4 21.9 22.6 22.9 22.5 24.1 19.8 P (kPa) 1.42 1.96 2.40 2.41 2.34 1.89 2.33 4.39 4.61 1.31 7.51 2.46 1.94 1.48 2.04 2.44 1.46 3.15 3.53 1.10 (kg/m) 1.22 1.22 1.18 1.25 1.22 1.23 1.22 1.23 1.26 1.22 1.29 1.21 1.24 1.21 1.22 1.22 1.21 1.23 1.23 1.21 Flow rate (kg/hr) 14.80 16.92 20.45 13.32 22.09 19.60 13.50 24.73 7.50 10.71 11.31 15.48 22.71 11.28 15.60 27.63 12.21 13.85 15.34 10.22

Dry gas outlet composition (vol%) H2 14.17 17.98 18.22 18.74 17.29 18.01 17.90 17.75 18.21 17.56 16.36 17.92 16.95 18.20 6.88 15.71 16.76 16.48 16.76 15.00 CO 19.93 19.02 19.24 18.45 18.50 18.30 18.83 18.99 18.98 18.09 15.91 18.79 18.79 17.51 25.44 21.05 19.00 17.79 19.82 19.68 CO2 12.51 12.51 14.33 15.11 12.99 12.23 14.29 15.06 18.18 13.54 17.00 14.00 12.00 15.33 8.48 12.43 14.97 14.29 11.68 15.15 CH4 1.30 1.42 1.49 1.70 1.57 1.44 1.33 1.50 1.53 1.43 1.08 1.34 1.27 1.46 1.00 1.10 1.31 1.48 1.25 1.53 N2 52.09 49.07 46.72 46.00 49.65 50.02 47.65 46.70 43.10 49.38 49.65 47.95 50.99 47.50 58.20 49.71 47.96 49.96 50.49 48.64

size (mm) 20 20 20 20 20 20 20 20 20 20 20 20 30 30 30 30 30 30 30 20

MC (%wb) 9.37 9.32 9.23 8.65 8.58 9.47 11.17 11.12 10.11 10.54 9.06 9.46 12.66 12.46 12.58 12.16 12.39 11.24 12.15 9.83

Gasa LHV (MJ/ Nm) 4.68 5.03 5.11 5.15 4.94 4.95 4.96 5.03 5.09 4.87 4.32 4.97 4.83 4.87 4.47 4.93 4.85 4.73 4.94 4.83

Outlet gas Tb (C) 468 488 460 473 464 483 397 580 393 407 488 424 526 471 514 596 441 501 437 438

Gas flow (Nm/ hr) 17.31 21.08 26.80 17.61 27.24 23.83 17.30 32.28 10.60 13.27 13.90 19.85 27.30 14.49 16.39 34.04 15.56 16.96 18.56 12.93

Tar (mg/ Nm) 58 34 47 37 22 20 50 16 61 65 48 35 88 45 23 62 28 33 59 964

wood rate (kg/hr) 8.00 9.46 10.61 7.53 12.29 7.78 7.28 13.10 5.25 6.11 6.15 7.03 11.72 6.29 6.95 13.78 7.31 7.03 7.91 5.82

AF ratio (-db) 2.04 1.97 2.12 1.94 1.97 2.78 2.09 2.12 1.59 1.96 2.02 2.43 2.22 2.05 2.57 2.28 1.91 2.22 2.21 1.95

Energy in (MJ/hr)c 135.1 159.9 179.5 128.2 209.4 131.3 120.5 217.0 88.0 101.9 104.2 118.6 190.8 102.6 113.3 225.6 119.4 116.3 129.5 97.8

Energy out (MJ/hr)b 81.0 106.1 137.0 90.6 134.6 117.9 85.9 162.4 54.0 64.6 60.0 98.6 131.9 70.6 73.4 167.7 75.5 80.1 91.6 62.4

Cold gas (%) 59.9 66.4 76.3 70.7 64.3 89.8 71.3 74.8 61.4 63.4 57.6 83.1 69.1 68.8 64.8 74.3 63.2 68.9 70.7 63.8

Run numbers were assigned by dates when experiments were conducted. [i.e. Run no 510 means that an experiment carried out on 10th of May (2003)]. LHV of producer gas was calculated from the gas composition by using the value of 11.2 MJ/Nm for H2, 13.1 MJ/Nm for CO and 37.1 MJ/Nm for CH4. (Reed &Agua 1988) Outlet temperatures of producer gas were average values of measured data between 30 -40 min after ignition. Input and output energy of the experimental gasifier was calculated by using the LHV of wood fuel and LHV of producer gas respectively. 170

Chapter 6 Experimental results and model validation

Table 6.2 Material balance summary


Run no 311 314 318 320 502 504 506 508 510 513 515 517 522 524 525 527 528 531 604 601 Air 14.80 16.92 20.45 13.32 22.09 19.60 13.50 24.73 7.50 10.71 11.31 15.48 22.71 11.28 15.60 27.63 12.21 13.85 15.34 10.22 Inputs (kg/hr) Dry wood Moisture Total Gas 7.25 0.75 22.80 20.36 8.58 0.88 26.38 23.76 9.63 0.98 31.06 30.43 6.88 0.65 20.85 20.10 11.24 1.05 34.38 30.95 7.04 0.74 27.38 26.94 6.47 0.81 20.78 19.77 11.64 1.46 37.83 37.16 4.72 0.53 12.75 12.39 5.47 0.64 16.82 15.11 5.59 0.56 17.46 16.43 6.36 0.67 22.51 22.48 10.24 1.48 34.43 30.94 5.51 0.78 17.57 16.62 6.08 0.87 22.55 20.15 12.10 1.68 41.41 39.24 6.40 0.91 19.52 18.05 6.24 0.79 20.88 19.61 6.95 0.96 23.25 21.11 5.25 0.57 16.04 15.16 Average closure (%) Standard deviation <95% mean Outputs (kg/hr) Tar H2O 1.00E-03 1.22 7.17E-04 1.43 1.26E-03 1.83 6.51E-04 1.21 5.99E-04 1.86 4.77E-04 1.08 8.65E-04 1.19 5.16E-04 1.11 6.46E-04 0.74 8.63E-04 0.91 6.67E-04 0.99 6.95E-04 0.45 2.40E-03 2.48 6.52E-04 1.00 3.77E-04 2.02 2.11E-03 2.35 4.36E-04 1.08 5.60E-04 1.18 1.09E-03 1.27 1.25E-02 0.91 Total 21.58 25.19 32.25 21.31 32.81 28.02 20.96 38.27 13.13 16.02 17.42 22.93 33.42 17.62 22.17 41.60 19.13 20.79 22.37 16.08 Closure (%) 94.7 95.5 103.8 102.2 95.4 102.3 100.8 101.2 103.0 95.2 99.8 101.9 97.1 100.3 98.3 100.4 98.0 99.6 96.2 100.3 99.3 2.84

Table 6.3 Carbon balance summary


Run no 311 314 318 320 502 504 506 508 510 513 515 517 522 524 525 527 528 531 604 601 Carbon inputs from wood (kg/hr) 3.84 4.54 5.10 3.64 5.95 3.73 3.43 6.17 2.50 2.90 2.96 3.37 5.42 2.92 3.22 6.41 3.39 3.31 3.68 2.78 Carbon outputs (kg/hr) from from From from CO CO2 CH4 tar 1.86 1.17 0.12 5.36E-10 2.16 1.42 0.16 3.83E-10 2.77 2.06 0.21 6.73E-10 1.76 1.44 0.16 3.48E-10 2.71 1.90 0.23 3.20E-10 2.36 1.58 0.19 2.54E-10 1.75 1.33 0.12 4.62E-10 3.31 2.62 0.26 2.76E-10 1.09 1.04 0.09 3.45E-10 1.29 0.97 0.10 4.61E-10 1.19 1.27 0.08 3.56E-10 1.99 1.49 0.14 3.71E-10 2.75 1.76 0.19 1.28E-09 1.37 1.20 0.11 3.48E-10 2.24 0.75 0.09 2.01E-10 3.85 2.27 0.20 1.13E-09 1.59 1.25 0.11 2.33E-10 1.62 1.30 0.13 2.99E-10 1.98 1.17 0.12 5.85E-10 1.36 1.05 0.11 6.66E-09 Average closure (%) Standard deviation <95% mean Closure Total 3.15 3.73 5.04 3.36 4.84 4.12 3.21 6.19 2.21 2.36 2.55 3.62 4.69 2.68 3.08 6.32 2.95 3.06 3.27 2.51 (%) 82.05 82.19 98.89 92.11 81.24 110.38 93.66 100.34 88.53 81.41 85.91 107.43 86.57 91.80 95.60 98.54 87.03 92.53 88.86 90.34 91.77 8.24

171

Chapter 6 Experimental results and model validation

Table 6.4 Oxygen balance summary


Run no 311 314 318 320 502 504 506 508 510 513 515 517 522 524 525 527 528 531 604 601 Oxygen inputs (kg/hr) Oxygen outputs (kg/hr) from dry from from From from from from Total Moistur tar H 2O wood air CO CO2 3.02 3.45 0.67 7.14 2.48 3.12 3.88E-10 1.09 3.58 3.94 0.78 8.30 2.87 3.78 2.77E-04 1.27 4.02 4.76 0.87 9.65 3.69 5.50 4.87E-04 1.62 2.87 3.10 0.58 6.55 2.34 3.84 2.52E-04 1.07 4.69 5.15 0.94 10.77 3.61 5.07 2.32E-04 1.65 2.94 4.57 0.65 8.16 3.14 4.20 1.84E-04 0.96 2.70 3.15 0.72 6.57 2.34 3.55 3.35E-04 1.05 4.86 5.76 1.29 11.91 4.41 6.99 2.00E-04 0.99 1.97 1.75 0.47 4.19 1.45 2.77 2.50E-04 0.66 2.28 2.50 0.57 5.35 1.72 2.57 3.34E-04 0.81 2.33 2.64 0.50 5.46 1.59 3.39 2.58E-04 0.88 2.65 3.61 0.59 6.85 2.66 3.96 2.69E-04 0.40 4.27 5.29 1.32 10.88 3.67 4.68 9.30E-04 2.20 2.30 2.63 0.70 5.62 1.82 3.19 2.52E-04 0.89 2.53 3.63 0.78 6.95 2.99 1.99 1.46E-04 1.79 5.05 6.44 1.49 12.98 5.13 6.06 8.17E-04 2.09 2.67 2.84 0.81 6.32 2.12 3.34 1.69E-04 0.96 2.60 3.23 0.70 6.53 2.16 3.47 2.17E-04 1.05 2.90 3.57 0.85 7.33 2.64 3.11 4.24E-04 1.13 2.19 2.38 0.51 5.08 1.81 2.79 4.82E-03 0.81 Average closure (%) Standard deviation <95% mean Closure Total 6.68 7.92 10.81 7.25 10.32 8.30 6.94 12.39 4.88 5.10 5.86 7.02 10.55 5.90 6.77 13.28 6.42 6.68 6.88 5.42 (%) 93.62 95.41 112.02 110.65 95.86 101.74 105.73 103.99 116.57 95.37 107.26 102.45 97.00 104.96 97.50 102.33 101.62 102.27 93.85 106.64 102.34 6.36

Table 6.5 Hydrogen balance summary


Run no 311 314 318 320 502 504 506 508 510 513 515 517 522 524 525 527 528 531 604 601 Hydrogen inputs (kg/hr) Hydrogen outputs (kg/hr) from from from from From From Total CH4 Tar H2O wood Moisture H2 0.38 0.08 0.47 0.22 0.04 7.93E-05 0.14 0.46 0.10 0.55 0.34 0.05 5.66E-05 0.16 0.51 0.11 0.62 0.44 0.07 9.95E-05 0.20 0.37 0.07 0.44 0.30 0.05 5.15E-05 0.13 0.60 0.12 0.71 0.42 0.08 4.73E-05 0.21 0.37 0.08 0.46 0.39 0.06 3.76E-05 0.12 0.34 0.09 0.43 0.28 0.04 6.83E-05 0.13 0.62 0.16 0.78 0.51 0.09 4.08E-05 0.12 0.25 0.06 0.31 0.17 0.03 5.11E-05 0.08 0.29 0.07 0.36 0.21 0.03 6.81E-05 0.10 0.30 0.06 0.36 0.20 0.03 5.27E-05 0.11 0.34 0.07 0.41 0.32 0.05 5.49E-05 0.05 0.54 0.16 0.71 0.41 0.06 1.90E-04 0.28 0.29 0.09 0.38 0.24 0.04 5.15E-05 0.11 0.32 0.10 0.42 0.10 0.03 2.98E-05 0.22 0.64 0.19 0.83 0.48 0.07 1.67E-04 0.26 0.34 0.10 0.44 0.23 0.04 3.44E-05 0.12 0.33 0.09 0.42 0.25 0.04 4.42E-05 0.13 0.37 0.11 0.48 0.28 0.04 8.65E-05 0.14 0.28 0.06 0.34 0.17 0.04 9.85E-04 0.10 Average closure (%) Standard deviation <95% mean Total 0.40 0.55 0.71 0.49 0.70 0.57 0.45 0.73 0.29 0.34 0.34 0.41 0.75 0.39 0.35 0.81 0.39 0.43 0.46 0.31 Closure (%) 84.75 99.69 114.69 110.89 98.68 124.65 103.98 93.05 92.20 94.92 94.94 100.60 105.98 101.62 84.42 97.37 88.65 101.65 96.98 90.57 99.0 9.87

172

Chapter 6 Experimental results and model validation

6.2.2 Temperature At the beginning, as the reactor warms up, the heat from the burning char goes to heating up the bottom char bed and the reactor itself (mainly the thick refractory wall), both of which absorb a large proportion of the heat produced at this stage. Under these conditions rates of gasification are insignificant to small. It was observed that only after the temperature at the grate level reached 350-400 C, was the outlet producer gas able to burn by itself without the LPG flare. Figure 6.1 shows a representative set of changes over time of temperatures at four different locations in the gasifier. The Figure shows clearly how the rate of temperature change at the base of the reactor is high initially and then the temperature levels off at its steady state value about 15-20 minutes after ignition starts. The temperature at the fuel ignition level is at a high value except at the very beginning of the run, but fluctuates within the range of 950-1270 C. These fluctuations are attributed to a combination of small reaction zone shifts and to changes in the micro-environment around the thermocouple as the solid bed moves down past it. Temperatures near the top of the gasifier (750 mm above the grate) were observed to stay about 10-20 C higher than the ambient temperature throughout the experimental run and therefore no physical or chemical changes would be expected in the fuel bed at that level.

1400 400 mm above grate (ignition level) 1200 1000 800 150 mm above grate 600 400 200 750 mm above grate 0 0 1000 2000 3000 4000 50 mm above grate

Temperature (C)

Time (sec)

Figure 6.1 Temperature record for experiment 314 (air flow 16.92 kg/hr, with 20 mm blocks)

173

Chapter 6 Experimental results and model validation

Temperature profiles measured in run 510 at three different levels above the ignition port are presented in Figure 6.2. It is evident that in this run the reaction zone moved upwards in the reactor. From the distance between corresponding sections of the profiles at different times the rate of reaction zone movement can be estimated. From the profiles shown in Figure 6.2, the rate of upward movement of the reaction zone was estimated to be 0.063 mm/s.

150 mm above ignition level 250 mm above ignition level 350 mm above ignition level 1200 1000

Temperature (C)

800 600 400 200 0 2000 4000 6000 8000

Time (sec)

Figure 6.2 Temperatures measured at three different levels above the ignition port during run 510 (Airflow 7.5 kg/hr, ignition started at 400 mm above the grate)

Figure 6.3 shows how temperatures vary along the gasifier once steady gas producing conditions have been reached. Air supply rates for the three curves shown were 10.7, 14.8 & 19.6 kg/hr and the temperatures shown are averages over the period 30 to 40 minutes after ignition. The profiles show temperatures increasing rapidly at the beginning of the combustion zone and then gradually decreasing in the gasification zone. This pattern is similar to that predicted by the model (model validation is discussed in detail in the next section (section 6.3). It can be seen that, as would be expected, the value of the peak temperature increases with increases in the airflow rate.

174

Chapter 6 Experimental results and model validation

Figure 6.3 also shows that peak temperatures for the three different airflow rates occur at different points along the reactor even though ignition was started at the same level (400 mm above ignition) in all cases. This illustrates the effect of airflow rate on movement of the reaction zone. As predicted by the numerical model, a higher airflow rate slows down the upward reaction front movement. This is discussed in more detail in Section 6.3.

1400 1200

Temperature (C)

1000 800 600 400 200 0 800 600 400 200 0

10.7 kg/hr 14.8 kg/hr 19.6 kg/hr

Distance from grate (mm)

Figure 6.3 Temperature profiles at three different air flow rates (average temperature values between 30 min and 40 min after ignition)

6.2.3 Gas composition Gas samples were taken from five different levels of the reactor at two different times during each experimental run and analysed by gas chromatography as discussed in Section 5.3.3. The outlet gas compositions were relatively consistent over the range of operating airflow rates used. The results show nitrogen to be the dominant constituent of the outlet gas (46-58 vol %). The concentration of CO2 was in the range of 11-17 vol%. Of the combustible gas components, CO has the highest concentration (16-21

175

Chapter 6 Experimental results and model validation

vol%) followed by H2 (14-19 vol%). A small fraction of CH4 (1.0-1.7 vol%) was also present. Net heating values of the outlet gas ranged from 4.3 to 5.2 MJ/Nm.

6.2.4 Tar content Tar contents of gas samples from five different levels in the gasifier were determined by following the procedures described in subsection 5.3.2. Table 6.1 shows that outlet gas tar contents for all experimental runs (except run 601) were less than 100 mg/Nm, which is about 20 times less than the tar content found in gas from a traditional downdraft wood gasifier. This can be attributed to two factors. Firstly, the initial loading of charcoal rather than wood into the gasifier below the level of the ignition ports ensured no tar formation occurred below the combustion zone. Secondly, there was uniform distribution of air across the reaction zone, since air was introduced to the system at the top of the reactor rather than from the sides of the reactor. The positive effect of the charcoal bed was well illustrated by what happened in run no 601 in which wood fuel only was used and there was no initial char bed. For this run the tar content was almost 1000 mg/Nm, which was about 15 times higher than that for an experimental run at a similar air supply rate for which a charcoal bed was used. Tar content in the outlet producer gas is shown in Figure 6.4 as a function of air supply rate for both wood block sizes. From Figure 6.4, it can be seen that tar content generally decreased as the air supply rate increased, as predicted by the model.

6.2.5 Energy output The total energy output (determined from the outlet gas composition and flow rate, and expressed in MJ/hr) is plotted against dry wood consumption rate in Figure 6.5. As can be seen from the figure, and not unexpectedly, total energy output increased linearly with the increase in fuel consumption. A similar result was also obtained in an experimental study in which a 600 mm internal diameter stratified downdraft gasifier was operated using wood chips and wood pellets (Walawender, Chern & Fan 1987).

176

Chapter 6 Experimental results and model validation

100

Outlet tar content (mg/Nm)

80

60 20 mm blocks 40 30 mm blocks

20

0 5 10 15 20 25 30

Air supply rate (kg/hr)

Figure 6.4 Outlet gas tar content as a function of air supply rate

180 160 Total energy output (MJ/hr) 140 120 100 80 60 40 20 0 4 6 8 10 12 14 20 mm blocks 30 mm blocks

Dry wood consumption rate (kg/hr)

Figure 6.5 Energy output versus dry wood consumption rate

177

Chapter 6 Experimental results and model validation

6.2.6 Cold gas efficiency Cold gas efficiency is plotted against the air supply rate in Figure 6.6. Within the range of air supply rates (7.5 kg/hr- 27.6 kg/hr) tested, average cold gas efficiency was found to be 70 %. As shown in Figure 6.6, similar values of cold gas efficiency were achieved for most of the runs. A slight increase in cold gas efficiency with increases in air supply rate is suggested by the results but, given the extent of scatter in the results, further work would be needed to confirm this apparent trend.

100

Cold gas efficiency (%)

80
20 mm blocks 30 mm blocks

60

40

20

0 5 10 15 20 25 30

Air supply rate (kg/hr)

Figure 6.6 Cold gas efficiency versus air supply rate

6.3

Comparison of model predictions and experimental results

To compare the model predictions with experimental results, the numerical model was run using the same operational input parameters as those used in the experimental runs. The predicted and experimentally measured values of various gasifier performance parameters (transient and steady temperature profiles, gas composition profiles, tar profile, wood consumption rate, air to fuel ratio, gas production rate, and reaction zone movement) were then compared.

178

Chapter 6 Experimental results and model validation

6.3.1 Temperature profiles In Figure 6.7, measured transient temperature profiles are compared with the temperature profiles predicted by the model. The numerical model predicts that within the first ten minutes after initial ignition heat released from the ignited char will cause the gasification zone temperature to increase. After 15-20 minutes of ignition, the rate at which the gasification zone heats up is predicted to slow and the temperature profiles to stabilise. Since gasification zone temperatures over the initial warm-up period are not high enough for gasification rates to be significant, the pyrolysis rate is predicted to be higher than the gasification rate, causing the reaction zone to move upwards. The trend of the experimental temperatures over time is quite similar to that predicted by the model. The model predicts that temperatures in the gasifier will increase sharply in the oxidation zone, peak, and then decrease more gradually along the gasification zone. However, the values of the peak gas temperature predicted by the model are higher than the experimentally measured values and experimentally measured temperatures rise to their peak values more gradually than those predicted by the model. Two reasons can be put forward to explain this discrepancy. Firstly, in the model oxidation of volatiles released from the pyrolysis zone is described by a kinetic expression alone, since mass transfer effects are very difficult to include. The effect of this is that the model predicts complete oxidation of the volatiles within too short a length of the reactor, too rapid attainment of the adiabatic flame temperature, and therefore too sharp an increase in the gas temperature. The second reason relates to the difficulties involved in obtaining accurate measurements of true flame temperatures in the gaseous oxidation zone within a packed bed of particles. Temperature gradients within the gas spaces are large, and the bed of particles in the reaction zone is continually moving, especially at the time of reactor start up. In addition, temperatures measured inside the reactor using the thermocouples inserted through the reactor wall are more likely to reflect solid surface temperatures or intermediate temperatures rather than true gas phase flame temperatures (Barriga & Essenhigh 1980). Resolution of this problem will have to wait until temperature sensors better suited to measuring the temperatures in the gas spaces of a moving packed bed are available.

179

Chapter 6 Experimental results and model validation

1800 1600 1400 Temperature (K) 1200 1000 800 600 400 200 800 600 400 Distance from grate (mm) 200 0

Model Experimental
Ts (2 min) 2 min 5 min 10 min 15 min 20 min Tg (2min) Ts (5 min) Tg (5 min) Ts (10 min) Tg (10 min) Ts (15 min) Tg (15 min) Ts (20 min) Tg (20 min)

Figure 6.7 Comparison of predicted and experimentally measured temperature profiles along the axis of the gasifier (air flow rate: 15.5 kg/hr, wood block size 20x20x20 mm)

Figure 6.8 shows the gas temperature at the reactor outlet (which was measured with thermocouple T1 in Figure 5.1 (Chapter 5)) as a function of air supply rate for both wood block sizes. It can be seen that there were no significant differences in gas outlet temperature for the two different wood block sizes, while gas temperature tends to increase with an increase in air supply rate. This is because a higher air supply rate

180

Chapter 6 Experimental results and model validation

leads to higher gas temperatures in the combustion zone, which also raises the gas outlet temperature. Figure 6.8 also compares the experimental results with model predictions. The figure shows that the outlet temperature predicted by the model increased over the range of air supply rates shown by an amount similar to that by which the experimental results increased. The predicted gas outlet temperatures, however, were generally higher than the experimentally measured values, especially at higher airflow rates. This discrepancy is largely attributable to the fact that the gas outlet temperature predicted by the model is the temperature at the level of the grate while the experimental gas outlet temperature was that measured inside a gas outlet pipe located about 250 mm downstream from the level of the grate (see Figure 5.1). It was not possible to insert a thermocouple at the level of the grate since clearance was required for rotation of the grate. Even though attempts to minimise heat losses downstream of the grate were made by wrapping the pipe in insulation, some loss of heat was inevitable.

700 Outlet gas tem perature (C) 600


20 mm blocks

500 400 300 200 5 10 15 20 25 30

30 mm blocks

Model (20 mm) Model (30 mm)

Air supply rate (kg/hr)

Figure 6.8 Predicted gas outlet temperatures compared with experimental results as a function of air supply rate (experimental points are average values over the 2 minute interval between 29 -31 minutes after ignition; predicted values are those at 30 min after ignition)

181

Chapter 6 Experimental results and model validation

6.3.2 Gas composition profiles Comparisons between the model predictions and experimentally measured values of gas compositions (dry basis) along the gasifier axis are presented in Figure 6.9. During the experimental runs it would have been ideal to carry out online measurement of gas composition at different points along the gasifier. However, this was not possible since there was no online gas analyser available. Therefore gas compositions were determined by taking gas samples and analysing them using gas chromatography as described in Chapter 5. For most of the experimental runs, gas samples were collected over two different intervals from five different levels in the gasifier. Both sets of samples were collected during the period of steady state operation (more than 30 min after ignition). Gas compositions during the gasifier start up period were determined for two runs (Run 504 and Run 531). In Figure 6.9, the data from experimental run no.504 is compared with compositions predicted by the model. It can be seen that the model was able to predict gas composition profiles reasonably well. For CO2, the experimental data confirmed the model prediction that CO2 levels would increase to a peak in the combustion zone and then decrease slightly in the gasification zone. For CO, H2 and CH4, there was a greater difference between model predictions and experimental data. The model predicted a much more rapid rate of increase in the levels of these gases and attainment of steady state values well upstream of the point where this occurred in the experimental runs. This is attributed to the fact that predicted temperatures distributions rose more sharply than experimental ones, as discussed in subsection 6.3.1. Predicted and observed steady-state values were in good agreement, however.

6.3.3 Tar profile As described in Chapter 5, tar contents of gas samples drawn from five different levels in the gasifier were measured. The top sampling point was located either 50 mm or 150 mm above the ignition level depending on where the ignition was started; the amount of tar in the sample from this sampling point was taken to be representative of gas

182

Chapter 6 Experimental results and model validation

phase tar content in the pyrolysis zone. The gases sampled from the other three sampling ports (each 100 mm apart along the gasifier below the top sampling port) enabled the tar profile through the gasification zone to be determined. The last sampling port enabled the tar content of the producer gas at the gasifier outlet to be determined.

25
Gas composition ( vol %)

20 15 10 5 0 700 600 500 400 300 200 100 0


Distance from grate (m m )

CO (model) H2 (model) CO2 (model) CH4 (model) CO (experimental) H2 (experimental) CO2 (experimental) CH4 (experimental)

Fig. 6.9 (a)

25
Gas composition (vol %)

20 15 10 5 0 700 600 500 400 300 200 100 0

CO (model) H2 (model) CO2 (model) CH4 (model) CO (experimental) H2 (experimental) CO2 (experimental) CH4 (experimental)

Distance from grate (m m )

Fig. 6.9 (b)

Figure 6.9 Comparison of model and experimental dry gas composition profiles along the gasifier for two different time intervals (air flow rate 19.6 kg/hr, small wood blocks) (a) 10-15 min after ignition (b) 30-40 min after ignition

183

Chapter 6 Experimental results and model validation

Figure 6.10 shows predicted and experimentally measured tar profiles along the gasifier for four different airflow rates. Both sets of results show the tar content in the pyrolysis zone to be much higher than that in the gasification zone and outlet gases. The highest level of tar (2500-4000mg/Nm) occurs in the pyrolysis zone and this decreases rapidly as the gas passes through the oxidation zone. Tar levels in the gasification zone are orders of magnitude less than that in the pyrolysis zone. Not unexpectedly, the lowest tar level occurs in the outlet gas samples.

experimental

model

experimental

model

4000 Tar content (mg/Nm) Tar content (mg/Nm)

4000

3000

3000

2000

2000

1000

1000

0 600 400 200 Distance from grate (mm) 0

0 600 400 200 Distance from grate (mm) 0

Fig. 6.10 (a) for air flow rate = 13.50 kg/hr

Fig. 6.10 (b) for air flow rate = 15.48 kg/hr

experimental

model

experimental

model

4000 Tar content (mg/Nm) Tar content (mg/Nm)

4000

3000

3000

2000

2000

1000

1000

0 600 400 200 Distance from grate (mm) 0

0 600 400 200 Distance from grate (mm) 0

Fig. 6.10 (c) for air flow rate = 16.92 kg/hr

Fig. 6.10 (d) for air flow rate = 19.60 kg/hr

Figure 6.10 Comparison of predicted and experimentally measured condensable tar profiles along the gasifier for different airflow rates under steady state operating conditions

184

Chapter 6 Experimental results and model validation

6.3.4 Fuel consumption Predicted and experimentally observed wood fuel consumption rates are presented in Figure 6.11 as a function of air supply rate. In line with expectations, it was found that the wood fuel consumption rate was directly dependent on the amount of air supplied to the gasifier. A higher air supply rate increases the rate of combustion because more oxygen is available to react with both char and volatile gases, and the gasification rate also increases since a higher amount of CO2 and H2O is released from the combustion zone. Figure 6.11 shows that model predictions agree very well with the experimental data.

Wood fuel consumption rate (kg/hr)

16 14 12 10 8 6 4 2 0 5 10 15 20 25 30
20 mm blocks 30 mm blocks

Model (20 mm) Model (30 mm)

Air supply rate (kg/hr)

Figure 6.11 Wood fuel consumption versus air supply rate

6.3.5 Air-fuel ratio Figure 6.12 shows the predicted effect of airflow rate on the air-fuel ratio (the ratio of mass of air supplied to the mass of fuel consumed) together with that calculated from experimental results. The figure shows that varying the air supply rate has no significant effect on the air-fuel ratio. This is consistent with the finding, discussed in subsection (6.3.4), that the rate of fuel consumption is a linear function of air supply

185

Chapter 6 Experimental results and model validation

rate. In other words, the air-fuel ratio is a parameter which cannot be controlled by the operator but which is controlled by the gasifier itself and by the properties of the fuel, as reported by other authors (Barriom, Fossum & Hustad 2001; Garca-Bacaicoa et al. 1994). In the case of the larger wood particle size the model predicts an air-fuel ratio higher than that for the smaller wood blocks. This is as expected given that the 30 mm cubes have a lower specific surface area than the 20 mm cubes. The fact that the observed air-fuel ratio for the 30 mm blocks lies below the predicted value and close to that for the 20 mm cubes suggests that the 30 mm blocks are cracking or breaking up into smaller pieces in the gasifier to a greater extent than allowed for in the model. This would be consistent with the observations of Salazar (1987) who observed a greater incidence of cracking as the diameter of the cylinders of Eucalyptus delegatensis he was pyrolysing was increased above 20 mm. For the 20 mm cubes, however, the agreement between predicted and observed air-fuel ratios is very good.

4 Air: fuel ratio (kg/kg d.b)

3
20 mm blocks

30 mm blocks

Model (20 mm) Model (30 mm) 1

0 5 10 15 20 25 30

Air supply rate (kg/hr)

Figure 6.12 Air-fuel ratio (kg/kg d.b) versus air supply rate

6.3.6 Gas production rate Figure 6.13 shows predicted and experimentally measured gas production rates as a function of air supply rate. The figure shows that there is a linear increase in gas

186

Chapter 6 Experimental results and model validation

production rate as the air supply rate increases. The model predicts a higher gas production rate when 20 mm blocks are used than when 30 mm blocks are used but there is no significant difference between the observed gas production rates for 20 mm and 30 mm blocks at the same air flow rate. The likely reasons for this divergence between model predictions and experimental results were discussed in subsection 6.3.5.

40 Gas production rate (Nm/hr)

30
20 mm blocks

20

30 mm blocks

Model (20 mm) 10 Model (30 mm)

0 5 10 15 20 25 30

Air supply rate (kg/hr)

Figure 6.13 Gas production versus air supply rate

6.3.7 Variation of reaction zone movement with air flow rate and wood block size By measuring temperature profiles along the gasifier at different times after ignition it is easy to establish if the reaction zone is changing position within the gasifier. Figure 6.14(a) shows temperature profiles measured at two different times for two different airflow rates but the same wood block sizes. Figure 6.14 (b) shows a similar pair of temperature profiles for two different wood block sizes and the same air flow rate. The profiles in Figure 6.14 (a) show that in the first case, where the airflow rate was 11.3 kg/hr, movement of the reaction front was in an upwards direction. However, in the case where the airflow rate was considerably higher (24.7 kg/hr) the reaction zone moved in the other direction, i.e. downwards.

187

Chapter 6 Experimental results and model validation


1400 1200 1000 Temperature (C) 800 600 400 (11.3 kg/hr, 20 mm block) 200 0 800 600 400 200 0 800 600 400 200 0 (24.7 kg/hr, 20 mm blcok)

1400 1200 1000 Temperature (C) 800 600 400 200 0

1000 sec 3000 sec

Distance from grate (mm)

Distance from grate (mm)

Figure 14.(a)

1400 1200 1000 Temperature (C) 800 600 400 200 0 800 600 400 200 0 (11.3 kg/hr, 20 mm block)

1400 1200 1000 Temperature (C) 800 600 400 200 0 800 600 400 200 0

1000 sec 3000 sec

(11.3 kg/hr, 30 mm block)

Distance from grate (mm)

Distance from grate (mm)

Figure 14.(b)

Figure 6.14 (a) Temperature profiles at two different times (1000 seconds and 3000 seconds respectively after ignition of the charcoal bed) for two different air flow rates (same fuel size) (b) Temperature profiles at the same two times for two different fuel sizes (same air supply rate) ir

The same effect of air supply rate on reaction front movement was observed in model predictions. As discussed in the sensitivity analysis of the model (Chapter 4), at lower air supply rates the predicted direction of reaction zone movement was upwards and, as

188

Chapter 6 Experimental results and model validation

the air supply rate increased, the rate of zone upward movement decreased. Beyond a certain limit (in this study this is about 20 kg/hr), further increases in air supply rate reverse the direction of movement, causing the reaction zone to move downwards towards the grate and the reactor operates in the gasification dominant mode. The effect of wood block size on reaction front movement is shown in Figure 6.14 (b). When using small wood blocks (20mm) at an airflow rate of 11.3 kg/hr, it is evident that the reaction front moves upwards. However, when big wood blocks (30mm) are used at the same air flow rate the reaction front moved downwards. A similar effect of fuel particle size on zone movement rate and direction is predicted by the model, as discussed in the section on sensitivity analysis in Chapter 4. Relatively good agreement was obtained between model predictions and experimental results. This is well illustrated in Figure 6.15 which compares predicted and observed rates of reaction zone movement for a range of airflow rates.

0.08 Upward zone movement rate (m m/sec)

0.06

0.04

0.02

Experimental model

0.00

10

12

14

16

18

20

22

Air flow rate (kg/hr)

Figure 6.15 Variation of reaction zone movement rate with air supply rate (comparison between model and measured values)

189

Chapter 6 Experimental results and model validation

6.4

Conclusions

A small scale stratified downdraft gasifier (206 mm ID) was run using two different sizes of red gum wood blocks and the effect of air supply rate on reactor behaviour was studied for the purpose of validating the model described and discussed in Chapters 3 and 4. The gasifier was run successfully at a fairly wide range of airflow rates (7.5 to 27.6 kg/hr) with corresponding energy outputs ranging from 54 to 168 MJ/hr. An average cold gas efficiency of 70% was observed. Of the combustible gas components, CO was present in the highest concentration (16-21 vol%) followed by H2 (14-19 vol%). A small fraction of CH4 (1.0-1.7 vol%) was also present in the producer gas. Net heating values of the outlet gas were in the range of 4.3-5.2 MJ/Nm. If, prior to ignition, the section of the gasifier below the ignition ports was filled with charcoal rather than with wood, a significant reduction in the tar content of the outlet producer gas was achieved. For all experimental runs where a charcoal bed was used during start up, tar contents of less than 100 mg/Nm were obtained. Such tar contents are about 20 times less than the tar content found in the gas from traditional downdraft wood gasifiers. At the beginning of an experimental run, temperatures rapidly increased at the ignition level of the charcoal bed and heat released from the combustion zone went largely to heating up the bottom char bed. It took about 20 minutes to reach a steady operating condition. The combustion zone temperature was dependent on the air supply rate and under steady operating conditions lay in the range of 1000-1300C. The observed rate of increase in gas temperature in the combustion zone was slower than that predicted by the model while the peak gas temperature predicted by the model was higher than that found experimentally. This discrepancy between predicted and measured temperatures is in part attributable to the difficulties in measuring gas temperatures in a moving packed bed and partly due to the nature of the kinetic expressions used to characterise gas phase combustion reactions. Experimental results showed that the outlet gas temperature increased (within the range of 400C to 600C) with an increase in air supply rate and a similar trend was observed in model predictions.

190

Chapter 6 Experimental results and model validation

The model was able to predict reasonably well the gas composition profiles at different times after start up. The rate of increase in the concentrations of two combustible components (H2 and CO) along the gasifier was observed to occur more gradually during experimental runs than was predicted by the model; however, there was good agreement between exit gas compositions predicted by the model and those obtained experimentally. As expected, the gravimetric tar content of the gas was found to be highest in the pyrolysis zone and reduced dramatically while passing through the oxidation zone. Good agreement between predicted and measured tar contents was obtained in the gasification region. Wood fuel consumption rate and gas production rate were both observed to increase linearly with an increase in air supply rate. Air-fuel ratio was not affected by changes in air supply rate. A higher airflow rate favoured the downward movement of reaction zones as did the use of larger wood block sizes. The value of the specific gasification rate for which stable operating conditions were achieved was predicted by the model to be in the range of 290-296 kg/hr/m (dry basis) and this agreed very well with the range of 290-308 kg/hr/m determined from experimental results.

191

Chapter 6 Experimental results and model validation

6.5

References

Barriom, M, Fossum, M & Hustad, JE 2001, A small-scale stratified downdraft gasifier coupled to a gas engine for combined heat and power production, Progress in Thermochemical Biomass Conversion ed. AV Bridgwater, pp. 217-54. Barriga, A & Essenhigh, RH 1980, A mathematical model of a combustion Pot: Comparison of theory and experiment, Paper 80- WA/HT-32, ASME Meeting, Chicago. Garca-Bacaicoa, P, Bilbao, R, Arauzo, J & Salvador, ML 1994, Scale-up of downdraft moving bed gasifier (25-300 kg/h) design, experimental aspects and results Bioresource Technology, vol. 48, pp. 229-35. Reed, TB & Agua, D 1988, Handbook of biomass downdraft gasifier engine systems, Biomass Energy Foundation Press, Washington. Salazar, CM 1987, The influence of particle size and shape on the mechanisms of decomposition of wood during pyrolysis, PhD Dissertation, The University of Melbourne, Australia. Walawender, WP, Chern, SM & Fan, LT 1987, Influence of operating parameters on the performance of a wood-fed downdraft gasifier, Energy from Biomass and Wastes X, ed. LK Donald, pp. 607-27.

192

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

Chapter 7

Analysis for tar reduction in stratified downdraft gasifier

7.1

Introduction

Previous chapters describe a theoretical and experimental investigation of the stratified downdraft wood gasifier. The model developed was able to predict reactor behaviour under both start-up and steady state operating conditions. Generally good agreement was obtained between predicted and experimentally measured trends and values. The model was also able to predict how the tar content of the gas changed along the gasifier and the stability of the reaction front. In this chapter use is made of the model developed in this project and the experimental results obtained in this study and by other researchers to determine the best operating conditions for a stratified downdraft gasifier, i.e. conditions under which it can operate stably for a long time, producing a low tar content gas.

7.2

Optimum length of reaction zones

One of the important steps in optimizing gasifier operation is to establish how long the reaction zones should be. Reaction zone length has significant effects on gas phase reactions and how far they proceed towards completion, and also on the extent of tar reduction. Generally, a longer reaction zone leads to higher cold gas efficiencies and better tar cracking, but at the same time it leads to increased heat losses to the surroundings and lower temperatures in the gasification zone. Besides, with a longer reactor it takes more time to reach steady conditions since a larger charcoal bed needs to be heated up. In this section, the effect of changes in the length of the reaction zone was investigated by analysing the cold gas efficiency profile and tar profile along the axis of the gasifier.

193

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

Figure 7.1 presents several cold gas efficiency profiles; these efficiencies were calculated using values predicted by the model and also values measured experimentally at different airflow rates. As was discussed in the previous chapter, although predicted and experimentally measured final gas compositions were similar, predicted rates of formation of CO and H2 were much faster than those observed experimentally. This discrepancy is reflected in Figure 7.1 where the cold gas efficiency predicted using the model rises to its final value much more quickly than the efficiencies calculated from the experimental data. Experimental observations showed that gas sampled at a point 550 mm above the grate was of very low quality, having an average CO and H2 content of 6%, which results in a very low cold gas efficiency of about 20%. An increase in efficiency was observed closer to the grate level. At a level of 350 mm above the grate a satisfactory fuel gas was produced, and below the level of 250 mm there was no further significant improvement. At that level (250 mm above the grate), the average measured cold gas efficiency was found to be 97% of the cold gas efficiency at the reactor outlet. Similarly, the cold gas efficiency predicted at that level by the model was 99.3% of the cold gas efficiency at the reactor outlet. Therefore, since there is so little change between the 250 mm level and the grate, 250 mm above the grate can be taken as the practically optimum end point for the reaction zones, at least from the standpoint of cold gas efficiency.

