Você está na página 1de 255

THESIS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

Principles and Models of Solid Fuel Combustion

Henrik Thunman

Department of Energy Conversion CHALMERS UNIVERSITY OF TECHNOLOGY Gteborg, Sweden 2001

Principles and Models of Solid Fuel Combustion

HENRIK THUNMAN ISBN 91-7291-058-5

HENRIK THUNMAN, 2001

Doktorsavhandlingar vid Chalmers tekniska hgskola Ny serie nr 1742 ISSN 0346-718X

Department of Energy Conversion CHALMERS UNIVERSITY OF TECHNOLOGY S-412 96 Gteborg Sweden Telephone +46 (0)31 772 1000

Cover: A burning piece (5cm long) of wood

Chalmers Reproservice Gteborg, Sweden 2001

ii

Principles and Models of Solid Fuel Combustion


Henrik Thunman Department of Energy Conversion Chalmers University of Technology S-412 96 Gteborg, Sweden

Abstract
Combustion of solid fuels stands for a substantial part of the heat and power production in the world. In this thesis different aspects related to the thermochemical conversion of solid fuels in fluidized and fixed beds are treated, such as: the combustion temperature of a fuel particle related to the surrounding temperature in a fluidized bed combustor (FBC); the fuel loading in a FBC; a transient description of the conversion of non-spherical fuel particles; the thermochemical properties and composition of volatiles released from wood; the thermal conductivity of wood during different stages of conversion; the modelling of a grate furnace; the general combustion behaviour of a fixed bed on a reciprocating or a travelling grate; and finally, the design and construction of experimental units for investigation of combustion in a fixed bed. The main result is a number of sub models validated by comparison with measurements. The sub models are to be used to describe and understand various combustion behaviours. Besides the sub models derived, it is concluded that the generally accepted behaviour of combustion on a reciprocating or a travelling grate is not correct for wet biofuels. Measurements and supplementary modelling show that the ignition of the bed is not caused by reflection of radiation from specially designed arches in the furnace. Instead ignition takes place at the bottom of the bed, close to the surface of the grate. This is important for new designs of reciprocating and travelling grates. Keywords: Biofuels, Coal, Combustion, Fixed bed, Fluidized bed, Fuel loading, Grate, Measurements, Modelling, Single particle, Thermal conductivity, Thermochemical conversion, Thermochemical properties, Wood

iii

iv

The thesis is based on the following Papers: I. G.I., Palchonok, C., Breitholtz, H., Thunman, B., Leckner, Impact of heat and mass transfer on combustion of a fuel particle in CFB boilers, 14th International concerance on FBC, Ed. F.D.S. Preto, ASME, 1997, pp. 871-878. (Awarded as best Paper) H. Thunman, B. Leckner, Fuel loading of a fluidized bed combustor, 4th European Conference on Industrial Furnaces and Boilers, Eds. W., Leuckel, J.,Ward, R.,Collin, A., Reis, INFUB, Esphinho-Porto, Portugal 1-4 April, 1997. (An appendix to the Paper is included from, H, Thunman, Loading and size distribution of fuel in a fluidized bed combustor, Thesis for the degree of licentiate of engineering, Department of Energy Conversion, Chalmers University of Technology, Gteborg, May 1997). Henrik Thunman, Bo Leckner, Fredrik Niklasson, Filip Johnsson, Combustion of wood particles a model for Eulerian calculations, Submitted for publication June 2001. H., Thunman, F., Niklasson, F., Johnsson, B., Leckner, Composition of volatile gases and thermo-chemical properties of wood for modeling of fixed or fluidized beds, Accepted for publication in Energy & Fuels, August 2001. H., Thunman, B., Leckner, Thermal conductivity of wood during different combustion stages, Submitted for publication February 2001. H., Thunman, L-E., mand, F., Ghirelli, B., Leckner, Modelling and verifying experiments of the whole furnace, Report to the European Commission JOR 3CT96 0059, Department of Energy Conversion, Chalmers University of Technology, Gteborg, Sweden, 1999. H., Thunman, B., Leckner, Ignition and propagation of a reaction front in crosscurrent bed combustion of wet biofuels, Fuel, 2001, 80, 473-481. M., Rnnbck, M., Axell, L., Gustavsson, H., Thunman, B., Leckner, Combustion process in a biomass fuel bed Experimental results, Progress in Thermochemical Biomass Conversion, Ed. Bridgwater, Tyrol, September, 2000.

II.

III.

IV.

V.

VI.

VII.

VIII.

Contribution by Henrik Thunman Paper I. Henrik Thunman together with Timo Joutsenoja from Tampere University of Technology carried out the pyrometer measurements. Henrik Thunman carried out the analysis of these measurements and the fundamental part of the modelling showing the ignition phenomena. Clas Breitholtz and Gennadij Palchonok improved the model concerning heat and mass transfer to the single particle. Gennadij Palchonok summarised and wrote the article. Paper II, V and VII. Henrik Thunman is the principal author. Paper III and IV. Henrik Thunman carried out all work, except for the measurements and the evaluation of the measurements, which was carried out by Fredrik Niklasson.

Paper VI. Henrik Thunman acted as a project leader, planed the survey measurements in the grate furnace, evaluated the first measurement campaign, derived the bed model, supervised the CFD-calculation, and summarised the report. Lars-Erik mand had an active part in the measurements, evaluated the second measurement campaign and made the conclusions concerning the NOx formation. Federico Ghirelli carried out the CFD-calculations as a part of his master degree thesis work. Paper VIII. Henrik Thunman suggested the design of the two furnaces and the measurements. Monica Axell and Marie Rnnbck constructed the furnaces and carried out the measurements. Marie Rnnbck wrote the paper. Bo Leckner is the supervisor of this work and had an active role in the writing of all papers. Filip Johnsson is Fredrik Niklassons supervisor, and Lennart Gustavsson is Monica Axells and Marie Rnnbcks supervisor. All the papers, I to VIII, were thoroughly discussed and edited by all co-authors.