100

Cold gas efficiency (%)

80

60 Air flow rate 40 13.5 kg/hr 16.92 kg/hr 22.09 kg/hr model (15 kg/hr)

20

0 500 400 300 200 100 0

Distance from grate (mm)

Figure 7.1 Cold gas efficiency profiles along the axis of the gasifier

194

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

For the three different airflow rates referred to in Figure 7.1, experimental measurements showed that the temperature rose from ambient levels to reaction temperature (higher than 700C) at a point between 600 mm and 650 mm above the grate. An average level of 625 mm above the grate was taken as the starting point for the reaction zones. From this it can be inferred that a reaction zone length of around 375 mm is sufficient to provide desired cold gas efficiency levels under a variety of operating conditions.

However, it is not enough just to accept the optimum reaction zone depth determined on the basis of cold gas efficiency. This depth also needs to be checked against the tar content profile along the gasifier axis. The average value of the tar content measured at a point 250 mm above the grate was found to be about twice that found in the outlet producer gas. The model predictions also showed that tar reduction continued below the 250mm level for a wide range of air supply rates. Therefore, a reaction zone depth selected solely with reference to cold gas efficiency is not appropriate where tar minimization is desirable. Since experimentally measured gas compositions and gravimetric tar measurements were only available for five different levels, and since predicted and experimentally measured tar profiles were in good agreement, the tar profile predicted by the model was used to determine the optimum reaction zone length.

As discussed in the previous chapter, the model predicts that the stratified downdraft gasifier should operate stably over the limited range of air flow rates between 19.5 kg/hr- 20.0 kg/hr. Therefore, tar profiles in this air flow rate range were used in selecting the optimum reaction zone length. Figure 7.2 shows predicted tar reduction rates along the gasification zone and inert char zone under stable operating conditions at the above airflow rates. The figure shows that tar reduction still continued along the gasification and inert char zone. The amount of tar removed per unit length of reactor diminishes steadily as one moves down the gasifier. A value of 0.025 mg/mm was felt to be an appropriate cut-off value for the tar reduction rate and this value was used in establishing the optimum end point for the reaction zones from a tar minimization point of view. The starting point of the reaction zone was defined as the level where

195

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

the solid temperature reached 1000 K and char oxidation could be expected to start in earnest.

Table 7.1 presents estimates of optimum reaction zone length for the two airflow rates bounding the region of stable operation. It was found that 425 mm is a suitable length for reaction zones where tar minimization is concerned. This length is greater than that determined from cold gas efficiency considerations.

0.6

0.5

Tar reduction rate (mg/mm)

0.4

Air flow rate 19.5 kg/hr 20.0 kg/hr

0.3

0.2

0.1

0.0 400

300

200

100

Distance from grate (mm)

Figure 7.2 Predicted tar reduction rates along the gasification zone and inert char zone

Table 7.1 Estimation of optimum reaction zone length based on tar reduction rate Air flow rate (kg/hr) 19.5 20.0 Starting point of reaction zone (mm above the grate) 593 589 Starting point where tar reduction rate lower than 0.025 mg/mm 176 157

Optimum reaction zones length (mm) 417 432 425

Average optimum reaction zones length (mm)

196

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

7.3

Investigation of low tar gas production for the stratified downdraft gasifier

As discussed in chapters 2 & 3, the stratified downdraft type of gasifier was selected for study in this project since it has a number of advantages over the traditional throated gasifier. Of particular importance is that the scale up problems associated with the throated gasifier can be largely eliminated. In Table 7.2, the tar content level in gas produced in this study when using wood blocks only (no initial charcoal bed) is compared with results from other experimental throatless downdraft gasifier studies. Table 7.2 shows that tar contents found in all these experimental studies are beyond the acceptable limit for engine applications. For the producer gas to be acceptable as a fuel for an internal combustion engine, the tar content needs to be lower than 100 mg/Nm and preferably less than 50 mg/Nm (Knoef 2000 & Stassen 1995). This shows that when working with stratified downdraft wood gasifiers operating in the conventional way one cannot be assured of producing gas with a tar content low enough for power applications. Thus, it is necessary to investigate the factors which influence final tar content in order to establish how tar contents can be reduced.

Table 7.2 Measured tar content levels from different throatless downdraft gasifier studies Fuel type & size (mm) Reactor Diameter (mm) Wood blocks (10-15 mm) 200 3,600 Bui, Loof & Bhattacharya (1994) Air was supplied from sides Wood chips (6.4-12.7 mm) Hard wood chips (14 mm) Red gum hard wood blocks (20 mm) 206 964 600 776-2,445 75 522 Milligan, Evans & Bridgwater (1997) Walawender, Chern & Fan (1985) This study _ Tar content (mg/Nm) Source Remarks

197

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

From Table 7.2, it can be seen that the tar content reported by Bui, Loof & Bhattacharya (1994) is higher than those reported by others. This is probably due to the nature of their air supply system. In their experimental set-up, air was introduced to the reactor through nozzles which were located circumferentially around the reactor. Generally, the oxygen in air supplied from the sides of the reactor is not uniformly distributed through the fuel since combustion tends to take place predominantly in the area close to the air nozzles and temperatures fall away as one moves towards the centre of the fuel bed. The effect of this radially non-uniform distribution of temperatures in the gasifier is that some wood particles pass through the combustion zone without completing devolatilization and therefore evolve tarry materials downstream of the combustion zone. In addition, some of the pyrolysis products passing down the centre of the gasifier will not be fully cracked. This problem will be aggravated in larger diameter gasifiers where temperatures at the centre of gasifier will be much lower than those near the walls. This helps confirm the superiority of a top air supply system over a side air supply system for the downdraft gasifier.

7.3.1 Performance of the reactor when using wood fuel only

As has already been discussed the stratified downdraft gasifier operating just with wood fuel cannot produce gas with a tar content low enough for engine applications. When the experimental work using wood blocks only was carried out, excessive smoke release from the stack was observed within the first 15 minutes after ignition since the volatiles released were not being completely oxidised in the burner. At this stage, the reactor was operating like an inverted updraft gasifier. About 15 minutes after

ignition, no more visible smoke was emitted and after a further 15 minutes a steady condition was reached where the producer gas was able to burn by itself without the LPG flare.

The effect of loading the lower part of the gasifier with charcoal rather than wood is well illustrated by Figure 7.3. This shows how the solid temperature at the level of the grate predicted by the model varies with time for the cases where an initial charcoal bed is and is not used. The figure shows that when no charcoal is used the gasifier can

198

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

be expected to take 15 to 20 minutes longer to reach a steady operating temperature than when an initial charge of charcoal is used. In the latter case, no significant reactions occur during the start up phase and most of the heat released in the combustion zone goes to raise the sensible heat content of the charcoal bed. However, in the case of the wood only operation, there is a much slower rate of increase in solid temperatures in that section of the bed below the ignition port. This is because much of the heat transferred to the solid particles is used to provide the energy needs of the endothermic wood decomposition reactions. In other words, for wood only operations, pyrolysis is the dominant process during the reactor start up period, and a longer time is needed for the solid temperature to reach its final steady value.

900

Solid temperature at grate level (K)

800 700 600 500 wood only 400 300 0 10 20 30 40 50 60 with charcoal bed

Time after ignition (min)

Figure 7.3 Dynamic behaviour of solid temperature at reactor outlet

As discussed above, for wood only operations devolatilization is the dominant process during the early part of the reactor warming up stage. From a practical point of view, it is helpful to know how long it takes before carbonization of that part of the fuel bed below the ignition port is complete. (This is because tar levels in the gas are likely to remain high until carbonisation is complete.) Figure 7.4 presents predicted volatile matter density profiles along the axis of gasifier at various times during the first 2 hours of gasifier operation. It can be inferred from Figure 7.4 that the rate of

199

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

devolatilization in the lower part of the gasifier is reasonably fast in the first hour of operation. After one hour of operation, devolatilization is largely complete except in the bottommost 100 mm of the bed. Even after two hours of operation a small amount of volatile matter is still being evolved within the zone 100 mm above the grate showing that completion of devolatilization has still not been achieved.

800

Density of volatile matter (kg/m)

600

20 min 40 m in 60 min

400

80 min 100 min 120 min

200

0 700 600 500 400 300 200 100 0

Distance from grate (mm)

Figure 7.4 Volatile matter profile along the gasifier axis

This failure to achieve complete devolatilization significantly affects the final tar content of the gas. Figure 7.5 shows tar levels along the axis of the experimental gasifier measured at steady operating condition during run no 601, in which no initial charcoal charge was used. Also shown are the tar profiles predicted by the model for the period 125 to 150 minutes after ignition. The figure shows that the final tar content is about 20 times higher than the average values of the tar content for the case when an initial charcoal charge was used. It is evident from Figure 7.5 that the high tar levels are attributable to the processes occurring in the lowest part of the bed. Within the combustion and gasification zones tar levels soon drop off to a level similar to those observed when charcoal is used. However, for wood only operations the low temperatures in the lowest part of the bed mean that volatile matter continues to evolve

200

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

over a long period of time and since it cannot be cracked efficiently at these low temperatures it is carried out in the final gas stream.

4000 125 min 130 min 135 min 140 min 145 min 150 min experimental

Tar content (mg/Nm 3)

3000

2000

1000

0 700

600

500

400

300

200

100

Distance from grate (mm)

Figure 7.5 Tar profiles along the axis of the gasifier for wood only operation

These observations show that to achieve effective tar reduction within the gasifier tar formation in the region downstream of the oxidation zone needs to be avoided as far as possible. They also show that running the gasifier for a number of hours will not necessarily achieve low enough tar levels in the gas. However, using an initial charge of charcoal in the gasifier has been shown to overcome these problems and yield gas with a tar level that is acceptable for use in engines.

7.3.2 Alternative operating procedures for the stratified downdraft gasifier

Despite its advantages, using a gasifier operating procedure that employs an initial charge of charcoal in the lower part of the fuel bed has the drawback that suitably sized charcoal pieces have to be available when needed. There are two ways of getting round this problem. The first is to charge the lower part of the gasifier with charcoal to the height of the ignition port during the first experimental run only, run the gasifier until all the wood has been consumed or converted to charcoal, and then use this residual charcoal in the next run. This was not the procedure followed during the experimental

201

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

runs undertaken in this project. In this study, the gasifier was unloaded after each run so that the tar sampling pipe could be cleaned by blowing pressurized air through it. A new charge of charcoal was then loaded before the next run. However, in real world applications, the charcoal remaining from a previous run should be able to be used directly in subsequent runs without the gasifier having to be emptied and a new charcoal charge loaded.

A second approach is to use the gasifier itself to produce the charcoal needed. As discussed in previous chapters, by supplying air at a low flow rate the gasifier can be made to run in a pyrolysis dominant mode and the reaction front will move upwards. Using this approach, the gasifier can be initially run as a charcoal producing unit rather than as a gas producer and then switched to gas production. The proposed operating procedure is as follows:

1) Fill the gasifier with wood blocks and ignite the bottommost 50-100 mm of bed. 2) After ignition is well under way, air is introduced from the top of the reactor at a very low airflow rate (5-7 kg/hr) [0.042-0.058 kg/msec]. 3) The reaction front will move upwards, and when it reaches a point 400-450 mm above the grate (the desirable reaction zone height - see section 7.2) the air flow rate is increased to 19-20 kg/hr [0.158-0.167 kg/msec] to achieve stable operation. Figure 7.6 shows a schematic diagram of the proposed operating procedure.

The proposed procedure was simulated using the model developed in this project and the results are shown in Figure 7.7 and Figure 7.8. Figure 7.7 shows how the temperature profile along the gasifier axis varies with time under the proposed procedure. The model was run assuming the reactor was initially operated at a low airflow (6 kg/hr) under which condition reaction zones were predicted to move upwards at an average velocity of 0.058 mm/sec. After a 90 minute period the reaction front was predicted to reach the required zone height and the airflow rate was increased to a value which is suitable for stable operation. Figure 7.7 shows that the temperature profiles at 105 minutes and 120 minutes after start up are practically coincident indicating that increasing the air flow rate had been successful in achieving stable operation.

202

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

Air 0.042-0.058 kg/msec Wood

Air 0.042-0.058 kg/msec Wood

Air 0.158-0.167 kg/msec Wood

Wood blocks Wood blocks Combustion front Char Ignited fuel bed bed 50 100 mm 400 450 mm

Wood blocks Combustion front Char bed 400 450 mm

Initial conditions

Reaction front moves upward

Reaches required zone height

Increase in air supply rate

Stable operation

Figure 7.6 Experimental procedure that obviates the need to use an initial charge of charcoal

2000 1800 1600

120 min

t = 15 min

15 min

Temperature (K)

1400 1200 1000 800 600 400 200 500 Ts Tg

400

300

200

100

Distance from grate (mm)

Figure 7.7 Changes over time of temperature profiles in the gasifier for the procedure of producing charcoal inside the gasifier

203

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

3500 3000

70

Tar content (mg/Nm 3)

2500 2000 1500 1000 500 0 20 40 60 80 100 120 66 Tar Efficiency 68

67

65

Time (min)

Figure 7.8 Variations over time of outlet gas tar content and cold gas efficiency for the procedure of producing charcoal inside the gasifier

Figure 7.8 shows the variation over time of outlet gas tar content and cold gas efficiency when the gasifier was run so as to produce the required charcoal bed inside the gasifier. It can be seen that high amounts of tar are predicted to be present in the outlet gas during the period when airflow rates are low. This occurs because the solid and gas phase peak temperatures are not high enough for effective tar cracking and also because the initial bed height is less than that required for good tar reduction to be achieved. After the airflow rate is increased to normal operating levels, predicted solid and gas phase temperatures increase rapidly and the tar content is dramatically reduced, to a value less than 10 mg/Nm. An increase in cold gas efficiency also occurs at this time, as the gas composition alters. The model predictions appear to confirm that under the proposed procedure the stratified downdraft gasifier can be brought to a point where it is running stably and producing a low tar content producer gas. However, the time required to reach a steady operating condition is rather long and this, coupled with the problem of how to deal with the tar rich gas produced over the initial period, makes the practical usefulness of this procedure rather doubtful.

204

Cold gsa efficiency (%)

69

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

7.3.3 Two different concepts on two-stage gasification

As discussed above, producer gas with a low tar content and which is suitable for engine applications can be produced in a stratified downdraft gasifier as long as the gasifier is initially filled with charcoal to the height of the ignition port, and then operated at an appropriate air flow rate. Attempts have also been made to produce a low tar content gas by modifying gasifier designs. Two research groups have investigated the use of what they have termed two-stage gasification. The first group (TUD, Denmark) studied a system involving two separate reactors while the other (AIT, Thailand) conducted their work with a single reactor. The underlying principles of the systems developed by both groups are illustrated in Figure 7.9.

Biomass

Partial oxidation of pyrolysis gas Pyrolytic gases + charcoal

Biomass

Pyrolysis unit

Air, steam

Flaming pyrolysis zone (First stage) Products gases, tar Reduction zone (Second stage)

Two-stage gasifier

Primary air

(a) Concept of two-stage gasifier (TUD, Denmark)

Figure 7.9 Two-stage gasification systems proposed by TUD (Denmark) and AIT (Thailand)

Gasification unit

Charcoal bed

Secondary air

Low tar content gas

Low tar content gas (b) Concept of two-stage gasifier (AIT, Thailand)

205

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

The concept put forward by the first group involves a separate initial pyrolysis unit coupled with a downdraft char gasifier (Bentzen et al. 2001; Brandt, Larsen & Henriksen 2000; Henriksen & Christensen 1994). The pyrolysis unit must be air tight and needs to be heated externally, and an electrical heating system was used for this purpose. Henriksen and Christensen (1994) suggested that the sensible heat content of the exhaust gas from the engine could be made use of to replace the electrical heating system in a future commercial plant. To transport the charcoal produced in the pyrolysis unit to the gasifier unit, a conveyor such as a piston feeder or a screw conveyor is needed. As well as this, an external steam producing system is also required to provide the steam to the gasifier unit. Since pyrolysis is carried out in a separate unit and the tars partially oxidized before entering the gasifier unit, the tar content of the gas from this system was below 25 mg/Nm, which is acceptably low. (Bentzen et al. 2001). However, the complexity of this system makes it far less attractive than the stratified downdraft gasifier operated using an initial charge of charcoal.

The two-stage gasifier proposed by the second group, at AIT, is much simpler, both to make and to operate. The main modification introduced by the second group is the use of a secondary air intake as well as the usual primary one. The addition of secondary air to the reduction zone not only introduces additional oxygen but also causes an increase in temperature in this zone. This has the effect of reducing tar levels considerably; Bui, Loof and Bhattacharya (1994) reported that the tar content of the gas from their gasifier was as low as 50 mg/Nm, about 40 times less than that in the gas from a single-stage reactor under similar operating conditions.

Due to the relative simplicity of the reactor used by the second group, it was possible to use the model developed in this study to investigate the performance of their system theoretically. The behaviour of this two-stage gasification system (AIT) was simulated using the model and the results are discussed in the next subsection.

206

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

7.3.4 Simulation on two-stage gasification

To simulate the two-stage gasifier concept, the model was modified such that air was supplied simultaneously at two different levels in the gasifier, as indicated in Figure 7.9 (b). Except in the case of the air supply rate, the values of the operating parameters used in the two-stage gasifier simulation were the same as the baseline parameter values for the single stage gasifier (Table 4.1). Fuel was assumed to be ignited through an ignition port located 500 mm above the grate and primary air was supplied through the top of the reactor. In the simulation it was assumed that secondary air could be uniformly distributed across the gasifier at the level where it was introduced. Since an air supply rate of 19-20 kg/hr enabled stable operating conditions to be achieved in a single stage reactor, the combined primary and secondary airflow rate was set at 20 kg/hr and different combinations of primary and secondary air flows tried. In addition, the effect of changing the level at which secondary air was introduced was also explored.

(a) Temperature profiles

Figure 7.10 shows how the solid and gas temperature profiles are predicted to vary with time along the two-stage gasifier. The temperature profiles at the ignition port level are quite similar to those of the single-stage reactor which was discussed in Chapter 4. However, a marked difference from the single- stage profile occurs at the level where secondary air is introduced, with both solid and gas phase temperatures increasing to a second, rather smaller peak. As is the case with the first peak, the gas phase temperature at the point where the second peak occurs is higher than the corresponding solid temperature. Since the introduction of secondary air leads to higher temperatures in the region close to the grate, the rate of increase in fuel bed temperatures at the level of the grate is faster than that observed in a single-stage unit. As a result, overall gasification rates in the two-stage unit increase faster during the reactor warming up period, and the model predicts that gas produced in a two-stage gasifier reaches the desired heating value within a shorter time than gas from a singlestage unit.

207

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

1800 1600 1400 Ts (15 sec) Tg (15 sec) Ts (45 sec) Tg (45 sec) Ts (75 sec) Tg (75 sec) Ts (120 sec) Tg (120 sec) Ts (210 sec) Tg (210 sec) Ts (360 sec) Tg (360 sec) Ts (690 sec) Tg (690 sec) Ts (1000 sec) Tg (1000 sec) 700 600 500 400 300 200 100 0

Temperature (K)

1200 1000 800 600 400 200

Distance from grate (mm)

Figure 7.10 Variation with time of temperature profiles along the axis of the gasifier for the first 1000 sec after ignition (Ignition port: 500 mm above grate; secondary air supply: 300 mm above grate; primary air flow rate = 15 kg/hr; secondary air flow rate = 5 kg/hr)

(b) Reaction rates profile

Figure 7.11 (a) shows how the oxidation rates of various gaseous components are predicted to vary along the first stage combustion zone. Patterns of behaviour similar to those in a single stage system are observed. Char oxidation reaches its peak 15 to 20 mm further up the gasifier than the gas phase oxidation reactions. In addition rates of char oxidation are significant over a greater distance than the gas phase reactions. CO oxidation rates show a higher peak than the other gaseous oxidation rates.

As shown in Figure 7.11 (b), patterns in the second stage are very different. Here both solid and gas phase oxidation rates are predicted to take off and peak at the same level. It should be noted that the scale on the X axis of Figure 7.11 (b) is a very expanded one, covering a distance of just one millimetre. The implication of this is that very high rates of reaction occur at the level where secondary air is introduced, with the injected oxygen reacting almost instantaneously.

208

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

1.6 1.4 C6H6 oxidation CO oxidation x 0.1 H2 oxidation Char oxidation

Oxidation rate (kg/m 3/s)

1.2 1.0 0.8 0.6 0.4 0.2 0.0 580 570

560

550

540

Distance from grate (mm) Fig. 7.11 (a) Solid and gas phase oxidation reaction rates in the first stage

1.6 1.4 1.2 C6H6 oxidation x 0.02 CO oxidation x 0.04 1.0 0.8 0.6 0.4 0.2 0.0 H2 oxidation x 0.2 C2H4 oxidation x 0.02 Char oxidation x 20

Oxidation rate (kg/m/s)

300.0

299.8

299.6

299.4

299.2

299.0

Distance from grate (mm) Fig. 7.11 (b) Solid and gas phase oxidation reaction rates in the second stage

Figure 7.11 Gas and solid phase oxidation rate profiles for the first and second stage, at 1000 sec after ignition (Ignition port: 500 mm above grate; secondary air supply: 300 mm above grate; primary air = 15 kg/hr; secondary air = 5 kg/hr)

209

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

Figure 7.12 shows how the rates of heterogeneous gasification reactions are predicted to vary along the axis of the reactor. As well as the expected peaks below the main oxidation zone there is a secondary set of peaks just below the point where the secondary air is introduced. However, gasification rates in this second stage are lower than those in the first stage. This is attributable to the lower concentrations of CO2 and water present and to the fact that the solid phase temperature in the second stage is lower than that in the first stage.

0.12 0.10 0.08 0.06 0.04 0.02 0.00 H2O gasification CO2 gasification x 10 H2 gasification x 500

Char gasification rate (kg/m/s)

600

500

400

300

200

100

Distance from grate (mm)

Figure 7.12 Heterogenous gasification reaction rates for two-stage gasification (Ignition port: 500 mm above grate; secondary air supply: 300 mm above grate; primary air flow rate = 15 kg/hr, secondary air flow rate = 5 kg/hr)

(c) Gas composition profile

The gas composition profiles for two-stage gasification are shown in Figure 7.13. It can be seen from Figure 7.13(a) that better gas quality was achieved by introducing secondary air, with concentrations of H2 and CO in particular increasing substantially immediately downstream of the second air supply point. Figure 7.13 (a) also shows that the secondary O2 supplied is consumed very rapidly due to the high gas temperatures and the consequent high gas phase oxidation rates. The further slow

210

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

increase in CO and H2 levels downstream of the secondary air supply point can be attributed to heterogenous gasification reactions involving CO2 and H2O formed at the point of secondary air injection.
25

Gas composition (vol%)

20 O2 15 CO CO2 H2 H2O 5

10

0 700 600 500 400 300 200 100 0

Distance from grate (mm) Fig. 7.13 (a)

6 5

Gas composition (vol%)

4 3 2 1 0 700 600 500 400 300 200 100 0

CH4 C2H4 primary tar C6H6

Distance from grate (mm) Fig. 7.13 (b)

Figure 7.13 Gas composition profiles along the axis of the gasifier for two-stage gasification (a) major gas species profile (b) minor gas species profile (Ignition port: 500 mm above grate; secondary air supply point: 300 mm above grate; primary air flow rate = 15 kg/hr, secondary air flow rate = 5 kg/hr)

211

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

Figure 7.13 (b) shows the predicted variation in concentrations of minor gas species. A rapid decrease in concentration of C6H6 and C2H4 is evident at the point where the secondary air is introduced. Methane levels also fall but to a much smaller extent.

(d) Condensable tar profile

In Figure 7.14, the primary tar profile for two-stage gasification is shown compared to that for a single stage gasifier without secondary air supply. The total amount of air supplied was kept the same for the single-stage (20 kg/hr of primary air) and two-stage systems (15 kg/hr of primary air and 5 kg/hr of secondary air). Since the primary air supply rate in the two-stage gasification unit is less than the (primary) air supply rate for the single-stage system, it is not unexpected that above the secondary air supply level the tar content in the two-stage unit is predicted to be higher than that in the single-stage unit. (As discussed in Chapter 4, higher air supply rates lead to higher solid and gas phase temperatures and consequently tar-cracking rates increase.) However, once the secondary air input level is reached in the two stage unit, the tar cracking rate increases dramatically due to the increase in temperature at this point. As a result, as shown in Figure 7.14, final gas phase tar levels in the two-stage unit end up well below that for gas from the single-stage system.

Table 7.3 presents the results obtained from a simulation of the two-stage gasifier for different primary air and secondary air combinations but the same total air supply rate of 20 kg/hr. It can be seen from Table 7.3 that as the amount of secondary air is increased, smaller amounts of tar and higher cold gas efficiencies are predicted. It was also observed that the closer to the oxidation zone the secondary air is introduced, the lower is the predicted final tar content and the higher the cold gas efficiency. There are two reasons for this: the first is that when secondary air is supplied at a point closer to the primary reaction zones, the gas temperature is higher and more of the oxygen utilized is consumed in gas phase oxidation reactions. The second reason is that at higher temperatures, proportionally less H2 is consumed in gas phase oxidation reactions and a higher cold gas efficiency results.

212

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

4000

Tar content (mg/Nm)

3000 Single stage 2000 Two-stage

1000

0 700 600 500 400 300 200 100 0

Distance from grate (mm)

Figure 7.14 Tar profile along the gasifier for both single-stage and two-stage units, at 1000 sec after ignition (Ignition port: 500 mm above grate for both cases; single-stage air flow rate: 20 kg/hr; two-stage total air supply rate: 20 kg/hr (primary air flow rate = 15 kg/hr; secondary air flow rate = 5 kg/hr); secondary air supply: at 300 mm above grate)

7.4

Conclusions

The dynamic model developed in this study was used to investigate and establish under what operating conditions a stratified downdraft gasifier can operate stably and produce a low tar gas which is acceptable for engine applications.

First an attempt was made to establish what was the best reaction zone length to use. Since reaction zone length affects both final tar content and cold gas efficiency, the effects on both these two parameters of variations in reaction zone length were determined. A reaction zone length of 375 mm was found to be adequate from a cold gas efficiency stand point since measured and predicted cold gas efficiency values at this point were 97% and 99.3% respectively of values considerably further downstream. However, from a tar reduction point of view, a longer reaction zone was necessary since tar reduction still continued beyond a 375 mm zone length. The

213

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

required zone length was determined by assuming that gains in gas quality would be insignificant once tar reduction rates fell to less than 0.025 mg/mm. This gave a zone length of 425 mm.

Table 7.3 Results from simulation of two-stage gasification concept (at 1000 sec after ignition)
Primary air Supplied rate (kg/hr) Secondary air Location [distance above grate (mm)] H2 CO CH4 CO2 H2O Gas composition (vol %)

Final tar content (mg/Nm)

Cold gas Efficiency (%)

Supplied rate (kg/hr)

19 18 17 16 15 19 18 17 16 15 19 18 17 16 15

1 2 3 4 5 1 2 3 4 5 1 2 3 4 5 200 300 400

12.89 13.65 14.66 15.98 17.86 12.23 12.54 13.23 14.21 15.67 11.82 11.80 12.87 13.10 14.39

15.03 16.07 17.21 18.25 19.28 14.69 15.51 16.47 17.40 18.34 14.52 15.14 15.94 16.90 17.74

2.30 13.49 11.25 2.37 13.65 11.17 2.45 13.82 10.81 2.50 14.26 10.20 2.41 15.13 9.26 2.28 13.44 11.87 2.35 13.57 12.16 2.42 13.72 12.16 2.50 14.01 11.94 2.55 14.53 11.32 2.24 12.16 11.82 2.27 13.55 11.80 2.33 13.73 12.28 2.39 13.91 12.81 2.46 14.39 12.34

76.26 49.77 31.32 13.29 2.19 91.33 68.16 49.39 32.04 15.39 98.63 77.27 58.66 43.66 25.30

65.75 68.75 72.09 75.36 78.15 64.34 66.50 69.41 72.52 75.74 63.44 64.82 67.23 70.45 73.77

Operating the reactor with wood fuel only was observed to have considerable disadvantages. A large amount of smoke formed within the first 15 minutes of ignition and it took 15 to 20 minutes longer for the gasifier to reach a state where producer gas of an acceptable heating value was being produced. The model predicted that complete devolatilization of the fuel bed under the ignition level would not be achieved even after two hours of operation since a small fraction of volatile matter was still being evolved in the region up to 100 mm above the grate. This was predicted to have a

214

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

significant effect on final tar content and this was confirmed by experimental results that showed that the final tar content of the producer gas was about 20 times higher than the average values of tar content measured when an initial charge of charcoal was used. It was concluded that avoiding tar formation in the region downstream of the oxidation zone is very important for producing a tar free gas.

Avoiding devolatilization below the oxidation zone can be simply achieved by charging the gasifier with charcoal prior to each run. This, however, requires that sufficient charcoal of a suitable quality be available on site. To avoid having to load charcoal before every run, the reactor can be operated using the charcoal left over from the previous run. A charcoal bed can also, at least in theory, be produced inside the gasifier by appropriately controlling the air flow rate: in this approach the wood bed is ignited at the level of the grate and an initially low air flow rate used; under these conditions the reaction front moves upward. Once the front reaches the required level, the air flow rate is increased to a level where stable operation occurs. By following this procedure, a low tar gas can be produced under stable operating conditions. However, more than an hour is required to generate a charcoal bed of the required depth and to achieve stable operation. There is also the problem of what to do with the tar rich gas produced during the initial period.

The concept of two-stage gasification was investigated using the dynamic model developed in this study. In two-stage gasification a secondary oxidation zone is created by introducing secondary air to the reactor at a level below the primary oxidation zone. This leads to local increases in both solid and gas phase temperatures, and substantial tar reduction is achieved at the second stage. Higher cold gas efficiencies are predicted to be achieved by the supplying the secondary air at levels close to the ignition port level.

In the simulation of two-stage gasification it was assumed that distribution of secondary air was uniform across the gasifier cross-section. This is perhaps a dubious assumption since it has been the difficulties associated with supplying air to gasifiers through ports on their sides that has necessitated the use of throated gasifiers in the past. Since it is secondary air, not primary air, that is being introduced, achieving

215

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

uniform distribution of air is not as important as it is in single-stage gasifier and even if relatively little of the injected air reaches the centre of the bed, improvements in tar levels in the gas should still be achievable in smaller diameter units.

216

Chapter 7 Analysis for tar reduction in stratified downdraft gasifier

7.5

References

Bentzen, JD, Henriksen, U, Hindsgaul, C & Brandt, P 2001, Optimised two-stage gasifier, Proceedings of the 1st world conference on biomass for energy and industry, eds. Kyritsis, S, Beenackers, AACM, Helm, P, Grassi, A & Chiaramonti, D, held in Sevilla, Spain, 5-9 June 2000, pp. 1-4. Brandt, P, Larsen, E & Henriksen, U 2000, High tar reduction in a two-stage gasifier Energy & Fuels, vol. 14, pp. 816-9. Bui, T, Loof, R & Bhattacharya, SC 1994, Multi-stage reactor for thermal gasification of wood, Energy, vol. 19, pp. 397-404. Henriksen, U & Christensen, O 1994, Gasification of straw in a two-stage 50 kW gasifier, Proceedings of the 8th European conference on biomass for energy, environment, agriculture and industry, eds. Chartier, PH, Beenackers, AACM & Grassi, G, Pergamon, Elsevier Science Ltd, Oxford, U.K, pp. 1568-78. Knoef, HAM 2000, The UNDP/WB monitoring program on small scale biomass gasifier (BTGs experience on tar measurements) Biomass & Bioenergy, vol. 18, pp. 39-54. Milligan, JB, Evans, GD & Bridgwater, AV 1997, Results from a transparent open-core downdraft gasifier, Developments in Thermochemical Biomass Conversion, vol.1, eds. Bridgwater AV& Boocock DGB, pp. 175-85. Stassen, HE 1995, UNDP/WB small-scale biomass gasifiers for heat and power, A global review, Energy sector management assistance programme, WBTP 296-Washington, DC. Walawender, WP, Chern, SM & Fan, LT 1985, Wood chip gasification in a commercial downdraft gasifier, Fundamentals of Thermochemical Biomass Conversion, eds. RP Overend, TA Milne & LK Mudge, pp. 911-21.

217

Chapter 8 Conclusions and recommendations

Chapter 8

Conclusions and recommendations

8.1

Introduction

In this project theoretical and experimental studies have been carried out on a stratified downdraft gasifier using red gum (Eucalyptus camaldulensis) wood blocks as a fuel. Several significant contributions have been made towards increasing our understanding of wood gasification and towards solving some of the associated problems. In this chapter the major conclusions of the theoretical and experimental investigations are summarized and recommendations for future research are presented.

8.2

Summary and conclusions for the theoretical study

A dynamic model for the stratified downdraft wood gasifier, which incorporates the phenomena of wood drying, pyrolysis, cracking of tars and oxidation of pyrolysis products, has been developed. The model describes the complex physical and chemical processes taking place in the reactor by the use of mass and energy balances, together with information about rates of chemical reactions and physical transport processes. Balance equations were written for the three solid phase components (biomass, char, and bound and absorbed moisture) and the ten gaseous species [O2, N2, CO2, CO, H2O, H2, CH4, C2H4, primary tar (C6H10.71O3.264) and secondary tar (C6H6)]. There were thirteen reactions considered in the model including those for drying and pyrolysis processes.

The main features taken into account by this model were temperature differences between the gas and solid surface, two-step pyrolysis processes, oxidation of volatile gases and secondary tars, particle shrinkage during heterogeneous reactions, the effect of reaction temperature on the ratio of (CO/CO2) during heterogeneous char

218

Chapter 8 Conclusions and recommendations

combustion, and the effective thermal conductivity, which enables the importance of radiation at high reactor temperatures to be properly accounted for. The heterogeneous reactions of char combustion and gasification were described by a chemical reaction rate expression that included the resistances due to external mass transfer and to intraparticle diffusion of the gaseous reagents. A two-step pyrolysis mechanism used in the model incorporated a primary pyrolysis step in which biomass decomposed to primary tar, volatile gases and char, and a secondary pyrolysis step in which primary tar cracked into secondary tar and combustible gases. An important initial condition assumed in most simulation runs using this model was that prior to a run commencing, the gasifier was charged with charcoal up to the ignition port level. This was done to avoid the excessive tar formation and smoky gas release that would otherwise occur during the reactor warming up stage.

In this study, a parametric sensitivity analysis was carried out to understand the effects of changes in operating and model parameters on model performance. In the sensitivity analysis, the effects of varying three operating parameters and eight model parameters were investigated.

The major conclusions of this theoretical investigation can be summarized as follows:

1) When the reactor is started by igniting the fuel at the top of the charcoal bed, the combustion zone temperature increases rapidly but the charcoal bed temperature at the level of the grate remains at ambient or near ambient temperatures for quite a while afterwards. In this way the simulation showed that a substantial warming up time is required before temperatures throughout the inert charcoal bed reach the levels at which gasification processes can take place at the required rate.

2) During the period immediately after ignition, the quality of the producer gas is very poor. This happens because temperatures are not high enough for the two main combustible gases, CO and H2, to be formed in significant quantities; the heterogeneous gasification reactions which produce these gases proceed only at a very slow rate until a gasification zone temperature higher than 800C is reached. The amount of H2 and CO in the producer gas increases gradually with time, but it

219

Chapter 8 Conclusions and recommendations

was found to take 20 to 25 minutes to achieve steady operating conditions. Under such conditions the heating value of the producer gas was about 5 MJ/Nm.

3) At the front end of the combustion zone solid surface temperatures were higher than gas temperatures but gas temperatures rise rapidly further down the combustion zone due to the large amount of heat released by gaseous phase oxidations, and gas temperatures exceed solid surface temperatures throughout the rest of the gasifier.

4) The superficial gas velocity, whilst the same as the inlet air velocity at the top of the gasifier, thereafter increases rapidly along the drying, pyrolysis and combustion zone. This occurs both because of mass transfers from the solid phase to the gas phase and also because of increases in gas phase temperatures.

5) The drying zone is typically short in a stratified downdraft gasifier, being of the order of only a few millimetres long due to the steep solids temperature rise in the upper part of the gasifier.

6) At steady state conditions, some CO is present in the gas leaving the oxidation zone but its concentration increases substantially in the gasification zone. Any hydrogen formed during pyrolysis processes is completely consumed in the combustion zone and so the hydrogen content of the producer gas results from char-H2O reactions in the char gasification zone. A virtually constant value of CH4 is found along the gasification zone, which shows that reactions involving CH4 occur to only a limited extent in this zone.

7) Primary tar release occurs over a relatively narrow pyrolysis zone in the reactor. Much of this tar is cracked into combustible gases and burnt while passing through the hot combustion zone but some passes uncracked into the gasification zone. Here cracking of the primary tar continues, giving mostly secondary tar (hydrocarbon). Decomposition of condensable tar is not complete within the

gasifer and a small amount is present in the outlet producer gas. More secondary tar than primary tar (oxygenate) is found in the outlet gas.

220

Chapter 8 Conclusions and recommendations

8) During the reactor warming up period, the reaction zone usually moves upwards since the pyrolysis process is dominant at that stage and gasification rates are very low. Once otherwise steady operating conditions are attained, three modes of operation are possible. At low air flow rates, the gasifier operates in the pyrolysis dominant mode and the reaction zone moves upwards. By increasing the air flow rate, the upward movement of the reaction zones can be slowed down. Once the air flow rate exceeds a certain critical value, the gasifier starts to operate in the gasification dominant mode and the reaction zones start to move downwards. The gasifier was predicted to operate stably, with the reaction zones stationary, when the specific gasification rate is in the range of 290-296 kg/hr/m.

9) The air flow rate to the gasifier was confirmed to be a most important operating parameter, strongly affecting temperatures, tar content and the stability of the reaction zones. A higher air supply rate promotes oxidation rates, and brings about increased gas and solid temperatures which in turn enhance the cracking of primary tar and the oxidation of secondary tar.

10) Wood particle size affects the composition of producer gas significantly. With an increase in particle size, a poorer quality producer gas is predicted: CO and H2 concentrations decrease while the CO2 concentration increases. Smaller amounts of tar are predicted for a larger wood particle size due to an increase in gas phase temperature. These differences occur in the case of larger particles since the pyrolysis rate is lower (due to slower rates of heat transport to the particle core) and the gasifier operates in a gasification dominant mode.

11) The magnitude of the solid-gas heat and mass transfer coefficient has significant effects on gas phase temperatures and outlet tar content. For higher solid-gas heat transfer coefficients or lower mass transfer coefficients, the oxidation rate of volatiles was enhanced, increasing the gas peak temperature and decreasing the amount of tar in the outlet producer gas.

221

Chapter 8 Conclusions and recommendations

12) Changing the heat and mass transport exchange area affects the tar content in the outlet gas. A larger heat and mass exchange area increases the final tar content.

13) The kinetics of tar formation and tar cracking directly control the final tar content of the gas. A higher tar formation rate and a lower tar cracking rate result in a higher tar content in the producer gas.

14) At a higher effective thermal conductivity, both solids and gas temperatures decrease, the amounts of tar in the outlet gases decrease and the reaction zone moves upwards.

15) The model predicts that a reduced tar content of the producer gas is achievable by the effective use of insulation. For high wall heat loss (insufficient insulation), a higher air supply rate is required to maintain a stable reaction zone.

8.3

Summary and conclusions for the experimental study

To characterize the performance of a stratified downdraft gasifier in practice and to validate the developed computer model, a small scale stratified downdraft gasifier (206 mm ID) was run using two different sizes of red gum wood blocks as fuel. A total of 19 experimental runs was carried out over a range of air supply rates. Temperatures at 16 different levels in the reactor were measured, gas from five different levels in the reactor was sampled, and both the composition of the gas and its tar content were measured.

The major conclusions of this experimental investigation can be summarized as follows:

1) The overall material balance for the experimental stratified downdraft gasifier closed very well for all experimental runs. Reasonably good elemental mass balance closures were also achieved.

222

Chapter 8 Conclusions and recommendations

2) A substantial reduction in the tar content of the outlet gases could be achieved by initially filling the gasifier with charcoal up to the ignition port level. The average measured tar content of the outlet gas for all experimental runs using an initial charcoal bed was found to be 44 mg/Nm, which is about 40 times less than the tar content found in gas from a traditional downdraft gasifier.

3) The experimental gasifier could be operated over a fairly wide range from 7.5 to 27.6 kg/hr with a corresponding energy output ranging from 54-168 MJ/hr. However (see point no 10 below), long term stable operation is only possible within a limited range of air flow rates.

4) The producer gas composition did not vary significantly over the range of air flow rates used. Of the combustible gas components, CO had the highest concentration (16-21 vol %), followed by H2 (14-19 vol %). Small amounts of CH4 (1.0-1.7 vol %) were also present.

5) The variation over time of the measured temperature profile along the axis of the gasifier agreed reasonably well with model predictions. It was found that 20 to 25 minutes was required to reach a steady temperature profile. The combustion zone temperature increased with an increase in airflow rate and peak values were in the range of 1000-1300C. As predicted in the numerical model, outlet gas temperature also increased at a higher air flow rate.

6) A discrepancy between the predicted gas temperatures and measured values within the combustion zone was observed. This is attributable to the difficulties involved in experimentally measuring the true flame temperature within the gas spaces of a packed bed. Also, in the model gas phase oxidation reactions were characterized using only the intrinsic kinetic rate, and neglecting mass transfer effects. This leads to a situation where complete oxidation is predicted to occur within a very limited zone length in the reactor and overestimates the gas phase temperature.