vi

Contents
Abstract ..................................................................................................................................... iii Acknowledgements................................................................................................................... ix Introduction................................................................................................................................ 1 Furnaces and Reactors used for Experiments ............................................................................ 2 Summary of Papers .................................................................................................................... 5 Conclusions.............................................................................................................................. 19 Future Work ............................................................................................................................. 21 References................................................................................................................................ 21 Errata list for the included Papers ............................................................................................ 24 Populrvetenskaplig sammanfattning ( In Swedish )............................................................... 25 Paper I to VIII

vii

viii

Acknowledgements
I would like to thank my supervisor Prof. Bo Leckner for his dedication in my work and for providing me the opportunity to accomplish this thesis work. As well, I would like to thank him for always being open for discussions and sharing his wide knowledge in the field of energy conversion. I also would like to express my gratitude to all co-authors (Gennadij Palchonok, Clas Breitholtz, Lars-Erik mand, Fredrik Niklasson, Filip Johnsson, Monica Axell, Marie Rnnbck, Lennart Gustavsson) and especially to Sven Andersson who introduced me into the subject. Finally, I would like to thank all colleagues that are or have been working at the department during these years for providing a friendly and inspiring atmosphere. This work was supported by a scholarship from the Nordic Energy Research Program for Combustion of Solid Fuels, and by grants from Swedish National Board for Industrial and Technical Development (NUTEK), and Small scale Combustion Program of the National Energy Administration. Part of the work was carried out under the European Union contract JOR3-CT96-0059. All of them are gratefully acknowledged.

ix

Introduction
The qualitative behaviour of the thermo-chemical conversion of solid fuels follows the same path, independent of the combustion situation drying, devolatilisation, char combustion, or gasification. Depending on particle size and heating rate these processes can occur sequentially or simultaneously in the solid fuel particle. Different aspects of the thermochemical conversion of solid fuels in fixed and fluidized beds are studied in this thesis, where a number of sub models are derived for comprehensive models of conversion, and for evaluation and understanding of measurements. In the papers assembled, the following has been investigated: excess temperatures of fuel related to the bed temperature (1123 K) in a circulating fluidized bed (CFB) combustor; fuel loading in a CFB combustor; the conversion rate of fuel particles in a fixed or fluidized bed; modelling of the combustion process in a reciprocating or travelling grate furnace; and finally, general combustion behaviour of a fuel bed supported by a reciprocating or travelling grate. The conditions (i.e. oxygen concentration, particle size) under which fuel particles can achieve a higher temperature than that of the surrounding in a CFB furnace, which can cause production of NOx or indicate bad mixing of the gas, are investigated in Paper I. For the investigation, measurements of surface temperatures on and heat and mass transfer to particles inside the CFB combustor were carried out, supported by a model. The importance of high excess temperatures can have on NOx production, for example, is strongly connected to the size distribution and concentration of particles on different levels in the furnace. These quantities determine the amount of particles that can achieve high temperatures. The size distribution and the number of particles on different levels are a direct measure of the fuel loading, which is the subject investigated in Paper II. The interest in the fuel loading is mainly for operation control; a small fuel loading gives a fast response to a change in the operating condition, whereas the opposite results for a large fuel loading. Work has been carried out previously in this area [1 - 4], but mostly in small experimental units. The measurements in Chalmers 12 MWth CFB furnace provide data for comparison with the results from these small units. For modelling of fuel loading a more extensive description is needed of the conversion of the solid fuel particles than for the description of the surface temperature in Paper I. The model is therefore extended by the introduction of model elements for drying, devolatilisation, fragmentation and attrition. However, significant simplifications can be made due to the well defined surrounding of the single particles in a fluidized bed. For example, the time t for drying and devolatilisation of spherical particle with the diameter d can be described with a simple empirical correlation of the form t = a d b, where a and b are empirical coefficients, [4-6]. This description of drying and devolatilisation is appropriate for homogenous fuels with spherical form and can therefore be used for most coals. For non-homogenous fuels or/and particles exposed to a transient environment, such as, in a fixed bed or during changes in operation conditions in a fluidized bed, a transient particle model is needed. For example, biofuels entering the furnace have moisture contents that often vary with time and, therefore, require a particle model that describes the progress of the drying and devolatilisation inside the particles. Furthermore, most biofuels have shapes that are far from spherical. In Paper III, a transient model was derived for the combustion of single fuel particles or groups of fuel particles in different beds. A great effort was made to minimise the computational effort, without losing important features influencing the combustion, such as shrinkage and particle shape.

A transient combustion model requires a large number of fuel properties, among which particle size, shape and composition, thermal conductivity, specific heat and shrinkage during conversion are the most important ones. In the assembled papers, for simplicity, fuel properties are taken for trunk wood. However, the sub-models derived are general and can be used for various solid fuels. In Paper IV thermo-chemical data for wood during different stages of combustion are summarised together with a model for the determination of the composition of the volatile gases leaving a fuel particle. These data are complemented in Paper V with a model for the thermal conductivity of wood during different stages of conversion. A fixed bed that is ignited from one end and with air introduced from the other, can be seen as a large fuel particle. The difference is that also air, introduced for the conversion of the fuel bed, follows the moisture and the volatiles through the bed and gives rise to internal heat generation by combustion. This combustion behaviour has been applied frequently because it is the generally accepted behaviour for combustion of a fuel bed on reciprocating or travelling grates, [7-11]. In Paper VI a simplified bed model is derived, based upon this assumption and including the main features of conversion of the fuel bed along the grate. The purpose is to investigate the possibility to model the entire grate furnace by connecting a model of the fuel bed to models included in an existing CFD-software, for description of the gaseous combustion above the fuel bed. For validation, extensive measurements in a 31 MWth reciprocating grate furnace were carried out. As a consequence of the validation of the model in Paper VI the assumed combustion behaviour of the bed model is questioned. This is further investigated in Paper VII. To achieve a more detailed knowledge of the combustion behaviour in fixed beds, two additional small-scale combustors were built, Paper VIII. The validity of the measurements from these units is assured by comparison with measurements from similar units found in the literature. For supplementary analyses an additional model of the conversion in a fixed fuel bed is derived. This model is optimised for the determination of the maximum possible propagation rate of the reaction front in a fuel bed, ignited from one end and with air introduced from the other end, Paper VII.

Furnaces and Reactors used for Experiments


In Paper I to VIII measurements were carried out in a number of test units: a 12 MWth Circulating Fluidized Bed (CFB) combustor located at Chalmers University of Technology, a 31 MWth reciprocating grate furnace located in Trollhttan (Sweden), three laboratory furnaces for investigation of the propagation rate of a reaction front in fixed beds, two at SP (Swedish National Testing and Research Centre, Bors Sweden), and one at VTT (Technical Research Centre of Finland, VTT-Energy Jyveskyl, Finland). A laboratory scale fluidized bed located at SP was used for investigations of single particles. A general outline of the CFB combustor used in the experiments for Paper I and II is given in Figure 1. The primary air for combustion of the fuel and fluidization of the inert bed is introduced in the bottom of the riser. The fuel is fed onto the inert bed and kept there until it is converted. Circulating fluidized beds are operating at a gas velocity that is higher than the

Flue gas Furnace Wall heattransfer surface Air Fuel inlet Cyclone
Heat Transfer Surfaces Flue Gas Secondary Air Flue Gas Circulation Fuel Secondary Air Flue Gas Circulation

Air
Primary Air Ashes

Figure 1 General outline of the 12 MWth CFB combustor at Chalmers University of Technology. Cross-section of riser 1.71.44 m, height 13.5 m.

Figure 2 General outline of the 31 MWth reciprocating grate furnace in Trollhttan, fabricated by Kvaerner Pulping AB. Grate cross-section 85 m, height 12.5 m.

terminal velocity of the inert particles, resulting in entrainment of inert and fuel particles by the gas up through and out of the riser. The particles leaving the riser are separated from the gas in a cyclone and are reintroduced to the riser. The main advantage with fluidized bed combustors is that primary measures can be taken for reduction of harmful emissions, facilitated by operation at a nearly constant temperature, and furthermore, that they are fuel flexible, as the fuel is kept in the inert bed until it is fully converted. The predominant primary measure is introduction of limestone for sulphur capture. The Chalmers CFB combustor is specially designed for research and is prepared with a large number of measurement holes, allowing measurement probes of different kinds to be introduced into the combustor. Furthermore, the combustor is equipped with a large amount of measurement equipment and a permanent system for gas analysis. For the measurements referred to in Paper I and II, probes were used for gas and particle sampling, heat transfer to a single particle, surface temperature of burning fuel particles and for recording reducing and oxidising conditions. An outline of the reciprocating grate furnace referred to in Paper VI and VIII is shown in Figure 2. This furnace is manufactured by Kvaerner Pulping AB and is operated by Trollhttan Energi AB as a part of the district heating system in the city of Trollhttan. In a reciprocating grate the fuel is introduced from one end of the furnace and transported by reciprocating rods and burned along the grate. When the fuel reaches the other end of the grate its conversion is finished and the remaining ash leaves the grate. The primary air for conversion is distributed in five zones along the grate to optimise the conversion of the fuel. For this project the furnace in Trollhttan was specially provided with measurement holes for probes, and during the measurements the furnace was prepared with temporary equipment to measure the primary airflow and the temperature of the surface of the grate. Four types of measurement probes where specially designed for the measurements, two for gas sampling, where one was combined with the measurement of gas temperature, one to measure the

fluctuation between reducing and oxidising conditions, one for detection of the gas flow direction. The combustion of a fixed bed of biofuel was investigated in three units, Papers VII and VIII. Two of the units, one large and one small, where designed and built at SP, Figure 3 and Figure 4, and the third was designed and built at VTT, Figure 5, within the same project as the measurements in Trollhttan. The work was focused on measurement of the propagation rate of the reaction front in a fixed fuel bed, and special experiments were carried out using the same fuel as in Trollhttan. All these three units operate in the same way; a batch of fuel is placed on a grate, air is introduced from one end, and the bed is ignited from the opposite end of the fuel bed. The propagation rate is evaluated from temperature measurements, which are made in different positions, mainly along the height in the fuel bed. The large unit at SP and the unit at VTT are prepared for gas analysis. There are differences between the units related to size and design, but the only major difference is in the design of the large unit at SP, Figure 3. This unit is up side down concerning the ignition and introduction of air compared to the other two designs, Figure 4 and Figure 5. For measurements on single particles, Paper III and IV, a small fluidized bed reactor, Figure 6, operating under bubbling conditions (gas velocity lower than the terminal velocity of the inert bed particles), was used. The reactor is electrically heated to the temperature of the experiments and particles were dropped into the reactor from the top. Gas leaving the reactor is sampled and analysed.
Air Holes for Measurement Probes

Flue Gas Fuel Thermo Couples

Distributor Plates Fuel

Flue gas

Scale

Air

Figure 3 General outline of the large experimental furnace at SP. Combustion chamber cross-section 0.520.48 m, height 0.7 m

Figure 4 General outline of the small experimental furnace at SP. Combustion chamber diameter 0.2m, height 0.45 m.

Flue Gas Gas sampling for analysis Fuel Thermocouples

Flue gas Gas sampling for analysis Fluidised bed


III II III II

Electrical heaters

Air Scale

Quartz reactor Air

Figure 5 General outline of the experimental furnace at VTT. Combustion chamber diameter 0.244 m, height 0.3 m.

Figure 6 General outline of small fluidized bed reactor for single particle experiments at SP. Reactor diameter 0.056 m, height 1.4 m.