7) Gas composition profiles along the reactor were measured during the reactor warm up phase and during steady state operation. It was found that the concentration of

223

Chapter 8 Conclusions and recommendations

the two main combustible components (CO & H2) of the outlet producer gas were low initially and increased over time reaching their final steady values after around 20 to 25 minutes. Model predictions of gas composition at two different times (during reactor warm up and steady operating conditions) agreed satisfactorily with experimental values. However, the experimentally observed CO and H2 formation rate along the gasification zone was found to be more gradual than that predicted by the model (for the same reasons as those discussed in point no 6 above).

8) The tar content of gas samples from five different levels in the reactor was measured and it was found that the amount of tar in the gas from sampling ports located 650 mm and 550 mm above the grate was as high as 4000 mg/Nm. This value was assumed representative of the amount of tar in the pyrolysis zone. The average tar content along the gasification zone was found to be 25 times less than in the pyrolysis zone. The average tar content in the outlet producer gas was around 80 times less than that in the gases in the pyrolysis zone.

9) Wood consumption rate and gas production rate increased linearly with increases in air flow rate. Model predictions of wood consumption and gas production rates agreed very well with experimental values. The air-fuel ratio was found to be constant and independent of air supply rate.

10) As the model predicted, experimental results showed that the air flow rate is the main operating parameter controlling gasifier stability. Higher airflow rates favoured the gasification dominant mode of operation as did the use of a larger wood block size. The specific gasification rate under stable operating conditions (reaction zones stationary within the gasifier) was in the range of 290-308 kg/mhr (dry basis) and this agreed well with model predictions.

Overall, there was an encouragingly good level of agreement between the results of model simulations and experimental measurements; however, deficiencies in our understanding of combined kinetic, heat transfer and mass transfer processes in the gas spaces of packed beds at high temperatures mean that some discrepancies remain between predicted and observed changes in the gasifier combustion zone.

224

Chapter 8 Conclusions and recommendations

8.4

Summary and conclusions related specifically to tar reduction

The model developed has been used to predict suitable operating conditions for a stratified downdraft gasifier that can operate stably over prolonged periods producing low tar gas, and these predictions have been confirmed experimentally.

1) A reaction zone length of 375 mm, for which a 97 % cold gas efficiency at the reactor outlet was observed in experimental results, and a 99.3% efficiency in model results, appears to be an optimum reaction zone length for the stratified downdraft gasifier on the basis of cold gas efficiency considerations. However, from a tar minimization point of view, a longer reaction zone length of 425 mm is a more suitable length for the stratified downdraft gasifier studied in this project.

2) Operating the reactor with wood fuel alone produced an outlet gas with a final tar content that was about 20 times higher than the average tar content value measured when using an initial charge of charcoal. This is attributable to incomplete devolatilization of the fuel in the region of the gasifier below the ignition level. This problem persisted, especially in the region less than 100 mm above the grate, even after two hours operation. It was inferred from this that to produce a low tar gas tar formation needs to be avoided in the region downstream of the oxidation zone.

3) Avoidance of devolatilization and the associated tar formation can be simply achieved by initially charging the gasifier with charcoal to the ignition port level. The requirement to have extra charcoal on hand for every run can be eliminated by using the charcoal left over from the previous run. The required charcoal bed could also, in theory, be produced inside the gasifier by operating the gasifier initially at a low air flow rate and then increasing the air flow rate to its normal operating level. However, the model predictions suggest that more than an hour would be required to produce a charcoal bed of sufficient depth (425 mm). There is also the problem of what to do with the high tar content gas produced during the charcoal-forming stage.

225

Chapter 8 Conclusions and recommendations

4) The dynamic model developed was used to simulate the behaviour of a two-stage gasifier in which air was introduced at two different levels in the reactor. The secondary air was found to create a secondary oxidation zone and a local increase in both solid and gas phase temperatures was predicted, as well as significant tar reduction.

The major finding was that the stratified downdraft gasifier using an initial charge of charcoal can produce a product gas with a tar content low enough to be acceptable for use in engine applications. Further reduction of tar content appears achievable by introducing secondary air at a suitable point downstream of the primary oxidation zone, provided that the secondary air can be distributed effectively across the gasifier cross-section.

8.5

Recommendations for further research works

The model developed is able to predict the dynamic behaviour of a stratified downdraft gasifier, including the changes in tar content with time along the gasifier, and the model predictions generally agreed well with experimental observations. However, further research needs to be carried out in order to improve the modelling of the biomass gasification process and to increase our understanding of the gasification process and reactor optimisation. Recommendations for further research work are as follows:

1) The kinetic coefficients used in modelling studies, including the one undertaken in this project, are generally obtained from experimental investigations in which reaction rates are determined by thermogravimetric analysis. The conditions employed in such experiments differ significantly from real gasification conditions. For instance, the kinetic expressions for the heterogeneous reactions (both for oxidation and gasification) used in this model were obtained from fine particle experiments where intraparticle heat and mass transfer processes are of very limited importance. This is not the case with larger particles for which intraparticle heat and mass transfer processes cannot be ignored (Salazar 1987). Effects of

226

Chapter 8 Conclusions and recommendations

particle size and biomass charcoal properties on oxidation and gasification reaction rates therefore need fuller investigation so that representative apparent kinetic rates can be found. Only once single particle processes are well characterized, and the resulting kinetic data integrated into gasification models, will predictions of reactor performance be improved sufficiently to eliminate the discrepancies between model predictions and experimental observations referred to earlier.

2) Tar reduction can occur by both homogeneous and heterogenous thermal cracking. However, all of the kinetic rates expressions for tar reduction were derived from homogenous tar cracking studies and no data were available for heterogenous tar cracking mechanisms. In this study, to avoid the overpredicting of tar content, a correction factor approach was applied. However, it is highly desirable that the need for this simplified assumption should be overcome by the determination of heterogeneous tar cracking data in which the effects of hot particle size are also considered.

3) In this study, in the modelling of primary pyrolysis processes a major simplifying assumption was made. Instead of temperature gradients within particles being allowed for an equivalent temperature for the whole particle was estimated as a function of surface temperature and particle diameter and used in kinetic expressions. To correctly estimate rates of formation and evolution of volatile gases (including tar) in the pyrolysis zone, a sub-model needs to be developed in which heat transport and mass transport processes within the particle are accounted for. This sub-model should be able to be incorporated into the main gasifier program and this will result in a more precise devolatilization rate in the pyrolysis zone.

4) In the dynamic model developed, the size of particle was assumed to be the same both before and after the pyrolysis process. However, experimental studies have shown that particle size is reduced during the pyrolysis process (Salazar 1987). Therefore an experimental correlation which covers different sizes and different temperature ranges needs to be developed so that particle shrinkage due to pyrolysis can be properly included in model formulations.

227

Chapter 8 Conclusions and recommendations

5) Solid-to-gas heat and mass transport processes are very important in modelling studies of gasification. In this study, a somewhat arbitrary correction factor had to be used to convert solid-gas heat transfer coefficients obtained for a non-reacting surface to ones applicable to a reacting surface, since in real gasification processes, heat and mass transfer between the solid surface and the gas phase occurs under strongly reactive conditions. To avoid the inaccuracies inherent in the use of an arbitrary correction factor further experimental research is needed on reactive fixed beds to investigate the influences of reaction processes on the interphase heat and mass transfer.

6) The dynamic model developed in this study has been shown to be able to predict separate solid and gas phase temperature profiles. However, no specific experimental techniques appear to be available to measure separately the temperatures of the solid and gas phases in a packed bed of particles at high temperatures. To compare the temperature profiles from modelling further work to develop a temperature measurement system capable of better determining solid and gas temperatures along the gasifier is highly desirable.

7) In the model, gas phase oxidation reactions were modelled using reaction rate expressions for combustible gases that were determined under conditions where oxygen availability was higher than in a gasifier. More reliable gas phase oxidation kinetic coefficients are required to avoid the overestimation of gas peak temperatures.

8) By modifying the existing experimental set-up, the performance of a two-stage gasifier should be investigated.

9) For two-stage gasification, better air supply methods, which are able to uniformly distribute the secondary air across the gasifier cross section without disturbing the solid fuel flow, should be explored.

228

Chapter 8 Conclusions and recommendations

8.6

Reference

Salazar, CM 1987, The influence of particle size and shape on the mechanisms of decomposition of wood during pyrolysis, PhD Dissertation, The University of Melbourne, Australia.

229

Appendix A Average particle temperature for use in primary pyrolysis process calculations

Appendix A

Average particle temperature for use in primary pyrolysis process calculations

In the transient model developed, the primary pyrolysis rate was estimated by using an average particle temperature in the primary pyrolysis kinetic rate equation. Average particle temperature was estimated by an equation derived from the correlation of Reed and Markson (1983) who determined the time required for completion of the flaming pyrolysis process and correlated this with particle size, temperature, moisture content and oxygen fraction. Details of the derivation are presented below:

Reed and Markson (1983) proposed the following equations to estimate the time required for completion of the flaming pyrolysis process:

t fp =

(3.481E 2) (1 + 1.76M )80.6 D(1 + 80.6 0.61D ) exp(2201T ) (1 + 3.4 FO 2 )

(A.1)

where tfp M D T = time required to complete the flaming pyrolysis process (sec) = apparent density of fuel (kg/m) = moisture content (wt fraction -) = diameter of fuel (m) = surface temperature (K)

FO2 = oxygen volume fraction (-)

Equation (A.1) can be rewritten as

(3.481E 2) (1 + 1.76M ) exp(2201T ) (1 + 3.4 FO 2 ) 80.6 D(1 + 80.6 0.61D )


t fp =

(A.2)

230

Appendix A Average particle temperature for use in primary pyrolysis process calculations

For the same , M, FO2 & T, and for different diameters D1 and D2

t D1 t D2 = 80.6 D1 (1 + 80.6 0.61D1 ) 80.6 D2 (1 + 80.6 0.61D2 )

(A.3)

For fine particles,


E Kinetic rate of primary pyrolysis (kg/ms) = Aexp RT

(A.4)

where

A E R

= pre-exponential factor for primary pyrolysis (1/sec) = activation energy for primary pyrolysis (J/mol) = universal gas constant (J/mol/K)

Rate of primary pyrolysis (kg/ms) can be also written as

Rp =

t fp

(A.5)

From equation A.4 & A.5, the time required to complete the pyrolytic decomposition of a particle of diameter D can be estimated: tfp =

E A exp RT D

1 E A exp RT D

(A.6)

where

TD = average temperature of particle in K

(Assuming D1 1 mm, fine particle)

tD1 =

1 E A exp RT D1

(A.7)

TD1 = average temperature of fine particle which is assumed to be equal to surface temperature of particle (K)

From equation (A.3) the time required to complete the pyrolytic decomposition of a particle of diameter D2 (sec) is:

231

Appendix A Average particle temperature for use in primary pyrolysis process calculations

tD2 = t D1

D2 (1 + 80.6 0.61D2 ) D1 (1 + 80.6 0.61D1 )

(A.8)

Since D1 = 0.001 m, equation A.8 becomes


t D 2 = t D1 D2 (1 + 80.6 0.61D2 ) = t D1 (953.138 D2 + 46861.984 D2 ) 0.001(1 + 80.6 0.61 0.001)

(A.9)

Substituting tD1 (eq. A.7) in equation A.8,

t D2 =

1 (953.138D2 + 46861.984 D2 ) E A exp RT D1

(A.10)

tD2 can be estimated from the equivalent (average) temperature for D2, (TD2) by using equation A.6

t D2 =

1 E A exp RT D2

(A.11)

Therefore,

1 1 (953.138D2 + 46861.984 D2 ) = E E A exp RT A exp RT D1 D2


E exp RT E D1 = exp RTD 2 (953.138D2 + 46861.984 D2 ) E E RT = RT ln(953.138D2 + 46861.984 D2 ) D 2 D1 1 1 R T = T + E ln(953.138D2 + 46861.984 D2 ) D 2 D1

TD 2 =

1 1 R T + E ln(953.138D2 + 46861.984 D2 ) D1

(A.12)

where TD1 is the temperature for a fine particle which is assumed as the surface temperature of a wood particle

232

Appendix A Average particle temperature for use in primary pyrolysis process calculations

Therefore, the general form of the equation to estimate the average temperature of the particle with diameter D can be written using equation A.12 as follows:

Tavg =

1 1 T S R + ln(953.138D + 46861.984 D ) E

(A.13)

where Tavg = average wood particle temperature (K) TS R E D = solid surface temperature (K) = universal gas constant (J/mol/K) = activation energy for primary pyrolysis (J/mol) = particle diameter (m)

Reference

Reed, TB & Markson, M 1983, A predictive model for stratified downdraft gasification of biomass, Progress in biomass conversion, vol. 4, eds. DA Tillman & EC John, pp. 217-54.

233

Appendix B Properties of gas species

Appendix B

Properties of gas species

Table B1 Atomic Diffusion Volumes (Reid, Prausnitz & Poling 1987) Atomic and structural diffusion volume increments C H O N 15.9 2.31 6.11 4.54

Diffusion volumes of molecules H2 N2 O2 Air CO CO2 H2O 6.12 18.5 16.3 19.7 18.0 26.9 13.1

234

Appendix B Properties of gas species

Table B2 Curve-fit coefficients for absolute enthalpy of selected gas species


0 a a a a h = a1T + 2 T 2 + 3 T 3 + 4 T 4 + 5 T 5 + a 6 R (Turns, 2000) 2 3 4 5

Species CO CO2 H2 H2O O2

T(K) 1000-5000 300-1000 1000-5000 300-1000 1000-5000 300-1000 1000-5000 300-1000 1000-5000 300-1000

a1 0.03025078E+02 0.03262451E+02 0.04453623E+02 0.02275724E+02 0.02991423E+02 0.03298124E+02 0.02672145E+02 0.03386842E+02 0.03697578E+02 0.03212936E+02
0

a2 0.14426885E-02 0.15119409E-02 0.03140168E-01 0.09922072E-01 0.07000644E-02 0.08249441E-02 0.03056293E-01 0.03474982E-01 0.06135197E-02 0.11274864E-02

a3 -0.05630827E-05 -0.03881755E-04 -0.12784105E-05 -0.10409113E-04 -0.05633828E-06 -0.08143015E-05 -0.08730260E-05 -0.06354696E-04 -0.1258842E-06 -0.05756150E-05

a4 0.10185813E-09 0.05581944E-07 0.02393996E-08 0.06866686E-07 -0.09231578E-10 -0.09475434E-09 0.12009964E-09 0.06968581E-07 0.01775281E-09 0.13138773E-08

a5

a6

-0.06910951E-13 -0.14268350E+05 -0.02474951E-10 -0.14310539E+05 -0.16690333E-13 -0.04896696E+06 -0.02117280E-10 -0.04837314E+06 0.15827519E-14 -0.08350340E+04 0.04134872E-11 -0.10125209E+04 -0.06391618E-13 -0.02989921E+06 -0.02506588E-10 -0.03020811E+06 -0.11364354E-14 -0.12339301E+04 -0.08768554E-11 -0.10052490E+04

h = a1T + a 2T 2 + a 3T 3 + a 4T 4 + a 5T 5 + a 6 (Turns, 2000)

Species C

T(K) 1000-5000 300-1000

a1 1.23870304E+01 5.59755420E+00

a2 6.91227177E-03 2.98573864E-02

a3 -1.85438496E-06 -1.54884907E-05

a4 2.68524262E-10 4.26136365E-09

a5

a6

-1.53215420E-14 -5.88033398E+03 -6.08974359E-13 -6.05750000E+02

235

Appendix B Properties of gas species

0 0 B C 2 3 h = h ref + A(T Tref ) + T 2 Tref + T 3 Tref 2 3

) R (Annamalai and Puri, 2001)

Species CH4 C2H4 C6H6

href0 (J/mol) (Thrns, 2000) -74831 +52283 +82927

A 1.702E+00 1.424E+00 -0.206E+00

B 9.081E-03 14.394E-03 39.064E-03

C -2.164E-06 -4.392E-06 -13.301E-06

where Tref =298.15 K

References
Annamalai, K & Puri, IK 2001, Advanced thermodynamics engineering, Boca Raton, FL: CRC Press. Reid, RC, Prausnitz, JM & Poling, RE 1987, The properties of gases and liquids, 4th edn, McGraw-Hill. Turns, SR 2000, An introduction to combustion: Concepts and application, 2nd edn, McGrawHill, New York.

236

Appendix C Discretization of model partial differential equations

Appendix C

Discretization of model partial differential equations

In this section, the step-by-step development of algebraic equations from the model partial differential equations in Chapter 3 (eq. 3.46-3.62) is presented.

The following diagram, which is the same as Figure 3.3, shows the grid-point cluster for the one-dimensional problem used for discretization.

xi

Control volume

Ti-1 i-1 i-

Ti i i+

Ti+1 i+1

xi-

xi+

Figure C1 A generic node point i (non-uniform grid) and the control volume in one dimension

Applying the methods and procedures (Pantakar 1980) described in Chapter 3, the model partial differential equations are discretized as follows:

237

Appendix C Discretization of model partial differential equations

C.1 Mass balance equations (solid phase)

Char:

By assuming constant char density along the reactor, which means that the time dependent char density term in equation 3.46 can be eliminated, an expression for the solid velocity Vs can be developed.

After eliminating the transient char density terms, equation 3.46 becomes

(1 )

4 CVS = Ri x i =1

whence
(1 ) c V S = (R1 + R2 + R3 + R4 ) x

(1 ) c

(VS ) i (VS ) i 1 i = (R1 + R2 + R3 + R4 ) xi 1


2

(1 ) c (VS ) i 1 (1 ) c (VS ) i = (R1 + R2 + R3 + R4 ) xi 1


i

This can be rewritten in the form VS i 1 = where A = (1 ) c B = (1 ) c

BVS i + C A

C = (R1 + R2 + R3 + R4 ) xi 1
i

Volatile matter:

The mass balance equation (eq. 3.47) can be discretized as follows:

238

Appendix C Discretization of model partial differential equations

(1 )

(V )i (V )i0
t

+ (1 )

(V VS )i (V VS )i 1
xi 1
2

= (R p1 )i

(V )i xi 1
t

(V )i0 xi 1
t

+ ( V )i (V S )i ( V )i 1 (V S )i 1 =

((1 )k

p1

V )i xi 1

(1 )

This can be rewritten in the form V i = where


A= xi 1 t
2

B V i 1 + C A

+ (VS )i + (k p1 )i xi 1

B = (VS )i 1

C=

(V )i0 xi 1
t

Moisture :

The mass balance equation (eq. 3.48) can be discretized as follows:


(1 )

( m )i ( m )i0
t

+ (1 )

( mVS )i ( mVS )i 1
xi 1
2

= (Rm )i

( m )i xi 1
t

( m )i0 xi 1
t

+ ( m )i (VS )i ( m )i 1 (VS )i 1 =

((1 )k m m )i xi 1 2 (1 )

This can be rewritten in the form m i = where


A= xi 1 t
2

B m i 1 + C A

+ (VS )i + (k m )i xi 1

B = (VS )i 1

C=

( m )i0 xi 1
t

239

Appendix C Discretization of model partial differential equations

Particle Diameter

The solid fuel particle diameter in each horizontal slice of differential thickness x within the gasifier can be determined by discretizing eq. 3.49, as shown below:
(1 ) (1 )V s + t 3 x 3 d d 6 6 =0

1 Vs + =0 t d 3 x d 3
Vs Vs 1 1 3 3 3 3 d i d i d i d i 1 + =0 t xi 1
0 2

(Vs ) 1 1 3 xi 1 3 xi 1 + 3 i d 2 2 d i d i i
0

( )
0 i

(V ) t s3 i 1 t = 0 d i 1

( )

1 (V ) 3 xi 1 + s3 i d d 2 i i

( )

( )

1 t = d3

( )

(Vs )i 1 x t 1 + i 2 d 3 i 1

( )

( )

1 (V ) 1 x xi 1 + (Vs )i t = + s3 i 1 t 3 0 i 1 d d3 2 2 d i i 1 i

( )

( )

(d )i

3 xi 1 + (Vs )i t 2 = xi 1 (V ) t 2 + s i 1 (d 3 )0 (d 3 )i 1 i

C.2 Mass balance equation (gaseous phase)

The mass balance equations for the gaseous phase contain nonlinear terms, which need be linearized using eq. 3.68. Linearization is carried out on equations written in terms of the molar fractions of the particular species involved. In this case, before

240

Appendix C Discretization of model partial differential equations

discretizing the differential equations, all of the reaction rate equations need to be converted to the molar fraction form by using the ideal gas law:

j =

pjM j RTg

(or) C j =

pj RTg

where, p j = PY j and Y j = mole fraction of species j

After applying the ideal gas law, the heterogeneous gas-char reaction rate equations (eq. 3.9) become:
AM Heterogeneous char oxidation . R1 = v C ( )RT g AM R2 = v C RT g PYO2 1 1 + k m k1 PYCO2 1 1 + k m k 2

Boudouard reaction

Water gas reaction

AM R3 = v C RT g

PYH 2O 1 1 + k m k 3

AM Methane formation reaction R4 = v C 2 RT g

PYH 2 1 1 + k m k 4

Similarly, the gas phase homogenous reactions are changed to their molar fraction form.

By applying the ideal gas law, the forward water-gas shift reaction rate equation (eq. 3.14) becomes:

241

Appendix C Discretization of model partial differential equations

E5 F R5 F = ( ) A5 F exp RT g

2 2 P Tg YCO YH O 2
kmolK 2 m 3 kPa 2 s

where A5 F = 4.34 10 4

The reverse water-gas shift reaction rate equation (eq.3.15) becomes:


E5 F A5 F exp RT g = KC P 2T 2Y
g

R5 R

CO2

YH 2

The rate equation (eq. 3.18) for hydrocarbon oxidation becomes:


E6 R6 = ( ) A6 exp RT g 0.5 1.8 0.5 Tg P YC H YO 2 6 6

kmolK 0.5 where, A6 = 6858 3 m kPa1.8 Sec

Since YC6H6 is nonlinear in the above equation, linearization is performed with respect to YC6H6. The equation becomes . R6 = R + 0.5 R
* 6 * 6

YC6 H 6 YC*6 H 6 YC*6 H 6

The superscript, *, denotes the previous-iteration values.

In the above linearization, the factor 0.5 is changed to 1.2 which makes the slope steeper. The steeper slope results in stable iteration even though it also leads to slower convergence.

Thus, the final form of the hydrocarbon oxidation equation (used for the secondary tar balance) is:
YC6 H 6 YC*6 H 6 YC*6 H 6

R 6 = R + 1 .2 R
* 6

* 6

The rate equation for carbon monoxide oxidation (3.20) becomes:


E7 R7 = ( ) A7 exp RT g 2 2 0 P Tg YCO YO0.5YH .5 2 2O

where, A7 = 1.88 10 9

kmolK 2 , m 3 kPa 2 s

242

Appendix C Discretization of model partial differential equations

Since YO2 is nonlinear in the above equation, linearization is performed with respect to YO2. The equation becomes . R7 = R + 0.5 R
* 7 * 7 * YO2 YO2 * YO2

In the above linearization, the factor 0.5 is also changed to 1.2 to make the slope steeper and achieve stable iteration.

Thus, the final form of the carbon monoxide oxidation equation (used for the oxygen balance) is:
R 7 = R + 1 .2 R
* 7 * 7 * YO2 YO2 * YO2

The rate equation for oxidation of methane (3.22) becomes:


1.5 E8 1.5 1.5 P Tg YCH 0.7YO 2 0.8 where, A8 = 2.09 1010 kmolK , R8 = ( ) A8 exp 4 RT m 3 kPa 1.5 s g

Since YCH4 and YO2 are nonlinear in the above equation, linearization is performed with respect to YCH4 and YO2. Linearizing with respect to YCH4 ,
* * R8 = R8 + 0.7 R8 * YCH 4 YCH 4 * YCH 4

Linearizing with respect to YO2 ,

R8 = R + 0.8 R
* 8

* 8

* YO2 YO2 * YO2

In the above two linearizations, the factors 0.7 and 0.8 are changed to 1.1 and 1.2 to make the slope steeper and thus achieve stable iteration.

Thus, the final form of the methane oxidation equation (used for the methane balance) is:
R8 = R + 1.1R
* 8 * 8 * YCH 4 YCH 4 * YCH 4

243

Appendix C Discretization of model partial differential equations

Thus, the final form of the methane oxidation equation (used for the oxygen balance) is:
R8 = R + 1.2 R
* 8 * 8 * YO2 YO2 * YO2

Using a process similar to that used for eq. 3.18, the rate equation for ethylene oxidation (eq. 3.24) can be transformed to
E9 R9 = ( ) A9 exp RT g 0.5 1.8 0.5 Tg P YC H YO 2 2 4

kmolK 0.5 where, A9 = 6858 3 m kPa 1.8 s

Since YC2H4 is nonlinear in the above equation, linearization is performed with respect to YC2H4. The equation becomes . R9 = R + 0.5 R
* 9 * 9

YC2 H 4 YC*2 H 4 YC*2 H 4

Similarly in the case of C6H6, the factor 0.5 is also changed to 1.2 to make the slope steeper.

Thus, the final form of the ethylene oxidation equation (used for the ethylene balance) is:
R9 = R + 1.2 R
* 9 * 9

YC2 H 4 YC*2 H 4 YC*2 H 4

The rate equation for hydrogen oxidation (3.26) becomes:


E10 R10 = ( ) A10 exp RT g 2.2 2.2 Tg P (YH 2 )1.1 (YO 2 )1.1
kmolK 2.2 m 3 kPa 2.2 s

where A10 = 8.4 10 6

Since YH2 and YO2 are nonlinear in the above equation, linearization is performed with respect to YH2 and YO2. Linearizing with respect to YH2 ,
R10 = R + 1.1R
* 10 * 10 * YH 2 YH 2 * YH 2 * YO2 YO2 * YO2

Linearizing with respect to YO2 ,

R10 = R + 1.1R
* 10

* 10

244

Appendix C Discretization of model partial differential equations

After converting the reaction rate equations to molar fraction forms and after linearization has been done, the mass balance differential equations (gaseous phase) are discretized as follows:

Oxygen:

The mass balance equation for oxygen (eq. 3.51) is discretized as shown below:
M O2 M = O2 R1 3M O R6 0.5M O R7 C O2 M O 2 + m YO2 g 2 2 t x Mg MC 1.5M O2 R8 M O2 R9 0.5M O2 R10

M O2 YO2 P M O 2 + mYO2 g x Mg t RTg

M = O2 R1 3M O R6 0.5M O R7 2 2 MC

1.5M O2 R8 M O2 R9 0.5M O2 R10

( )YO2 PM O2 RTg

( )YO2 PM O2 RTg i t

M O2 mYO g 2 M g

M O2 mYO g 2 M g i xi 1
2

i 1

M O2 Av M C M C RTg

PYO2 3M O2 ( )k 6Tg0.5 P 1.8YC06.5 6 YO 2 H 1 1 + k m k1


* * Y YO * O 1.5M O2 R8 + 1.2 R8 2 * 2 YO2

* * Y YO * O 0.5M O2 R7 + 1.2 R7 2 * 2 YO2

0.5 1.8 0.5 M O2 ( )k 9Tg P YC6 H 6 YO 2

* * YO2 YO2 * 0.5M O2 R10 + 1.1R10 * YO2

This can be rewritten in the form

(YO 2 )i

B(YO 2 )i 1 + C A

245

Appendix C Discretization of model partial differential equations

where
( ) PM O2 xi 1 2 A= RTg t + m M O2 g Mg i M + O2 MC i Av M C RT g Pxi 1 2 1 1 + k m k1
2

+ 3M O2 ( )k 6T + M O2 ( )k 9T

0.5 g

P Y

1.8

0.5 C6 H 6

xi 1 +
2

* 0.6M O2 R7 xi 1

* O2

* 1.8M O2 R8 xi 1

* O2

0.5 g

P Y

1.8

0.5 C6 H 6

xi 1 +
2

* 0.55M O2 R10 xi 1

* O2

M O2 B = m g Mg

i 1
* * + 0.1M R * x + 0.3M O2 R8 xi 1 + 0.05M O2 R10 xi 1 O2 7 i 1 2 2 2 i
0

( )YO2 PM O2 xi 1 2 C = RTg t

Carbon monoxide:

The mass balance equation for carbon monoxide (eq. 3.49) is discretized as shown below:
= 2(1 ) M CO R1 + 2 M CO R2 + M CO R3 MC MC MC 3.20 27.62 R p1 + R p2 + 6 M CO R6 + M CO R8 + 2 M CO R9 + M CO R5 R M CO R5 F M CO R7 + 29.50 52.80
( )YCO P ( )YCO P mYCO M CO M CO M CO RT g Mg g i i RTg + t = 2(1 )
0

M YCO P M CO + mYCO CO g x Mg t RTg

mYCO M CO g Mg i xi 1
2

i 1

M CO M M R1 + 2 CO R2 + CO R3 + 6 M CO R6 + M CO R8 + 2 M CO R9 + M CO R5 R MC MC MC

0 0 M CO ( )k 5 F P 2Tg 2 YCO YH 2O M CO ( )k 7 P 2Tg 2YCO YO2.5YH .25 + O

3.20 27.62 R p1 + R p2 29.50 52.80

246

Appendix C Discretization of model partial differential equations

( )YCO P ( )YCO P mYCO M CO M CO M CO RT RT g Mg g g i i + t = 2(1 )

mYCO M CO g Mg i x i 1
2

i 1

M CO M M R1 + 2 CO R2 + CO R3 + 6 M CO R6 + M CO R8 + 2 M CO R9 + M CO R5 R MC MC MC

0 0 M CO ( )k 5 P 2Tg 2 YCO YH 2O M CO ( )k 7 P 2Tg 2YCO YO2.5YH .25 + O

3.20 27.62 R p1 + R p2 29.50 52.80

This can be rewritten in the form where

(YCO )i

B(YCO )i 1 + C A

( ) PM CO xi 1 2 A= RTg t

+ m M CO g Mg i
2

+ M CO ( )k 5 P 2Tg 2YH O x 1 i 2 2 i

0 + M CO ( )k 7 P 2Tg 2YO02.5YH .25 xi 1 O

M B = m CO g Mg

i 1
0

( )YCO PM CO xi 1 M 3.20 27.62 2 C = R p1 xi 1 + R p 2 xi 1 + 2(1 ) CO R1 xi 1 + 2 2 2 RTg t MC 29.50 52.80 i M M + 2 CO R2 xi 1 + CO R3 xi 1 + M CO R5 R xi 1 + 6 M CO R6 xi 1 2 2 2 2 MC MC + M CO R8 xi 1 + 2M CO R9 xi 1


2 2

Carbon dioxide:

The mass balance equation for carbon dioxide (eq. 3.53) is discretized as shown below:

CO2 t

dmCO2 dx

= (2 1)

M CO2 MC

R1 + M CO2 R5 F + M CO2 R7

M CO2 MC

R2 +

6.80 R p1 79.50

6.40 R p 2 M CO2 R5 R 52.80

247

Appendix C Discretization of model partial differential equations

M CO2 YCO2 P M CO2 + m YCO2 g x t RTg Mg M CO2 MC R2 +

M = (2 1) CO2 R1 + M CO R5 F + M CO R7 2 2 MC

6.80 6.40 R p1 + R p 2 M CO2 R5 R 79.50 52.80

( )YCO2 PM CO2 RTg

( )YCO2 PM CO2 RTg i t

M CO2 mYCO g 2 Mg M CO2 Av M C M C RTg

M CO2 mYCO g 2 Mg i xi 1
2

i 1

(2 1)

M CO2 MC

R1 + M CO2 R5 F + M CO2 R7

PYCO2 6.80 R + 1 1 79.50 p1 + k m k 2

6.40 R p2 52.80

E5 F A5 F exp RT g M CO2 KC

P 2T 2Y
g

CO2

YH 2

This can be rewritten in the form (YCO 2 )i = where


( ) PM CO2 xi 1 M + m CO2 A= g RTg t Mg i E5 F A5 F exp RT g + M CO2 KC M CO2 B = m g Mg i 1
0

B(YCO 2 )i 1 + C A
Pxi 1 2 1 1 + k m k 2

M CO2 + MC i

Av M C RT g

P 2T 2Y x g H2 i 1

( )YCO2 PM CO2 xi 1 6.80 6.40 2 C = R p1 xi 1 + R p 2 xi 1 + 2 2 RTg t 79.50 52.80 i M CO2 + (2 1) R1 xi 1 + M CO2 R5 F xi 1 + M CO2 R7 xi 1 2 2 2 MC

248

Appendix C Discretization of model partial differential equations

Hydrogen:

The mass balance equation for hydrogen (eq. 3.54) is discretized as shown below: H 2 t m 2 H x

M H2 MC

R3 + M H 2 R5 F + 3R6 M H 2 + 2 R9 M H 2 M H 2 R5 R 2

M H2 MC

R4

R10 M H 2 +

1.63 R p2 52.80

YH 2 PM H 2 t RTg

M H 2 R5 R

M H2 M H2 = + m YH R3 + M H 2 R5 F + 3 R 6 M H 2 + 2 R9 M H 2 g 2 z M g MC M H2 1.63 2 R4 R10 M H 2 + R p2 MC 52.80

( )YH 2 PM H 2 RTg

( )YH 2 PM H 2 RTg i t

M H2 mYH g 2 Mg

M H2 mYH g 2 Mg i xi 1
2

i 1

M H2 MC

R3

+ M H 2 R5 F + 3R6 M H 2 + 2 R9 M H 2 M H 2 Av M C 2 M C RTg PYH 2 1 1 + k m k 4

E5 F A5 F exp RT g M H2 KC

P 2T 2Y
g

CO2

YH 2

* * YH 2 YH 2 * M H 2 R10 + 1.1R10 * YH 2

1.63 R p2 + 52.80

This can be rewritten in the form (YH 2 )i = where


( ) PM H 2 xi 1 2 A= RTg t + m M H 2 g Mg i P 2T 2Y
g

B(YH 2 )i 1 + C A
Pxi 1 2 1 1 + k m k 4

M + 2 H2 MC i

Av M C RT g

E5 F A5 F exp RT g + M H2 KC M H2 B = m g Mg i 1

CO2

xi 1 +
2

* 1.1M H 2 R10 xi 1

* H2

249

Appendix C Discretization of model partial differential equations

( )YH 2 PM H 2 xi 1 2 C = RTg t
2

M H2 + 1.63 R x R3 xi 1 + M H 2 R5 F xi 1 p2 1 + i 2 2 2 52.80 MC i
2 2 2

* * + 3R6 M H 2 xi 1 + 2 R9 M H 2 xi 1 M H 2 R10 xi 1 + 1.1M H 2 R10 xi 1

Water vapor:

The mass balance equation for water vapor (eq. 3.55) is discretized as shown below:

M M + mYH O H 2O = 2 M H O R8 + M H O R10 H 2O R3 g 2 2 2 x MC Mg 16.30 M H 2O ( )k 5 F P 2Tg 2YCO YH 2O + Rm + R p1 + M H 2O R5 R 79.50

YH 2O PM H 2O t RTg

( )YH 2O PM H 2O RTg

( )YH 2O PM H 2O RTg i t M H 2O MC

M mYH O H 2O g 2 Mg

M m YH O H 2O g 2 Mg i xi 1
2

i 1

= 2M H 2O R8 + M H 2O R10 + 16.30 R p1 + M H 2O R5 R 79.50

R3 M H 2O ( )k 5 F P 2Tg 2YCO YH 2O + Rm

This can be rewritten in the form (YH 2O )i =

B(YH 2O )i 1 + C A

where
( ) PM H 2O xi 1 2 A= RTg t + m M H 2O g Mg i M H 2O + MC i Av M C RT g Px i 1 2 1 1 + k m k 3

+ M H 2O ( )k 5 F P 2Tg 2YCO xi 1

M H 2O B = m g Mg

i 1

250

Appendix C Discretization of model partial differential equations

( )YH 2O PM H 2O xi 1 16.30 2 C = + R m x i 1 + R p1 xi 1 + M H 2O R5 R xi 1 2 2 2 RTg t 79.50 i + 2 M H 2O R8 xi 1 + M H 2O R10 xi 1


2 2

Methane:

The mass balance equation for methane (eq. 3.56) is discretized as shown below:
YCH 4 PM CH 4 RTg t M CH 4 + m YCH g 4 x Mg M CH 4 0.40 5.10 = R4 M CH 4 R8 + R p1 + R p2 MC 79.50 52.80

( )YCH 4 PM CH 4 RTg

( )YCH 4 PM CH 4 RTg i t

M CH 4 m YCH g 4 M g

M CH 4 m YCH g 4 M g i xi 1
2

i 1

M CH 4 MC

* Y Y * CH 4 R4 M CH 4 R8 + 1.1R8 * YCH 4

* CH 4

0.40 5.10 R p1 + R p2 + 52.80 79.50

This can be rewritten in the form (YCH 4 )i = where


( ) PM CH 4 xi 1 2 A= RTg t + m M CH 4 g Mg i

B(YCH 4 )i 1 + C A

* 1.1M CH 4 R8 xi 1 2 + * YCH 4 i

M CH 4 B = m g Mg

i M CH 4 5.10 + 0.40 R x + R p 2 xi 1 + R4 xi 1 79.50 p1 i 12 52.80 2 2 MC i

( )YCH 4 PM CH 4 xi 1 2 C = RTg t
* + 0.1M CH 4 R8 xi 1 2

251

Appendix C Discretization of model partial differential equations

Ethylene:

The mass balance equation for ethylene (eq. 3.57) is discretized as shown below:
YC2 H 4 PM C2 H 4 t RTg M + YC H m C2 H 4 x 2 4 g M g
( )YC2 H 4 PM C2 H 4 RTg i t
* C2 H 4

5.40 = R R9 M C2 H 4 52.80 p 2
0

( )YC2 H 4 PM C2 H 4 RTg

M YC H m C2 H 4 g 2 4 Mg

M YC H m C2 H 4 g 2 4 Mg i xi 1
2

i 1

* Y Y * C H M C2 H 4 R9 + 1.2 R9 2 4 * YC2 H 4

5.40 + R 52.80 p 2

This can be rewritten in the form (YC 2 H 4 )i =


where

B(YC 2 H 4 )i 1 + C A

( ) PM C2 H 4 xi 1 2 A= RTg t
M C2 H 4 B = m g Mg i1

+ m M C2 H 4 g Mg i

M R* + 1.2 C2 H 4 9 x 1 i 2 YC*2 H 4 i

( )YC2 H 4 PM C2 H 4 xi 1 2 C = RTg t

* + 5.40 R x + 0.2M C2 H 4 R9 xi 1 p2 i 1 2 2 52.80 i

Primary Tar :

The mass balance equation for primary tar (oxygenate) (eq. 3.58) is discretized as shown below:
Ytar1 PM tar1 + Ytar1 m M tar1 = R p 2 + 52.80 R p1 g x 79.50 Mg t RTg

252

Appendix C Discretization of model partial differential equations

( )Ytar1 PM tar1 ( )Ytar1 PM tar1 Ytar1 m M tar1 Ytar1 m M tar1 g g RTg RTg Mg Mg i i i i 1 + = t xi 1
2

(( ) PT

1 g

k p 2Ytar1

52.80 R p1 79.50

This can be rewritten in the form (Ytar1 )i = where


( ) PM tar1 xi 1 2 A= RTg t M B = m tar1 g Mg i 1
( )Ytar1 PM tar1 xi 1 2 C = RTg t + 52.80 R x p1 i 1 2 79.50 i
0

B(Ytar1 )i 1 + C A

+ m M tar1 + ( ) PT 1 k x g p2 i 1 g Mg 2 i i

Secondary Tar :

The mass balance equation for secondary tar (hydrocarbon, benzene) (eq. 3.59) is discretized as shown below:
Ytar 2 PM tar 2 RTg t + Ytar 2 m M tar 2 g x Mg = R6 M tar 2 + 6.65 R p 2 52.80

( )Ytar 2 PM tar 2 RTg

( )Ytar 2 PM tar 2 RTg i t


* tar 2

Ytar 2 m M tar 2 g Mg

Ytar 2 m M tar 2 g Mg i xi 1
2

i 1

* Y * Y M tar 2 R6 + 1.2 R6 tar 2 * Ytar 2

6.65 + 52.80 R p 2

This can be rewritten in the form (Ytar 2 )i = where

B(Ytar 2 )i 1 + C A

253

Appendix C Discretization of model partial differential equations

( ) PM tar 2 xi 1 2 A= RTg t M Btar 2 = m tar 2 g Mg


C tar 2

+ m M tar 2 g Mg i

* + 1.2 M tar 2 R6 x 1 * i 2 Ytar 2 i

i 1
* + 6.65 R x + 0.2 M tar 2 R6 xi 1 p2 i 1 2 2 52.80 i
0

( )Ytar 2 PM tar 2 xi 1 2 = RTg t

Total gas :

The mass balance equation for total gas (eq. 3.60) is also discretized as shown below:
PM g t RTg + (m ) = R p1 + Rm + R1 + R2 + R3 + R4 x g
i
0

( ) PM g RT g

( ) PM g RT g i t

(m ) (m )
g i

g i 1

x i 1

= R p1 + Rm + R1 + R2 + R3 + R4

This can be rewritten in the form mg i = where


A =1 B =1

Bmg i 1 + C A

( ) PM g xi 1 ( ) PM g xi 1 2 2 C = (R p1 + Rm + R1 + R2 + R3 + R4 )xi 1 + 2 RTg t RTg t i i


0

C.3 Energy balance equation (solid phase)

The solid energy balance differential equation is discretized as follows:

254

Appendix C Discretization of model partial differential equations

(1 )

T ( s hs ) + (m S hs ) = k eff s hsg Av(Ts Tg ) x x x t

(eq. 3.61)

4 4h sw (Ts Tw ) + R m H m + R p1 H p1 + R j H j D j =1

- Central difference scheme for second order terms - Backward difference scheme for first order terms

where hs = Cp s (Ts Ta ) , ms = (1 )v s s

(1 ) ( s Cp s (Ts Ta ) ) i ( s Cp s (Ts Ta ) )i0 + t (1 ) [( s v s Cp s (Ts Ta ) )i ( s v s Cp s (Ts Ta ) )i 1 ] = x i 1


2

1 (K eff )i+ 12 (Ts )i+1 (Ts )i (K eff )i 12 (Ts )i (Ts )i1 (hsg Av (Ts Tg ) )i x i 1 x i x i + 1 2 2 4 4h sw (Ts Tw ) + Rm H m + R p1 H p1 + (R j H j )i D i j =1
(1 ) ( s Cp s Ts )i (1 ) ( s Cp s Ta )i (1 ) ( s Cp s Ts )i0 + (1 ) ( s Cp s Ta )i0 t t t t (1 )( s v s Cp s Ts )i (1 )( s v s Cp s Ta ) i (1 )( s v s Cp s Ts ) i 1 + x i 1 x i 1 x i 1 (K ) + (K ) (1 )( s v s Cp s Ta ) i 1 (K eff )i +1 + (K eff )i eff i ( ) eff i +1 + = (Ts )i Ts i +1 2x x x i 1 2 x i x i + 1 i i+ 1 2 2 2 (K ) + (K ) (K ) + (K ) eff i eff i 1 eff i 1 (Ts )i + eff i (Ts )i 1 (hsg Av Ts )i + (hsg Av T g )i 2 x x 2 x x i i i 1 i 1 2 2 4 4 4 (hsw Ts )i + (hsw Tw )i + R m H m + R p1 H p1 + (R j H j )i D D j =1
2 2 2

This can be rewritten in the form Ai (Ts )i = Bi (Ts )i +1 + Ci (Ts )i 1 + Di

255

Appendix C Discretization of model partial differential equations

Ai =

(K ) + (K ) (1 ) eff i +1 eff i ( m Cpm + bio Cpbio + char Cpchar ) i + 2x x t i i+ 1 2

(K ) + (K ) (1 )( Cp + Cp + Cp ) (v ) eff i eff i 1 m m bio bio char char i s i + + 2x x xi 1 i i 1 2 2 4(hsw )i + (hsg Av )i + D

(K ) + (K eff i +1 eff Bi = 2x x i i+ 1 2

)
i

(K ) + (K ) (1 )( Cp + Cp + Cp ) (v ) eff i eff i 1 m m bio bio char char i 1 s i 1 Ci = 2x x + xi 1 i i 1 2 2


(1 ) (( m Cp m + bio Cp bio + char Cp char )Ta ) i t (1 ) (( m Cp m + bio Cp bio + char Cp char )Ts )i0 + t (1 ) (( m Cp m + bio Cp bio + char Cp char )Ta )i0 + t (1 )(( m Cp m + bio Cp bio + char Cp char )v s Ta ) i x i 1 Di = (1 )(( m Cp m + bio Cp bio + char Cp char )v s Ta ) i 1 x i 1
2 4 4 + (hsg Av Tg ) + (TW hsw )i + Rm H m + R p1 H p1 + (R j H j )i D j =1 2

C.4 Energy balance equation (gaseous phase)

As discussed in 3.9.3, underrelaxation is carried out for all source terms related to gas oxidation reactions as follows:

10 10 10 R j H j = ( ) R j H j + (1 ) R j H j j =6 i j =6 i j =6 i

256

Appendix C Discretization of model partial differential equations

where is the underrelaxation factor. = 0.3 is used in the computer program.

After underrelaxation is done, the gas phase energy balance differential equation is discretized as follows:
T ( g hg ) + (mg hg ) = k g g + hsg Av(Ts Tg ) x x t x 10 4h gw (Tg Tw ) + R p 2 H p 2 + R j H j D j =5

( )

(eq 3.62)

- Central difference scheme for second order terms - Backward difference scheme for first order terms where hg = Cp g (Tg Ta )

[( Cp (T T )) ( Cp (T T )) ]+ t 1 [(m Cp (T T )) (m Cp (T T )) ] x

0 g g g a
i

i 1

i+ 1

(T ) (T ) (T ) (T ) 1 (K g )i+ 12 g i+1 g i (K g )i 12 g i g i1 xi xi + 1 xi 1 2 2 4(hgw )i + (hsg Av (Ts Tg ) )i (Tg Tw ) i + (R p 2 H p 2 )i + (R5 F H 5 F )i D = + (R5 R H 5 R )i 10 10 + ( ) R j H j + (1 ) R j H j j =6 i j =6 i


*

257

Appendix C Discretization of model partial differential equations

( g Cp g Ta )i = t t t t (K ) + (K ) (K ) + (K ) (K ) + (K ) g i 1 g i +1 g i +1 g i g i g i (Tg ) (Tg )i 2x x 2x x (Tg )i +1 2x x i i i i 1 i+ 1 i+ 1 2 2 2 (K ) + (K ) (mg Cp g Tg )i (mg Cp g Ta )i (mg Cp g Tg )i1 g i 1 g i (Tg )i1 + + i + xi 1 xi 1 xi 1 2xi xi 1 2 2 2 2 i 1 4(hgwTg )i 4(hgwTw )i (mg Cp g Ta ) ( + hsg AvTs )i (hsg AvTg )i + + (R p 2 H p 2 )i xi 1 D D
0 0 g g g 2

Cp g Tg )i

Cp g Ta )i

Cp g Tg )i +

+ (R5 F H 5 F )i + (R5 R H 5 R )i

10 10 + ( ) R j H j + (1 ) R j H j j =6 i j =6 i

This can be rewritten in the form Ai (Tg )i = Bi (Tg )i +1 + C i (Tg )i 1 + Di (K ) + (K ) (K ) + (K ) (m Cp ) 4(hgw )i g i g i +1 g i 1 g g i g i + + (hsg Av ) + 2x x + 2x x + x D i i i+ 1 i 1 i 1 2 2 2

Ai =

( Cp ) t
g g

(K ) + (K ) g i g i +1 Bi = 2x x i i+ 1 2 (K ) + (K ) (m Cp ) g i g i 1 g g i 1 Ci = 2x x + x i i 1 i 1 2 2
Di =

g Cp g Ta ) i +

g Cp g T g )i 0

g Cp g Ta )i + 0

(m Cp T ) (m Cp T )
g g a
i

+ (hsg Av Ts )i +

4(hgw )i Tw D

xi 1

i 1

xi 1

+ (R p 2 H p 2 )i + (R5 F H 5 F )i + (R5 R H 5 R )i
*

10 10 + ( ) R j H j + (1 ) R j H j j =6 i j =6 i

Reference

Pantankar, SV 1980, Numerical heat transfer and fluid flow, New York: McGraw-Hill.

258

Appendix D Computer code listing

Appendix D

Computer code listing

Given below is the computer code for the model developed duing this project. It is written in the Fortran programming language. The Fortran code was developed by using ELF90 (Lahey, 1998) compiler under Windows NT 4.0 operating system.

It was tested on the following compiler and operating system pairs:

Essential Lahey Fortran 90 Compiler Release 4.00b and Windows NT 4.0 service pack 5, Compaq Visual Fortran 6.1 and Windows 2000 service pack 4, f90 and UNIX (IRIX).

!========================================================================================== ! START MODULE !========================================================================================== MODULE gasification_data ! Purpose: ! The main purpose of this module is for data sharing for Transient wood gasifier model. IMPLICIT NONE ! List of parameters: INTEGER, PARAMETER :: DP = KIND(1.0D0) ! ********************************** ! List of main user input parameters ! ********************************** ! Gasifier Dimensions ! =================== REAL(DP):: D_rec REAL(DP):: L_rec ! Fuel Properties ! =============== REAL(DP):: rho_biomass REAL(DP):: moisture REAL(DP):: d_pwood REAL(DP):: air_rate REAL(DP):: Pre REAL(DP):: T_a REAL(DP):: T_gas REAL(DP):: void ! Double precision (DP)

! Inner diameter of the reactor, [m] ! Effective length of the reactor, [m]

! ! ! ! ! ! ! !

Apparent density of biomass in dry basis, [kg/m3] Moisture content of the fuel in dry basic, [wt %] Initial wood particle diameter, [m] Air supply rate, [kg/hr] Pressure of gas mixture, [KPa] Ambient temperature, [K] Gas Inlet temperature, [K] Bed voidage fraction, [-]

259

Appendix D Computer code listing

REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP)::

T_ign depth_ign loc_ign depth_char secondair loc_secair

! ! ! ! ! !

Ignition temperature, [K] Depth of the ignition, [m] Location of the ignition from the reactor top, [m] Depth of char at initial condition, [-] Secondary air supply rate, [kg/hr] Location of secondary air from the reactor top, [m]

! Main Model parameters ! ===================== REAL(DP):: zeta REAL(DP):: km_max REAL(DP):: Av_mf REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: Ap1_mf Ap2_mf A6_mf keff_mf wall_loss

! ! ! ! ! ! ! ! ! !

Reacting to non-reacting solid-to-gas heat transfer coefficient ratio, [-] Defined maximum mass transfer coefficient, [m/s] Ratio to effective and actural heat and mass transfer surface area, [-] Multiplication factor of primary pyrolysis rate, [-] Multiplication factor of secondary pyrolysis rate, [-] Multiplication factor of secondary tar oxidation rate, [-] Multiplication factor to Effictive thermal condutivity, [-] Multiplication factor of wall heat transfer coefficient, [-]

! Ultimate Analysis of the fuel ! ============================= REAL(DP):: carbon = 52.20D0 REAL(DP):: hydrogen = 6.10D0 REAL(DP):: oxygen = 41.70D0 REAL(DP):: nitrogen = 0.00D0 REAL(DP):: sulfur = 0.00D0 REAL(DP):: ash = 0.00D0 ! Main operational parameters ! =========================== REAL(DP):: d_pinitial = 0.020D0 REAL(DP):: depth_bio

! ! ! ! ! !

Carbon content of the fuel in dry basic, [wt %] Hydrogen content of the fuel in dry basic, [wt %] Oxygen content of the fuel in dry basic, [wt %] Nitrogen content of the fuel in dry basic, [wt %] Sulfur content of the fuel in dry basic, [wt %] Ash content of the fuel in dry basic, [wt %]

! Initial char diameter, [m] ! Depth of biomass, [m]

! Defined Parameters for numerical soultion ! ========================================== INTEGER, parameter :: N = 700 ! Total number of slice, [-] INTEGER :: N_top = 200 ! Initial number of slices of top unreacted zone,[-] INTEGER :: N_ign = 50 ! Initial number of slices of ignition depth, [-] REAL(DP):: delta_x ! Depth of slice for central difference ! (x(i+1/2)-x(i-1/2)), [m] REAL(DP):: delta_x1 ! Depth of slice for backward difference ! (x(i)-x(i-1)), [m] REAL(DP):: delta_x2 ! Depth of slice for forward difference ! (x(i+1)-x(i)), [m] ! Main Model parameters ! ===================== REAL(DP):: Ep2_mf = REAL(DP):: P_alpha =

0.850D0 0.5D0

REAL(DP):: wall_coeff

0.0D0

! ! ! ! ! !

Multiplication factor of Activation energy of tar cracking rate, [-] Adjustable factor for effitive tar cracking temperature at pyrolysis zone, [-] 0 for nomrmal case, 1 for T_s=T_g Inital wall temperature estimation gradient, [-]

! **************************** ! List of reactions' constants ! **************************** ! Pre_exponential frequency factors (PEFF) ! ======================================== REAL(DP):: Am = 5.13D+06 ! PEFF REAL(DP):: Ap1 = 3.20D+05 ! PEFF REAL(DP):: Ap2 = 1.150D+04 ! PEFF REAL(DP):: A1 = 0.6581 ! PEFF REAL(DP):: A2 = 589.0D+0 ! PEFF REAL(DP):: A3 = 1.67*3.42D+0 ! PEFF REAL(DP):: A4 = 3.42D-03 ! PEFF REAL(DP):: A5 = 4.34D-04 ! PEFF REAL(DP):: A6 = 6.858D+03 ! PEFF REAL(DP):: A7 = 1.88D+09 ! PEFF REAL(DP):: A8 = 2.09D+10 ! PEFF REAL(DP):: A9 = 6.858D+03 ! PEFF REAL(DP):: A10 = 8.40D+06 ! PEFF REAL(DP)::A_gamma = 4.3D0 ! PEFF

of drying, [1/s] of primary pyrolysis, [1/s] of secondary pyrolysis, [kmolK/m3kPaSec] of char oxidation, [m/sK] of C-CO2 reaction, [m/sK] of C-H2O reaction, [m/sK] of C_H2 reaction, [m/sK] of H2O_gas shift reaction, [kmolK2/m3kPa2sec] of tar oxidation, [kmolK0.5/mkPa1.8Sec] of CO oxidation, [kmolK2/m3kPa2Sec] of CH4 oxidation, [kmolK1.5/kN1.5Sec] of C2H4 oxidation, [kmolK0.5/mkPa1.8Sec] of H2 oxidation, [kmolK2/m3kPa2Sec] for (CO/CO2) in char oxidation, [-]

260

Appendix D Computer code listing

! Activation energy ! ================= REAL(DP):: Em = REAL(DP):: Ep1 = REAL(DP):: Ep2 = REAL(DP):: E1 = REAL(DP):: E2 = REAL(DP):: E3 = REAL(DP):: E4 = REAL(DP):: E5 = REAL(DP):: E6 = REAL(DP):: E7 = REAL(DP):: E8 = REAL(DP):: E9 = REAL(DP):: E10 = REAL(DP)::E_gamma =

87900.0D0 113000.0D0 93300.0D0 74830.0D0 222825.0D0 129700.0D0 129700.0D0 60279.0D0 80230.0D0 125580.0D0 202600.0D0 80230.0D0 30514.0D0 28185.D0

! ! ! ! ! ! ! ! ! ! ! ! ! ! !

Activation energy of drying, [J/mol] Activation energy of primary pyrolysis, [J/mol] Activation energy of secondary pyrolysis, [J/mol] Activation energy of char oxidation, [J/mol] Activation energy of C-CO2 reaction, [J/mol] Activation energy of C-H2O reaction, [J/mol] Activation energy of C_H2 reaction, [J/mol] Activation energy of shift reaction, [J/mol] Activation energy of tar oxidation, [J/mol] Activation energy of CO oxidation, [J/mol] Activation energy of CH4 oxidation, [J/mol] Activation energy of C2H4 oxidation, [J/mol] Activation energy of H2 oxidation, [J/mol] Activation energy for (CO/CO2) in char oxidation, [J/mol]

! Heat of reactions ! ================= REAL(DP):: heatp1 = REAL(DP):: heatp2 = REAL(DP):: heatm =

0.D0 50.0D+03 -2257.0D+03

! Heat of reaction for reaction p1, [J/kg] ! Heat of reaction for reaction p2, [J/kg] ! Heat of drying, [J/kg]

! ************************************** ! List of the yield of pyrolysis process ! ************************************** ! Yield of primary pyrolysis ! ========================== REAL(DP):: X1_char = 0.205D0 REAL(DP):: X1_tar1 = 0.528D0 REAL(DP):: X1_CO = 0.032D0 REAL(DP):: X1_CO2 = 0.068D0 REAL(DP):: X1_CH4 = 0.004D0 REAL(DP):: X1_H2O = 0.163D0 REAL(DP):: X1_gas = 0.267D0 REAL(DP):: X1_volt = 0.795D0 ! Yield of secondary pyrolysis ! ============================ REAL(DP):: X2_tar2 = 0.0665D0 REAL(DP):: X2_CO = 0.2762D0 REAL(DP):: X2_CO2 = 0.064D0 REAL(DP):: X2_CH4 = 0.051D0 REAL(DP):: X2_H2 = 0.0163D0 REAL(DP):: X2_C2H4 = 0.054D0

! ! ! ! ! ! ! !

Mass Mass Mass Mass Mass Mass Mass Mass

Fraction Fraction Fraction Fraction Fraction Fraction Fraction Fraction

of of of of of of of of

Char yield of primary pyrolysis,[-] Tar yield of primary pyrolysis, [-] CO yield of primary pyrolysis, [-] CO2 yield of primary pyrolysis, [-] CH4 yield of primary pyrolysis, [-] H2O yield of primary pyrolysis, [-] Gas yield of primary pyrolysis, [-] volatile yield of pri pyrolysis,[-]

! ! ! ! ! !

Mass Mass Mass Mass Mass Mass

Fraction Fraction Fraction Fraction Fraction Fraction

of of of of of of

Tar yield of secondary pyrolysis, [-] CO yield of secondary pyrolysis, [-] CO2 yield of secondary pyrolysis, [-] CH4 yield of secondary pyrolysis, [-] H2O yield of secondary pyrolysis, [-] C2H4 yield of secondary pyrolysis,[-]

! **************************************** ! Constants values and physical properties ! **************************************** REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: pi R mole_C mole_O2 mole_CO2 mole_CO mole_H2O mole_H2 mole_CH4 mole_C2H4 mole_N2 mole_tar1 mole_tar2 mf_O2 mass_O2 sigma emiss Cp_m DV_O2 = DV_N2 = = 3.141592654D0 = 8.3144D0 = 12.0D0 = 32.0D0 = 44.0D0 = 28.0D0 = 18.0D0 = 2.0D0 = 16.0D0 = 28.0D0 = 28.0D0 = 134.934D0 = 78.0D0 = 0.21D0 = 0.233D0 = 5.6703D-08 = 0.85D0 = 4200.0D0 12.22D0 9.08D0 ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! Value of pi Universal gas constant, [J/mol K] Molecular weight of Char,[g/mol] Molecular weight of O2, [g/mol] Molecular weight of CO2, [g/mol] Molecular weight of CO, [g/mol] Molecular weight of H2O, [g/mol] Molecular weight of H2, [g/mol] Molecular weight of CH4, [g/mol] Molecular weight of C2H4, [g/mol] Molecular weight of N2, [g/mol] Molecular weight of primary tar, [g/mol] Molecular weight of secondary tar, [g/mol] Molar fraction of oxygen in air, [-] Mass fraction of oxygen in air, [-] Stefan-Boltzmann constant, [W/m2K4] Emissivity of char, [-] Specific heat of moisture,[J/kgK] Diffusion volumes of O2 Diffusion volumes of N2

261

Appendix D Computer code listing

REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP)::

DV_CO DV_CO2 DV_H2 DV_H2O DV_CH4 DV_C2H4 DV_tar1 DV_tar2

= 22.01D0 = 28.12D0 = 4.62D0 = 10.73D0 = 25.14D0 = 41.04D0 = 140.08314D0 = 109.26D0

! ! ! ! ! ! ! !

Diffusion Diffusion Diffusion Diffusion Diffusion Diffusion Diffusion Diffusion

volumes volumes volumes volumes volumes volumes volumes volumes

of of of of of of of of

CO CO2 H2 H2O CH4 C2H4 tar1 tar2

! ************************************ ! Calculation Parameters and variables ! ************************************ INTEGER :: INTEGER :: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: INTEGER :: INTEGER :: INTEGER :: i,iteration logic = time delta_t = alpha = error = sec_check = output_time check_stage ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! Index Check temperature index 0 for exit Gasifier operating time, [s] Time step, [s] Under relexation coefficent, [-] Maximum allownce error for covergence, [-] Check secondary air supply 0 for skip Time when output results are print out Check sigle stage of two stage (1 for single-stage, 2 for two-stage) Check for sensitivity performance (0 for w/o sensitivity analysis, 1 for with sensitivity anlysis) Selcect the parameter for sensitivity performance Location of the grids, [m] Mass Concentration of moisture, [kg/m] Mass Concentration of volatile, [kg/m] Mass Concentration of char, [kg/m] Mass Concentration of gas, [kg/m] Molecular weight of gas mixture, [g/mol] Molar Fraction of O2, [-] Molar Fraction of CO2, [-] Molar Fraction of CO, [-] Molar Fraction of H2O, [-] Molar Fraction of H2, [-] Molar Fraction of CH4, [-] Molar Fraction of N2, [-] Molar Fraction of C2H4, [-] Molar Fraction of primary tar, [-] Molar Fraction of secondary tar, [-] Mass flux of total gas, [kg/ms] Temperature of solid surface, [K] Temperature of gas, [K] Superfical velocity of gas, [m/s] Solid velocity, [m/s] Particle diameter, [m] Wall temperature, [K] Average wood temperaure for pyrolysis process, [K] Equivalent temperature for tar cracking, [K] Rate of drying, [kg/m/s] Reaction rate of primary pyrolysis, [kg/m/s] Reaction rate of secondary pyrolysis, [kg/m/s] Reaction rate of char oxidation, [kg/m/s] Reaction rate of C-CO2 reaction, [kg/m/s] Reaction rate of C-H2O reaction, [kg/m/s] Reaction rate of C_H2 reaction, [kg/m/s] Rate of forward shift reaction, [kmol/m/s] Rate of reverse shift reaction, [kmol/m/s] Reaction rate of tar oxidation, [kmol/m/s] Reaction rate of CO oxidation, [kmol/m/s] Reaction rate of CH4 oxidation, [kmol/m/s] Reaction rate of C2H4 oxidation, [kmol/m/s] Reaction rate of H2 oxidation, [kmol/m/s] Diffusion coefficient between O2 and gas mixture (m2/s) Diffusion coefficient between CO2 and gas mixture (m2/s) Diffusion coefficient between H2 and gas mixture (m2/s) Diffusion coefficient between

0 3.0D0 0.3D0 1.0D-04 0

INTEGER :: check_sensitivity

INTEGER :: check_analysis REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), REAL(DP), DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: DIMENSION(N):: x rho_m rho_volt rho_char rho_g mole_g Y_O2 Y_CO2 Y_CO Y_H2O Y_H2 Y_CH4 Y_N2 Y_C2H4 Y_tar1 Y_tar2 m_g T_s T_g v_g v_s d_p T_wall T_e T_p2 ratem rateP1 rateP2 rate1 rate2 rate3 rate4 rate5 rate5R rate6 rate7 rate8 rate9 rate10

REAL(DP):: diff_O2_gas REAL(DP):: diff_CO2_gas REAL(DP):: diff_H2_gas REAL(DP):: diff_H2O_gas

262

Appendix D Computer code listing

REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP)::

A_O2,B_O2,C_O2 A_CO,B_CO,C_CO A_CO2,B_CO2,C_CO2 A_H2O,B_H2O,C_H2O A_H2,B_H2,C_H2 A_CH4,B_CH4,C_CH4 A_C2H4,B_C2H4,C_C2H4 A_tar1,B_tar1,C_tar1 A_tar2,B_tar2,C_tar2 A_gas,B_gas,C_gas A_Vs,B_Vs,C_Vs A_rhom,B_rhom,C_rhom A_rhov,B_rhov,C_rhov A_dp,B_dp,C_dp

! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! !

H2O and gas mixture (m2/s) Variables used in O2 balance term Variables used in CO balance term Variables used in CO2 balance term Variables used in H2O balance term Variables used in H2 balance term Variables used in CH4 balance term Variables used in C2H4 balance term Variables used in primary tar balance term Variables used in primary tar balance term Variables used in overall gas mixture balance term Variables used in solid velocity calculation Variables used in moisture density calculation Variables used in volatile density calculation Variables used in particle diameter calculation Molar flux of gas mixture (kmol/m2sec) Ratem that used in moisture content balance Ratep1 that used in volatile matter balance Ratep1 that used in primary tar balance Rate1 that used in O2 balance Rate2 that used in CO2 balance Rate3 that used in H2O balance Rate4 that used in H2 balance Rate5 that used in CO balance Rate5 that used in H2O balance Reverse rate5 that used in CO2 balance Reverse rate5 that used in H2 balance Rate6 that used in O2 balance Rate7 that used in CO balance Rate9 that used in O2 balance The stoichiometric coefficient to identify the moles of oxidant per mole of Carbon for reaction 1 Inverse of Y_O2 Inverse of Y_tar2 Inverse of C2H4 Inverse of CH4 Inverse of CH4 Renolds number, [-] Prandtl number, [-] Schmidt number for reaction1, [-] Schmidt number for reaction2, [-] Schmidt number for reaction3, [-] Schmidt number for reaction4, [-] Film mass transfer coefficient of Film mass transfer coefficient of Film mass transfer coefficient of Film mass transfer coefficient of

REAL(DP):: molar_g REAL(DP), DIMENSION(N):: REAL(DP), DIMENSION(N):: REAL(DP), DIMENSION(N):: REAL(DP), DIMENSION(N):: REAL(DP), DIMENSION(N):: REAL(DP), DIMENSION(N):: REAL(DP), DIMENSION(N):: REAL(DP), DIMENSION(N):: REAL(DP), DIMENSION(N):: REAL(DP), DIMENSION(N):: REAL(DP), DIMENSION(N):: REAL(DP), DIMENSION(N):: REAL(DP), DIMENSION(N):: REAL(DP), DIMENSION(N):: REAL(DP):: gamma REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: Y_O2inverse Y_tar2inverse Y_C2H4inverse Y_CH4inverse Y_H2inverse Re Pr Sc_1 Sc_2 Sc_3 Sc_4 km_1 km_2 km_3 km_4 A_v h_sg k_ratio k_rs k_rg k_ro h_bw h_gw h_sw

ratem_m ratep1_volt ratep2_tar1 rate1_O2 rate2_CO2 rate3_H2O rate4_H2 rate5_CO rate5_H2O rate5R_CO2 rate5R_H2 rate6_O2 rate7_CO rate9_O2

reaction1, reaction2, reaction3, reaction4,

[m/s] [m/s] [m/s] [m/s]

Particle surface area/ Unit volume, [1/m] Solid to gas heat transfer coefficient, [W/mK] Ratio of solid and gas conductivity, [-] Solid effective radial conductivity, [W/mK] Gas effective radial conductivity, [W/mK] Static effective radial conductivity, [W/mK] Bed to wall heat transfer coefficient, [W/mK] Gas to wall heat transfer coefficient, [W/mK] Solid to wall heat transfer coefficient, [W/mK] Lower heating value of producer gas, [MJ/Nm3 d.b] Total energy output from the reactor, [MJ/hr] Air fuel (biomass) ratio, [kg/kgd.b] Equivalence ratio, [-] Fuel consumption, [kgw.b/hr] Gas yield in mass, [kg/hr] Gas yield in volume, [Nm3/hr] Specific gasification rate, [kg/m2sec] Condensable tar outlet, [mg/Nm3] Cold gas efficiency of the gasifier, [%]

REAL(DP):: LHV_gas REAL(DP):: Energy_out REAL(DP):: air_fuel REAL(DP):: ER REAL(DP):: fuel_flow REAL(DP):: gas_flow_mass REAL(DP):: gas_flow_vol REAL(DP):: spe_gasi_rate REAL(DP):: tar_outlet REAL(DP), DIMENSION(N):: efficiency REAL(DP):: C_in,H_in,O_in,N_in REAL(DP):: C_out, H_out,O_out,N_out REAL(DP):: masstotal_in

! C,H,O,N flux input to the reactor, [kg/m2/s] ! C,H,O,N flux output to the reactor, [kg/m2/s] ! Total mass flux input to the reactor, [kg/m2/s]

263

Appendix D Computer code listing

REAL(DP):: masstotal_out

! Total mass flux output from the reactor, [kg/m2/s]

! Variable used for temperature Calculation ! --------------------------------REAL(DP):: T_701 REAL(DP):: solid1, solid2, solid3, solid4, solid5 REAL(DP):: A_Ts, B_Ts, C_Ts, D_Ts REAL(DP):: gas1, gas2, gas3, gas4, gas5 REAL(DP):: A_Tg, B_Tg, C_Tg, D_Tg REAL(DP), DIMENSION(N):: P_Ts, Q_Ts REAL(DP), DIMENSION(N):: T_sold, T_gold REAL(DP), DIMENSION(N):: P_Tg, Q_Tg,gasheat=0. ! --------------------------------! Value from the previous time step ! --------------------------------REAL(DP), DIMENSION(N):: YP_O2 REAL(DP), DIMENSION(N):: YP_CO2 REAL(DP), DIMENSION(N):: YP_CO REAL(DP), DIMENSION(N):: YP_H2O REAL(DP), DIMENSION(N):: YP_H2 REAL(DP), DIMENSION(N):: YP_CH4 REAL(DP), DIMENSION(N):: YP_N2 REAL(DP), DIMENSION(N):: YP_C2H4 REAL(DP), DIMENSION(N):: YP_tar1 REAL(DP), DIMENSION(N):: YP_tar2 REAL(DP), DIMENSION(N):: rhoP_m REAL(DP), DIMENSION(N):: rhoP_volt REAL(DP), DIMENSION(N):: rhoP_g REAL(DP), DIMENSION(N):: mP_g REAL(DP), DIMENSION(N):: TP_s REAL(DP), DIMENSION(N):: TP_g REAL(DP), DIMENSION(N):: moleP_g REAL(DP), DIMENSION(N):: dP_p REAL(DP):: molarP_g

! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! !

Molar Fraction of O2, [-] Molar Fraction of CO2, [-] Molar Fraction of CO, [-] Molar Fraction of H2O, [-] Molar Fraction of H2, [-] Molar Fraction of CH4, [-] Molar Fraction of N2, [-] Molar Fraction of C2H4, [-] Molar Fraction of primary tar, [-] Molar Fraction of secondary tar, [-] Mass Concentration of moisture , [kg/m] Mass Concentration of volatile, [kg/m] Mass Concentration of gas mixture, [kg/m] Mass flux of total gas, [kg/ms] Temperature of solid surface, [K] Temperature of gas, [K] Molecular weight of gas mixture, [g/mol] Particle diameter, [m] Molar flux of gas mixture, [kmol/m2sec]

! --------------------------------------------! Variable for calculating program running time ! --------------------------------------------CHARACTER (len=10) :: date, t, zone INTEGER, DIMENSION(8):: dt_init, dt_final REAL :: run_time ! ---------------------------------! Variables used for adaptive method ! ---------------------------------REAL(DP):: Max_Tg ! Maximum gas temperature REAL(DP), DIMENSION(N):: Tg_nd ! Dimensionless gas temperature REAL(DP), DIMENSION(N):: x_nd ! Dimensionless grids distance REAL(DP), DIMENSION(N):: xnew_nd ! Dimensionless new grids distance REAL(DP), DIMENSION(N):: xnew ! New grids distance, [m] REAL(DP), DIMENSION(N):: rhonew_m ! Density of moisture at new grids, [kg/m] REAL(DP), DIMENSION(N):: rhonew_volt ! Density of volatile at new grids, [kg/m] REAL(DP), DIMENSION(N):: rhonew_char ! Density of char at new grids, [kg/m] REAL(DP), DIMENSION(N):: Ynew_O2 ! Molar of O2v_g at new grids, [-] REAL(DP), DIMENSION(N):: Ynew_CO2 ! Molar of CO2 at new grids, [-] REAL(DP), DIMENSION(N):: Ynew_CO ! Molar of CO at new grids, [-] REAL(DP), DIMENSION(N):: Ynew_H2O ! Molar of H2O at new grids, [-] REAL(DP), DIMENSION(N):: Ynew_H2 ! Molar of H2 at new grids, [-] REAL(DP), DIMENSION(N):: Ynew_CH4 ! Molar of CH4 at new grids, [-] REAL(DP), DIMENSION(N):: Ynew_N2 ! Molar of N2 at new grids, [-] REAL(DP), DIMENSION(N):: Ynew_C2H4 ! Molar of C2H4 at new grids, [-] REAL(DP), DIMENSION(N):: Ynew_tar1 ! Molar of primary tar at new grids, [-] REAL(DP), DIMENSION(N):: Ynew_tar2 ! Molar of secondary tar at new grids, [-] REAL(DP), DIMENSION(N):: mnew_g ! Mass flux of total gas at new grids, [kg/ms] REAL(DP), DIMENSION(N):: Tnew_s ! Temperature of solid surface at new grids, [K] REAL(DP), DIMENSION(N):: Tnew_g ! Temperature of gas at new grids, [K] REAL(DP), DIMENSION(N):: vnew_s ! Solid velocity at new grids [m/s] REAL(DP), DIMENSION(N):: dnew_p ! Particle diameter at new grids [m] ! -----------------------------------------! Variables used for reaction front movement ! -----------------------------------------REAL(DP)::maxTg_x ! peak gas temp location at time t [K] REAL(DP)::maxTg_x05 ! peak gas temp location at 5 min [K] REAL(DP)::maxTg_x10 ! peak gas temp location at 10 min [K]

264

Appendix D Computer code listing

REAL(DP)::maxTg_x15 REAL(DP)::maxTg_x20 REAL(DP)::maxTg_x25 REAL(DP)::maxTg_x30 INTEGER:: ierror CHARACTER(len=30) :: filename CONTAINS

! ! ! !

peak peak peak peak

gas gas gas gas

temp temp temp temp

location location location location

at at at at

15 20 25 30

min min min min

[K] [K] [K] [K]

! Status flag from I/O statements: 0 for success ! Name of file to open

! ***************************************************************************************** SUBROUTINE user_input() ! This subroutine is for user to enter the input parameter. Write(*,*)" Write(*,*)" Write(*,*)" Write(*,*)" Write(*,*)" Write(*,*)" Write(*,*)" ------------------------------------------------------------------------------" TRANSIENT MODEL FOR STRATIFIED DOWNDRAFT WOOD GASIFIER " ------------------------------------------------------------------------------" This program is a transient model of a stratified downdraft gasifier. Model is" able to predict the transient behivour of reactor namely, solid and gas phase" temperarue profile, velocity profile, reaction rates profile, gas composition" profile and condensable tar profile along the axis of gasifer."

Write(*,'(/T20, A35)') " -------------------------------- " Write(*,'(T20, A35)') " User defined gasifier Dimensions " Write(*,'(T20, A35)') " -------------------------------- " Write(*,*) "Enter the inner diameter of the reactor in metre (for base case = 0.206)" Read(*,*) D_rec Write(*,*) "Enter the effective length of the reactor in metre (for base case = 0.7)" Read(*,*) L_rec Write(*,'(/T20, A33)') " ---------------------------- " Write(*,'(T20, A33)') " User defined fuel properties " Write(*,'(T20, A33)') " ---------------------------- " Write(*,*) "Enter the apparent density of biomass in [kg d.b/m3] (for base case = 950)" Read(*,*) rho_biomass Write(*,*) "Enter the moisture content of the fuel in [wt d.b %] (for base case = 10)" Read(*,*) moisture Write(*,*) "Enter the initial wood particle diameter in metre (for base case = 0.02)" Read(*,*) d_pwood Write(*,*) "Enter the bed void fraction (for base case = 0.46) " Read(*,*) void Write(*,'(/T18, A40)') " ----------------------------------- " Write(*,'(T18, A40)') " User defined operational parameters " Write(*,'(T18, A40)') " ----------------------------------- " Write(*,*) "Model is able to run for single-stage and two-stage" DO Write(*,*) "(Enter '1' for signle stage or '2' for two-stage)" Read(*,*) check_stage IF (check_stage==1) THEN Write(*,*) "Enter the air supply rate, [kg/hr] (for base case = 15)" Read(*,*) air_rate secondair = 0.D0 loc_secair = 0.D0 ELSE IF (check_stage==2) THEN Write(*,*) "Enter the primary air supply rate, [kg/hr] (for base case = 15)" Read(*,*) air_rate Write(*,*) "Enter the secondary air supply rate, [kg/hr]" Write(*,*) "(in the range of 1-5 kg/hr, for base case = 5 kg/hr)" Read(*,*) secondair Write(*,*)"Enter the location of the secondary air supply from the reactor top in (m)" Write(*,*) "(it must be between less than 0.5, for base case = 0.4 m)" Read(*,*) loc_secair ELSE Write(*,*) "Error in enter value, it must be either 1 or 2" END IF IF(check_stage==1.or.check_stage==2)exit END DO Write(*,*) "Enter the pressure of inlet air, [KPa] (atmosperic condtion = 101.325)" Read(*,*) Pre Write(*,*) "Enter the ambient temperature in K (for normal condtion = 300)"

265

Appendix D Computer code listing

Read(*,*) T_a Write(*,*) "Enter the inlet air temperarue in K (for base case = 300)" Read(*,*) T_gas Write(*,*) "Enter the ignition temperature in K" Write(*,*) "(in the range of 900-1200 K, for base case = 1000)" Read(*,*) T_ign Write(*,*) "Enter the initial depth of the ignition in meter (for base case= 0.05)" Read(*,*) depth_ign Write(*,*) "Enter the location of the ignition from the reactor top in meter " Write(*,*) "(for base case = 0.2)" Read(*,*) loc_ign Write(*,*) "Enter the depth of char bed at initial condition in meter" Write(*,*) "(for base case = 0.5)" Read(*,*) depth_char Write(*,*) "Enter the time to write the output result in second" Write(*,*) "It can be divided by 3 (e.g, 300, 999, 1200) (base case for 1800)" Read(*,*) output_time DO Write(*,'(/T2, A)') "Enter '1' for sensitivity analysis of model parameter" Write(*,'(T2, A/)') "Enter '0' for to run the program without sensitivity analysis" Read(*,*) check_sensitivity ! Defined the model parameter for base run zeta = 0.60D0 ! Reacting to non-reacting heat transfer coefficient ratio [-] km_max = 0.045D0 ! Defined maximum mass transfer coefficient, [m/s] Av_mf = 1.0D0 ! Ratio to effective and actural transport exchanged area, [-] Ap1_mf = 1.00D0 ! Multiplication factor of primary pyrolysis rate, [-] Ap2_mf = 1.0D0 ! Multiplication factor of secondary pyrolysis rate, [-] A6_mf = 1.0D0 ! Multiplication factor of secondary tar oxidation rate, [-] keff_mf = 1.0D0 ! Multiplication factor to effictive thermal condutivity, [-] wall_loss = 1.0D0 ! Multiplication factor of wall heat transfer coefficient, [-] IF (check_sensitivity==0) THEN exit ELSE IF (check_sensitivity==1) THEN DO Write(*,'(/T2, A)')"Effect of eight model parameter can be examined by changing the" Write(*,'(T2, A)') "single parameter while all other parameter are kept constant." Write(*,'(/T2, A)')"Effect of which parameter do you want to analyse?" Write(*,'(/T2, A)')"Enter '1' to analyse the reacting to non-reacting" Write(*,'(T12, A)')"solid-to-gas heat transfer coefficient ratio" Write(*,'(T2, A)') "Enter '2' to analyse defined maximum mass transfer coefficient" Write(*,'(T2, A)') "Enter '3' to analyse the ratio to effective and actural" Write(*,'(T12, A)')"heat and mass transfer surface area" Write(*,'(T2, A)') "Enter '4' to analyse the primary pyrolysis rate" Write(*,'(T2, A)') "Enter '5' to analyse the secondary pyrolysis rate" Write(*,'(T2, A)') "Enter '6' to analyse the secondary tar oxidation rate" Write(*,'(T2, A)') "Enter '7' to analyse the effictive thermal condutivity" Write(*,'(T2, A/)')"Enter '8' to analyse the wall heat transfer coefficient" Read(*,*) check_analysis IF (check_analysis==1) THEN Write(*,*) "Enter the reacting to non-reacting heat transfer coefficient ratio" Write(*,*) " it must be less than '1', (for base case = 0.6)" read(*,*)zeta exit ELSE IF (check_analysis==2) THEN Write(*,*) "Enter the defined maximum mass transfer coefficient" Write(*,*) " it must be in a range of 0.04-0.06 (for base case = 0.045)" read(*,*)km_max exit ELSE IF (check_analysis==3) THEN Write(*,*) "Enter the ratio to effective and actural transport exchanged area" Write(*,*) " it should be in a range of 0.8-1.2 (for base case = 1)" read(*,*)Av_mf exit ELSE IF (check_analysis==4) THEN Write(*,*) "Enter the multiplication factor of primary pyrolysis rate" Write(*,*) " it should be in a range of 0.5-2 (for base case = 1)" read(*,*)Ap1_mf exit ELSE IF (check_analysis==5) THEN Write(*,*) "Enter the multiplication factor of secondary pyrolysis rate" Write(*,*) " it should be in a range of 0.7-1.3 (for base case = 1)" read(*,*)Ap2_mf exit

266

Appendix D Computer code listing

ELSE IF (check_analysis==6) THEN Write(*,*) "Enter the multiplication Write(*,*) " it should be in a range read(*,*)A6_mf exit ELSE IF (check_analysis==7) THEN Write(*,*) "Enter the multiplication Write(*,*) " it should be in a range read(*,*)keff_mf exit ELSE IF (check_analysis==8) THEN Write(*,*) "Enter the multiplication Write(*,*) " it should be in a range read(*,*)wall_loss exit ELSE Write(*,*) "Error in enter value, it END IF END DO exit ELSE Write(*,*) "Error in enter value, it must END IF END DO

factor of secondary tar oxidation rate" of 0.5-3 (for base case = 1)"

factor to Effictive thermal condutivity" of 0.5-3 (for base case = 1)"

factor of wall heat transfer coefficient" of 0.5-1.5 (for base case = 1)"

must be any interger value from '1' to '8'"

be either '0' or '1'"

Write(*,'(T2, A)') "Enter result output file name: It must be less than 30 characters." Write(*,'(T2, A)') "Please aware that your file name must be new in the folder.If your" Write(*,'(T2, A/)')"defined file name already exist in the folder, it will be replaced." READ (*,*) filename RETURN END SUBROUTINE user_input ! *********************END SUBROUTINE user_input******************************************* FUNCTION ln (value) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::value REAL(DP)::ln ! Calculate ln value ln=LOG(value)/LOG(2.7182818281D0) RETURN END FUNCTION ln ! ****************************End FUNCTION ln********************************************** FUNCTION h_C (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp REAL(DP)::h_C