Summary of Papers
Paper I Surprisingly high temperatures of burning particles of bituminous coal, temperature levels that can be suspected to cause a production of thermal NOx in the surrounding gasfilms, were measured with a two-colour optical pyrometer in the upper part of the Chalmers CFB combustor. Excess temperatures up to 600K above the bed temperature of 1123 K were registered, Figure 7a, while the cross-section average oxygen concentration was around 6%. The excess temperature in Figure 7a is plotted against the X-factor, which is an equivalent view factor of the particle in the view field of the pyrometer, obtained from the analysis of the measurement. The X-factor for particles of different sizes located along the centre position of the view-field is shown in Figure 7b. In order to investigate at which oxygen concentrations and particle sizes these high temperatures could exist, a diagram showing the ignition behaviour of a fuel was created, based on a simple heat balance of burning spherical particles, Figure 8. In the figure the Damkhler number, defined as the reaction rate divided by the mass transfer coefficient, characterizes the ignition behaviour, as combustion goes from kinetic control to diffusion control. By means of the diagram, it is possible to discuss the possibility and conditions of high excess temperatures. The diagram can be built-up for all types of combustion conditions, but the paper focuses on the upper part of the CFB combustor. From the diagram, Figure 8, it is seen that ignition takes place if the oxygen concentration is above 8% and for fuel particle sizes of around 1 mm. This means that the oxygen concentration must be at least 1/3 higher than the cross-sectional average value before ignition takes place. In order to get excess temperatures of more than 500K, the oxygen concentration needs to be at lest twice as high as the highest measured average concentration, and the particle must be between 0.5 and 2 mm. The only explanation for the high particle 5

800

10

di , mm

Excess temperature, K

600

10

-1

1 0.5

X-factor, -

400

10

-2

0.2 0.1

200

10

-3

0 -4 10

10

-3

10

-2

10

-1

10

10

-4

a)

X-factor, -

10

-3

10

-2

10

-1

b)

Distance from the probe tip, m

Figure 7 a) Temperature measured with two-colour pyrometry as a function of the X-factor. b) The X-factor as a function of distance from probe-tip, for different particle diameters. temperatures measured must therefore be a wide fluctuation of local oxygen concentration, resulting from insufficient mixing. The presence of such fluctuations is confirmed by simultaneous measurements of reducing and oxidising conditions. The time intervals during oxidising conditions are short, and in order to allow a particle to have a sufficiently long residence time in the oxygen-rich environment to be ignited, the particle must follow the gas up through the furnace. The ability of particles to follow the gas is given by the terminal velocity, which in a gas flow is a function of the drag coefficient, and in a circulating fluidized bed, enhanced by the collisions with surrounding small particles. In general, small particles are more likely to follow the gas than larger particles, and consequently it is more likely that the smaller fuel particles experience long enough periods in oxygen-rich environment to allow them to ignite and achieve high excess temperatures. Besides the analysis summarised here, Paper I includes a large work on heat and mass transfer to a single particle, which is covered extensively in the doctoral theses [12] and [13].
600
Da=5 10

Excess temperature, K

500
2

400
1

300
0.5 16 14

200 100 0 10

12

0.2 0.1

Y ,% 10 O2 9 8 7 6 4

-4

10

-3

10

-2

Particle diameter, m

Figure 8 Excess temperature versus particle diameter for different oxygen concentrations (solid lines), and Damkhler numbers (dashed lines).

Paper II The fuel loading of a fluidized bed combustors affects the control of operation; a small fuel loading results in a fast response to changes in operating condition, whereas the opposite is achieved for large fuel loading. Here, a general model of fuel loading and size distribution of fuel in a fluidized bed combustor has been developed. Specially, for the circulating fluidized bed combustor is that the fuel concentration and size distribution are modelled along the riser. Devolatilisation is modelled with the classical empirical particle diameter power law (t=adb) and char combustion as a shrinking sphere exposed to a surface reaction. During both devolatilisation and char combustion the fuel undergoes fragmentation and attrition, which for the fuel loading is nearly as important as the rate of devolatilisation and char combustion. The fragmentation and attrition are to some extent treated in a new way. The new treatment is in the definition of the rates of fragmentation and attrition, and size distribution of the fragments. A good agreement between model and measurements in Chalmers 12 MWth CFBC is obtained, Figure 9. The fragmentation rate constants obtained from the evaluation of experimental data also show a good agreement with rate constants from literature data. The model includes the most influencing parameters and is computationally efficient.

Fuel mass p.d.f. [1/mm]

250

Fuel mass p.d.f. [1/mm]

500

500

250

0 0.01 500

0.1 1 10 Particle diameter [mm]


Fuel concentration []

0 0.01

0.1 1 10 Particle diameter [mm]

Fuel mass p.d.f. [1/mm]

0.06

0.04

250

0.02

0 0.01

0.1 1 10 Particle diameter [mm]

0 0

5 Height [m]

10

Figure 9 Comparison between measured and modelled size distribution and fuel concentration on different heights.

Despite of the complexity of the fragmentation and attrition during combustion in a fluidized bed combustor the model derived in this paper gives a good agreement with measurement by adjusting four empirical coefficients. The values of these empirical coefficients are severely restricted by both the resulting fuel loading and size distribution, and by the conversion of the solid fuel. For example, in the measurements there are nearly no large particles found, indicating that the large fuel particles fragment almost immediately after they are introduced into the combustor. This significantly reduces the fuel loading, which is a severe restriction on the possibility to fit the fuel loading. Furthermore a large number of fuel particles found in the fly ash indicates the amount of particle produced by attrition, and limits the minimum rate of attrition. The attrition also reduces the fuel loading, and to keep the fuel loading at the measured level, which leaves the fragmentation rate during char combustion as the only parameter that are not directly restricted by the measurements. For the coal used in the measurements it is concluded that this fragmentation rate must be rather small. Since, a significantly larger fragmentation rate during char combustion would lead to a much lower fuel loading than the measured one. The parametric study shows that the fuel loading depends on type and size of fuel, fragmentation and gas pressure inside the combustor; all these parameters have a great influence on combustion. A rise in pressure and/or superficial velocity increases the power output from the combustor, and also the fuel loading. Bed temperature and air to fuel ratio also have an influence on the fuel loading, but not as great as that of the other parameters investigated. The size distribution of fuel in the bed is mostly dependent on the fragmentation behaviour of the fuel. For a CFBC, the superficial gas velocity mainly affects the variation of fuel concentration along the riser. Paper III A simplified model of thermochemical particle conversion, independent of particle size and shape, is derived. The model operates with a small number of variables and treats the most essential features of conversion of solid fuel particles, such as temperature gradients inside a particle, release of volatiles, shrinkage and swelling, considering also typical shapes (spheres, finite cylinders and parallelepipeds). The model treats the particle in one dimension, and the conversion can be described by the heat and mass transport to the surface of the particle. The derivation of a control surface of a finite cylinder is illustrated in Figure 10. When modelling a large combustion system, such as a fluidized bed or a fixed bed on a grate, this is a great advantage, as the model should not be limited to just a single particle. In fact, it can handle the conversion of a solid phase in a computational cell, where the conversion is related to surface area per unit volume instead of surface area of a single particle. The model divides the particle into four layers: moist (virgin) wood, dry wood, char residue and ash. The development and the temperature of these layers are computed as function of time, as shown in Figure 11. The model is validated numerically by a heat and mass balance, and it agrees well with measurements performed on more than 60 samples of particles of different sizes, wood species and moisture contents. A comparison with experimental time of devolatilisation, Figure 12, and time of char combustion, Figure 13, shows that the substantial simplifications made do not severely influence the overall agreement of the model. The model calculations predict the great influence from shrinkage on time of devolatilisation and char combustion. This is shown in Figure 14, which illustrates the influence of shrinkage for the largest particle size of spruce used in the experiments 102540 mm. In the figure the shrinkage by volume during drying and devolatilisation is chosen to 15, 35 and 45%. Larger shrinkage results in a

more compact char layer, and consequently higher thermal conductivity and a shorter time of devolatilisation. On the other hand, for char combustion the trends are the opposite as shown by the model, because increased shrinkage means a smaller external surface. The shrinkage coefficients used in the calculation are obtained from direct measurement, supported by a qualitative judgment of the fit between measured and calculated times of devolatilisation and char combustion.
l l

x 10 5
d/2

-3

r0=(d-l)/2
d

Particle radius [m]

l/2 l/2

d/2 r0=0

4 3 2 1
3 4 1 2

r d r

r r

200 500

400 Time [s] 700

600

T[K] 300

900 1100

Figure 10 Definition of the starting position of the radius r0 and the control surface (solid line) at radius r inside a finite cylinder (dashed line) having a length l, longer or shorter than the diameter d.

Figure 11 Example of simulated temperature and positions of drying (1), devolatilisation (2), char combustion (3) and particle surface (4) inside a wet spruce particle as a function of time.

200
Measured time [s]

400

20%
Measured time [s]

150 -20% 100 50 0

300 200 100 0

20% -20%

100 200 Calculated time [s]

200 Calculated time [s]

400

Figure 12 Comparison between calculated and measured time of devolatilisation for 61 samples of, wet (o) and dry ( ) spruce, and wet () and dry (+) birch.

Figure 13 Comparison between calculated and measured time of char combustion for 61 samples of, wet (o) and dry ( ) spruce, and wet () and dry (+) birch.

400

Measured time [s]

200

200 400 Calculated time [s]

Figure 14 Times of devolatilisation (o) and char combustion (). Arrows indicate the influence of increasing shrinkage. Paper IV For comprehensive models of combustion devices, for example fixed and fluidized beds, there are a need of a sub model providing a sufficient composition of the volatile gases leaving a fuel particle of typical sizes. The real composition of the volatile gases becomes too complex for these models, since, for example, for wood the composition consists of more than 100 species [14]. However, much work has been published on releases of volatiles [15-25], but a general model of the composition of volatile gases and a comprehensive presentation of the related thermo-chemical properties of the fuel are still missing. Due to the complexity of the transformation of the fuel, it becomes excessively time-consuming to model the gas leaving the thermally large particles by means of a set of reaction rates. Because of time restrictions a more simplified description of volatile release is needed for comprehensive bed models. Here, a sub model is presented, whose structure is valid for any solid fuel. The model solves a system of six equations to obtain the six gas concentrations. The system of equations consists of three mass balances, one energy balance and two empirical ratios. In some cases, dependent on the available input data, the energy balance can be replaced by an additional empirical ratio, and the energy balance can be used for validation. The model includes empirical coefficients, and they have to be specified for certain classes of fuel. Gases from devolatilization of the dry part of high-volatile fuel are assumed to consist of CO2, CO, H2O, H2, light hydrocarbons (mainly methane and ethylene), and lumped hydrocarbons (that is, the remaining hydrocarbons). The method proposed is for estimation of the quantities of these gas components in a case when no data are available, or for a measured set of data needed to be checked. In addition to the empirical ratios, thermo-chemical properties of the fuel during its various phases of conversion are needed. Therefore, measurements have been carried out on two types of wood to provide data that can be compared with data from literature. This collection of data is especially selected to suit combustion and high temperature gasification. The data are specialized to wood. The measurements concern softwood (spruce) and hardwood (birch), but similar collection of data can be established for other fuels as well. The measurements were carried out in a fluidized bed reactor operated at 1123 K and the results show a clear correlation between the size of the fuel, expressed as specific area (initial surface area divided by initial volume of the particle), and several quantities such as char yield, ratios of CO to CO2, and light hydrocarbon to CO2, Figure 15. As the specific area increases (size decreases) a nearly linear decrease of the char yield and linear rise of the ratios

10

1.8
THC/CO2 (by mass)

4.0
CO/CO2 (by mass)

x100

20 10 0

H x1000 CO x100 x100 C

1.6 1.4 1.2 1.0 0.8 300 1.8

3.5 3.0 2.5 2.0 300

20 10 0 20 10 0 20 10 0 300 600 900 Specific area [1/m]

2O

600 900 Specific Area [1/m]

600 900 Specific Area [1/m]

CO x50

20 10 0

H2O/CO2 (by mass)

Char Yield [%]

1.6 1.4 1.2 1.0 0.8 300 600 900 Specific Area [1/m]

22

C H x100

18 14 10 300 600 900 Specific Area [1/m]

20 10

0 300 600 900 Specific area [1/m]

Figure 15 Ratios of THC to CO2, CO to CO2 and H2O to CO2, and char yield, related to the specific area of the wood particles, birch (wet, +, dry, *, solid trend lines), spruce (wet, o, dry, , dashed trend line).