! value ! ln value

! Gas Temperature [K] ! Absolute enthalpy of C (Graphite) at temperature T [kJ/kmol]

! Calculate absolute enthalpy of O2 at temperature T IF (Temp>1000.) THEN h_C=(1.23870304D+01*Temp)+(6.91227177D-03*Temp**2.)-(1.85438496D-06*Temp**3.)+& (2.68524262D-10*Temp**4.)-(1.53215420D-14*Temp**5.)-(5.88033398D+03) ELSE h_C=-(5.59755420D+00*Temp)+(2.98573864D-02*Temp**2.)-(1.54884907D-05*Temp**3.)+& (4.26136365D-09*Temp**4.)-(6.08974359D-13*Temp**5.)-(6.05750000D+02) END IF RETURN END FUNCTION h_C ! *********************** END FUNCTION h_C ************************************************ FUNCTION h_O2 (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp REAL(DP)::h_O2

! Gas Temperature [K] ! Absolute enthalpy of O2 at temperature T [kJ/kmol]

267

Appendix D Computer code listing

! Calculate absolute enthalpy of O2 at temperature T IF (Temp>1000.) THEN h_O2= R*((0.03697578D+02*Temp)+(0.06135197D-02/2.D0*Temp**2.)-& (0.1258842D-06/3.D0*Temp**3.)+(0.01775281D-09/4.D0*Temp**4.)-& (0.11364354D-14/5.D0*Temp**5.)-(0.12339301D+04)) ELSE h_O2= R*((0.03212936D+02*Temp)+(0.11274864D-02/2.D0*Temp**2.)-& (0.05756150D-05/3.D0*Temp**3.)+(0.13138773D-08/4.D0*Temp**4.)-& (0.08768554D-11/5.D0*Temp**5.)-(0.1005249D+04)) END IF RETURN END FUNCTION h_O2 ! *********************** END FUNCTION h_O2 *********************************************** FUNCTION h_CO2 (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp REAL(DP)::h_CO2

! Gas Temperature [K] ! Absolute enthalpy of CO2 at temperature T [kJ/kmol]

! Calculate absolute enthalpy of CO2 at temperature T IF (Temp>1000.) THEN h_CO2= R*((0.04453623D+02*Temp)+(0.03140168D-01/2.D0*Temp**2.)-& (0.12784105D-05/3.D0*Temp**3.)+(0.02393996D-08/4.D0*Temp**4.)-& (0.16690333D-13/5.D0*Temp**5.)-(0.04896696D+06)) ELSE h_CO2= R*((0.02275724D+02*Temp)+(0.09922072D-01/2.D0*Temp**2.)-& (0.10409113D-04/3.D0*Temp**3.)+(0.06866686D-07/4.D0*Temp**4.)-& (0.0211728D-10/5.D0*Temp**5.)-(0.04837314D+06)) END IF RETURN END FUNCTION h_CO2 ! *********************** END FUNCTION h_CO2 ********************************************** FUNCTION h_CO (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp REAL(DP)::h_CO

! Gas Temperature [K] ! Absolute enthalpy of CO at temperature T [kJ/kmol]

! Calculate absolute enthalpy of CO at temperature T IF (Temp>1000.) THEN h_CO= R*((0.03025078D+02*Temp)+(0.14426885D-02/2.D0*Temp**2.)-& (0.05630827D-05/3.D0*Temp**3.)+(0.10185813D-09/4.D0*Temp**4.)-& (0.06910951D-13/5.D0*Temp**5.)-(0.1426835D+05)) ELSE h_CO= R*((0.03262451D+02*Temp)+(0.15119409D-02/2.D0*Temp**2.)-& (0.03881755D-04/3.D0*Temp**3.)+(0.05581944D-07/4.D0*Temp**4.)-& (0.02474951D-10/5.D0*Temp**5.)-(0.14310539D+05)) END IF RETURN END FUNCTION h_CO ! *********************** END FUNCTION h_CO *********************************************** FUNCTION h_H2 (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp REAL(DP)::h_H2

! Gas Temperature [K] ! Absolute enthalpy of H2 at temperature T [kJ/kmol]

! Calculate absolute enthalpy of H2 at temperature T IF (Temp>1000.) THEN h_H2= R*((0.02991423D+02*Temp)+(0.07000644D-02/2.D0*Temp**2.)-& (0.05633828D-06/3.D0*Temp**3.)-(0.09231578D-10/4.D0*Temp**4.)+& (0.15827519D-14/5.D0*Temp**5.)-(0.0835034D+04)) ELSE h_H2= R*((0.03298124D+02*Temp)+(0.08249441D-02/2.D0*Temp**2.)-& (0.08143015D-05/3.D0*Temp**3.)-(0.09475434D-09/4.D0*Temp**4.)+&

268

Appendix D Computer code listing

(0.04134872D-11/5.D0*Temp**5.)-(0.10125209D+04)) END IF RETURN END FUNCTION h_H2 ! *********************** END FUNCTION h_H2 *********************************************** FUNCTION h_H2O (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp REAL(DP)::h_H2O

! Gas Temperature [K] ! Absolute enthalpy of H2O at temperature T [kJ/kmol]

! Calculate absolute enthalpy of H2O at temperature T IF (Temp>1000.) THEN h_H2O= R*((0.02672145D+02*Temp)+(0.03056293D-01/2.D0*Temp**2.)-& (0.08730260D-05/3.D0*Temp**3.)+(0.12009964D-09/4.D0*Temp**4.)-& (0.06391618D-13/5.D0*Temp**5.)-(0.02989921D+06)) ELSE h_H2O= R*((0.03386842D+02*Temp)+(0.03474982D-01/2.D0*Temp**2.)-& (0.06354696D-04/3.D0*Temp**3.)+(0.06968581D-07/4.D0*Temp**4.)-& (0.02506588D-10/5.D0*Temp**5.)-(0.03020811D+06)) END IF RETURN END FUNCTION h_H2O ! *********************** END FUNCTION h_H2O ********************************************** FUNCTION h_CH4 (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp REAL(DP)::h_CH4

! Gas Temperature [K] ! Absolute enthalpy of CH4 at temperature T [kJ/kmol]

! Calculate absolute enthalpy of CH4 at temperature T h_CH4=(R*(((1.702D0)*(Temp-298.15D0))+((9.081D-03/2.D0)*(Temp**2.-298.15D0**2.))- & ((2.164D-06/3.D0)*(Temp**3.-298.15D0**3.))))-74831.D0 RETURN END FUNCTION h_CH4 ! *********************** END FUNCTION h_CH4 ********************************************** FUNCTION h_C2H4 (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp REAL(DP)::h_C2H4

! Gas Temperature [K] ! Absolute enthalpy of C2H4 at temperature T [kJ/kmol]

! Calculate absolute enthalpy of C2H4 at temperature T h_C2H4=(R*(((1.424D0)*(Temp-298.15D0))+((14.394D-03/2.D0)*(Temp**2.-298.15D0**2.))- & ((4.392D-06/3.D0)*(Temp**3.-298.15D0**3.))))+52283.D0 RETURN END FUNCTION h_C2H4 ! *********************** END FUNCTION h_C2H4 ********************************************* FUNCTION h_C6H6 (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp REAL(DP)::h_C6H6

! Gas Temperature [K] ! Absolute enthalpy of C6H6 at temperature T [kJ/kmol]

! Calculate absolute enthalpy of C6H6 at temperature T h_C6H6=(R*(-((0.206D0)*(Temp-298.15D0))+((39.064D-03/2.D0)*(Temp**2.-298.15D0**2.))- & ((13.301D-06/3.D0)*(Temp**3.-298.15D0**3.))))+82927.D0 RETURN END FUNCTION h_C6H6 ! *********************** END FUNCTION h_C6H6 *********************************************

269

Appendix D Computer code listing

FUNCTION heat1 (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp ! Gas Temperature [K] REAL(DP)::heat1 ! Calculate the heat of reaction of Reaction 1 [J/kg of Carbon] ! Calculate the heat of reaction of Reaction 1 heat1 = 1000.D0*(h_C(Temp)+gamma*h_O2(Temp)-(2.D0-2.D0*gamma)*h_CO(Temp)-& (2.D0*gamma-1.D0)*h_CO2(Temp))/mole_C RETURN END FUNCTION heat1 ! *********************** END FUNCTION heat1 ********************************************** FUNCTION heat2 (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp ! Gas Temperature [K] REAL(DP)::heat2 ! Calculate the heat of reaction of Reaction 2 [J/kg of Carbon] ! Calculate the heat of reaction of Reaction 2 heat2 = 1000.D0*(h_C(Temp)+h_CO2(Temp)-2.D0*h_CO(Temp))/mole_C RETURN END FUNCTION heat2 ! *********************** END FUNCTION heat2 ********************************************** FUNCTION heat3 (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp ! Gas Temperature [K] REAL(DP)::heat3 ! Calculate the heat of reaction of Reaction 3 [J/kg of Carbon] ! Calculate the heat of reaction of Reaction 3 heat3 = 1000.D0*(h_C(Temp)+h_H2O(Temp)-h_H2(Temp)-h_CO(Temp))/mole_C RETURN END FUNCTION heat3 ! *********************** END FUNCTION heat3 ********************************************** FUNCTION heat4 (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp ! Gas Temperature [K] REAL(DP)::heat4 ! Calculate the heat of reaction of Reaction 4 [J/kg of Carbon] ! Calculate the heat of reaction of Reaction 4 heat4 = 1000.D0*(h_C(Temp)+2.D0*h_H2(Temp)-h_CH4(Temp))/mole_C RETURN END FUNCTION heat4 ! *********************** END FUNCTION heat4 ********************************************** FUNCTION heat5 (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp ! Gas Temperature [K] REAL(DP)::heat5 ! Calculate the heat of reaction of forward reaction 5 [J/kmol] ! Calculate the heat of reaction of Reaction 5 heat5 = 1000.D0*(h_CO(Temp)+h_H2O(Temp)-h_H2(Temp)-h_CO2(Temp)) RETURN END FUNCTION heat5 ! *********************** END FUNCTION heat5 ********************************************** FUNCTION heat5R (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp ! Gas Temperature [K]

270

Appendix D Computer code listing

REAL(DP)::heat5R

! Calculate the heat of reaction of reverse Reaction 5 [J/kmol]

! Calculate the heat of reaction of reverse Reaction 5 heat5R = 1000.D0*(h_CO2(Temp)+h_H2(Temp)-h_H2O(Temp)-h_CO(Temp)) RETURN END FUNCTION heat5R ! *********************** END FUNCTION heat5 ********************************************** FUNCTION heat6 (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp ! Gas Temperature [K] REAL(DP)::heat6 ! Calculate the heat of reaction of Reaction 6 [J/kmol of C6H6] ! Calculate the heat of reaction of Reaction 6 heat6 = 1000.D0*(h_C6H6(Temp)+3.D0*h_O2(Temp)-3.D0*h_H2(Temp)-6.D0*h_CO(Temp)) RETURN END FUNCTION heat6 ! *********************** END FUNCTION heat6 ********************************************** FUNCTION heat7 (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp REAL(DP)::heat7

! Gas Temperature [K] ! Calculate the heat of reaction of Reaction 7 [J/kmol of CO]

! Calculate the heat of reaction of Reaction 7 heat7 = 1000.D0*(h_CO(Temp)+0.5D0*h_O2(Temp)-h_CO2(Temp)) RETURN END FUNCTION heat7 ! *********************** END FUNCTION heat7 ********************************************** FUNCTION heat8 (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp REAL(DP)::heat8

! Gas Temperature [K] ! Calculate the heat of reaction of Reaction 8 [J/kmol of CH4]

! Calculate the heat of reaction of Reaction 8 heat8 = 1000.D0*(h_CH4(Temp)+1.5D0*h_O2(Temp)-2.D0*h_H2O(Temp)-h_CO(Temp)) RETURN END FUNCTION heat8 ! *********************** END FUNCTION heat8 ********************************************** FUNCTION heat9 (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp ! Gas Temperature [K] REAL(DP)::heat9 ! Calculate the heat of reaction of Reaction 9 [J/kmol of C2H4] ! Calculate the heat of reaction of Reaction 9 heat9 = 1000.D0*(h_C2H4(Temp)+ h_O2(Temp)-2.D0*h_H2(Temp)-2.D0*h_CO(Temp)) RETURN END FUNCTION heat9 ! *********************** END FUNCTION heat9 ********************************************** FUNCTION heat10 (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp REAL(DP)::heat10

! Gas Temperature [K] ! Calculate the heat of reaction of Reaction 10 [J/kmol of H2]

! Calculate the heat of reaction of Reaction 10 heat10 = 1000.D0*(h_H2(Temp)+0.5D0*h_O2(Temp)-h_H2O(Temp))

271

Appendix D Computer code listing

RETURN END FUNCTION heat10 ! *********************** END FUNCTION heat10 ********************************************* FUNCTION K (A,E,Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::A REAL(DP), INTENT(IN)::E REAL(DP), INTENT(IN)::Temp REAL(DP)::K ! Calculate K K=A*exp(-E/(R*Temp)) RETURN END FUNCTION K ! ****************************END FUNCTION K*********************************************** FUNCTION visco (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp ! Gas Temperature, K REAL(DP)::visco ! Viscosity, kg/ms ! Calculate viscosity of the gas visco=1.98D-05*((Temp/300.D0)**(2.D0/3.D0)) RETURN END FUNCTION visco ! *********************END FUNCTION visco************************************************** FUNCTION Cp_g(Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp REAL(DP)::Cp_g ! Gas Temperature [K] ! Specific heat of gas mixture [J/kgK]

! ! ! !

Pre-exponential frequency factor for reaction, 1/s.K Activation energy for reaction, J/mol Temperature, K Kinetic reaction rate, 1/s

! Calculate specific heat of gas mixture Cp_g=-3.63930D-8*temp**3.+ 6.50858D-5*temp**2. + 1.66918D-1*temp + 9.41079D+2 RETURN END FUNCTION Cp_g ! *********************END FUNCTION Cp_g ************************************************** FUNCTION Cp_char(Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp REAL(DP)::Cp_char ! Calculate specific heat of char Cp_char=(1.39D0+0.00036D0*(temp+T_a)/2.D0)*1000.D0 RETURN END FUNCTION Cp_char ! *********************END FUNCTION Cp_char *********************************************** FUNCTION Cp_wood(Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp REAL(DP)::Cp_wood ! Solid Temperature [K] ! Specific heat of wood [J/kgK] ! Solid Temperature [K] ! Specific heat of char [J/kgK]

! Calculate specific heat of wood Cp_wood=(1.39D0+0.00036D0*(temp+T_a)/2.D0)*1000.D0

272

Appendix D Computer code listing

RETURN END FUNCTION Cp_wood ! *********************END FUNCTION Cp_wood *********************************************** FUNCTION K_wood (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp REAL(DP)::K_wood ! Solid Temperature [K] ! Thermal conductivity of wood [W/mK]

! Calculate Thermal conductivity of wood K_wood=0.13D0+3.D-4*Temp RETURN END FUNCTION K_wood ! *********************END FUNCTION K_wood ************************************************ FUNCTION K_g (Temp) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp REAL(DP)::K_g ! Gas Temperature [K] ! Thermal conductivity of gas [W/mK]

! Calculate Thermal conductivity of gas K_g=4.77D-04*Temp**0.717D0 RETURN END FUNCTION K_g ! *********************END FUNCTION K_g *************************************************** FUNCTION h_rs (Temp,emiss) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp REAL(DP), INTENT(IN)::emiss REAL(DP)::h_rs ! Solid Temperature [K] ! Emissivity ! Solid radiation coefficient [W/mK]

! Calculate the solid radiation coefficient h_rs=4.D0*sigma*(Temp**3.)*(emiss/(2.D0-emiss)) RETURN END FUNCTION h_rs ! *********************END FUNCTION h_rs ************************************************** FUNCTION h_rv (Temp,emiss,void) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::Temp REAL(DP), INTENT(IN)::emiss REAL(DP), INTENT(IN)::void REAL(DP)::h_rv ! ! ! ! Gas Temperature [K] Emissivity[-] Voidage fraction[-] Void to void radiation coefficient [W/m2K]

! Calculate the void to void radiation coefficient h_rv=4.D0*sigma*(Temp**3.)/(1.D0+((void/(2.D0*(1.D0-void)))*((1.D0-emiss)/emiss))) RETURN END FUNCTION h_rv ! *********************END FUNCTION h_rv ************************************************** FUNCTION K_eff(m_g,T_s,T_g,d_p,emiss,void) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::m_g REAL(DP), INTENT(IN)::T_s REAL(DP), INTENT(IN)::T_g REAL(DP), INTENT(IN)::d_p REAL(DP), INTENT(IN)::emiss REAL(DP), INTENT(IN)::void

! ! ! ! ! !

mass flux [kg/m2sec] Gas Temperature [K] Gas Temperature [K] Particle diameter[m] Emissivity[-] Voidage fraction[-]

273

Appendix D Computer code listing

REAL(DP) ::K_eff

! Effectivity thermal conductivity, [W/mK]

! Calculate the effectivity thermal conductivity in case of gas filled voids, K_eff=keff_mf* ((K_g(T_g)*((1.D0-void)/((2.D0/3.D0)*(K_g(T_g)/K_wood(T_s))+& (1.D0/(10.D0/void+(d_p*h_rs(T_s,emiss)/K_g(T_g)))))+& (void*d_p*h_rv(T_g,emiss,void)/k_g(T_g))))& +(K_g(T_g)*void)& +((0.14D0/(1.D0+46.D0*(d_p/D_rec)))*(d_p*m_g/(visco(T_g)))*& (Cp_g(T_g)*(visco(T_g))))) RETURN END FUNCTION K_eff ! *********************END FUNCTION K_eff************************************************** FUNCTION packp (k) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::k REAL(DP)::packp ! Calculate packing parameter packp=(0.3525D0*(((k-1.D0)/k)**2.)/(((ln(k-0.5431D0*(k-1))))-& (0.4569D0*(k-1.D0)/k)))-(2.D0/(3.D0*k)) RETURN END FUNCTION packp ! *********************END FUNCTION packp************************************************** FUNCTION HV_biomass(C,H,O,N,S,A,M) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN)::C,H,O,N,S,A,M REAL(DP) ::HHV_biomass REAL(DP) ::HV_biomass ! Ratio of solid conductivity and gas conductivity ! Packing parameter

! ! ! !

Carbon, Hydrogen, Oxygen, Nitrogen, Sulfur, Ash & Moisture content of the fuel in dry basic [%] Higher heating value of biomass [MJ/kg] Net or (lower) heating value of biomass [MJ/kg]

! Calculate Higher heating value of biomass in moisture and ash free basic [MJ/kg] HHV_biomass = (34.1D0*C+132.2D0*H+6.8D0*S-1.53D0*A-12.D0*(O+N))/(100.D0) ! Calculate Net (Lower) heating value of biomass in moisture and ash free basic [MJ/kg] HV_biomass = HHV_biomass-2.442D0*(M/100.D0+H/100.D0*9.D0) RETURN END FUNCTION HV_biomass ! *********************END FUNCTION HV_biomass********************************************* FUNCTION HV_gas(CO,H2,CH4,H2O) IMPLICIT NONE ! Declare calling arguments REAL(DP), INTENT(IN):: CO,H2,CH4,H2O REAL(DP) ::HV_gas

! Mole fraction of CO,H2,CH4,H2O [-] ! Lower heating value of Producer gas [MJ/kg]

! Calculate Net (Lower) heating value of producer gas in dry basic [MJ/Nm3] HV_gas = ((CO*100.D0/(1.D0-H2O)*13.1D0)+(H2*100.D0/(1.D0-H2O)*11.2D0)+& (CH4*100.D0/(1.D0-H2O)*37.1D0))/100.D0 IF (CO<0.00000001D0.AND.H2<0.00000001D0.AND.CH4<0.00000001D0) HV_gas=0.D0 RETURN END FUNCTION HV_gas ! *********************END FUNCTION HV_gas************************************************* FUNCTION Kp_wg(temp) IMPLICIT NONE REAL(DP), INTENT(IN)::temp REAL(DP)::Kp_wg ! Calculate the equilibrium constant of water gas shift reaction

274

Appendix D Computer code listing

kp_wg=2.7182818281D0**((1.D0/8.3144D0/temp)*(11321.D0-31.08D0*temp+3*temp*& ln(temp)-0.00028D0*temp**2.-91500.D0/temp)) RETURN END FUNCTION kp_wg ! *****************************END FUNCTION kp_wg****************************************** FUNCTION Diff_AB (Temp,pre,mole_A,mole_B,DV_A,DV_B) IMPLICIT NONE ! This function is to find the Diffusion coefficient between gas A and gas B ! at temperature T & Pressure P. ! Declare calling arguments REAL(DP), INTENT(IN):: Temp REAL(DP), INTENT(IN):: Pre REAL(DP), INTENT(IN):: mole_A REAL(DP), INTENT(IN):: mole_B REAL(DP), INTENT(IN):: DV_A REAL(DP), INTENT(IN):: DV_B REAL(DP):: Diff_AB

! ! ! ! ! ! !

Gas Temperature [K] Pressure of gas mixture, [bar] Molecular weight of A, [g/mol] Molecular weight of B, [g/mol] Diffusion volumes of A Diffusion volumes of B Diffusion O2efficient between gas A and gas B [m2/sec]

! Calculate Diffusion O2efficient between O2 and gas mixture at temperature T Diff_AB = (0.00143*Temp**1.75)*sqrt(1.D0/mole_A+1.D0/mole_B)*1.D-04 / & (pre*sqrt(2.D0)*(DV_A**(1./3.)+DV_B**(1./3.))**2.) RETURN END FUNCTION Diff_AB ! *********************** END FUNCTION Diff_AB ******************************************** SUBROUTINE diff_O2_gmix() ! This subroutine is to calculate the Diffusion coefficient between ! O2 and gas mixture at reactor temperature and pressure. REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: diff_O2_N2 diff_O2_CO diff_O2_CO2 diff_O2_H2 diff_O2_H2O diff_O2_CH4 diff_O2_C2H4 diff_O2_tar1 diff_O2_tar2 = = = = = = = = = Diff_AB Diff_AB Diff_AB Diff_AB Diff_AB Diff_AB Diff_AB Diff_AB Diff_AB ! ! ! ! ! ! ! ! ! Diffusion Diffusion Diffusion Diffusion Diffusion Diffusion Diffusion Diffusion Diffusion coefficient coefficient coefficient coefficient coefficient coefficient coefficient coefficient coefficient between between between between between between between between between O2 O2 O2 O2 O2 O2 O2 O2 O2 and and and and and and and and and N2 [m2/sec] CO [m2/sec] CO2 [m2/sec] H2 [m2/sec] H2O [m2/sec] CH4 [m2/sec] C2H4 [m2/sec] tar1 [m2/sec] tar2 [m2/sec]

diff_O2_N2 diff_O2_CO diff_O2_CO2 diff_O2_H2 diff_O2_H2O diff_O2_CH4 diff_O2_C2H4 diff_O2_tar1 diff_O2_tar2

(T_g(i),pre/100.D0,mole_O2,mole_N2,DV_O2,DV_N2) (T_g(i),pre/100.D0,mole_O2,mole_CO,DV_O2,DV_CO) (T_g(i),pre/100.D0,mole_O2,mole_CO2,DV_O2,DV_CO2) (T_g(i),pre/100.D0,mole_O2,mole_H2,DV_O2,DV_H2) (T_g(i),pre/100.D0,mole_O2,mole_H2O,DV_O2,DV_H2O) (T_g(i),pre/100.D0,mole_O2,mole_CH4,DV_O2,DV_CH4) (T_g(i),pre/100.D0,mole_O2,mole_C2H4,DV_O2,DV_C2H4) (T_g(i),pre/100.D0,mole_O2,mole_tar1,DV_O2,DV_tar1) (T_g(i),pre/100.D0,mole_O2,mole_tar2,DV_O2,DV_tar2)

diff_O2_gas =(1.D0-Y_O2(i))/(Y_N2(i)/diff_O2_N2+Y_CO(i)/diff_O2_CO+Y_CO2(i)/diff_O2_CO2+ & Y_H2(i)/diff_O2_H2+Y_H2O(i)/diff_O2_H2O+Y_CH4(i)/diff_O2_CH4+ & Y_C2H4(i)/diff_O2_C2H4+Y_tar1(i)/diff_O2_tar1+Y_tar2(i)/diff_O2_tar2) RETURN END SUBROUTINE diff_O2_gmix ! *********************END SUBROUTINE diff_O2_gas****************************************** SUBROUTINE diff_CO2_gmix() ! This subroutine is to calculate the Diffusion coefficient between ! CO2 and gas mixture at reactor temperature and pressure. REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: diff_CO2_N2 diff_CO2_O2 diff_CO2_CO diff_CO2_H2 diff_CO2_H2O diff_CO2_CH4 ! ! ! ! ! ! Diffusion Diffusion Diffusion Diffusion Diffusion Diffusion coefficient coefficient coefficient coefficient coefficient coefficient between between between between between between CO2 CO2 CO2 CO2 CO2 CO2 and and and and and and N2 [m2/sec] O2 [m2/sec] CO [m2/sec] H2 [m2/sec] H2O [m2/sec] CH4 [m2/sec]

275

Appendix D Computer code listing

REAL(DP):: diff_CO2_C2H4 REAL(DP):: diff_CO2_tar1 REAL(DP):: diff_CO2_tar2 diff_CO2_N2 diff_CO2_O2 diff_CO2_CO diff_CO2_H2 diff_CO2_H2O diff_CO2_CH4 diff_CO2_C2H4 diff_CO2_tar1 diff_CO2_tar2 = = = = = = = = = Diff_AB Diff_AB Diff_AB Diff_AB Diff_AB Diff_AB Diff_AB Diff_AB Diff_AB

! Diffusion coefficient between CO2 and C2H4 [m2/sec] ! Diffusion coefficient between CO2 and tar1 [m2/sec] ! Diffusion coefficient between CO2 and tar2 [m2/sec] (T_g(i),pre/100.D0,mole_CO2,mole_N2,DV_CO2,DV_N2) (T_g(i),pre/100.D0,mole_CO2,mole_O2,DV_CO2,DV_O2) (T_g(i),pre/100.D0,mole_CO2,mole_CO,DV_CO2,DV_CO) (T_g(i),pre/100.D0,mole_CO2,mole_H2,DV_CO2,DV_H2) (T_g(i),pre/100.D0,mole_CO2,mole_H2O,DV_CO2,DV_H2O) (T_g(i),pre/100.D0,mole_CO2,mole_CH4,DV_CO2,DV_CH4) (T_g(i),pre/100.D0,mole_CO2,mole_C2H4,DV_CO2,DV_C2H4) (T_g(i),pre/100.D0,mole_CO2,mole_tar1,DV_CO2,DV_tar1) (T_g(i),pre/100.D0,mole_CO2,mole_tar2,DV_CO2,DV_tar2)

diff_CO2_gas=(1.D0-Y_CO2(i))/(Y_N2(i)/diff_CO2_N2+Y_O2(i)/diff_CO2_O2+Y_CO(i)/diff_CO2_CO+& Y_H2(i)/diff_CO2_H2+Y_H2O(i)/diff_CO2_H2O+Y_CH4(i)/diff_CO2_CH4+ & Y_C2H4(i)/diff_CO2_C2H4+Y_tar1(i)/diff_CO2_tar1+Y_tar2(i)/diff_CO2_tar2) RETURN END SUBROUTINE diff_CO2_gmix ! *********************END SUBROUTINE diff_CO2_gmix**************************************** SUBROUTINE diff_H2O_gmix() ! This subroutine is to calculate the Diffusion coefficient between ! H2O and gas mixture at reactor temperature and pressure. REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: diff_H2O_N2 diff_H2O_O2 diff_H2O_CO diff_H2O_CO2 diff_H2O_H2 diff_H2O_CH4 diff_H2O_C2H4 diff_H2O_tar1 diff_H2O_tar2 = = = = = = = = = Diff_AB Diff_AB Diff_AB Diff_AB Diff_AB Diff_AB Diff_AB Diff_AB Diff_AB ! ! ! ! ! ! ! ! ! Diffusion Diffusion Diffusion Diffusion Diffusion Diffusion Diffusion Diffusion Diffusion coefficient coefficient coefficient coefficient coefficient coefficient coefficient coefficient coefficient between between between between between between between between between H2O H2O H2O H2O H2O H2O H2O H2O H2O and and and and and and and and and N2 [m2/sec] O2 [m2/sec] CO [m2/sec] CO2 [m2/sec] H2 [m2/sec] CH4 [m2/sec] C2H4 [m2/sec] tar1 [m2/sec] tar2 [m2/sec]

diff_H2O_N2 diff_H2O_O2 diff_H2O_CO diff_H2O_CO2 diff_H2O_H2 diff_H2O_CH4 diff_H2O_C2H4 diff_H2O_tar1 diff_H2O_tar2 diff_H2O_gas

(T_g(i),pre/100.D0,mole_H2O,mole_N2,DV_H2O,DV_N2) (T_g(i),pre/100.D0,mole_H2O,mole_O2,DV_H2O,DV_O2) (T_g(i),pre/100.D0,mole_H2O,mole_CO,DV_H2O,DV_CO) (T_g(i),pre/100.D0,mole_H2O,mole_H2O,DV_CO2,DV_CO2) (T_g(i),pre/100.D0,mole_H2O,mole_H2,DV_H2O,DV_H2) (T_g(i),pre/100.D0,mole_H2O,mole_CH4,DV_H2O,DV_CH4) (T_g(i),pre/100.D0,mole_H2O,mole_C2H4,DV_H2O,DV_C2H4) (T_g(i),pre/100.D0,mole_H2O,mole_tar1,DV_H2O,DV_tar1) (T_g(i),pre/100.D0,mole_H2O,mole_tar2,DV_H2O,DV_tar2)

= (1.D0-Y_H2O(i))/(Y_N2(i)/diff_H2O_N2+Y_O2(i)/diff_H2O_O2+ & Y_CO(i)/diff_H2O_CO+Y_CO2(i)/diff_H2O_CO2+Y_H2(i)/diff_H2O_H2+ & Y_CH4(i)/diff_H2O_CH4+Y_C2H4(i)/diff_H2O_C2H4+Y_tar1(i)/diff_H2O_tar1+ & Y_tar2(i)/diff_H2O_tar2)

RETURN END SUBROUTINE diff_H2O_gmix ! *********************END SUBROUTINE diff_H2O_gmix**************************************** SUBROUTINE diff_H2_gmix() ! This subroutine is to calculate the Diffusion coefficient between ! H2 and gas mixture at reactor temperature and pressure. REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: REAL(DP):: diff_H2_N2 diff_H2_O2 diff_H2_CO diff_H2_CO2 diff_H2_N2 diff_H2_O2 diff_H2_CO diff_H2_CO2 diff_H2_H2O diff_H2_CH4 diff_H2_C2H4 diff_H2_tar1 diff_H2_tar2 = = = = Diff_AB Diff_AB Diff_AB Diff_AB ! ! ! ! ! ! ! ! ! Diffusion Diffusion Diffusion Diffusion Diffusion Diffusion Diffusion Diffusion Diffusion coefficient coefficient coefficient coefficient coefficient coefficient coefficient coefficient coefficient between between between between between between between between between H2 H2 H2 H2 H2 H2 H2 H2 H2 and and and and and and and and and N2 [m2/sec] O2 [m2/sec] CO [m2/sec] CO2 [m2/sec] H2O [m2/sec] CH4 [m2/sec] C2H4 [m2/sec] tar1 [m2/sec] tar2 [m2/sec]

(T_g(i),pre/100.D0,mole_H2,mole_N2,DV_H2,DV_N2) (T_g(i),pre/100.D0,mole_H2,mole_O2,DV_H2,DV_O2) (T_g(i),pre/100.D0,mole_H2,mole_CO,DV_H2,DV_CO) (T_g(i),pre/100.D0,mole_H2,mole_H2,DV_CO2,DV_CO2)

276

Appendix D Computer code listing

diff_H2_H2O diff_H2_CH4 diff_H2_C2H4 diff_H2_tar1 diff_H2_tar2

= = = = =

Diff_AB Diff_AB Diff_AB Diff_AB Diff_AB

(T_g(i),pre/100.D0,mole_H2,mole_H2O,DV_H2,DV_H2O) (T_g(i),pre/100.D0,mole_H2,mole_CH4,DV_H2,DV_CH4) (T_g(i),pre/100.D0,mole_H2,mole_C2H4,DV_H2,DV_C2H4) (T_g(i),pre/100.D0,mole_H2,mole_tar1,DV_H2,DV_tar1) (T_g(i),pre/100.D0,mole_H2,mole_tar2,DV_H2,DV_tar2)

diff_H2_gas =(1.D0-Y_H2(i))/(Y_N2(i)/diff_H2_N2+Y_O2(i)/diff_H2_O2+Y_CO(i)/diff_H2_CO+ & Y_CO2(i)/diff_H2_CO2+Y_H2O(i)/diff_H2_H2O+Y_CH4(i)/diff_H2_CH4+ & Y_C2H4(i)/diff_H2_C2H4+Y_tar1(i)/diff_H2_tar1+Y_tar2(i)/diff_H2_tar2) RETURN END SUBROUTINE diff_H2_gmix ! *********************END SUBROUTINE diff_H2_gmix***************************************** SUBROUTINE initial() ! This subroutine is to initialize the solid and gas temperature profile, ! gas composition profile and species density profile. DO i=1,N IF (i<=N_top) THEN IF (i==1) THEN x(i)=0.D0 ELSE x(i)=x(i-1)+depth_bio/real(N_top-1) END IF rho_m(i) = rho_m(1) rho_volt(i) = rho_volt(1) rho_char(i) = rho_char(1) T_s(i) = T_a T_g(i) = T_gas T_wall(i) = T_a d_p(i) = d_pwood ELSE IF (i<=N_top+N_ign) THEN x(i) = x(i-1)+depth_ign/real(N_ign) rho_m(i) = 0.D0 rho_char(i) = rho_char(1) rho_volt(i) = 0.D0 T_s(i) = T_ign T_g(i) = T_gas T_wall(i) = T_a d_p(i) = d_pinitial ELSE x(i) rho_m(i) rho_char(i) rho_volt(i) T_s(i) T_g(i) T_wall(i) d_p(i) END IF Y_O2(i) Y_N2(i) Y_CO2(i) Y_CO(i) Y_H2O(i) Y_H2(i) Y_CH4(i) Y_C2H4(i) Y_tar1(i) Y_tar2(i) mole_g(i) m_g(i) v_s(i) rho_g(i) END DO RETURN END SUBROUTINE initial = = = = = = = = = = = = = =

= = = = = = = =

x(i-1)+(L_rec-depth_bio-depth_ign)/real(N-N_top-N_ign) 0.D0 rho_char(1) 0.D0 T_a T_gas T_a d_pinitial

Y_O2(1) Y_N2(1) Y_CO2(1) Y_CO(1) Y_H2O(1) Y_H2(1) Y_CH4(1) Y_C2H4(1) Y_tar1(1) Y_tar2(1) mole_g(1) m_g(1) 0.D0 mole_g(i)*pre/(R*T_g(i))

277

Appendix D Computer code listing

! *********************END SUBROUTINE initial********************************************** SUBROUTINE pretime() ! This subroutine is to save the speices denisty, gas composition and gas & ! solid temperature of the previous time. DO i=1,N YP_O2(i) YP_CO2(i) YP_CO(i) YP_H2O(i) YP_H2(i) YP_CH4(i) YP_N2(i) YP_C2H4(i) YP_tar1(i) YP_tar2(i) mP_g(i) TP_s(i) TP_g(i) moleP_g(i) rhoP_m(i) rhoP_volt(i) rhoP_g(i) dP_p(i) END DO RETURN END SUBROUTINE pretime ! *********************END SUBROUTINE pretime********************************************** SUBROUTINE mh_coefficient() ! This subroutine is to calculate the mass transfer coefficient and ! solid to gas heat transfer coefficient. ! Calculate the molecular weight of gas mixture, g/mol mole_g(i) = Y_O2(i)*mole_O2+Y_CO(i)*mole_CO +Y_CO2(i)*mole_CO2+Y_H2O(i)*mole_H2O+ & Y_H2(i)*mole_H2+Y_CH4(i)*mole_CH4+Y_C2H4(i)*mole_C2H4+Y_tar1(i)*mole_tar1+ & Y_tar2(i)*mole_tar2+Y_N2(i)*mole_N2 ! Calculate the mass concentration of gas mixture, [kg/m3] rho_g(i) = mole_g(i)*pre/(R*T_g(i)) ! Calculate the Particle surface area/ Unit volume. A_v=Av_mf*(6.D0/D_p(i))*(1.D0-void) ! Calculate the Renolds Number Re=d_p(i)*m_g(i)/(visco(T_g(i))) ! Calculate the Prandtl Number Pr=Cp_g(T_g(i))*(visco(T_g(i)))/K_g(T_g(i)) !CALL diff_C2H4_gmix() CALL diff_O2_gmix() CALL diff_CO2_gmix() CALL diff_H2O_gmix() CALL diff_H2_gmix() Sc_1=visco(T_g(i))/(rho_g(i)*diff_O2_gas) Sc_2=visco(T_g(i))/(rho_g(i)*diff_CO2_gas) Sc_3=visco(T_g(i))/(rho_g(i)*diff_H2O_gas) sc_4=visco(T_g(i))/(rho_g(i)*diff_H2_gas) ! Calculate the Solid to gas heat transfer coefficient,W/mK with Bird et al, 2002 fomula h_sg = zeta*Cp_g(T_g(i))*m_g(i)*(2.19D0*(Re**(-2.D0/3.D0))+& 0.78D0*(Re**(-0.381D0)))*(Pr**(-2.D0/3.D0)) ! Calculate the mass transfer coefficient, m/s, Bird et al, 2002 fomula km_1 = (m_g(i)/rho_g(i))*(2.19D0*(Re**(-2.D0/3.D0))+ & 0.78D0*(Re**(-0.381D0)))*(Sc_1**(-2.D0/3.D0)) = = = = = = = = = = = = = = = = = = Y_O2(i) Y_CO2(i) Y_CO(i) Y_H2O(i) Y_H2(i) Y_CH4(i) Y_N2(i) Y_C2H4(i) Y_tar1(i) Y_tar2(i) m_g(i) T_s(i) T_g(i) mole_g(i) rho_m(i) rho_volt(i) rho_g(i) d_p(i)

278

Appendix D Computer code listing

Km_2 = (m_g(i)/rho_g(i))*(2.19D0*(Re**(-2.D0/3.D0))+ & 0.78D0*(Re**(-0.381D0)))*(Sc_2**(-2.D0/3.D0)) km_3 = (m_g(i)/rho_g(i))*(2.19D0*(Re**(-2.D0/3.D0))+ & 0.78D0*(Re**(-0.381D0)))*(Sc_3**(-2.D0/3.D0)) km_4 = (m_g(i)/rho_g(i))*(2.19D0*(Re**(-2.D0/3.D0))+ & 0.78D0*(Re**(-0.381D0)))*(Sc_4**(-2.D0/3.D0)) IF IF IF IF (km_1>km_max) (km_2>km_max) (km_3>km_max) (km_4>km_max) km_1=km_max km_2=km_max km_3=km_max km_4=km_max

RETURN END SUBROUTINE mh_coefficient ! *********************END SUBROUTINE mh_coefficient*************************************** SUBROUTINE reaction_rates() ! This subroutine is to calculate the rates of drying, primary and secondary pyrolysis, ! hetrogeneous gas_char reactions, water-gas shift reaction and ! oxidation of CO and secondary tar. gamma = (2.D0+K(A_gamma,E_gamma,T_s(i)))/(2.D0+2.D0*K(A_gamma,E_gamma,T_s(i))) T_e(i) = 1/((1/T_s(i))+((R/Ep1)*ln(953.138*d_p(i)+46861.984*d_p(i)**2))) IF (T_e(i)<T_a) THEN T_e(i)=T_a ELSE T_e(i)=T_e(i) END IF ratem_m(i) = (1.D0-void)*K(Am,Em,T_s(i)) ratep1_volt(i) = (1.D0-void)*Ap1_mf*K(Ap1,Ep1,T_e(i)) T_p2 (i) = T_s(i)*p_alpha+T_g(i)*(1.D0-p_alpha) IF (T_s(i)>T_g(i)) THEN ratep2_tar1(i) = void*mole_tar1*pre*Ap2_mf*K(Ap2,Ep2*Ep2_mf,T_p2(i))/T_g(i) ELSE ratep2_tar1(i) = void*mole_tar1*pre*Ap2_mf*K(Ap2,Ep2*Ep2_mf,T_g(i))/T_g(i) END IF rate1_O2(i) = (mole_C*A_v*pre/(gamma*R*T_g(i)))/((1./km_1)+(1./(T_s(i)*K(A1,E1,T_s(i)))))