Figure 16 Estimated composition of volatile gases, expressed as mass fractions, versus specific area, for birch (* solid line) and spruce ( and dashed line). (Observe the different scaling factors).

of CO to CO2, and light hydrocarbon to CO2 can be seen. The lower char yield is accompanied by an increasing amount of carbon relative to hydrogen and oxygen in the volatile gases, which, together with the two ratios mentioned, affects the composition of the volatile gases, since it favors the lumped hydrocarbons on the behalf of the carbon dioxide. The two tested woods, hardwood (birch) and softwood (spruce), show the same trends. The only differences are in the level of char yield and in amount of H2O in the volatile gases. Further work could involve other fuels and a more detailed investigation on the influence of temperature. In the present work the most relevant temperature for the application was studied. Analysis of the measurement data with the model derived for the composition of the volatiles, Figure 16, shows that the mass fractions of water, hydrogen, carbon monoxide and light hydrocarbons are rather constant with specific area, but the level of carbon dioxide decreases on the behalf of a rise of the mass fraction of lumped hydrocarbons. Paper V The effective thermal conductivity is one of the most important parameters for modelling of the thermo-chemical conversion of wood. It varies both with temperature and conversion of the wood. There are measurements available in the literature and suggestions on modelling of this problem, especially for wet and dry wood, but for char the knowledge is poor. Here, two principal models of effective thermal conductivity on the basis of the pore structure in wood, [26] and [27], are validated by a comparison with direct numerical simulation of the fibre structure. The validation leads to a more general model, both for conductivity in the perpendicular and parallel direction relative to the fibres in the wood. A secondary result is that the thermal conductivity of the solid phase in the fibre wall of the wood can be evaluated for dry and wet fuel from measurement data on effective thermal conductivity. The effective thermal conductivity can be estimated from given values of temperature, density and moisture content of the wood. It can also be applied to pellets and chipboards. In addition, the general 11

nHmOk

i j

model expresses the effective thermal conductivity of char, since the wood material maintains its fibre structure during conversion. From the model, the thermal conductivity perpendicular to the fibre is estimated to 0.52 W/mK for the solid material in the dry wood by a least-square fit to measurement data from [28] and other authors, reviewed in [29]. Parallel to the fibre, the corresponding thermal conductivity is estimated to 0.73 W/mK, based on measurements reviewed in [29]. The agreement between the model derived here and the models of [26] and [27] and measurement data can be seen in Figure 17, for dry wood, and Figure 18, for wet wood. There is good agreement between all of the models and the measurement data for dry wood, but the difference becomes noticeable if one regards the heat conductivity of the solid in the fibre. In [26] this heat conductivity is suggested to be 0.43W/mK, but in [27] it is suggested to 0.6W/mK. If the values mentioned are inserted into their respective equations, there is an even better agreement with measurements than the one shown in Figure 17, but for high density fuels the difference becomes noticeable. If these values are used, for example, to estimate the heat conductivity of pellets, which have a high density (1000-1300 kg/m3), the difference between the models can be more than 15 %. When char is produced at a high heating rate and with a high final temperature, the resulting porous structure is nearly pure carbon. Here, it is proposed that carbon in the porous structure has the same thermal conductivity and density as amorphous carbon.

0.8 0.7
Measured thermal conductivity [W/mK]

0.5

Heat conductivity [W/mK]

0.6 0.5 0.4 0.3 0.2 0.1 0 0 500 1000 Dry density [kg/m3] 1500

0.4

0.3

0.2

0.1

0.1 0.2 0.3 0.4 0.5 Calculated thermal conductivity [W/mK]

Figure 17 Heat conductivity of dry wood as a function of density. Conductivity parallel to fibre (solid line with dots), measured data, [29] (stars). Conductivity perpendical to ibre, model derived here (solid line), model derived in [27] (dashed line), [26] (dotted line). Measured, [28] (circles), [29] (rhombs).

Figure 18 Comparison of measured, [28], and calculated effective thermal conductivity perpendicular to the fibres in wood, with a moisture content of 7-45% (based on wet wood).

12

Paper VI During the last decade the gaseous combustion in the free-room above the grate in a reciprocating grate furnace has been modelled with different CFD (computational fluid dynamics) tools, where the combustible gas leaving the bed is distributed along the grate based on the assumed combustion behaviour of the fuel bed. In the present work, an independently derived bed model is connected to a CFD calculation of the gaseous combustion above the bed, in order to model an entire furnace. The main task of the work is to investigate if the combustion of the fuel bed and the gaseous combustion above the bed can be calculated separately, and then combined efficiently in an iterative calculation. For this purpose a simplified bed model was derived, which separates the bed in four layers; surface layer, char layer, reaction front layer and virgin fuel layer, based on an assumed combustion behaviour of the bed, Figure 19. For each layer a heat and mass balance is established, which describes the conversion of the fuel bed. The layers represent: the nonreacting layer of virgin fuel, the propagation of the drying, devolatilisation and gaseous reaction, the conversion of the char residue, the surface of the bed that is interacting with the surrounding. Input to the model are data on the fuel fed, such as fuel properties, bed height and porosity, and along the grate, such as equivalent radiation temperature for the radiation exchange with the surface of the bed, gas flow and velocity of the bed. The bed model gives resulting position of the reaction front, bed height, composition and temperature of the gas, and temperatures inside the bed, Figure 20. Heat transfer and gas exchange between the bed and the free-room above the bed is used as boundary condition to the two models, which results in an iterative calculation procedure, where the equivalent radiation temperature above the grate is assumed first, and the conversion of the fuel bed is calculated. This is followed by a CFD-calculation of the gaseous combustion above the bed, where surface temperature of the bed and temperature and composition of the gas leaving the fuel are used as boundary conditions. The heat flux to the bed resulting from the CFD calculation is recalculated to an equivalent radiation temperature, and used as a boundary condition for the fuel bed, and so on. The experience from the calculations made is that the solution reaches convergence in just a few iterations.
Heat Flux

xs

Bed height

xc xr
Re act
Unreac l layer ted fue

Su
Ch

rfa

ce

lay

er

Figure 19 Schematic figure of the four layers describing the bed. The thicknesses of the layers are indicated by x and a subscript related to the layer.

ion fro nt

ar

xU

Primary air Time or Length of grate

la r ye

13

2000

Temperature [K]

1600 1200 800 400 0 10 Char layer Surface layer Surrounding (Input)

Outgoing Molar Flow [moles/s]

CH4 CO O2 CO2 H O 2 N
2

Bed height [m]

0.4 0.3 0.2 0.1 0 10

Bed surface Reaction front Mass loss

0.8 0.6 0.4 0.2 0 7.5

Ingoing Molar Flow [moles/s] (Input)

Bed vel. (Input)

4.5

1.5

4 Bed length [m]

Velocity change bed Velocity change gas

Figure 20 Input data and result of modelling. The grey areas indicate the given changes in the primary airflow and velocity of the fuel layer along the grate.

14

Bed velocity [mm/s] (Input)

N2 O2 H O

(Input) (Input) (Input)

Relative mass loss []

0.5

In order to validate the models, extensive measurements on a commercial reciprocating grate furnace were carried out for two different operating conditions. In these measurements gas sampling and temperature measurements were made in more than 80 positions. Furthermore, continuous sampling was made of more than 30 operating parameters. Average data from most of these measurements are reported in Paper VI, but some additional data can be found in Paper VII. The extensive documentation of the measurement data in Paper VI and VII makes the data useful for validation of new and existing models and for comparison with other furnaces. The measurements in the lower part of the combustor show on a major combustion zone in the first part along the grate, where the volatiles enter the furnace from the bed, and on a final char burnout in the later part along the grate. The measurements also show that the final burnout of the gases leaving the bed takes place in a narrow region around the introduction of the secondary air. Furthermore, an internal circulation of gases in the upper part of the furnace is seen from the measurements. When comparing the model calculations with the measurements, there are clear indications on some general errors in the assumed combustion behaviour made in the bed model, which is discussed in paper VII. However, in the upper part above the secondary air inlet the gases from the lower part of the furnace are well mixed and a better agreement between calculation and measurements is observed in this region. Paper VII Grate firing is the most common way to burn bio-fuels in small-scale units. Different combustion modes are achieved depending on how fuel and primary air are introduced. In continuous systems fuel and air are usually fed in cross-current and counter-current flow. In Paper VI the generally assumed combustion behaviour of wet biofuels was questioned and, therefore, the combustion in the mentioned 31 MW reciprocating grate furnace (a crosscurrent flow combustor) is further studied. For the investigation, additional experiments were carried out in batch-fired pot furnaces, using the same fuel as in the reciprocating grate, forest waste with a moisture content of approximately 50%. The generally accepted suggestion on the combustion in a cross-current flow furnace is that it start by ignition on the surface of the bed, followed by a reaction front propagating from the surface down to the grate, [7-11]. Measurements and visual observations presented here show, however, that in the case of wet fuels the ignition takes place close to the grate, followed by a reaction front propagating from the grate up to the surface of the bed. Hence, the progress of combustion in the bed is opposite to the expected one. The measurement positions in the grate furnace are shown in Figure 21.

Figure 21 Position of measurements in the reciprocating grate furnace. Gas sampling ( ), Thermocouples ( ) (31 MW furnace, Kvaerner Pulping AB)

15

Figure 22 Principle combustion behaviours in a cross-current flow unit. (a) if the bed is ignited from the surface of the bed, or (b) from the surface of the grate. A position indicates of ignition, B vertical view, C reaction front, I unreacted fuel, II drying, III devolatilisation, IV char combustion or gasification, V ash.

Figure 23 Measured time-average temperature on the reciprocating grate. Filled circles indicate measurement with pure wood chips and unfilled circles are measurement with forest waste.

Figure 24 Total hydrocarbon concentration above the bed of the reciprocating grate. Filled circles indicate the second measurement row and unfilled circles the first measurement row (Cf. Fig 3). Measurements from operation with pure wood chips.

Figure 25 Calculated maximum propagation rate of the reaction front for 10mm wood particles, dotted lines. Solid lines are curve fits of the measurements. The measurements are indicated by capital letters and calculation with small letters, the different moisture contents are indicated by, A,a 10% [8], B,b 30% [8], C,c 40.3% (VTT) and D,d 56.6% (SP).

16

The result from the investigation is that the progress of combustion can not be the one generally expected, shown in Figure 22a, according to following observation: 1. In the pot furnaces the reaction front could not propagate through the bed for the same type of fuel and air flows as in the furnace, where the fuel burned without problems. This observation is supported by the model calculations. 2. Even the highest velocity of the reaction front, 0.35 mm/s, measured for wood cubes with a moisture content of 10%, is not sufficient for the front to reach the surface of the reciprocating grate within the first 4 m, if the bed is ignited from the top. For a bed height of 500 mm, it would take 1430 s for the reaction front to propagate from the surface of the bed to the grate. With the velocity of the grate, 6 mm/s, the reaction front would reach the grate 8.6 m after ignition on the surface. The corresponding distance at velocities measured for wet fuels is 18 to 60 m. In contrast, in the furnace case, according to Figure 23, the high temperatures clearly indicate that combustion takes place at or close to the surface of the grate after less than 1.8 m. 3. The temperature on the grate, Figure 23, shows that heat was generated in the bed before ignition was visually observed on the surface of the bed. 4. There was a heavy evolution of smoke from the bed before the position of ignition on the surface of the bed and before the visible flame zone, according to visual observations in the reciprocating grate furnace. 5. Volatiles were found far from the fuel inlet, Figure 24. A moist wood-chip particle (50% moisture) dries and devolatilises completely in two minutes at a temperature of 973K, [30]. The temperature inside the bed is around 1300K, [8] and Paper VIII, during char combustion, and the actual time for complete drying and devolatilisation should be less than two minutes. The fuel is transported 0.7 m along the grate in two minutes. Ignition is measured within the first 1.8 m for pure wood chips on the reciprocating grate, Figure 23, and all moisture and volatiles should have left the bed before 2.5 m (1.8+0.7 m) distance from the feed point if the reaction had propagated from the top of the bed to the grate. However, Figure 24 shows that volatiles leave the bed up to 4-5 m from the beginning of the grate. The conclusion from these facts is that the ignition has not taken place on the surface of the bed, but in the bed, most likely on the surface of the grate. The arguments supporting this observation and that reaction propagates up through the bed, Figure 22b, are: 1. In such a case, the theoretical maximum moisture content is attained when the lower heating value for moist fuel is equal to zero. This occurs at a moisture content of 85 to 90%, excluding heat losses. The reaction takes place inside the bed in a location that is quite well insulated from the surrounding. Moisture content of up to 70 or 80 % should therefore generate sufficient heat for drying and devolatilisation. The heat is generated at the bottom of the bed by char combustion and is transported up through the bed by the gas, which dries and devolatilises fresh fuel. For a counter-current bed, which is ignited from the top, on the other hand, this moisture content is too high, since in this case, devolatilisation and combustion takes place in a narrow front. The calculation shows that the maximum moisture content the range of air flows considered is between 35 and 45%. It is also clear from the calculation that the velocity of the reaction front becomes very slow for these moisture contents. 2. Combustion and heat release in the bottom part of the bed generates gas, which transfers its heat to the fresh fuel and exits the bed at a temperature, not much higher than that of entering fuel particles. This explains the heavy evolution of smoke, observed to leave the first 1 to 1.5 m of the bed. 17