IF (Y_O2(i)<0.0000001) THEN rate2_CO2(i) = (mole_C*A_v*pre/(R*T_g(i)))/((1./km_2)+(1./(T_s(i)*K(A2,E2,T_s(i))))) rate3_H2O(i) = (mole_C*A_v*pre/(R*T_g(i)))/((1./km_3)+(1./(T_s(i)*K(A3,E3,T_s(i))))) rate4_H2(i) =(mole_C*A_v*pre/(2.D0*R*T_g(i)))/((1./km_4)+(1./(T_s(i)*K(A4,E4,T_s(i))))) ELSE rate2_CO2(i) = 0.D0 rate3_H2O(i) = 0.D0 rate4_H2(i) = 0.D0 END IF rate5_CO(i) rate5_H2O(i) rate5R_CO2(i) rate5R_H2(i) rate6_O2(i) rate7_CO(i) rate9_O2(i) ratem(i) ratep1(i) ratep2(i) rate1(i) rate2(i) rate3(i) rate4(i) rate5(i) rate5R(i) rate6(i) rate7(i) rate8(i) rate9(i) = = = = = = = = = = = = = = void*K(A5,E5,T_g(i))*(pre/T_g(i))**2.*Y_H2O(i) = void*K(A5,E5,T_g(i))*(pre/T_g(i))**2.*Y_CO(i) = void*K(A5,E5,T_g(i))*(pre/T_g(i))**2.*Y_H2(i)/Kp_wg(T_g(i)) = void*K(A5,E5,T_g(i))*(pre/T_g(i))**2.*Y_CO2(i)/Kp_wg(T_g(i)) = void*A6_mf*K(A6,E6,T_g(i))*(pre**1.8)*(Y_tar2(i))**0.5/(T_g(i)**0.5) = void*K(A7,E7,T_g(i))*(pre/T_g(i))**2.*(Y_O2(i))**0.5*(Y_H2O(i))**0.5 = void*K(A9,E9,T_g(i))*(pre**1.8)*(Y_C2H4(i))**0.5/(T_g(i)**0.5)

ratem_m(i)*rho_m(i) ratep1_volt(i)*rho_volt(i) Y_tar1(i)*ratep2_tar1(i) rate1_O2(i)*Y_O2(i) rate2_CO2(i)*Y_CO2(i) rate3_H2O(i)*Y_H2O(i) rate4_H2(i)*Y_H2(i) rate5_CO(i)*Y_H2O(i) rate5R_CO2(i)*Y_H2(i) rate6_O2(i)*(Y_O2(i)) rate7_CO(i)*(Y_CO(i)) void*K(A8,E8,T_g(i))*(pre/T_g(i))**1.5*(Y_O2(i))**0.8*(Y_CH4(i))**0.7 rate9_O2(i)*(Y_O2(i))

279

Appendix D Computer code listing

rate10(i) = void*K(A10,E10,T_g(i))*(pre/T_g(i))**2.2*(Y_O2(i))**1.1*(Y_H2(i))**1.1 RETURN END SUBROUTINE reaction_rates ! *********************END SUBROUTINE reaction_rates*************************************** SUBROUTINE gas_balance() ! This subroutine is to calculate the gas composition by applying mass balance. delta_x1 molar_g molarP_g = x(i)-x(i-1) = pre*delta_x1/(R*T_g(i)*delta_t) = pre*delta_x1/(R*TP_g(i)*delta_t)

! By applying overall gas mixture balance ! --------------------------------------A_gas = 1.D0 B_gas = 1.D0 C_gas = (ratem(i)+ratep1(i)+rate1(i)+rate2(i)+rate3(i)+rate4(i))*delta_x1+ & (void*molarP_g*moleP_g(i))-(void*molar_g*mole_g(i)) m_g(i)=(B_gas*m_g(i-1)+C_gas)/A_gas ! By applying O2 balance ! ---------------------Y_O2inverse = 0.D0 IF(Y_O2(i)>1.0D-300)Y_O2inverse=1.D0/Y_O2(i) A_O2 = (void*molar_g*mole_O2)+(m_g(i)*mole_O2/mole_g(i))+ & (gamma)*(mole_O2/mole_C)*rate1_O2(i)*delta_x1+ & 3.D0*mole_O2*rate6_O2(i)*delta_x1+0.60D0*mole_O2*rate7(i)*delta_x1*Y_O2inverse+& 1.8D0*mole_O2*rate8(i)*delta_x1*Y_O2inverse+mole_O2*rate9_O2(i)*delta_x1+& 0.55D0*mole_O2*rate10(i)*delta_x1*Y_O2inverse = m_g(i-1)*mole_O2/mole_g(i-1) = (void*molarP_g*mole_O2*YP_O2(i))+0.1D0*mole_O2*rate7(i)*delta_x1+& 0.3D0*mole_O2*rate8(i)*delta_x1+0.05D0*mole_O2*rate10(i)*delta_x1

B_O2 C_O2

! By applying CO balance ! ---------------------A_CO = (void*molar_g*mole_CO)+(m_g(i)*mole_CO/mole_g(i))+mole_CO*rate7_CO(i)*delta_x1+& mole_CO*rate5_CO(i)*delta_x1 B_CO = m_g(i-1)*mole_CO/mole_g(i-1) C_CO = (void*molarP_g*mole_CO*YP_CO(i))+(X1_CO/X1_volt)*ratep1(i)*delta_x1+ & (X2_CO/X1_tar1)*ratep2(i)*delta_x1+ & (2.D0-2.D0*gamma)*mole_CO*rate1(i)*delta_x1/mole_C + & 2.D0*mole_CO*rate2(i)*delta_x1/mole_C+ mole_CO*rate3(i)*delta_x1/mole_C+ & 6.D0*mole_CO*rate6(i)*delta_x1+mole_CO*rate8(i)*delta_x1+ & 2.D0*mole_CO*rate9(i)*delta_x1 + mole_CO*rate5R(i)*delta_x1 ! By applying CO2 balance ! ----------------------A_CO2 = (void*molar_g*mole_CO2)+(m_g(i)*mole_CO2/mole_g(i))+ & (mole_CO2/mole_C)*rate2_CO2(i)*delta_x1+mole_CO2*rate5R_CO2(i)*delta_x1 B_CO2 = m_g(i-1)*mole_CO2/mole_g(i-1) C_CO2 = (void*molarP_g*mole_CO2*YP_CO2(i))+(X1_CO2/X1_volt)*ratep1(i)*delta_x1+ & (X2_CO2/X1_tar1)*rateP2(i)*delta_x1+ & (2.D0*gamma-1.D0)*mole_CO2*rate1(i)*delta_x1/mole_C+ & mole_CO2*rate5(i)*delta_x1+ mole_CO2*rate7(i)*delta_x1 ! By applying H2O balance ! ----------------------A_H2O = (void*molar_g*mole_H2O)+(m_g(i)*mole_H2O/mole_g(i))+ & (mole_H2O/mole_C)*rate3_H2O(i)*delta_x1 + mole_H2O*rate5_H2O(i)*delta_x1 B_H2O = m_g(i-1)*mole_H2O/mole_g(i-1) C_H2O = (void*molarP_g*mole_H2O*YP_H2O(i))+ratem(i)*delta_x1+ & (X1_H2O/X1_volt)*ratep1(i)*delta_x1+2.D0*mole_H2O*rate8(i)*delta_x1+ & mole_H2O*rate10(i)*delta_x1 +mole_H2O*rate5R(i)*delta_x1 ! By applying H2 balance ! ---------------------Y_H2inverse = 0.D0 IF(Y_H2(i)>0.D0)Y_H2inverse=1.D0/Y_H2(i) A_H2 = (void*molar_g*mole_H2)+(m_g(i)*mole_H2/mole_g(i))+ &

280

Appendix D Computer code listing

B_H2 C_H2

2.D0*(mole_H2/mole_C)*rate4_H2(i)*delta_x1+ & 1.1D0*mole_H2*rate10(i)*delta_x1*Y_H2inverse + mole_H2*rate5R_H2(i)*delta_x1 = m_g(i-1)*mole_H2/mole_g(i-1) = (void*molarP_g*mole_H2*YP_H2(i))+(X2_H2/X1_tar1)*ratep2(i)*delta_x1+ & (mole_H2/mole_C)*rate3(i)*delta_x1+mole_H2*rate5(i)*delta_x1+& 3.D0*rate6(i)*mole_H2*delta_x1+2.D0*rate9(i)*mole_H2*delta_x1 & +0.1D0*mole_H2*rate10(i)*delta_x1

! By applying CH4 balance ! ----------------------Y_CH4inverse = 0.D0 IF(Y_CH4(i)>0.D0)Y_CH4inverse=1.D0/Y_CH4(i) A_CH4 = (void*molar_g*mole_CH4)+(m_g(i)*mole_CH4/mole_g(i))+& 1.1D0*mole_CH4*rate8(i)*delta_x1*Y_CH4inverse B_CH4 = m_g(i-1)*mole_CH4/mole_g(i-1) C_CH4 = (void*molarP_g*mole_CH4*YP_CH4(i))+(X1_CH4/X1_volt)*ratep1(i)*delta_x1+ & (X2_CH4/X1_tar1)*ratep2(i)*delta_x1+(mole_CH4/mole_C)*rate4(i)*delta_x1+& 0.1D0*mole_CH4*rate8(i)*delta_x1 ! By applying C2H4 balance ! -----------------------Y_C2H4inverse = 0.D0 IF(Y_C2H4(i)>0.D0)Y_C2H4inverse=1.D0/Y_C2H4(i) A_C2H4 = (void*molar_g*mole_C2H4)+(m_g(i)*mole_C2H4/mole_g(i))+& 1.2D0*mole_C2H4*rate9(i)*delta_x1*Y_C2H4inverse B_C2H4 = m_g(i-1)*mole_C2H4/mole_g(i-1) C_C2H4 = (void*molarP_g*mole_C2H4*YP_C2H4(i))+(X2_C2H4/X1_tar1)*ratep2(i)*delta_x1+& 0.2D0*mole_C2H4*rate9(i)*delta_x1

! By applying primary tar balance ! ------------------------------A_tar1 = (void*molar_g*mole_tar1)+(m_g(i)*mole_tar1/mole_g(i))+ & ratep2_tar1(i)*delta_x1 B_tar1 = m_g(i-1)*mole_tar1/mole_g(i-1) C_tar1 = (void*molarP_g*mole_tar1*YP_tar1(i))+(X1_tar1/X1_volt)*ratep1(i)*delta_x1

! By applying secondary tar balance ! --------------------------------Y_tar2inverse = 0.D0 IF(Y_tar2(i)>0.D0)Y_tar2inverse=1.D0/Y_tar2(i) A_tar2 = (void*molar_g*mole_tar2)+(m_g(i)*mole_tar2/mole_g(i))+ & 1.2D0*mole_tar2*rate6(i)*delta_x1*Y_tar2inverse B_tar2 = m_g(i-1)*mole_tar2/mole_g(i-1) C_tar2 = (void*molarP_g*mole_tar2*YP_tar2(i))+(X2_tar2/X1_tar1)*ratep2(i)*delta_x1+ & 0.2D0*mole_tar2*rate6(i)*delta_x1 ! Calculate the gas compostions Y_O2(i) = (B_O2*Y_O2(i-1)+C_O2)/A_O2 IF (Y_O2(i)<1.0D-300) Y_O2(i)=0.D0 IF (NINT(time)>3.AND.x(i)>loc_secair.AND.sec_check==1) THEN Y_O2(i)= (Y_O2(i)*m_g(i)+ (mf_O2* secondair*4.D0)/(pi*D_rec**2.D0*3600.D0))/& (m_g(i) + (secondair*4.D0)/(pi*D_rec**2.D0*3600.D0)) sec_check=0 END IF Y_CO(i) = (B_CO*Y_CO(i-1)+C_CO)/A_CO IF (Y_CO(i)<0.D0) Y_CO(i)=0.D0 Y_H2(i) = (B_H2*Y_H2(i-1)+C_H2)/A_H2 IF (Y_H2(i)<0.D0) Y_H2(i)=0.D0 Y_CH4(i) = (B_CH4*Y_CH4(i-1)+C_CH4)/A_CH4 IF (Y_CH4(i)<0.D0) Y_CH4(i)=0.D0 Y_C2H4(i)= (B_C2H4*Y_C2H4(i-1)+C_C2H4)/A_C2H4 IF (Y_C2H4(i)<0.D0) Y_C2H4(i)=0.D0 Y_tar1(i) = (B_tar1*Y_tar1(i-1)+C_tar1)/A_tar1

281

Appendix D Computer code listing

IF (Y_tar1(i)<0.D0) Y_tar1(i)=0.D0 Y_tar2(i) = (B_tar2*Y_tar2(i-1)+C_tar2)/A_tar2 IF (Y_tar2(i)<0.D0) Y_tar2(i)=0.D0 Y_CO2(i) = (B_CO2*Y_CO2(i-1)+C_CO2)/A_CO2 Y_H2O(i) = (B_H2O*Y_H2O(i-1)+C_H2O)/A_H2O Y_N2(i) = 1.D0-(Y_O2(i)+Y_CO(i)+Y_CO2(i)+Y_H2O(i)+Y_H2(i)+ & Y_CH4(i)+Y_C2H4(i)+Y_tar1(i)+Y_tar2(i))

RETURN END SUBROUTINE gas_balance ! *********************END SUBROUTINE gas_balance****************************************** SUBROUTINE solid_velocity() delta_x1 = x(i)-x(i-1) ! This subroutine is to calculate the solid velocity. A_Vs = (1.D0-void)*(rho_char(1)) B_Vs = A_Vs C_Vs = (rate1(i)+rate2(i)+rate3(i)+rate4(i))*delta_x1 V_s(i-1)=(B_Vs*V_s(i)+C_Vs)/A_Vs RETURN END SUBROUTINE solid_velocity ! *********************END SUBROUTINE solid_velocity*************************************** SUBROUTINE particle_diameter() delta_x1 = x(i)-x(i-1) ! This subroutine is to calculate the particle_diameter A_dp=delta_x1+V_s(i)*delta_t B_dp=delta_x1/dP_p(i)**3.D0 C_dp=V_s(i-1)*delta_t/d_p(i-1)**3.D0 d_p(i)=(A_dp/(B_dp+C_dp))**(1.D0/3.D0) RETURN END SUBROUTINE particle_diameter ! *********************END SUBROUTINE particle_diameter************************************ SUBROUTINE solid_balance() ! This subroutine is to calculate the solid density by applying mass balance. delta_x1 = x(i)-x(i-1) A_rhom = delta_x1/delta_t+V_s(i)+ratem_m(i)*delta_x1/(1.D0-void) B_rhom = V_s(i-1) C_rhom = rhoP_m(i)*delta_x1/delta_t rho_m(i)=(B_rhom*rho_m(i-1)+C_rhom)/A_rhom IF (rho_m(i)<0.D0) THEN rho_m(i)=0.D0 ratem(i)=0.D0 END IF IF (rho_m(i)>rho_m(1)) THEN rho_m(i)=rho_m(1) ratem(i)=0.D0 END IF A_rhov = delta_x1/delta_t+V_s(i)+ratep1_volt(i)*delta_x1/(1.D0-void) B_rhov = V_s(i-1) C_rhov = rhoP_volt(i)*delta_x1/delta_t rho_volt(i)=(B_rhov*rho_volt(i-1)+C_rhov)/A_rhov IF (rho_volt(i)<0.D0) THEN rho_volt(i)=0.D0 ratep1(i)=0.D0 END IF

282

Appendix D Computer code listing

IF (rho_volt(i)>rho_volt(1)) THEN rho_volt(i)=rho_volt(1) ratep1(i)=0.D0 END IF RETURN END SUBROUTINE solid_balance ! *********************END SUBROUTINE solid_balance**************************************** SUBROUTINE wallh() ! This subroutine is to calculate the heat transfer coefficient of ! solid surface and gas to wall. ! Calculate the conductivity ratio k_ratio=k_wood(T_s(i))/k_g(T_g(i)) ! Calculate the solid effective radial conductivity k_rs=k_wood(T_s(i))*(1.0-void)/((2.D0/(3.D0*k_ratio))+(1.D0/((1.D0/packp(k_ratio))+& (h_rs(T_s(i),emiss)*d_p(i)/k_wood(T_s(i)))))) ! Calculate the gas effective radial conductivity k_rg=k_g(T_g(i))*((void*(1.D0+(d_p(i)*h_rv(T_g(i),emiss,void)/& k_g(T_g(i)))))+(0.14D0*Pr*Re/(1.D0+46.D0*((d_p(i)/D_rec)**2.D0)))) ! Calculate the static effective radial conductivity k_ro=(k_g(T_g(i))*void*(1.D0+(d_p(i)*h_rv(T_g(i),emiss,void)/k_g(T_g(i)))))+& ((k_g(T_g(i))*(1.D0-void))/((2.D0/(3.D0*k_ratio))+ & (1.D0/((1.D0/packp(k_ratio))+(h_rs(T_s(i),emiss)*d_p(i)/k_g(T_g(i))))))) ! Calculate the bed to wall heat transfer coefficient h_bw=wall_loss*(2.44D0*k_ro*(D_rec**(-4.D0/3.D0)))+(0.033D0*k_g(T_g(i))*Pr*Re/d_p(i)) ! Calculate the gas to wall heat transfer coefficient h_gw=h_bw*k_rg/(k_rg+k_rs) ! Calculate the solid to wall heat transfer coefficient h_sw=h_bw*k_rs/(k_rg+k_rs) RETURN END SUBROUTINE wallh ! *********************END SUBROUTINE wallh************************************************ SUBROUTINE thomasenergy_balance() ! This subroutine is to find coefficients by applying energy balance for the calculation of ! the solid and gas temperature. IF (NINT(time)<=300)THEN wall_coeff=0.D0 ELSE IF (NINT(time)>300.AND.NINT(time)<=600)THEN wall_coeff=0.2D0 ELSE IF (NINT(time)>600.AND.NINT(time)<=900)THEN wall_coeff=0.4D0 ELSE IF (NINT(time)>900.AND.NINT(time)<=1200)THEN wall_coeff=0.6D0 ELSE IF (NINT(time)>1200.AND.NINT(time)<=1500)THEN wall_coeff=0.75D0 ELSE wall_coeff=0.750D0 END IF T_wall(i) = T_a+(T_s(i)-T_a)*wall_coeff IF (T_s(i)>T_g(i)) THEN T_wall(i) = T_a+(T_g(i)-T_a)*wall_coeff ELSE T_wall(i) = T_a+(T_s(i)-T_a)*wall_coeff END IF IF (T_s(N)<620.0) THEN T_701 = T_s(N) -(-2.08999D-12*T_s(N)**6.0 + 6.22059D-09*T_s(N)**5.0 - & 7.55365D-06*T_s(N)**4.0 +4.78617D-03*T_s(N)**3.0 - &

283

Appendix D Computer code listing

1.66904D+00*T_s(N)**2.0+ 3.04147D+02*T_s(N) - 2.26665D+04) IF(T_701>T_s(N))T_701=T_s(N) ELSE T_701 = T_s(N) -( 5.67296D-13*T_s(N)**6.0 - 2.69520D-09*T_s(N)**5.0 + & 5.29841D-06*T_s(N)**4.0 - 5.51380D-03*T_s(N)**3.0 + & 3.20165D+00*T_s(N)**2.0 - 9.82866D+02*T_s(N) + 1.24565D+05) END IF ! ----------! Soild Phase ! ----------IF (i==N) THEN delta_x1 = x(i)-x(i-1)

solid1 = (1.D0-void)*(rho_m(i)*Cp_m+(rho_volt(i)*Cp_wood(T_s(i))+ & rho_char(i)*Cp_char(T_s(i)))) solid2 = (1.D0-void)*(rho_m(i-1)*Cp_m+(rho_volt(i-1)*Cp_wood(T_s(i-1))+ & rho_char(i-1)*Cp_char(T_s(i-1)))) solid3 = (1.D0-void)*(rhoP_m(i)*Cp_m+(rhoP_volt(i)*Cp_wood(T_s(i))+ & rho_char(i)*Cp_char(T_s(i)))) solid4 = h_rs(T_s(i),emiss)*(1.D0-void) solid5 = K_eff(m_g(i),T_s(i),T_g(i),d_p(i),emiss,void)+ & K_eff(m_g(i-1),T_s(i-1),T_g(i-1),d_p(i-1),emiss,void) A_Ts = (1.D0/delta_t+v_s(i)/delta_x1)*solid1+2.D0*solid4/(delta_x1)+ & solid5/(delta_x1**2.0)+h_sg*A_v+4.D0*h_sw/D_rec B_Ts = 0.D0 C_Ts = solid5/(delta_x1**2.0)+solid2*v_s(i-1)/delta_x1 D_Ts = (1.D0/delta_t)*(solid1*T_a+solid3*TP_s(i)-solid3*T_a)+ & solid1*V_s(i)*T_a/delta_x1-solid2*V_s(i-1)*T_a/delta_x1+ h_sg*A_v*T_g(i)+ & 4.D0*h_sw*T_wall(i)/D_rec+ratep1(i)*heatp1+ & rate1(i)*heat1(T_s(i))+rate2(i)*heat2(T_s(i))+rate3(i)*heat3(T_s(i))+ & rate4(i)*heat4(T_s(i))+ heatm*ratem(i)+2.D0*solid4*T_701/(delta_x1) P_Ts(i) = B_Ts/(A_Ts-C_Ts*P_Ts(i-1)) Q_Ts(i) = (D_Ts+C_Ts*Q_Ts(i-1))/(A_Ts-C_Ts*P_Ts(i-1)) ELSE delta_x1 delta_x2 delta_x = x(i)-x(i-1) = x(i+1)-x(i) = (x(i+1)-x(i-1))/2.D0

solid1 = (1.D0-void)*(rho_m(i)*Cp_m+(rho_volt(i)*Cp_wood(T_s(i))+ & rho_char(i)*Cp_char(T_s(i)))) solid2 = (1.D0-void)*(rho_m(i-1)*Cp_m+(rho_volt(i-1)*Cp_wood(T_s(i-1))+ & rho_char(i-1)*Cp_char(T_s(i-1)))) solid3 = (1.D0-void)*(rhoP_m(i)*Cp_m+(rhoP_volt(i)*Cp_wood(T_s(i))+ & rho_char(i)*Cp_char(T_s(i)))) solid4 = (K_eff(m_g(i),T_s(i),T_g(i),d_p(i),emiss,void)+ & K_eff(m_g(i+1),T_s(i+1),T_g(i+1),d_p(i+1),emiss,void))/2.D0 solid5 = (K_eff(m_g(i),T_s(i),T_g(i),d_p(i),emiss,void)+ & K_eff(m_g(i-1),T_s(i-1),T_g(i-1),d_p(i-1),emiss,void))/2.D0 A_Ts = (1.D0/delta_t+v_s(i)/delta_x1)*solid1+solid4/(delta_x*delta_x2)+ & solid5/(delta_x*delta_x1)+ h_sg*A_v+4.D0*h_sw/D_rec B_Ts = solid4/(delta_x*delta_x2) C_Ts = solid5/(delta_x*delta_x1)+solid2*v_s(i-1)/delta_x1 D_Ts = (1.D0/delta_t)*(solid1*T_a+solid3*TP_s(i)-solid3*T_a)+ & solid1*V_s(i)*T_a/delta_x1-solid2*V_s(i-1)*T_a/delta_x1+ h_sg*A_v*T_g(i)+ & 4.D0*h_sw*T_wall(i)/D_rec+ratep1(i)*heatp1+ & rate1(i)*heat1(T_s(i))+rate2(i)*heat2(T_s(i))+rate3(i)*heat3(T_s(i))+ & rate4(i)*heat4(T_s(i)) + heatm*ratem(i) P_Ts(i) = B_Ts/(A_Ts-C_Ts*P_Ts(i-1)) Q_Ts(i) = (D_Ts+C_Ts*Q_Ts(i-1))/(A_Ts-C_Ts*P_Ts(i-1)) END IF ! --------! Gas Phase ! ---------

284

Appendix D Computer code listing

IF (i==N) THEN delta_x1 gas1 gas2 gas3 gas4 gas5 = = = = = = x(i)-x(i-1)

rho_g(i)*Cp_g(T_g(i)) m_g(i-1)*Cp_g(T_g(i-1)) rhoP_g(i)*Cp_g(TP_g(i)) 2.D0*void*h_rv(T_g(N),emiss,void)/(1.0-void) K_g(T_g(i))+K_g(T_g(i-1))

gasheat(i) = alpha*(rate6(i)*heat6(T_g(i))+rate7(i)*heat7(T_g(i))+& rate8(i)*heat8(T_g(i))+rate9(i)*heat9(T_g(i))+& rate10(i)*heat10(T_g(i)))+(1.-alpha)*gasheat(i) A_Tg = void*gas1/delta_t+gas4/(delta_x1)+gas5/(delta_x1**2)+& m_g(i)*Cp_g(T_g(i))/delta_x1+h_sg*A_v+ 4.D0*h_gw/D_rec-& ratep2(i)*heatp2*Ep2*ep2_mf/(R*T_g(i)**2.D0) B_Tg = 0.D0 C_Tg = gas5/(delta_x1**2)+gas2/delta_x1 D_Tg = (void/delta_t)*(gas1*T_a+gas3*TP_g(i)-gas3*T_a)+ & m_g(i)*Cp_g(T_g(i))*T_a/delta_x1-gas2*T_a/delta_x1+h_sg*A_v*T_s(i)+ & 4.D0*h_gw*T_wall(i)/D_rec+rate5(i)*heat5(T_g(i))+rate5R(i)*heat5R(T_g(i))+& gasheat(i) +heatp2*ratep2(i)*(1.D0-Ep2*ep2_mf/(R*T_g(i)))+gas4*T_701/(delta_x1) P_Tg(i) = B_Tg/(A_Tg-C_Tg*P_Tg(i-1)) Q_Tg(i) = (D_Tg+C_Tg*Q_Tg(i-1))/(A_Tg-C_Tg*P_Tg(i-1)) ELSE delta_x1 delta_x2 delta_x gas1 gas2 gas3 gas4 gas5 = = = = = = x(i)-x(i-1) = x(i+1)-x(i) = (x(i+1)-x(i-1))/2.D0

rho_g(i)*Cp_g(T_g(i)) m_g(i-1)*Cp_g(T_g(i-1)) rhoP_g(i)*Cp_g(TP_g(i)) (K_g(T_g(i))+K_g(T_g(i+1)))/2.D0 (K_g(T_g(i))+K_g(T_g(i-1)))/2.D0

gasheat(i) = alpha*(rate6(i)*heat6(T_g(i))+rate7(i)*heat7(T_g(i))+& rate8(i)*heat8(T_g(i))+rate9(i)*heat9(T_g(i))+& rate10(i)*heat10(T_g(i)))+(1.-alpha)*gasheat(i) A_Tg = void*gas1/delta_t+gas4/(delta_x*delta_x2)+gas5/(delta_x*delta_x1)+& m_g(i)*Cp_g(T_g(i))/delta_x1+h_sg*A_v+ 4.D0*h_gw/D_rec B_Tg = gas4/(delta_x*delta_x2) C_Tg = gas5/(delta_x*delta_x1)+gas2/delta_x1 D_Tg = (void/delta_t)*(gas1*T_a+gas3*TP_g(i)-gas3*T_a)+ & m_g(i)*Cp_g(T_g(i))*T_a/delta_x1-gas2*T_a/delta_x1+h_sg*A_v*T_s(i)+ & 4.D0*h_gw*T_wall(i)/D_rec+rate5(i)*heat5(T_g(i))+& rate5R(i)*heat5R(T_g(i))+ gasheat(i) +heatp2*ratep2(i) P_Tg(i) = B_Tg/(A_Tg-C_Tg*P_Tg(i-1)) Q_Tg(i) = (D_Tg+C_Tg*Q_Tg(i-1))/(A_Tg-C_Tg*P_Tg(i-1)) END IF RETURN END SUBROUTINE thomasenergy_balance ! *********************END SUBROUTINE thomasenergy_balance********************************* SUBROUTINE thomas(N,P,Q,U) IMPLICIT NONE ! This subroutine is to calculate the solution of a numbers of linear algebraic equations ! in the form of tridiagonal matrix by Thomas Algorithm. INTEGER , REAL(DP), REAL(DP), REAL(DP), INTENT(IN) INTENT(IN) INTENT(IN) INTENT(OUT) :: :: :: :: N P(N) Q(N) U(N) ! ! ! ! Number of equations Coefficient Solution at old mesh points Solution of a numbers of linear equations

INTEGER :: i DO i=N,1,-1 IF (i==N) THEN

285

Appendix D Computer code listing

U(i)=Q(i) ELSE U(i)=P(i)*U(i+1)+Q(i) END IF END DO RETURN END SUBROUTINE thomas ! *********************END SUBROUTINE thomas*********************************************** SUBROUTINE temperatures() ! This subroutine is to calculate the solid and gas temperature by Thomas Algorithm. ! ----------! Soild Phase ! ----------IF (i==N) THEN T_s(i)=Q_Ts(i) ELSE T_s(i)=P_Ts(i)*T_s(i+1)+Q_Ts(i) END IF ! --------! Gas Phase ! --------IF (i==N) THEN T_g(i)=Q_Tg(i) ELSE T_g(i)=P_Tg(i)*T_g(i+1)+Q_Tg(i) END IF RETURN END SUBROUTINE temperatures ! *********************END SUBROUTINE temperatures***************************************** SUBROUTINE mass_balance() ! This subroutine is to check the mass balance of the system. mole_g(N) = Y_O2(N)*mole_O2+Y_CO(N)*mole_CO +Y_CO2(N)*mole_CO2+Y_H2O(N)*mole_H2O+ & Y_H2(N)*mole_H2+Y_CH4(N)*mole_CH4+Y_C2H4(N)*mole_C2H4+Y_tar1(N)*mole_tar1+ & Y_tar2(N)*mole_tar2+Y_N2(N)*mole_N2 masstotal_in = masstotal_out = C_in H_in O_in N_in V_s(1)*(1.D0-void)*(rho_m(1)+rho_volt(1)+rho_char(1))+m_g(1) m_g(N)

= (12.D0/23.D0)*V_s(1)*(1.D0-void)*(rho_volt(1)+rho_char(1)) = (1.4D0/23.D0)*V_s(1)*(1.D0-void)*(rho_volt(1)+rho_char(1))+ & V_s(1)*(1.D0-void)*rho_m(1)*mole_H2/mole_H2O = (9.6D0/23.D0)*V_s(1)*(1.D0-void)*(rho_volt(1)+rho_char(1))+& V_s(1)*(1.D0-void)*rho_m(1)*0.5D0*mole_O2/mole_H2O+m_g(1)*mole_O2*Y_O2(1)/mole_g(1) = m_g(1)*mole_N2*Y_N2(1)/mole_g(1)

C_out = mole_C*m_g(N)*Y_CO(N)/mole_g(N)+ mole_C*m_g(N)*Y_CO2(N)/mole_g(N)+ & mole_C*m_g(N)*Y_CH4(N)/mole_g(N)+(2.D0*mole_C)*m_g(N)*Y_C2H4(N)/mole_g(N) + & (6.D0*mole_C)*m_g(N)*Y_tar1(N)/mole_g(N) +(6.D0*mole_C)*m_g(N)*Y_tar2(N)/mole_g(N) H_out = m_g(N)*mole_H2*Y_H2(N)/mole_g(N)+mole_H2*m_g(N)*Y_H2O(N)/mole_g(N) + & (2.D0*mole_H2)*m_g(N)*Y_CH4(N)/mole_g(N) + & (2.D0*mole_H2)*m_g(N)*Y_C2H4(N)/mole_g(N) + (10.71D0*0.5D0*mole_H2)* & m_g(N)*Y_tar1(N)/mole_g(N)+(3.D0*mole_H2)*m_g(N)*Y_tar2(N)/mole_g(N) O_out = m_g(N)*mole_O2*Y_O2(N)/mole_g(N)+(mole_O2*0.5D0)*m_g(N)*Y_CO(N)/mole_g(N)+ & mole_O2*m_g(N)*Y_CO2(N)/mole_g(N)+(mole_O2*0.5D0)*m_g(N)*Y_H2O(N)/mole_g(N) + & (mole_O2*0.5D0*3.264D0)*m_g(N)*Y_tar1(N)/mole_g(N) N_out = m_g(N)*mole_N2*Y_N2(N)/mole_g(N) RETURN END SUBROUTINE mass_balance ! *********************END SUBROUTINE mass_balance***************************************** SUBROUTINE Tg_peak() ! This subroutine calculate the location of gas peak temperature at a particular time

286

Appendix D Computer code listing

IF (NINT(time)==300) THEN CALL peakTg_x () maxTg_x05 = maxTg_x END IF IF (NINT(time)==600) THEN CALL peakTg_x () maxTg_x10 = maxTg_x END IF IF (NINT(time)==900) THEN CALL peakTg_x () maxTg_x15 = maxTg_x END IF IF (NINT(time)==1200) THEN CALL peakTg_x () maxTg_x20 = maxTg_x END IF IF (NINT(time)==1500) THEN CALL peakTg_x () maxTg_x25 = maxTg_x END IF IF (NINT(time)==1800) THEN CALL peakTg_x () maxTg_x30 = maxTg_x END IF RETURN END SUBROUTINE Tg_peak ! *********************END SUBROUTINE Tg_peak********************************************** FUNCTION IMPLICIT ! ! ! ! ! sign_test(arg1,arg2) NONE This is the sign test routine Returns: -1. if ARG1 and ARG2 are of opposite sign. 0. if either argument is zero. +1. if ARG1 and ARG2 are of the same sign. :: arg1 :: arg2

REAL(DP), INTENT(IN) REAL(DP), INTENT(IN) REAL(DP) :: sign_test

sign_test = SIGN(1.D0,arg1) * SIGN(1.D0,arg2) IF ((arg1 == 0.D0) .OR. (arg2 == 0.D0)) sign_test = 0.D0 RETURN END FUNCTION sign_test ! ****************************END FUNCTION sign_test*************************************** SUBROUTINE rezone(N,xold,f,xnew) IMPLICIT NONE ! ! This subroutine is to distribute the new mesh nodes by global static rezoning methods based on the equidistribution principle and arclength monitor. ! ! ! ! ! ! ! ! Number of mesh points Location of old mesh points Solution at old mesh points Location of new mesh points Mesh monitor function by the arclength Total arclength sloution from xold(1) to xold(i) Total arclength from xnew(1) to xnew(i) Equal arclength based on equidistribution principle

INTEGER , INTENT(IN) :: N REAL(DP), INTENT(IN) :: xold(N) REAL(DP), INTENT(IN) :: f(N) REAL(DP), INTENT(OUT) :: xnew(N) REAL(DP):: moni (N-1) REAL(DP):: arc (N) REAL(DP):: newarc (N) REAL(DP):: egrid INTEGER :: i,j,upper, lower ! Find the mesh monitor function

DO i=1,N-1 moni(i)= SQRT(1.d0+((f(i+1)-f(i))/(xold(i+1)-xold(i)))**2.) END DO

287

Appendix D Computer code listing

! Calculate the total arclength of curve arc(1)=0.D0 DO i=2,N arc(i)=moni(i-1)*(xold(i)-xold(i-1))+arc(i-1) END DO ! Find the equal arclength by equidistribution principle egrid=arc(N)/real(N-1) ! Estimate the total arclength of new mesh newarc(1)=arc(1) DO i=2,N newarc(i)=real(i-1)*egrid END DO ! Evaluate the new grids location lower=2 xnew(1)=xold(1) DO j=2,N-1 DO i=lower,N IF(arc(i)-newarc(j)>=0.D0)THEN lower=i-1 upper=i EXIT END IF END DO xnew(j)= xold(lower)+(newarc(j)-arc(lower))* & ((xold(upper)-xold(lower))/(arc(upper)-arc(lower))) END DO xnew(N)=xold(N) RETURN END SUBROUTINE rezone ! *********************************END SUBROUTINE rezone*********************************** SUBROUTINE deriv (N, x, f, D ) ! ! This subroutine is to set derivatives needed to determine a monotone piecewise cubic Hermite interpolant to the data given in X and F.

INTEGER, INTENT(IN) REAL (DP), INTENT(IN) REAL (DP), INTENT(IN) REAL (DP), INTENT(OUT)

:: :: :: ::

N x(N) f(N) D(N)

! ! ! !

Number of data points. Independent variable values Dependent variable values to be interpolated. Derivative values at the data points.

REAL (DP) :: slope1, slope2,W1,W2,Dmax,Dmin INTEGER::i slope1 = (f(2)-f(1))/(x(2)-x(1)) slope2 = (f(3)-f(2))/(x(3)-x(2)) W1= (x(3)+x(2)-2*x(1))/(x(3)-x(1)) W2= (X(1)-X(2))/(x(3)-x(1)) D(1)=W1*slope1+W2*slope2 IF (sign_test(D(1),slope1)<0.D0) THEN D(1)=0.D0 ELSE IF (sign_test(slope1,slope2)<=0.D0) THEN Dmax=slope1*3.D0 IF (ABS(D(1))>ABS(Dmax)) D(1)=DMax END IF DO i=2,N-1 IF (i>2) THEN

288

Appendix D Computer code listing

slope1=slope2 slope2=(f(i+1)-f(i))/(x(i+1)-x(i)) END IF IF (sign_test(slope1,slope2)<=0.D0) THEN D(i)=0.D0 ELSE W1= (x(i+1)+x(i)-2*x(i-1))/(3.D0*(x(i+1)-x(i-1))) W2= (2*x(i+1)-x(i)-x(i-1))/(3.D0*(x(i+1)-x(i-1))) Dmax=max(ABS(slope1),ABS(slope2)) Dmin=min(ABS(slope1),ABS(slope2)) D(i)=Dmin/((W1*slope1/Dmax)+(W2*slope2/Dmax)) END IF END DO W1=(X(N-1)-x(N))/(X(N)-X(N-2)) W2=(2*x(N)-x(N-1)-x(N-2))/(X(N)-X(N-2)) D(N)=W1*slope1+W2*slope2 IF (sign_test(D(N),slope2)<=0.D0) THEN D(N)=0.D0 ELSE IF (sign_test(slope1,slope2)<0.D0) THEN Dmax=slope2*3.D0 IF (ABS(D(N))>ABS(Dmax))D(N)=DMax END IF RETURN END SUBROUTINE deriv ! ***************************END SUBROUTINE deriv****************************************** SUBROUTINE interp (x_left,x_right,f_left,f_right,d_left,d_right,Ne,xe,fe) IMPLICIT NONE ! ! This subroutine is to evaluate a cubic polynomial given in Hermite form at an array of points. :: :: :: :: :: :: :: :: x_left x_right f_left f_right d_left d_right NE xe(NE) ! ! ! ! ! ! ! ! ! ! ! ! Left point of interval of definition of cubic. Right point of interval of definition of cubic. Values of function at X_left Values of function at X_right Values of derivative at X_left Values of derivative at X_right Number of evaluation points Array of points at which the function is to be evaluated. Array of values of the cubic function defined by X_left,X_right, F_left, F_right, D_left,D_right at the points XE.

REAL(DP), INTENT(IN) REAL(DP), INTENT(IN) REAL(DP), INTENT(IN) REAL(DP), INTENT(IN) REAL(DP), INTENT(IN) REAL(DP), INTENT(IN) INTEGER, INTENT(IN) REAL(DP), INTENT(IN)

REAL(DP), INTENT(OUT) :: fe(NE)

INTEGER :: i REAL(DP) :: x_diff, K1, K2, slope slope=(f_right-f_left)/(x_right-x_left) K1= (d_left+d_right-2.*slope)/((x_right-x_left)**2) K2= -(2.*d_left+d_right-3.*slope)/(x_right-x_left) DO i=1, NE x_diff= XE(i)-x_left fe(i)=f_left+x_diff*(D_left+x_diff*(K2+x_diff*K1)) END DO RETURN END SUBROUTINE interp ! *****************************END SUBROUTINE interp*************************************** SUBROUTINE pch_intp (N,x,f,Ne,xe,fe) IMPLICIT NONE ! ! This subroutine is to evaluate a piecewise cubic Hermite function at an array of points. :: N :: x(N) :: f(N) ! ! ! Number of data points. Independent variable values Dependent variable values to be interpolated.

INTEGER, INTENT(IN) REAL (DP), INTENT(IN) REAL (DP), INTENT(IN)

289

Appendix D Computer code listing

INTEGER, INTENT(IN) REAL(DP), INTENT(IN) REAL(DP), INTENT(OUT)

:: NE :: xe(NE) :: fe(NE)

! ! ! ! ! !

Number of evaluation points Array of points at which the function is to be evaluated. Array of values of the cubic function defined by X_left,X_right, F_left, F_right, D_left,D_right at the points XE.

REAL(DP) :: D(N) INTEGER :: i,j,k, left,right,log=0 CALL deriv (N, x, f, D ) k=2 DO i=1,ne DO j=k,Ne IF (xe(i)-x(j-1)==0.D0)THEN fe(i)=f(j-1) log=1 EXIT ELSE IF (x(j)-xe(i)==0.D0)THEN fe(i)=f(j) log=2 EXIT ELSE IF(xe(i)<x(j)) THEN left=j-1 right=j EXIT END IF END DO k=j IF (log==0) call interp(x(left),x(right),f(left),f(right),d(left),d(right),1,xe(i),fe(i)) IF (log==2)k=j+1 log=0 END DO RETURN END SUBROUTINE pch_intp ! ***********************END SUBROUTINE pch_intp******************************************* SUBROUTINE NON_DIMENT_Tg () ! This subroutine is to non dimentionalize the gas temperature and grids distance. max_Tg = T_g(1) DO i=2,N IF (T_g(i)>T_g(i-1)) max_Tg=T_g(i) END DO DO i=1,N Tg_nd(i)=T_g(i)/max_Tg x_nd(i) = x(i)/L_rec END DO RETURN END SUBROUTINE NON_DIMENT_Tg ! *************************END SUBROUTINE NON_DIMENT ************************************** SUBROUTINE static_rezone () ! This subroutine is to find the new grids and the variables at the new grid points. CALL rezone (N,x_nd,Tg_nd,xnew_nd) ! Change the dimensional system DO i=1,N xnew(i)=xnew_nd(i)*L_rec END DO !Find 18 variables at the new grid points CALL pch_intp (N,x,T_s,N,xnew,Tnew_s) CALL pch_intp (N,x,T_g,N,xnew,Tnew_g) CALL pch_intp (N,x,Y_O2,N,xnew,Ynew_O2)

290

Appendix D Computer code listing

CALL CALL CALL CALL CALL CALL CALL CALL CALL CALL CALL CALL CALL CALL CALL

pch_intp pch_intp pch_intp pch_intp pch_intp pch_intp pch_intp pch_intp pch_intp pch_intp pch_intp pch_intp pch_intp pch_intp pch_intp

(N,x,Y_N2,N,xnew,Ynew_N2) (N,x,Y_CO2,N,xnew,Ynew_CO2) (N,x,Y_CO,N,xnew,Ynew_CO) (N,x,Y_H2O,N,xnew,Ynew_H2O) (N,x,Y_H2,N,xnew,Ynew_H2) (N,x,Y_CH4,N,xnew,Ynew_CH4) (N,x,Y_C2H4,N,xnew,Ynew_C2H4) (N,x,Y_tar1,N,xnew,Ynew_tar1) (N,x,Y_tar2,N,xnew,Ynew_tar2) (N,x,rho_m,N,xnew,rhonew_m) (N,x,rho_char,N,xnew,rhonew_char) (N,x,rho_volt,N,xnew,rhonew_volt) (N,x,m_g,N,xnew,mnew_g) (N,x,v_s,N,xnew,vnew_s) (N,x,d_p,N,xnew,dnew_p)

DO i=1,N x(i) T_s(i) T_g(i) Y_O2(i) Y_N2(i) Y_CO2(i) Y_CO(i) Y_H2O(i) Y_H2(i) Y_CH4(i) Y_C2H4(i) Y_tar1(i) Y_tar2(i) rho_m(i) rho_volt(i) rho_char(i) m_g(i) v_s(i) d_p(i) END DO RETURN END SUBROUTINE static_rezone ! ************************* END SUBROUTINE static_rezone SUBROUTINE peakTg_x () ! This subroutine is to find the location of peak gas temperature. max_Tg = T_g(1) DO i=2,N IF (T_g(i)>T_g(i-1)) THEN max_Tg=T_g(i) maxTg_x=x(i) END IF END DO RETURN END SUBROUTINE peakTg_x ! *************************END SUBROUTINE peakTg_x **************************************** SUBROUTINE coldgas_efficiency() ! This subroutine is to find the location of peak gas temperature. DO i=2,N efficiency(i)=m_g(i)*pi*D_rec**2.*900.D0*273.15D0/(T_g(i)*rho_g(i))* & HV_gas(Y_CO(i),Y_H2(i),Y_CH4(i),Y_H2O(i))*100.D0/ & ((m_g(i)*pi*D_rec**2.*900.D0-air_rate)*100.D0/(moisture+ash+100.D0)*& HV_biomass(Carbon,Hydrogen,Oxygen,Nitrogen,Sulfur,Ash,Moisture)) END DO RETURN END SUBROUTINE coldgas_efficiency = = = = = = = = = = = = = = = = = = = xnew(i) Tnew_s(i) Tnew_g(i) Ynew_O2(i) Ynew_N2(i) Ynew_CO2(i) Ynew_CO(i) Ynew_H2O(i) Ynew_H2(i) Ynew_CH4(i) Ynew_C2H4(i) Ynew_tar1(i) Ynew_tar2(i) rhonew_m(i) rhonew_volt(i) rhonew_char(i) mnew_g(i) vnew_s(i) dnew_p(i)

*********************************

291

Appendix D Computer code listing

! *************************END SUBROUTINE coldgas_efficiency ****************************** SUBROUTINE gasifier_performance() ! This subroutine calculate the gasifier performance summary ! Calculate the fuel consumption in wet basis (kg/hr) fuel_flow = V_s(1)*(1.D0-void)*(rho_m(1)+rho_volt(1)+rho_char(1))*pi*D_rec**2.*900.D0 ! Calculate the specific gasification rate in dry basic (kg/m2hr) spe_gasi_rate = fuel_flow*4.0D0*100.D0/((pi*D_rec**2.)*(moisture+100.D0)) ! Calculate the gas mass production rate (kg/hr) gas_flow_mass = m_g(N)*pi*D_rec**2.*900.D0 ! Calculate the gas volume production rate (Nm/hr) gas_flow_vol = gas_flow_mass*273.15D0/(T_g(N)*rho_g(N)) ! Calculate the outlet condensable tar content (mg/Nm) tar_outlet = Y_tar1(N)*Mole_tar1*Pre*1.D+06/(R*273.15D0) ! Calculate air fuel ratio (- kg/kg d.b) air_fuel= m_g(1)*(moisture+100.D0)*pi*D_rec**2.*900.D0/(fuel_flow*100.D0) ! Calculate eqivalent ratio ER = air_fuel*0.233D0/1.463D0 ! (Note:mass of oxygen required for complete combustion of biomass/ mass of biomass= 1.463) ! (Note:mass of Oxygen in 1 kg of air = 0.233) ! Calcuate the net heating value of producer gas in dry basic (MJ/Nm3) LHV_gas = HV_gas(Y_CO(N),Y_H2(N),Y_CH4(N),Y_H2O(N)) ! Calculate the total energy output to the gasifier ! based on the net heating value to producer gas in dry basic [MJ/hr] Energy_out = gas_flow_vol*LHV_gas RETURN END SUBROUTINE gasifier_performance ! *********************END SUBROUTINE gasifier_performance********************************* SUBROUTINE input() ! This subroutine print out the input parameter. Write(7,'(/T2, A34)') " --------------------------------- " Write(7,'(T2, A34)') " USER INPUT OPERATIONAL PARAMETERS " Write(7,'(T2, A34)') " --------------------------------- " Write(7,'(/T2, A34,F8.2)') Write(7,'(T2, A34,F8.2)') Write(7,'(T2, A34,F8.2)') Write(7,'(T2, A34,F8.2)') Write(7,'(T2, A34,F8.2)') Write(7,'(T2, A34,F8.2)') Write(7,'(T2, A34,F8.2)') Write(7,'(T2, A34,F9.3)') Write(7,'(T2, A34,F8.2)') Write(7,'(T2, A34,F8.2)') Write(7,'(T2, A34,F8.2//)') "Reactor daiameter [mm] "Wood apparent density [kg/hr] "Moisture content [%Db] "Paticle diameter [mm] "Bed voidage fraction [-] "Initial depth of char bed [mm] "Feed air temperature [K] "Feed air pressure [kPa] "Ignition temperature [K] "Ignited location (mm below top) "Initial ignition depth [mm] =", =", =", =", =", =", =", =", =", =", =", D_rec*1000.0 rho_biomass moisture d_pwood*1000.0 void depth_char*1000.0 T_gas Pre T_ign loc_ign*1000.0 depth_ign*1000.0

IF (check_stage==1) THEN Write(7,'(T2, A55)') "Gasifier is operated by supplying air from reactor top" Write(7,'(T2, A55)') "------------------------------------------------------" Write(7,'(T2, A34,F8.2)') "Air supply rate [kg/hr] =", air_rate ELSE IF (check_stage==2) THEN Write(7,'(T2, A47)') "Gasifier is operated with two-stage air supply" Write(7,'(T2, A47)') "----------------------------------------------" Write(7,'(T2, A36,F8.2)') "Primary air supply rate [kg/hr] =", air_rate Write(7,'(T2, A36,F8.2)') "Secondary air supply rate [kg/hr] =", secondair Write(7,'(T2, A33)') "Location of secondary air supply" Write(7,'(T2, A36,F8.2//)')"(mm below reactor top) =", loc_secair*1000.0 END IF Write(7,'(/T2, A24)') " ---------------------- " Write(7,'(T2, A24)') " INPUT MODEL PARAMETERS "

292

Appendix D Computer code listing

Write(7,'(T2, A24)')

" ---------------------- "

Write(7,'(/T2, A62,F5.2)') & "zeta (reacting to non-reacting ratio of heat transfer) [-] Write(7,'(T2, A62,F6.3)') & "Defined maximum mass transfer coefficient [m/s] Write(7,'(T2, A62,F5.2)') & "Ratio to effective and actural transport surface area [-] Write(7,'(T2, A62,F5.2)') & "Multiplication factor of primary pyrolysis rate [-] Write(7,'(T2, A62,F5.2)') & "Multiplication factor of secondary pyrolysis rate [-] Write(7,'(T2, A62,F5.2)') & "Multiplication factor of secondary tar oxidation rate [-] Write(7,'(T2, A62,F5.2)') & "Multiplication factor to Effictive thermal condutivity [-] Write(7,'(T2, A62,F5.2/)') & "Multiplication factor of wall heat transfer coefficient [-]

=", =", =", =", =", =", =", =",

zeta km_max Av_mf Ap1_mf Ap2_mf A6_mf Keff_mf wall_loss

RETURN END SUBROUTINE input ! *********************END SUBROUTINE input************************************************ SUBROUTINE output_profile() ! This subroutine print out the results at specific time. ! Write the out put ! ----------------Write(7,'(/T50, A)') "=============================================" Write(7,'(T50, A20, T70, I5, T76, A19)') & "Model performance at", output_time, "sec after ignition" Write(7,'(T50, A)') "=============================================" 100 FORMAT (' ',T3,F10.6,T14,F8.2,T24,F8.2,T34,F6.1,T40,F6.1,T48,F9.4,T58,F9.4,T68,& F9.4,T78,F9.4,T88,F9.4,T98,F9.4,T108,F9.4,T118,F9.5,T128,F9.4,T138,& F9.4,T148,F9.5,T160,F12.10,T172,F15.5,T189,F9.5) 200 FORMAT (/,T6,A5,T14,A8,T24,A7,T34,A6,T40,A6,T48,A7,T58,A7,T68,A8,T78,A8,T88,& A7,T98,A8,T108,A9,T118,A9,T128,A9,T138,A7,T148,A7,T160,A6,T172,A15,T189,A10) 300 FORMAT (' ',T6,A5,T14,A8,T24,A7,T34,A6,T40,A6,T48,A7,T58,A7,T68,A8,T78,A8,T88,& A7,T98,A8,T108,A9,T118,A9,T128,A9,T138,A7,T148,A7,T160,A6,T172,A15,T189,A10) Write(7,200) "no","T_solid","T_gas","rho_m","rho_V","Y_O2","Y_CO","Y_CO2","Y_H2O","Y_H2",& "Y_CH4","Y_C2H4","Y_tar1","Y_tar2","Y_N2","v_g","v_s","tar_content","efficiency" Write(7,300) "==","=======","=====","=====","=====","====","====","=====","=====","====",& "=====","======","======","======","====","===","===","===========","==========" DO i=1,N WRITE(7,100)700.-x(i)*1000.,T_s(i),T_g(i),rho_m(i),rho_volt(i),Y_O2(i)*100.,& Y_CO(i)*100.,Y_CO2(i)*100.,Y_H2O(i)*100.,Y_H2(i)*100.,Y_CH4(i)*100.,& Y_C2H4(i)*100.,Y_tar1(i)*100.,Y_tar2(i)*100.,Y_N2(i)*100.,& m_g(i)/rho_g(i),v_s(i),Y_tar1(i)*Mole_tar1*Pre*1.D+06/(R*273.15D0),efficiency(i) END DO RETURN END SUBROUTINE output_profile ! *********************END SUBROUTINE output_profile*************************************** SUBROUTINE rates_output() ! This subroutine print out the reaction rates profile (kg/m3/sec). Write(7,'(/T50, A)') "==========================================" Write(7,'(T50, A17, T67, I5, T73, A19)') & "Reaction rates at", output_time, "sec after ignition" Write(7,'(T50, A)') "==========================================" 1000 FORMAT (' ',T3,F10.6,14ES15.5) 2000 FORMAT (' ',T6,A5,14A15) Write(7,2000) "no","drying","pri_pyro","sec_pyro","rate1","rate2","rate3","rate4",& "rate5","rate5R","rate6","rate7","rate8","rate9","rate10" Write(7,2000) "==","======","========","========","=====","=====","=====","=====",& "=====","======","=====","=====","=====","=====","======"

293

Appendix D Computer code listing

DO i=1,N write(7,1000)700.-x(i)*1000.0,ratem(i),ratep1(i),ratep2(i),rate1(i),rate2(i),rate3(i),& rate4(i),rate5(i)*28.D0,rate5R(i)*28.D0,rate6(i)*78.D0,rate7(i)*28.D0,& rate8(i)*16.D0,rate9(i)*28.D0,rate10(i)*2.D0 END DO RETURN END SUBROUTINE rates_output ! *********************END SUBROUTINE rates_output***************************************** SUBROUTINE output_summary() ! This subroutine print out the results of reactor performance. ! Write the output summary ! -----------------------Write(*,'(//T3,A)') "----------------------------------" Write(*,'(T3,A)') "REACTOR PERFORMANCE RESULT SUMMARY" Write(*,'(T3,A)') "----------------------------------" Write(*,'(/T2, A46,F10.2)')"Fuel consumption rate (kgw.b/hr) write(*,'(T2, A46,F10.2)') "Specific gasification rate (kg d.b/mhr) Write(*,'(T2, A46,F10.2)') "Net heating value of gas (MJ/Nm3 d.b) Write(*,'(T2, A46,F10.2)') "Gas production rate (Nm/hr) Write(*,'(T2, A46,F10.2)') "Primary tar Out let (mg/Nm3) Write(*,'(T2, A46,F10.3)') "Air_fuel ratio [(-) kg/kgd.b] Write(*,'(T2, A46,F10.3)') "Equivalence ratio (-) Write(*,'(T2, A46,F10.2)') "Total Energy out(MJ/hr) Write(*,'(T2, A46,F10.2/)')"Cold gas efficiency (%) =", =", =", =", =", =", =", =", =", fuel_flow spe_gasi_rate LHV_gas gas_flow_vol tar_outlet air_fuel ER Energy_out Efficiency(N)

RETURN END SUBROUTINE output_summary ! *********************END SUBROUTINE output_summary*************************************** SUBROUTINE output_summary7() ! This subroutine print out the results of reactor performance in output file. ! Write the output summary ! -----------------------Write(7,'(//T4,A)') "----------------------------------" Write(7,'(T4,A)') "REACTOR PERFORMANCE RESULT SUMMARY" Write(7,'(T4,A)') "----------------------------------" Write(7,'(/T2, A47,F10.2)')"Fuel consumption rate (kgw.b/hr) write(7,'(T2, A47,F10.2)') "Specific gasification rate (kg d.b/mhr) Write(7,'(T2, A47,F10.2)') "Net heating value of gas (MJ/Nm3 d.b) Write(7,'(T2, A47,F10.2)') "Gas production rate (Nm/hr) Write(7,'(T2, A47,F10.2)') "Primary tar Out let (mg/Nm3) Write(7,'(T2, A47,F10.3)') "Air_fuel ratio [(-) kg/kgd.b] Write(7,'(T2, A47,F10.3)') "Equivalence ratio (-) Write(7,'(T2, A47,F10.2)') "Total Energy out(MJ/hr) Write(7,'(T2, A47,F10.2/)')"Cold gas efficiency (%) IF (check_stage==1) THEN IF(output_time>600) Write(7,'(T3, A50,F10.6)') & "Zone movement between 05 min and 10 min (mm/sec) =", (maxTg_x05-maxTg_x10)/0.3D0 IF (output_time>900) Write(7,'(T3, A50,F10.6)') & "Zone movement between 10 min and 15 min (mm/sec) =", (maxTg_x10-maxTg_x15)/0.3D0 IF (output_time>1200) Write(7,'(T3, A50,F10.6)') & "Zone movement between 15 min and 20 min (mm/sec) =", (maxTg_x15-maxTg_x20)/0.3D0 IF (output_time>1500) Write(7,'(T3, A50,F10.6)') & "Zone movement between 20 min and 25 min (mm/sec) =", (maxTg_x20-maxTg_x25)/0.3D0 IF (output_time>=1800) Write(7,'(T3, A50,F10.6)') & "Zone movement between 25 min and 30 min (mm/sec) =", (maxTg_x25-maxTg_x30)/0.3D0 IF (output_time==1800) Write(7,'(T3, A50,F10.6/)') & "Average Zone movement between(mm/sec) =", (maxTg_x05-maxTg_x30)/1.5D0 =", =", =", =", =", =", =", =", =", fuel_flow spe_gasi_rate LHV_gas gas_flow_vol tar_outlet air_fuel ER Energy_out Efficiency(N)

294

Appendix D Computer code listing

END IF RETURN END SUBROUTINE output_summary7 ! *********************END SUBROUTINE output_summary*************************************** SUBROUTINE READ_TIME (run_time,dt_init,dt_final) REAL, INTENT(OUT):: run_time INTEGER, DIMENSION(8), INTENT(IN):: dt_init, dt_final ! This subroutine is to estimate the running time of program IF (dt_final(5)<dt_init(5)) THEN run_time = real(((dt_final(7)+dt_final(6)*60+(dt_final(5)+24)*3600)*1000+dt_final(8))-& ((dt_init(7)+dt_init(6)*60+dt_init(5)*3600)*1000+dt_init(8)))/1000. ELSE run_time = real(((dt_final(7)+dt_final(6)*60+dt_final(5)*3600)*1000+dt_final(8))-& ((dt_init(7)+dt_init(6)*60+dt_init(5)*3600)*1000+dt_init(8)))/1000. END IF RETURN END SUBROUTINE READ_TIME ! ************************* END SUBROUTINE READ_TIME ************************************** END MODULE gasification_data !========================================================================================== ! END MODULE !==========================================================================================

295

Appendix D Computer code listing

!========================================================================================== ! START MAIN PROGRAM !========================================================================================== PROGRAM gasification ! This program is a transient model for stratified downdraft wood gasifier. ! Outputs of the model are transient solid and gas temperature profile along the gasifier & ! density of different species, gas composition (Including tar) porfile along the gasifier. USE gasification_data CALL user_input() call date_and_time (date, t, zone, dt_init) ! Open output file. OPEN ( UNIT=7, FILE=filename, ACCESS='SEQUENTIAL', & FORM='FORMATTED', STATUS='replace',IOSTAT=ierror ) IF (ierror==0) THEN ! Open successful ! Print out the input data CALL input() ! ************************************* ! Initialize the operational parameters ! ************************************* depth_bio = L_rec-depth_char ! For mass balance ! ---------------rho_char(1) = rho_biomass * (1.D0-ash/100.D0)* X1_char rho_m(1) = rho_biomass * (moisture/100.D0) rho_volt(1) = rho_biomass * (1.D0-ash/100.D0)* (1.D0-X1_char) m_g(1) Y_O2(1) Y_N2(1) Y_CO(1) Y_CO2(1) Y_H2O(1) Y_H2(1) Y_CH4(1) Y_C2H4(1) Y_tar1(1) Y_tar2(1) = air_rate*4.D0/(pi*D_rec**2.D0*3600.D0) = mf_O2 = 1.D0-mf_O2 = = = = = = = = 0.D0 0.D0 0.D0 0.D0 0.D0 0.D0 0.D0 0.D0

mole_g(1) = Y_O2(1)*mole_O2+Y_CO(1)*mole_CO +Y_CO2(1)*mole_CO2+Y_H2O(1)*mole_H2O+ & Y_H2(1)*mole_H2+Y_CH4(1)*mole_CH4+Y_C2H4(1)*mole_C2H4+Y_tar1(1)*mole_tar1+& Y_tar2(1)*mole_tar2+Y_N2(1)*mole_N2 ! For Energy balance ! -----------------P_Ts(1) = 0.D0 Q_Ts(1) = T_a P_Tg(1) = 0.D0 Q_Tg(1) = T_gas CALL initial() DO time=time+delta_t CALL pretime() iteration = 0 DO i=2,N gasheat(i)=0.D0 END DO

296

Appendix D Computer code listing

DO logic = 0 DO i=2,N CALL mh_coefficient() CALL reaction_rates() END DO DO i=N,2,-1 CALL solid_velocity() END DO DO i=2,N CALL particle_diameter() END DO sec_check=1 DO i=2,N CALL solid_balance() CALL gas_balance() END DO T_sold(1)=T_s(1) T_gold(1)=T_g(1) DO i=2,N CALL wallh() CALL thomasenergy_balance() T_sold(i)=T_s(i) T_gold(i)=T_g(i) END DO CALL thomas(N,P_Tg,Q_Tg,T_g) CALL thomas(N,P_Ts,Q_Ts,T_s) DO i=1,N IF (ABS((T_s(i)-T_sold(i))/T_s(i))>error.OR.& ABS((T_g(i)-T_gold(i))/T_g(i))>error) logic=1 END DO iteration = iteration + 1 IF (logic==0) EXIT IF (iteration>1000)THEN Write(*,*)'------- CONVERGENCE PROBLEM ----------' EXIT END IF END DO Write(*,"(' ',T2,A7,T10,I6,T18,A4,T25,A18,T45,I5)")& "time =",NINT(time),"sec,","no of iteration =", iteration CALL NON_diment_Tg () CALL static_rezone () CALL coldgas_efficiency() IF IF IF IF IF IF (NINT(time)==300) call (NINT(time)==600) call (NINT(time)==900) call (NINT(time)==1200)call (NINT(time)==1500)call (NINT(time)==1800)call Tg_peak() Tg_peak() Tg_peak() Tg_peak() Tg_peak() Tg_peak()

IF (NINT(time)==output_time) EXIT END DO CALL date_and_time (date, t, zone, dt_final) CALL mass_balance() CALL gasifier_performance() CALL output_summary()

297

Appendix D Computer code listing

CALL READ_TIME (run_time,dt_init,dt_final) Write(*,'(T2, A15, T17, I5, T22, A4, T27, F6.3, T33, A4)') & "Running time =",INT(run_time/60.) ,"min", mod(run_time,60.),"sec" Write(*,'(/T2, A58, T63, A)') & "See the result details in your defined output file called",filename CALL output_profile() CALL rates_output() CALL output_summary7() ELSE Write(*,*)'Error opening file:error=',ierror END IF ! Close file CLOSE (UNIT=7) STOP END PROGRAM gasification !========================================================================================== ! END MAIN PROGRAM !==========================================================================================

References

Fritsch, FN & Carlson, RE 1980, Monotone piecewise cubic interpolation, Siam J Numer Anal, vol. 17, no. 2, pp. 238-46. Lahey Computer Systems 1998, Essential Lahey Fortran 90, Revision D, Lahey Computer Systems, Inc., USA, <http://www.lahey.com/>. Hyman, JM & Naughton, MJ 1983, Static rezone methods for tensor-product grids, Report LAUR-83-3245, Los Alamos National Laboratory, Los Alamos, NM. Pantankar, SV 1980, Numerical heat transfer and fluid flow, New York: McGraw-Hill.

298

Appendix E Air mass flow measurement using orifice plate with D & D/2 tappings

Appendix E

Air mass flow measurement using orifice plate with D & D/2 tappings

A schematic of the orifice plate with D& D/2 tappings is presented in Figure E1. In this section, details of the air mass flow measurement procedure using the orifice plate with D & D/2 tappings is described.

D D

D/2

d
t

Figure E1 Constructional arrangement of experimental orifice plate with D and D/2 tappings

Air mass flow in a D & D/2 tappings orifice plate can be determined using the following equation and its auxiliary equations:

W = 163.27CZEd 2 p (Page 23, BSI 1965)


where W C Z

= mass flow rate (kg/hr) = basic coefficient [Value read from C vs m graph (Page 120, BSI 1965)] = correction factor [Z = ZR * ZD] = expansibility factor [Value read from graph (Page 122, BSI 1965)]

299

Appendix E Air mass flow measurement using orifice plate with D & D/2 tappings

E d p

= velocity approach factor [ = 1

(1 m 2 ) ]

= orifice or throat diameter (inches) = pressure difference (inches H2O) = density of fluid at upstream tapping (lb/ft)

= 2.7[ (P pv ) / KT + 0.62 pv / T ] (Page 40, BSI 1965)

where P T K pv RH pvs

= specific gravity of the dry gas [for air =1 (Page 41, BSI 1965)] = the absolute pressure at the upstream tapping (lb/in) = the absolute temperature at upstream tapping in degrees Rankine (F + 456.67) = gas law deviation coefficient at the temperature T and pressure P [taken as unity (Page 50, BSI 1965)] = partial pressure of water vapour (lb/in) [ = RHpvs ] = relative humidity = saturation vapour pressure at the temperature at the upstream tapping (lb/in) [values can read from table 6, Page 48, BSI 1965] [ = 9.6892E - 06T 3 + 1.9970E - 05T 2 + 8.3783E - 03T + 8.4482E - 02 ]

m ZR ZD

= area ratio (d/D) = Reynold number correction factor [Value read from ZR vs m graph (Page 120, BSI 1965)] = Pipe size correction factor [Value read from ZD vs m graph (Page 121, BSI 1965)]

Reference British Standards Institution 1965, Methods for the measurement of fluid flow in pipes: Part1. Orifice Plates, Nozzles and Venturi Tubes, B.S. 1042 : Part 1 : 1964.

300

Appendix F Experimental test summaries

Appendix F

Experimental test summaries

301

Appendix F Experimental test summaries

Run no: 311

Duration of run: 66 min

Operating Variables
Fuel (wood blocks) properties Size (mm) Moisture content (%wb) 20 x 20 x 20 9.37 Inlet air (Average values) Temperature (C) 21.30 Gauge pressure (kPa) 1.42 Density (kg/m) 1.22 Flow rate (kg/hr) 14.80 Experimental setting Height of charcoal 400 (mm above grate) Ignition level 400 (mm above grate)

Results
Dry gas composition (vol%) first sampling [30 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 3.76 4.05 3.05 0.33 88.81 450 mm above grate 9.76 9.50 11.14 1.10 68.50 350 mm above grate 15.35 16.8 11.79 1.55 54.51 250 mm above grate 15.52 17.36 11.78 1.49 53.85 Outlet gas 14.17 19.93 12.51 1.30 52.09 Dry gas composition (vol%) - Second sampling [45 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 4.61 5.82 4.48 0.50 84.59 450 mm above grate 11.67 12.61 12.89 1.27 61.56 350 mm above grate 13.47 17.38 11.59 1.37 56.19 250 mm above grate 13.74 18.21 11.18 1.37 55.50 Outlet gas 15.50 17.46 11.33 1.33 54.38 Tar content Material Balance Level Tar (mg/Nm) Input (kg/hr) Output (kg/hr) 550 mm above grate 3416 Air 14.80 Gas 20.36 450 mm above grate 129 Dry wood 7.25 Tar 1.00E-03 350 mm above grate 117 Moisture 0.75 H2O 1.22 250 mm above grate 46 Total 22.8 Total 21.58 Outlet gas 58 Closure (%) 94.7 Elemental Balance Input (kg/hr) Output (kg/hr) Closure (%) C 3.84 C 3.15 82.1 O 7.14 O 6.68 93.6 H 0.47 H 0.40 84.8 N 11.35 N 11.35 100.0 Temperature profile(average) Temperature profile(average) Reactor performance summary (30-40) min after ignition (45-55) min after ignition Net HV of 4.68 outlet gas Level above Temperature Level above Temperature (MJ/Nm) grate (mm) (C) grate (mm) (C) 800 22.4 800 22.8 Gas flow rate 17.31 750 22.7 750 23.9 (Nm/hr) 700 23.0 700 38.9 650 25.7 650 57.4 Consumption of 8.00 600 59.1 600 97.0 wood (kg/hr) 550 527.1 550 804.4 500 1188.9 500 1159.4 Air/Fuel ratio 2.04 450 1084.6 450 1183.5 (-) 400 1023.3 400 1140.0 350 939.7 350 976.6 Energy In 135.15 300 859.4 300 816.4 (MJ/hr) 250 777.5 250 743.4 200 717.2 200 702.6 Energy Out 80.99 150 654.1 150 671.9 (MJ/hr) 100 581.0 100 626.7 50 533.0 50 553.8 Cold gas 59.93 efficiency (%) Outlet gas 467.9 Outlet gas 484.1

302

Appendix F Experimental test summaries

Run no: 314

Duration of run: 68 min

Operating Variables
Fuel (wood blocks) properties Size (mm) Moisture content (%wb) 20 x 20 x 20 9.32 Inlet air (Average values) Temperature (C) 21.70 Gauge pressure (kPa) 1.96 Density (kg/m) 1.22 Flow rate (kg/hr) 16.92 Experimental setting Height of charcoal 400 (mm above grate) Ignition level 400 (mm above grate)

Results
Dry gas composition (vol%) first sampling [30 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 1.55 3.63 6.60 1.18 87.05 450 mm above grate 8.49 8.99 13.95 1.82 66.75 350 mm above grate 13.54 14.70 14.06 1.71 56.00 250 mm above grate 17.10 18.68 14.58 1.63 48.03 Outlet gas 17.98 19.02 12.51 1.42 49.07 Dry gas composition (vol%) first sampling [45 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 2.91 3.87 7.00 0.51 85.71 450 mm above grate 11.79 12.69 14.34 1.81 59.36 350 mm above grate 15.26 14.39 12.75 1.62 55.98 250 mm above grate 16.86 16.32 12.82 1.59 52.41 Outlet gas 17.83 18.71 11.49 1.50 50.46 Tar content Material Balance Level Tar (mg/Nm) Input (kg/hr) Output (kg/hr) 550 mm above grate 2994 Air 16.92 Gas 23.76 450 mm above grate 250 Dry wood 8.58 Tar 7.17E-04 350 mm above grate 133 Moisture 0.88 H2O 1.43 250 mm above grate 129 Total 26.38 Total 25.19 Outlet gas 34 Closure (%) 95.5 Elemental Balance Input (kg/hr) Output (kg/hr) Closure (%) C 4.54 C 3.73 82.19 O 8.30 O 7.92 95.41 H 0.55 H 0.55 99.69 N 12.98 N 12.98 100.0 Temperature profile(average) Temperature profile(average) Reactor performance summary (30-40) min after ignition (45-55) min after ignition Net HV of 5.03 outlet gas Level above Temperature Level above Temperature (MJ/Nm) grate (mm) (C) grate (mm) (C) 800 23.6 800 26.1 Gas flow rate 21.08 750 24.1 750 28.7 (Nm/hr) 700 25.2 700 31.2 650 28.2 650 36.4 Consumption of 9.46 600 66.9 600 122.3 wood (kg/hr) 550 504.0 550 854.4 500 994.8 500 1170.9 Air/Fuel ratio 1.97 450 1210 450 1127.6 (-) 400 1018.7 400 1046.7 350 920.9 350 977.0 Energy In 159.90 300 844.1 300 819.9 (MJ/hr) 250 759.3 250 736.2 200 683.0 200 680.8 Energy Out 106.10 150 636.8 150 642.5 (MJ/hr) 100 554.9 100 599.2 50 542.6 50 573.0 Cold gas 66.35 efficiency (%) Outlet gas 487.6 Outlet gas 509.8

303

Appendix F Experimental test summaries

Run no: 318

Duration of run: 72 min

Operating Variables
Fuel (wood blocks) properties Size (mm) Moisture content (%wb) 20 x 20 x 20 9.23 Inlet air (Average values) Temperature (C) 32.60 Gauge pressure (kPa) 2.40 Density (kg/m) 1.18 Flow rate (kg/hr) 20.45 Experimental setting Height of charcoal 500 (mm above grate) Ignition level 500 (mm above grate)

Results
Dry gas composition (vol%) first sampling [30 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 4.69 12.00 12.77 1.65 68.89 450 mm above grate 7.64 14.95 13.84 1.56 62.01 350 mm above grate 14.29 18.90 14.36 1.24 51.20 250 mm above grate 17.44 18.13 14.46 1.46 48.52 Outlet gas 18.22 19.24 14.33 1.49 46.72 Dry gas composition (vol%) first sampling [40 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 6.84 11.34 17.57 1.30 62.95 450 mm above grate 10.67 14.05 13.63 1.45 60.20 350 mm above grate 14.72 16.19 13.15 1.11 54.83 250 mm above grate 15.30 17.11 13.18 1.33 53.08 Outlet gas 17.21 18.95 13.98 1.30 48.56 Tar content Material Balance Level Tar (mg/Nm) Input (kg/hr) Output (kg/hr) 550 mm above grate 1547 Air 20.45 Gas 30.43 450 mm above grate 322 Dry wood 9.63 Tar 1.26E-03 350 mm above grate 233 Moisture 0.98 H2O 1.83 250 mm above grate 105 Total 31.06 Total 32.25 Outlet gas 47 Closure (%) 103.8 Elemental Balance Input (kg/hr) Output (kg/hr) Closure (%) C 5.10 C 5.04 98.89 O 9.65 O 10.81 112.02 H 0.62 H 0.71 114.69 N 15.68 N 15.68 100.0 Temperature profile(average) Temperature profile(average) Reactor performance summary (30-40) min after ignition (40-50) min after ignition Net HV of 5.11 outlet gas Level above Temperature Level above Temperature (MJ/Nm) grate (mm) (C) grate (mm) (C) 800 32.1 800 32.7 Gas flow rate 26.80 750 32.3 750 36.8 (Nm/hr) 700 76.6 700 81.9 650 476.7 650 569.0 Consumption of 10.61 600 1017.4 600 1050.4 wood (kg/hr) 550 1205.7 550 1189.8 500 1112.1 500 1035.7 Air/Fuel ratio 2.12 450 950.4 450 883.2 (-) 400 822.7 400 766.8 350 726.6 350 695.1 Energy In 179.52 300 678.2 300 620.9 (MJ/hr) 250 630.4 250 571.4 200 619.3 200 587.4 Energy Out 137.04 150 604.1 150 573.7 (MJ/hr) 100 545.9 100 572.4 50 513.7 50 538.9 Cold gas 76.34 efficiency (%) Outlet gas 460.4 Outlet gas 483.8

304

Appendix F Experimental test summaries

Run no: 320

Duration of run: 98 min

Operating Variables
Fuel (wood blocks) properties Size (mm) Moisture content (%wb) 20 x 20 x 20 8.65 Inlet air (Average values) Temperature (C) 17.50 Gauge pressure (kPa) 2.41 Density (kg/m) 1.25 Flow rate (kg/hr) 13.32 Experimental setting Height of charcoal 500 (mm above grate) Ignition level 500 (mm above grate)

Results
Dry gas composition (vol%) first sampling [30 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 11.83 11.24 8.14 1.44 67.35 450 mm above grate 16.32 18.55 16.17 1.50 47.46 350 mm above grate 17.50 20.57 12.53 1.20 48.21 250 mm above grate 18.33 20.79 12.32 1.32 47.24 Outlet gas 18.74 18.45 15.11 1.70 46.00 Dry gas composition (vol%) first sampling [45 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 8.87 11.10 6.26 1.22 72.55 450 mm above grate 16.83 18.10 13.70 1.67 49.69 350 mm above grate 19.07 21.13 13.60 1.43 44.78 250 mm above grate 18.40 20.49 13.47 1.37 46.27 Outlet gas 18.28 18.21 14.32 1.43 47.75 Tar content Material Balance Level Tar (mg/Nm) Input (kg/hr) Output (kg/hr) 550 mm above grate 637 Air 13.32 Gas 20.10 450 mm above grate 151 Dry wood 6.88 Tar 6.51E-04 350 mm above grate 134 Moisture 0.65 H2O 1.21 250 mm above grate 107 Total 20.85 Total 21.31 Outlet gas 37 Closure (%) 102.2 Elemental Balance Input (kg/hr) Output (kg/hr) Closure (%) C 3.64 C 3.36 92.11 O 6.55 O 7.25 110.65 H 0.44 H 0.49 110.89 N 10.22 N 10.22 100.0 Temperature profile(average) Temperature profile(average) Reactor performance summary (30-40) min after ignition (45-55) min after ignition Net HV of 5.15 outlet gas Level above Temperature Level above Temperature (MJ/Nm) grate (mm) (C) grate (mm) (C) 800 21.2 800 31.1 Gas flow rate 17.61 750 23.9 750 39.2 (Nm/hr) 700 355.9 700 575.7 650 912.5 650 971.4 Consumption of 7.53 600 1140.9 600 1192.9 wood (kg/hr) 550 1103.3 550 1044.3 500 1050.4 500 928.7 Air/Fuel ratio 1.94 450 941.8 450 842.1 (-) 400 769.7 400 714.2 350 685.6 350 622.5 Energy In 128.22 300 591.2 300 565.2 (MJ/hr) 250 536.4 250 530.1 200 531.2 200 521.9 Energy Out 90.61 150 526.8 150 501.6 (MJ/hr) 100 491.8 100 490.0 50 496.3 50 482.9 Cold gas 70.67 efficiency (%) Outlet gas 472.9 Outlet gas 471.1

305

Appendix F Experimental test summaries

Run no: 502

Duration of run: 80

Operating Variables
Fuel (wood blocks) properties Size (mm) Moisture content (%wb) 20 x 20 x 20 8.58 Inlet air (Average values) Temperature (C) 22.40 Gauge pressure (kPa) 2.34 Density (kg/m) 1.22 Flow rate (kg/hr) 22.09 Experimental setting Height of charcoal 400 (mm above grate) Ignition level 400 (mm above grate)

Results
Dry gas composition (vol%) first sampling [30 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 4.69 7.00 13.77 1.65 72.89 450 mm above grate 7.20 11.01 14.97 1.30 65.52 350 mm above grate 11.78 14.65 13.98 1.57 58.02 250 mm above grate 16.49 18.95 10.99 1.62 51.94 Outlet gas 17.29 18.50 12.99 1.57 49.65 Dry gas composition (vol%) first sampling [45 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 3.78 5.65 13.97 1.41 75.19 450 mm above grate 10.35 13.70 13.75 1.21 60.99 350 mm above grate 12.46 17.11 11.73 1.31 57.38 250 mm above grate 15.86 17.42 12.88 1.56 52.28 Outlet gas 16.71 18.15 13.77 1.50 49.87 Tar content Material Balance Level Tar (mg/Nm) Input (kg/hr) Output (kg/hr) 550 mm above grate 3465 Air 22.09 Gas 30.95 450 mm above grate 219 Dry wood 11.24 Tar 5.99E-04 350 mm above grate 152 Moisture 1.05 H2O 1.86 250 mm above grate 80 Total 34.38 Total 32.81 Outlet gas 22 Closure (%) 95.40 Elemental Balance Input (kg/hr) Output (kg/hr) Closure (%) C 5.95 C 4.84 81.24 O 10.77 O 10.32 95.86 H 0.71 H 0.70 98.68 N 16.94 N 16.94 100.0 Temperature profile(average) Temperature profile(average) Reactor performance summary (30-40) min after ignition (45-55) min after ignition Net HV of 4.94 outlet gas Level above Temperature Level above Temperature (MJ/Nm) grate (mm) (C) grate (mm) (C) 800 18.7 800 20.9 Gas flow rate 27.24 750 19.0 750 23.0 (Nm/hr) 700 19.3 700 24.5 650 24.6 650 47.2 Consumption of 12.29 600 166.7 600 195.5 wood (kg/hr) 550 938.0 550 1163.2 500 1185.9 500 1090.4 Air/Fuel ratio 1.97 450 1075.5 450 897.8 (-) 400 853.6 400 792.5 350 809.6 350 747.7 Energy In 209.43 300 785.5 300 701.0 (MJ/hr) 250 756.1 250 696.8 200 652.8 200 667.7 Energy Out 134.64 150 627.8 150 650.1 (MJ/hr) 100 583.2 100 607.7 50 519.0 50 535.7 Cold gas 64.29 efficiency (%) Outlet gas 464.4 Outlet gas 483.1

306

Appendix F Experimental test summaries

Run no: 504

Duration of run: 78 min

Operating Variables
Fuel (wood blocks) properties Size (mm) Moisture content (%wb) 20 x 20 x 20 9.47 Inlet air (Average values) Temperature (C) 20.70 Gauge pressure (kPa) 1.89 Density (kg/m) 1.23 Flow rate (kg/hr) 19.60 Experimental setting Height of charcoal 400 (mm above grate) Ignition level 400 (mm above grate)

Results
Dry gas composition (vol%) first sampling [10 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 0.00 2.84 18.35 0.00 73.56 450 mm above grate 6.05 9.04 17.1 0.00 64.24 350 mm above grate 9.89 11.89 16.81 2.03 59.38 250 mm above grate 10.83 12.99 14.61 2.16 59.41 Outlet gas 11.18 13.21 14.91 2.32 58.38 Dry gas composition (vol%) first sampling [30 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 4.00 7.12 16.9 0.42 71.56 450 mm above grate 10.66 13.20 15.13 1.16 59.85 350 mm above grate 15.70 16.77 13.8 1.67 52.06 250 mm above grate 16.66 17.11 14.58 1.76 49.89 Outlet gas 18.01 18.30 12.23 1.44 50.02 Tar content Material Balance Level Tar (mg/Nm) Input (kg/hr) Output (kg/hr) 550 mm above grate 2874 Air 19.60 Gas 26.94 450 mm above grate 419 Dry wood 7.04 Tar 4.77E-04 350 mm above grate 81 Moisture 0.74 H2O 1.08 250 mm above grate 94 Total 27.38 Total 28.02 Outlet gas 20 Closure (%) 102.3 Elemental Balance Input (kg/hr) Output (kg/hr) Closure (%) C 3.73 C 4.12 110.38 O 8.16 O 8.30 101.74 H 0.46 H 0.57 124.65 N 15.03 N 15.03 100.0 Temperature profile(average) Temperature profile(average) Reactor performance summary (10-15) min after ignition (30-40) min after ignition Net HV of 4.95 outlet gas Level above Temperature Level above Temperature (MJ/Nm) grate (mm) (C) grate (mm) (C) 800 15.5 800 22.1 Gas flow rate 23.83 750 16.0 750 24.0 (Nm/hr) 700 17.3 700 25.2 650 17.4 650 28.3 Consumption of 7.78 600 17.7 600 43.4 wood (kg/hr) 550 59.3 550 85.9 500 698.5 500 725.3 Air/Fuel ratio 2.78 450 1047.0 450 1240.3 (-) 400 1138.1 400 1151.9 350 1035.2 350 1000.8 Energy In 131.29 300 955.7 300 919.5 (MJ/hr) 250 859.8 250 818.5 200 708.8 200 748.6 Energy Out 117.91 150 628.4 150 691.9 (MJ/hr) 100 523.7 100 602.8 50 404.6 50 553.8 Cold gas 89.81 efficiency (%) Outlet gas 342.5 Outlet gas 482.8

307

Appendix F Experimental test summaries

Run no: 506

Duration of run: 113 min

Operating Variables
Fuel (wood blocks) properties Size (mm) Moisture content (%wb) 20 x 20 x 20 11.17 Inlet air (Average values) Temperature (C) 23.20 Gauge pressure (kPa) 2.33 Density (kg/m) 1.22 Flow rate (kg/hr) 13.50 Experimental setting Height of charcoal 400 (mm above grate) Ignition level 400 (mm above grate)

Results
Dry gas composition (vol%) first sampling [30 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 7.79 10.69 14.34 2.01 65.16 450 mm above grate 15.26 14.39 12.75 1.91 55.69 350 mm above grate 16.86 18.32 12.82 1.89 50.11 250 mm above grate 17.33 18.71 11.49 1.60 50.86 Outlet gas 17.90 18.83 14.29 1.33 47.65 Dry gas composition (vol%) first sampling [45 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 8.67 11.64 13.46 1.03 65.20 450 mm above grate 14.86 15.45 13.71 1.46 54.53 350 mm above grate 16.32 17.77 14.75 1.63 49.52 250 mm above grate 16.55 17.68 13.66 1.29 50.82 Outlet gas 17.63 18.55 13.96 1.47 48.38 Tar content Material Balance Level Tar (mg/Nm) Input (kg/hr) Output (kg/hr) 550 mm above grate 3982 Air 13.50 Gas 19.77 450 mm above grate 228 Dry wood 6.47 Tar 8.65E-04 350 mm above grate 79 Moisture 0.81 H2O 1.19 250 mm above grate 115 Total 20.78 Total 20.96 Outlet gas 50 Closure (%) 100.80 Elemental Balance Input (kg/hr) Output (kg/hr) Closure (%) C 3.43 C 3.21 93.66 O 6.57 O 6.94 105.73 H 0.43 H 0.45 103.98 N 10.35 N 10.35 100.00 Temperature profile(average) Temperature profile(average) Reactor performance summary (30-40) min after ignition (45-55) min after ignition Net HV of 4.96 outlet gas Level above Temperature Level above Temperature (MJ/Nm) grate (mm) (C) grate (mm) (C) 800 18.2 800 21.3 Gas flow rate 17.30 750 19.5 750 29.0 (Nm/hr) 700 29.5 700 98.6 650 381.5 650 831.6 Consumption of 7.28 600 861.6 600 1125.2 wood (kg/hr) 550 1088.4 550 1188.6 500 1145.9 500 1089.2 Air/Fuel ratio 2.09 450 1072.4 450 1005.8 (-) 400 940.6 400 887.1 350 859.5 350 806.5 Energy In 120.54 300 739.7 300 690.3 (MJ/hr) 250 629.3 250 588.7 200 556.8 200 540.0 Energy Out 85.89 150 532.6 150 524.9 (MJ/hr) 100 482.5 100 474.1 50 438.7 50 445.5 Cold gas 71.25 efficiency (%) Outlet gas 396.8 Outlet gas 405.0

308

Appendix F Experimental test summaries

Run no: 508

Duration of run (min): 74 min

Operating Variables
Fuel (wood blocks) properties Size (mm) Moisture content (%wb) 20 x 20 x 20 11.12 Inlet air (Average values) Temperature (C) 27.10 Gauge pressure (kPa) 4.39 Density (kg/m) 1.23 Flow rate (kg/hr) 24.73 Experimental setting Height of charcoal 400 (mm above grate) Ignition level 400 (mm above grate)

Results
Dry gas composition (vol%) first sampling [30 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 0.00 0.00 2.21 0.00 97.79 450 mm above grate 4.49 5.75 15.80 1.10 72.87 350 mm above grate 14.43 16.74 14.16 1.12 53.55 250 mm above grate 16.39 17.81 14.75 1.39 49.67 Outlet gas 17.75 18.99 15.06 1.50 46.70 Dry gas composition (vol%) first sampling [45 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 0.00 0.00 0.63 0.00 99.37 450 mm above grate 2.11 4.68 11.22 0.00 81.99 350 mm above grate 12.43 14.99 15.28 1.19 56.11 250 mm above grate 16.56 18.59 14.70 1.29 48.86 Outlet gas 17.63 18.55 13.96 1.47 48.38 Tar content Material Balance Level Tar (mg/Nm) Input (kg/hr) Output (kg/hr) 550 mm above grate 406 Air 24.73 Gas 37.16 450 mm above grate 1475 Dry wood 11.64 Tar 5.16E-04 350 mm above grate 185 Moisture 1.46 H2O 1.11 250 mm above grate 206 Total 37.83 Total 38.27 Outlet gas 16 Closure (%) 101.20 Elemental Balance Input (kg/hr) Output (kg/hr) Closure (%) C 6.17 C 6.19 100.34 O 11.91 O 12.39 103.99 H 0.78 H 0.73 93.05 N 18.97 N 18.97 100.0 Temperature profile(average) Temperature profile(average) Reactor performance summary (30-40) min after ignition (45-55) min after ignition Net HV of 5.03 outlet gas Level above Temperature Level above Temperature (MJ/Nm) grate (mm) (C) grate (mm) (C) 800 22.9 800 26.5 Gas flow rate 32.28 750 24.0 750 27.7 (Nm/hr) 700 25.7 700 29.9 650 39.1 650 34.2 Consumption of 13.10 600 44.3 600 36.3 wood (kg/hr) 550 68.4 550 55.4 500 448.4 500 81.7 Air/Fuel ratio 2.12 450 999.0 450 665.7 (-) 400 1198.5 400 1061.2 350 932.9 350 1263.8 Energy In 217.03 300 895.2 300 1064.8 (MJ/hr) 250 757.6 250 898.4 200 724.6 200 807.9 Energy Out 162.42 150 675.1 150 658.1 (MJ/hr) 100 661.7 100 631.6 50 622.5 50 609.3 Cold gas 74.84 efficiency (%) Outlet gas 580.4 Outlet gas 598.9

309

Appendix F Experimental test summaries

Run no: 510

Duration of run (min): 137 min

Operating Variables
Fuel (wood blocks) properties Size (mm) Moisture content (%wb) 20 x 20 x 20 10.11 Inlet air (Average values) Temperature (C) 20.90 Gauge pressure (kPa) 4.61 Density (kg/m) 1.26 Flow rate (kg/hr) 7.50 Experimental setting Height of charcoal 400 (mm above grate) Ignition level 400 (mm above grate)

Results
Dry gas composition (vol%) first sampling [30 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 2.61 4.25 11.88 0.78 80.48 450 mm above grate 11.90 12.17 16.78 1.07 58.07 350 mm above grate 14.92 16.95 16.90 1.43 49.80 250 mm above grate 17.30 18.10 17.13 1.31 46.17 Outlet gas 18.21 18.98 18.18 1.53 43.10 Dry gas composition (vol%) first sampling [45 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 5.30 7.62 10.91 1.38 74.79 450 mm above grate 12.82 13.51 16.27 1.25 56.14 350 mm above grate 15.84 16.14 15.15 1.41 51.45 250 mm above grate 16.70 17.49 18.00 1.61 46.20 Outlet gas 17.59 17.56 17.83 1.44 45.58 Tar content Material Balance Level Tar (mg/Nm) Input (kg/hr) Output (kg/hr) 550 mm above grate 3046 Air 7.50 Gas 12.39 450 mm above grate 307 Dry wood 4.72 Tar 6.46E-04 350 mm above grate 211 Moisture 0.53 H2O 0.74 250 mm above grate 135 Total 12.75 Total 13.13 Outlet gas 61 Closure (%) 103.00 Elemental Balance Input (kg/hr) Output (kg/hr) Closure (%) C 2.50 C 2.21 88.53 O 4.19 O 4.88 116.57 H 0.31 H 0.29 92.20 N 5.75 N 5.75 100.0 Temperature profile(average) Temperature profile(average) Reactor performance summary (30-40-) min after ignition (45-55) min after ignition Net HV of 5.09 outlet gas Level above Temperature Level above Temperature (MJ/Nm) grate (mm) (C) grate (mm) (C) 800 17.5 800 18.9 Gas flow rate 10.60 750 18.0 750 22.6 (Nm/hr) 700 20.9 700 28.3 650 22.0 650 53.5 Consumption of 5.25 600 37.0 600 523.5 wood (kg/hr) 550 467.0 550 872.2 500 1070.2 500 1052.1 Air/Fuel ratio 1.59 450 1049.3 450 1051.7 (-) 400 983.8 400 915.5 350 746.9 350 748.5 Energy In 87.97 300 671.1 300 661.1 (MJ/hr) 250 578.0 250 609.4 200 534.8 200 552.0 Energy Out 53.97 150 497.5 150 502.1 (MJ/hr) 100 440.5 100 448.3 50 426.3 50 431.5 Cold gas 61.35 efficiency (%) Outlet gas 393.4 Outlet gas 404.6

310

Appendix F Experimental test summaries

Run no: 513

Duration of run (min): 128 min

Operating Variables
Fuel (wood blocks) properties Size (mm) Moisture content (%wb) 20 x 20 x 20 10.54 Inlet air (Average values) Temperature (C) 19.70 Gauge pressure (kPa) 1.31 Density (kg/m) 1.22 Flow rate (kg/hr) 10.71 Experimental setting Height of charcoal 400 (mm above grate) Ignition level 400 (mm above grate)

Results
Dry gas composition (vol%) first sampling [30 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 1.33 4.71 17.24 1.37 75.35 450 mm above grate 10.02 10.58 16.50 1.78 61.13 350 mm above grate 13.79 14.93 14.73 1.68 54.87 250 mm above grate 16.15 16.64 12.45 1.40 53.35 Outlet gas 17.56 18.09 13.54 1.43 49.38 Dry gas composition (vol%) first sampling [45 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 2.50 4.62 13.14 0.96 78.78 450 mm above grate 10.87 11.04 15.76 1.56 60.76 350 mm above grate 14.85 14.08 14.06 1.52 55.49 250 mm above grate 17.38 16.52 13.66 1.15 51.29 Outlet gas 17.24 18.68 13.24 1.78 49.07 Tar content Material Balance Level Tar (mg/Nm) Input (kg/hr) Output (kg/hr) 550 mm above grate 3332 Air 10.71 Gas 15.11 450 mm above grate 182 Dry wood 5.47 Tar 8.63E-04 350 mm above grate 161 Moisture 0.64 H2O 0.91 250 mm above grate 182 Total 16.82 Total 16.02 Outlet gas 65 Closure (%) 95.20 Elemental Balance Input (kg/hr) Output (kg/hr) Closure (%) C 2.90 C 2.36 81.41 O 5.35 O 5.10 95.37 H 0.36 H 0.34 94.92 N 8.21 N 8.21 100.0 Temperature profile(average) Temperature profile(average) Reactor performance summary (30-40) min after ignition (45-55) min after ignition Net HV of 4.87 outlet gas Level above Temperature Level above Temperature (MJ/Nm) grate (mm) (C) grate (mm) (C) 800 15.3 800 17.0 Gas flow rate 13.27 750 15.3 750 17.6 (Nm/hr) 700 18.3 700 25.7 650 48.9 650 62.7 Consumption of 6.11 600 610.4 600 759.0 wood (kg/hr) 550 1035.5 550 1083.8 500 1130.8 500 1138.9 Air/Fuel ratio 1.96 450 1055.8 450 1000.5 (-) 400 938.5 400 881.0 350 798.3 350 793.6 Energy In 101.89 300 756.8 300 750.3 (MJ/hr) 250 660.1 250 662.3 200 590.1 200 608.4 Energy Out 64.59 150 543.0 150 568.4 (MJ/hr) 100 499.5 100 534.1 50 441.8 50 479.1 Cold gas 63.39 efficiency (%) Outlet gas 406.9 Outlet gas 436.8

311

Appendix F Experimental test summaries

Run no: 515

Duration of run: 111 min

Operating Variables
Fuel (wood blocks) properties Size (mm) Moisture content (%wb) 20 x 20 x 20 9.06 Inlet air (Average values) Temperature (C) 21.30 Gauge pressure (kPa) 7.51 Density (kg/m) 1.29 Flow rate (kg/hr) 11.31 Experimental setting Height of charcoal 400 (mm above grate) Ignition level 400 (mm above grate)

Results
Dry gas composition (vol%) first sampling [30 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 0.84 1.25 2.91 0.15 94.85 450 mm above grate 9.19 10.62 15.26 1.04 63.88 350 mm above grate 13.50 12.37 16.02 1.22 56.88 250 mm above grate 15.47 15.07 16.71 1.29 51.47 Outlet gas 16.36 15.91 17.00 1.08 49.65 Tar content Material Balance Level Tar (mg/Nm) Input (kg/hr) Output (kg/hr) 550 mm above grate 674 Air 11.31 Gas 16.43 450 mm above grate 1016 Dry wood 5.59 Tar 6.67E-04 350 mm above grate 145 Moisture 0.56 H2O 0.99 250 mm above grate 126 Total 17.46 Total 17.42 Outlet gas 48 Closure (%) 99.80 Elemental Balance Input (kg/hr) Output (kg/hr) Closure (%) C 2.96 C 2.55 85.91 O 5.46 O 5.86 107.26 H 0.36 H 0.34 94.94 N 8.67 N 8.67 100.0 Temperature profile(average) Temperature profile(average) Reactor performance summary (30-40) min after ignition (45-55) min after ignition Net HV of outlet gas 4.32 Level above Temperature Level above Temperature (MJ/Nm) grate (mm) (C) grate (mm) (C) 800 14.3 800 14.4 Gas flow rate 13.90 750 14.4 750 14.9 (Nm/hr) 700 15.2 700 16.6 650 15.2 650 17.2 Consumption of 6.15 600 18.2 600 25.3 wood (kg/hr) 550 49.0 550 62.3 500 608.4 500 750.1 Air/Fuel ratio 2.02 450 1039.5 450 1074.4 (-) 400 1148.1 400 1129.7 350 1076.9 350 991.2 Energy In 104.25 300 921.9 300 871.6 (MJ/hr) 250 802.4 250 754.2 200 761.8 200 740.7 Energy Out 60.02 150 659.1 150 652.9 (MJ/hr) 100 596.1 100 599.1 50 534.6 50 559.1 Cold gas 57.57 efficiency (%) Outlet gas 488.4 Outlet gas 504.3

312

Appendix F Experimental test summaries

Run no: 517

Duration of run: 117 min

Operating Variables
Fuel (wood blocks) properties Size (mm) Moisture content (%wb) 20 x 20 x 20 9.46 Inlet air (Average values) Temperature (C) 23.90 Gauge pressure (kPa) 2.46 Density (kg/m) 1.21 Flow rate (kg/hr) 15.48 Experimental setting Height of charcoal 500 (mm above grate) Ignition level 500 (mm above grate)

Results
Dry gas composition (vol%) first sampling [30 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 5.79 7.19 15.08 1.19 70.75 450 mm above grate 12.84 13.40 15.30 1.50 56.96 350 mm above grate 16.06 16.32 15.96 0.74 50.92 250 mm above grate 15.11 18.68 13.77 1.11 51.33 Outlet gas 17.92 18.79 14.00 1.34 47.95 Tar content Material Balance Level Tar (mg/Nm) Input (kg/hr) Output (kg/hr) 550 mm above grate 3278 Air 15.48 Gas 22.48 450 mm above grate 380 Dry wood 6.36 Tar 6.95E-04 350 mm above grate 84 Moisture 0.67 H2O 0.45 250 mm above grate 77 Total 22.51 Total 22.93 Outlet gas 35 Closure (%) 101.90 Elemental Balance Input (kg/hr) Output (kg/hr) Closure (%) C 3.37 C 3.62 107.43 O 6.85 O 7.02 102.45 H 0.41 H 0.41 100.60 N 11.87 N 11.87 100.0 Temperature profile(average) Temperature profile(average) Reactor performance summary (30-40) min after ignition (45-55) min after ignition Net HV of outlet gas 4.97 Level above Temperature Level above Temperature (MJ/Nm) grate (mm) (C) grate (mm) (C) 800 18.9 800 22.5 Gas flow rate 19.85 750 19.3 750 23.3 (Nm/hr) 700 23.0 700 26.9 650 48.3 650 301.0 Consumption of 7.03 600 408.5 600 907.0 wood (kg/hr) 550 1027.7 550 1155.7 500 1189.7 500 1093.4 Air/Fuel ratio 2.43 450 1009.0 450 1006.0 (-) 400 873.3 400 863.7 350 773.4 350 763.4 Energy In 118.64 300 767.9 300 769.2 (MJ/hr) 250 702.2 250 704.2 200 659.4 200 666.2 Energy Out 98.57 150 593.3 150 607.8 (MJ/hr) 100 553.1 100 570.8 50 515.5 50 534.5 Cold gas 83.08 efficiency (%) Outlet gas 423.7 Outlet gas 445.9

313

Appendix F Experimental test summaries

Run no: 522

Duration of run: 67 min

Operating Variables
Fuel (wood blocks) properties Size (mm) Moisture content (%wb) 30 x 30 x 30 12.66 Inlet air (Average values) Temperature (C) 16.20 Gauge pressure (kPa) 1.94 Density (kg/m) 1.24 Flow rate (kg/hr) 22.71 Experimental setting Height of charcoal 500 (mm above grate) Ignition level 500 (mm above grate)

Results
Dry gas composition (vol%) first sampling [30 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 0.00 0.00 0.00 0.00 100.00 450 mm above grate 0.59 2.60 17.06 1.04 78.71 350 mm above grate 8.89 8.62 15.99 1.12 65.38 250 mm above grate 15.38 17.33 13.13 1.44 52.72 Outlet gas 16.95 18.79 12.00 1.27 50.99 Dry gas composition (vol%) first sampling [45 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 0.00 0.00 0.00 0.00 100.00 450 mm above grate 1.48 2.09 5.41 0.68 90.34 350 mm above grate 9.23 10.98 15.50 1.05 63.24 250 mm above grate N/A N/A N/A N/A N/A* Outlet gas 15.63 19.41 13.68 1.06 50.22 Tar content Material Balance Level Tar (mg/Nm) Input (kg/hr) Output (kg/hr) 550 mm above grate 2245 Air 22.71 Gas 30.94 450 mm above grate 376 Dry wood 10.24 Tar 2.40E-03 350 mm above grate 168 Moisture 1.48 H2O 2.48 250 mm above grate 174 Total 34.43 Total 33.42 Outlet gas 88 Closure (%) 97.10 Elemental Balance Input (kg/hr) Output (kg/hr) Closure (%) C 5.42 C 4.69 86.57 O 10.88 O 10.55 97.00 H 0.71 H 0.75 105.98 N 17.42 N 17.42 100.0 Temperature profile(average) Temperature profile(average) Reactor performance summary (30-40) min after ignition (45-55) min after ignition Net HV of 4.83 outlet gas Level above Temperature Level above Temperature (MJ/Nm) grate (mm) (C) grate (mm) (C) 800 20.1 800 23.1 Gas flow rate 27.30 750 21.5 750 26.0 (Nm/hr) 700 25.4 700 31.5 650 28.5 650 32.2 Consumption of 11.72 600 42.5 600 66.2 wood (kg/hr) 550 117.1 550 103.3 500 517.5 500 296.4 Air/Fuel ratio 2.22 450 962.7 450 618.0 (-) 400 1094.6 400 958.5 350 1186.4 350 1215.0 Energy In 190.80 300 1195.2 300 1177.1 (MJ/hr) 250 968.5 250 1101.8 200 903.3 200 958.9 Energy Out 131.88 150 783.4 150 830.4 (MJ/hr) 100 735.9 100 761.9 50 689.9 50 694.9 Cold gas 69.12 efficiency (%) Outlet gas 525.8 Outlet gas 572.6 * Gas sample composition was not measurable due to a leakage in gas sampling container.

314

Appendix F Experimental test summaries

Run no: 524

Duration of run: 134 min

Operating Variables
Fuel (wood blocks) properties Size (mm) Moisture content (%wb) 30 x 30 x 30 12.46 Inlet air (Average values) Temperature (C) 23.40 Gauge pressure (kPa) 1.48 Density (kg/m) 1.21 Flow rate (kg/hr) 11.28 Experimental setting Height of charcoal 400 (mm above grate) Ignition level 400 (mm above grate)

Results
Dry gas composition (vol%) first sampling [30 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 0.00 0.00 2.09 0.00 97.91 450 mm above grate 8.63 10.23 10.75 1.20 69.19 350 mm above grate 12.54 14.27 16.88 1.13 55.17 250 mm above grate 15.67 16.87 14.60 1.69 51.17 Outlet gas 18.20 17.51 15.33 1.46 47.50 Dry gas composition (vol%) first sampling [40 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 0 0 0 0 100.00 450 mm above grate 11.48 11.98 14.35 1.86 60.33 350 mm above grate 17.20 15.28 18.09 1.49 47.94 250 mm above grate 16.96 15.89 16.06 1.61 49.48 Outlet gas 16.87 16.71 15.10 1.52 49.80 Tar content Material Balance Level Tar (mg/Nm) Input (kg/hr) Output (kg/hr) 550 mm above grate 784 Air 11.28 Gas 16.62 450 mm above grate 395 Dry wood 5.51 Tar 6.52E-04 350 mm above grate 267 Moisture 0.78 H2O 1.00 250 mm above grate 124 Total 17.57 Total 17.62 Outlet gas 45 Closure (%) 100.30 Elemental Balance Input (kg/hr) Output (kg/hr) Closure (%) C 2.92 C 2.68 91.80 O 5.62 O 5.90 104.96 H 0.38 H 0.39 101.62 N 8.65 N 8.65 100.0 Temperature profile(average) Temperature profile(average) Reactor performance summary (30-40) min after ignition (45-55) min after ignition Net HV of outlet gas 4.87 Level above Temperature Level above Temperature (MJ/Nm) grate (mm) (C) grate (mm) (C) 800 18.5 800 20.9 Gas flow rate 14.49 750 19.3 750 19.8 (Nm/hr) 700 19.9 700 21.4 650 23.4 650 26.8 Consumption of 6.29 600 38.5 600 38.9 wood (kg/hr) 550 216.6 550 156.0 500 660.5 500 665.7 Air/Fuel ratio 2.05 450 876.9 450 921.1 (-) 400 1089.7 400 1116.3 350 965.2 350 1037.9 Energy In 102.64 300 922.0 300 901.1 (MJ/hr) 250 809.8 250 838.8 200 732.5 200 755.6 Energy Out 70.61 150 707.4 150 710.4 (MJ/hr) 100 603.1 100 697.7 50 545.2 50 607.9 Cold gas 68.79 efficiency (%) Outlet gas 471.4 Outlet gas 493.2

315

Appendix F Experimental test summaries

Run no: 525

Duration of run: 107 min

Operating Variables
Fuel (wood blocks) properties Size (mm) Moisture content (%wb) 30 x 30 x 30 12.58 Inlet air (Average values) Temperature (C) 21.90 Gauge pressure (kPa) 2.04 Density (kg/m) 1.22 Flow rate (kg/hr) 15.60 Experimental setting Height of charcoal 400 (mm above grate) Ignition level 400 (mm above grate)

Results
Dry gas composition (vol%) first sampling [30 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 0.00 0.00 0.00 0.00 100.00 450 mm above grate 0.00 7.05 13.02 1.77 78.16 350 mm above grate 3.31 8.86 11.94 1.18 74.72 250 mm above grate 7.85 21.30 10.43 1.12 59.30 Outlet gas 6.88 25.44 8.48 1.00 58.20 Dry gas composition (vol%) first sampling [45 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 0.00 0.00 0.00 0.00 100.00 450 mm above grate 0.00 0.00 0.00 0.00 100.00 350 mm above grate 3.40 7.99 16.58 1.32 70.71 250 mm above grate 8.07 18.33 14.15 1.54 57.91 Outlet gas 7.21 22.39 11.89 1.10 57.41 Tar content Material Balance Level Tar (mg/Nm) Input (kg/hr) Output (kg/hr) 550 mm above grate 122 Air 15.60 Gas 20.15 450 mm above grate 1772 Dry wood 6.08 Tar 3.77E-04 350 mm above grate 200 Moisture 0.87 H2O 2.02 250 mm above grate 167 Total 22.55 Total 22.17 Outlet gas 23 Closure (%) 98.30 Elemental Balance Input (kg/hr) Output (kg/hr) Closure (%) C 3.22 C 3.08 95.60 O 6.95 O 6.77 97.50 H 0.42 H 0.35 84.42 N 11.97 N 1.97 100.0 Temperature profile(average) Temperature profile(average) Reactor performance summary (30-40) min after ignition (45-55) min after ignition Net HV of 4.47 outlet gas Level above Temperature Level above Temperature (MJ/Nm) grate (mm) (C) grate (mm) (C) 800 26.5 800 26.6 Gas flow rate 16.39 750 26.0 750 26.6 (Nm/hr) 700 25.6 700 27.1 650 27.5 650 28.9 Consumption of 6.95 600 33.3 600 35.6 wood (kg/hr) 550 57.3 550 53.4 500 122.8 500 97.7 Air/Fuel ratio 2.57 450 462.8 450 362.9 (-) 400 775.7 400 629.7 350 979.9 350 874.0 Energy In 113.25 300 1138.7 300 1131.1 (MJ/hr) 250 1105.4 250 1103.5 200 1058.0 200 1093.5 Energy Out 73.35 150 1009.0 150 1012.2 (MJ/hr) 100 863.1 100 950.5 50 752.8 50 833.8 Cold gas 64.77 efficiency (%) Outlet gas 514.2 Outlet gas 615.6

316

Appendix F Experimental test summaries

Run no: 527

Duration of run: 68 min

Operating Variables
Fuel (wood blocks) properties Size (mm) Moisture content (%wb) 30 x 30 x 30 12.16 Inlet air (Average values) Temperature (C) 22.60 Gauge pressure (kPa) 2.44 Density (kg/m) 1.22 Flow rate (kg/hr) 27.63 Experimental setting Height of charcoal 400 (mm above grate) Ignition level 400 (mm above grate)

Results
Dry gas composition (vol%) first sampling [30 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 0.00 0.00 0.00 0.00 100.00 450 mm above grate 3.91 7.92 12.95 1.97 73.25 350 mm above grate 11.41 15.57 12.37 1.27 59.38 250 mm above grate 14.48 19.26 13.79 1.36 51.11 Outlet gas 15.71 21.05 12.43 1.10 49.71 Tar content Material Balance Level Tar (mg/Nm) Input (kg/hr) Output (kg/hr) 550 mm above grate 366 Air 27.63 Gas 39.24 450 mm above grate 2492 Dry wood 12.10 Tar 2.11E-03 350 mm above grate 233 Moisture 1.68 H2O 2.35 250 mm above grate 97 Total 41.41 Total 41.60 Outlet gas 62 Closure (%) 100.40 Elemental Balance Input (kg/hr) Output (kg/hr) Closure (%) C 6.41 C 6.32 98.54 O 12.98 O 13.28 102.33 H 0.83 H 0.81 97.37 N 21.19 N 21.19 100.0 Temperature profile(average) Temperature profile(average) Reactor performance summary (30-40) in after ignition (45-55) min after ignition Net HV of outlet gas 4.93 Level above Temperature Level above Temperature (MJ/Nm) grate (mm) (C) grate (mm) (C) 800 14.0 800 21.8 Gas flow rate 34.04 750 13.3 750 22.3 (Nm/hr) 700 14.5 700 23.3 650 17.1 650 38.6 Consumption of 13.78 600 56.2 600 51.2 wood (kg/hr) 550 157.3 550 65.9 500 264.9 500 204.2 Air/Fuel ratio 2.28 450 883.8 450 427.0 (-) 400 1201.2 400 965.7 350 1155.8 350 1237.6 Energy In 225.63 300 1092.3 300 1136.2 (MJ/hr) 250 1076.8 250 1086.0 200 1061.0 200 1073.7 Energy Out 167.66 150 993.9 150 1057.6 (MJ/hr) 100 882.3 100 943.2 50 740.3 50 877.9 Cold gas 74.31 efficiency (%) Outlet gas 596.0 Outlet gas 675.8

317

Appendix F Experimental test summaries

Run no: 528

Duration of run: 107 min

Operating Variables
Fuel (wood blocks) properties Size (mm) Moisture content (%wb) 30 x 30 x 30 12.39 Inlet air (Average values) Temperature (C) 22.90 Gauge pressure (kPa) 1.46 Density (kg/m) 1.21 Flow rate (kg/hr) 12.21 Experimental setting Height of charcoal 500 (mm above grate) Ignition level 500 (mm above grate)

Results
Dry gas composition (vol%) first sampling [30 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 0.00 0.00 0.00 0.00 100.00 450 mm above grate 0.14 2.67 17.13 1.55 78.51 350 mm above grate 10.48 14.09 14.66 0.86 59.91 250 mm above grate 14.77 17.88 15.12 1.06 51.17 Outlet gas 16.76 19.00 14.97 1.31 47.96 Dry gas composition (vol%) first sampling [45 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 0.00 0.00 0.00 0.00 100.00 450 mm above grate 0.00 0.00 16.28 0.00 83.72 350 mm above grate 11.08 13.21 14.83 1.18 59.70 250 mm above grate 14.55 16.96 13.62 1.08 53.80 Outlet gas 16.13 18.26 12.35 0.18 53.08 Tar content Material Balance Level Tar (mg/Nm) Input (kg/hr) Output (kg/hr) 550 mm above grate 3847 Air 12.21 Gas 18.05 450 mm above grate 243 Dry wood 6.40 Tar 4.36E-04 350 mm above grate 149 Moisture 0.91 H2O 1.08 250 mm above grate 88 Total 19.52 Total 19.13 Outlet gas 28 Closure (%) 98.0 Elemental Balance Input (kg/hr) Output (kg/hr) Closure (%) C 3.39 C 2.95 87.03 O 6.32 O 6.42 101.62 H 0.44 H 0.39 88.65 N 9.37 N 9.37 100.0 Temperature profile(average) Temperature profile(average) Reactor performance summary (30-40) min after ignition (45-55) min after ignition Net HV of 4.85 outlet gas Level above Temperature Level above Temperature (MJ/Nm) grate (mm) (C) grate (mm) (C) 800 21.6 800 27.8 Gas flow rate 15.56 750 13.4 750 18.9 (Nm/hr) 700 14.1 700 19.9 650 15.4 650 23.1 Consumption of 7.31 600 26.1 600 37.2 wood (kg/hr) 550 71.4 550 88.5 500 397.9 500 469.0 Air/Fuel ratio 1.91 450 772.1 450 778.5 (-) 400 983.1 400 985.9 350 1074.8 350 1068.0 Energy In 119.38 300 953.9 300 974.8 (MJ/hr) 250 880.7 250 878.6 200 799.6 200 833.9 Energy Out 75.49 150 718.1 150 760.2 (MJ/hr) 100 627.4 100 662.4 50 567.1 50 613.1 Cold gas 63.24 efficiency (%) Outlet gas 441.3 Outlet gas 474.5

318

Appendix F Experimental test summaries

Run no: 531

Duration of run: 118 min

Operating Variables
Fuel (wood blocks) properties Size (mm) Moisture content (%wb) 30 x 30 x 30 11.24 Inlet air (Average values) Temperature (C) 22.50 Gauge pressure (kPa) 3.15 Density (kg/m) 1.23 Flow rate (kg/hr) 13.85 Experimental setting Height of charcoal 400 (mm above grate) Ignition level 400 (mm above grate)

Results
Dry gas composition (vol%) first sampling [10 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 0.00 0.00 0.21 0.00 99.79 450 mm above grate 2.35 5.91 14.05 0.00 77.69 350 mm above grate 3.29 6.88 15.81 1.02 73.00 250 mm above grate 5.72 8.06 15.30 1.40 69.53 Outlet gas 5.42 9.14 15.71 1.52 68.21 Dry gas composition (vol%) first sampling [30 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 0.00 0.83 3.69 0.00 95.49 450 mm above grate 3.98 4.57 15.87 0.68 74.90 350 mm above grate 9.42 11.94 15.71 1.52 61.41 250 mm above grate 14.51 15.58 13.94 1.12 54.85 Outlet gas 16.48 17.79 14.29 1.48 49.96 Tar content Material Balance Level Tar (mg/Nm) Input (kg/hr) Output (kg/hr) 550 mm above grate 1198 Air 13.85 Gas 19.61 450 mm above grate 708 Dry wood 6.24 Tar 5.60E-04 350 mm above grate 188 Moisture 0.79 H2O 1.18 250 mm above grate 86 Total 20.88 Total 20.79 Outlet gas 33 Closure (%) 99.6 Elemental Balance Input (kg/hr) Output (kg/hr) Closure (%) C 3.31 C 3.06 92.53 O 6.53 O 6.68 102.27 H 0.42 H 0.43 101.65 N 10.62 N 10.62 100.0 Temperature profile(average) Temperature profile(average) Reactor performance summary (10-15) min after ignition (30-40) min after ignition Net HV of 4.73 outlet gas Level above Temperature Level above Temperature (MJ/Nm) grate (mm) (C) grate (mm) (C) 800 19.0 800 20.4 Gas flow rate 16.96 750 18.9 750 20.4 (Nm/hr) 700 19.4 700 21.8 650 20.8 650 26.3 Consumption of 7.03 600 19.1 600 43.2 wood (kg/hr) 550 143.7 550 100.3 500 419.9 500 429.1 Air/Fuel ratio 2.22 450 818.1 450 782.1 (-) 400 976.7 400 1092.6 350 928.7 350 1177.6 Energy In 116.31 300 875.4 300 1092.6 (MJ/hr) 250 732.5 250 1082.3 200 619.9 200 1011.9 Energy Out 80.14 150 580.3 150 952.1 (MJ/hr) 100 542.7 100 868.2 50 452.9 50 777.3 Cold gas 68.90 efficiency (%) Outlet gas 355.1 Outlet gas 500.6

319

Appendix F Experimental test summaries

Run no: 604

Duration of run: 129 min

Operating Variables
Fuel (wood blocks) properties Size (mm) Moisture content (%wb) 30 x 30 x 30 12.15 Inlet air (Average values) Temperature (C) 24.10 Gauge pressure (kPa) 3.53 Density (kg/m) 1.23 Flow rate (kg/hr) 15.34 Experimental setting Height of charcoal 500 (mm above grate) Ignition level 500 (mm above grate)

Results
Dry gas composition (vol%) first sampling [30 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 0.00 0.13 3.71 0.00 96.16 450 mm above grate 7.84 10.81 15.84 1.20 64.31 350 mm above grate 16.08 17.65 13.25 1.31 51.71 250 mm above grate 14.44 18.72 11.82 1.17 53.84 Outlet gas 16.76 19.82 11.68 1.25 50.49 Dry gas composition (vol%) first sampling [45 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 0.00 0.00 6.57 0.00 93.43 450 mm above grate 3.00 5.34 18.29 0.51 72.86 350 mm above grate 12.45 13.43 15.91 1.53 56.68 250 mm above grate 14.40 16.41 13.35 1.03 54.81 Outlet gas 15.44 18.72 11.82 1.17 52.84 Tar content Material Balance Level Tar (mg/Nm) Input (kg/hr) Output (kg/hr) 550 mm above grate 2782 Air 15.34 Gas 21.11 450 mm above grate 381 Dry wood 6.95 Tar 1.09E-03 350 mm above grate 245 Moisture 0.96 H2O 1.27 250 mm above grate 162 Total 23.25 Total 22.37 Outlet gas 59 Closure (%) 96.2 Elemental Balance Input (kg/hr) Output (kg/hr) Closure (%) C 3.68 C 3.27 88.86 O 7.33 O 6.88 93.85 H 0.48 H 0.46 96.98 N 11.77 N 11.77 100.0 Temperature profile(average) Temperature profile(average) Reactor performance summary (30-40) min after ignition (45-55) min after ignition Net HV of 4.94 outlet gas Level above Temperature Level above Temperature (MJ/Nm) grate (mm) (C) grate (mm) (C) 800 21.0 800 25.8 Gas flow rate 18.56 750 21.5 750 26.1 (Nm/hr) 700 25.1 700 30.3 650 38.1 650 47.5 Consumption of 7.91 600 151.4 600 100.2 wood (kg/hr) 550 481.7 550 257.9 500 912.0 500 711.0 Air/Fuel ratio 2.21 450 1054.3 450 959.9 (-) 400 1042.6 400 1106.1 350 938.9 350 1048.7 Energy In 129.53 300 830.8 300 913.7 (MJ/hr) 250 767.1 250 798.1 200 727.3 200 768.2 Energy Out 91.61 150 658.7 150 708.3 (MJ/hr) 100 610.4 100 669.2 50 525.3 50 587.5 Cold gas 70.73 efficiency (%) Outlet gas 436.7 Outlet gas 483.7

320

Appendix F Experimental test summaries

Run no: 601

Duration of run: 139 min

Operating Variables
Fuel (wood blocks) properties Size (mm) Moisture content (%wb) 20 x 20 x 20 9.83 Inlet air (Average values) Temperature (C) 19.80 Gauge pressure (kPa) 1.10 Density (kg/m) 1.21 Flow rate (kg/hr) 10.22 Experimental setting Height of charcoal (mm above grate) Ignition level 400 (mm above grate)

Results
Dry gas composition (vol%) first sampling [30 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 0.00 0.00 0.00 0.00 100.00 450 mm above grate 2.99 4.66 10.56 0.28 81.51 350 mm above grate 9.65 11.08 14.84 1.19 63.24 250 mm above grate 12.45 13.23 13.38 1.54 59.41 Outlet gas 14.72 14.35 14.67 1.59 54.66 Dry gas composition (vol%) first sampling [45 minutes after ignition] Level H2 CO CO2 CH4 Trace (O2+N2) 550 mm above grate 0.00 0.13 3.71 0.00 96.16 450 mm above grate 7.84 10.81 16.51 1.20 63.65 350 mm above grate 16.08 17.65 15.22 1.31 49.74 250 mm above grate 14.44 18.72 15.27 1.17 50.40 Outlet gas 15.00 19.68 15.15 1.53 48.64 Tar content Material Balance Level Tar (mg/Nm) Input (kg/hr) Output (kg/hr) 550 mm above grate 3309 Air 10.22 Gas 15.16 450 mm above grate 318 Dry wood 5.25 Tar 1.25E-02 350 mm above grate 310 Moisture 0.57 H2O 0.91 250 mm above grate 153 Total 16.04 Total 16.08 Outlet gas 964 Closure (%) 100.3 Elemental Balance Input (kg/hr) Output (kg/hr) Closure (%) C 2.78 C 2.51 90.34 O 5.08 O 5.42 106.64 H 0.34 H 0.31 90.57 N 7.84 N 7.84 100.0 Temperature profile(average) Temperature profile(average) Reactor performance summary (30-40) min after ignition (45-55) min after ignition Net HV of 4.83 outlet gas Level above Temperature Level above Temperature (MJ/Nm) grate (mm) (C) grate (mm) (C) 800 25.9 800 25.2 Gas flow rate 12.93 750 25.3 750 25.9 (Nm/hr) 700 25.5 700 27.2 650 27.0 650 28.3 Consumption of 5.82 600 30.0 600 35.3 wood (kg/hr) 550 37.1 550 53.6 500 68.5 500 472.7 Air/Fuel ratio 1.95 450 467.8 450 873.1 (-) 400 892.1 400 936.0 350 967.9 350 894.2 Energy In 97.82 300 906.8 300 830.0 (MJ/hr) 250 810.6 250 755.7 200 783.8 200 742.8 Energy Out 62.39 150 690.9 150 650.7 (MJ/hr) 100 650.0 100 620.2 50 571.6 50 554.4 Cold gas 63.78 efficiency (%) Outlet gas 437.9 Outlet gas 413.1

321

Appendix G Photo documentation

Appendix G Photo documentation

Figure G.1 Experimental rig

Figure G.2 The lower, middle and upper sections of the stratified downdraft gasifier

322

Appendix G Photo documentation

Figure G.3 Technique of installing castable insulation in the middle section of the experimental gasifier

323

Appendix G Photo documentation

(a) Chromel-Alumel (K-Type) thermocouples

(b) Thermocouples inside reactor

(c) Data logger connection

Figure G.4 Temperature measurement system

324

Appendix G Photo documentation

(a) Water bath and impingers for tar collection

(b) Sample gas volume measurement system

Figure G.5 Gas sampling train

325

Appendix G Photo documentation

(a) Gas chromatograph (GC) with thermal conductivity detector

(b) PVC gas sampling container

Figure G.6 Apparatus used for gas composition measurement

326

Appendix G Photo documentation

(a) Collected condensate from five different levels in the gasifier

(a) Rotary evaporator

(b) Rotary evaporator

(c) Water bath

(d) Condensable tar collected

Figure G.7 Tar measurement system

327

Appendix G Photo documentation

(a) Wood drying

(b) Wood cutting

Figure G.8 Wood fuel preparation

328

Appendix H

Soft copy of the computer code on CD

329

You might also like