3. The reaction front propagates up through the bed and the devolatilisation continues until the devolatilisation front reaches the surface of the bed. This explains the volatile release far down on the grate, see Figure 22. 4. Ignition of the fuel on the surface of the grate takes place as soon as the fuel has reached the ignition temperature. Larger particles need longer time to heat up and ignite, and consequently there is a later ignition of the pure wood chips than of the forest waste that includes a large portion of fine saw-dust, although the moisture content was 10% higher, Figure 23. Large scale mixing of the fuel bed could be another reason for the temperature raise before 1.8 m from the fuel inlet, but ignition or stirring caused by the transport of the bed along the grate were not observed on the surface during the first 1 to 1.5 m from the fuel inlet. Another fact that talks against large scale mixing is the tendency of wood chips to stick together. The low measured propagation rates of the reaction front in a wet fuel bed are confirmed to be close to the theoretical maximum. This means that the propagation rate in a reciprocating grate cannot be that much higher in the same combustion situation, even if two-dimensional effects are present. The theoretical maximum is determined by a calculation where all ingoing parameters are chosen in favour of the propagation of the reaction front. In Figure 25 it is seen that all measurement data, [8] and measurements carried out in the present work in the large unit at SP and at VTT, are within or close to the theoretical maximum given by the model. For the highest moisture contents there can be observed a slightly higher propagation rate than the theoretical maximum one. The reason for this is most likely that the moisture content of the fuel was somewhat lowered during ignition of the bed. A large part of the fuel had to be dried by the heat added for ignition before the fuel ignited, and during this time also the fuel below dried by heat from the same heating source and from the convective air flow through the bed. This result questions some design principles for grate furnaces: the ignition arches designed to reflect radiation from the flames to the bed surface to ignite the bed are not needed. However, the design of the furnaces may not change after all, since these arches also have positive affects on the gas flow in the furnace. When publishing this paper we did not find any support for this combustion behaviour in the literature. However, later on it was found that the same observation was made in [31], where combustion of wet cellulose on a laboratory travelling grate was investigated. Furthermore, after submission of the paper measurements from burning wood chips on a travelling grate was presented in [32], where the same combustion behaviour is indicated in a small grate furnace. However, no conclusion was drawn except that further investigation is needed. Paper VIII In order to further investigate combustion of a fuel bed placed on a grate, two small batchfired furnaces were designed and built. In the first furnace, the larger unit, air is introduced from the top end and ignition take place in the bottom end. This unit has the advantage that heat transfer and thereby the ignition can be controlled. Furthermore, the relatively large surface area of the grate gives a one-dimensional combustion behaviour and allows larger particle sizes. The larger size also gives the opportunity to supply the unit with more measurement equipment and to penetrate the fuel bed with measurement probes without a significant disturbance of the combustion behaviour, thus creating a better possibility to follow the progress of the reaction front. The disadvantages consist in the difficulties in interpreting the measurement results from inside the bed, as the relative position of the location of the measurement inside the bed changes as the bed shrinks away, as the reaction

18

0.12 Ignition rate (kg/ms) 0.10 0.08 0.06 0.04 0.02 0.00 0.0

0.12 0.10 0.08 0.06 0.04 0.02 0.00 0.0


wood 8 mm wood 10 mm, Gort wood 5-20 mm, Horttanainen

Large test rig, pellet 8mm Small test rig, pellet 8mm

0.1

0.2

0.3 0.4 0.5 0.6 Air mass flow (kg/ms)

Ignition rate (kg/ms)

0.1

0.2

0.3 0.4 0.5 0.6 Air mass flow (kg/ms)

Figure 26 Measured ignition rate (coincide with the propagation rate) in a bed of 8 mm wood pellets as a function of airflow, in the large and the small unit.

Figure 27 Ignition rate for 8-mm wood cylinders, the small unit, 10-mm wood cubes, [8], and 5-20-mm, [33], wood chips as function of air mass flow rate.

front in the bed propagates from the grate to the surface of the bed. Furthermore, the grate, on which the fuel bed rests, must be provided with a few rather large holes, in order to avoid the grate to clog. This results in an unwanted gas flow pattern close to the surface of the grate. In the second furnace, the smaller unit, the air is instead introduced from the bottom and the ignition takes place at the top. In this furnace the ignition and heat transfer to fuel bed surface is not controlled, but an even gas flow through the bed can easily be arranged and the relative bed position of the measurements inside the furnace is easier to approximate, as a position inside the bed is fixed until the reaction front passes. The smaller unit is also much easier to operate, which makes it more sufficient for initial and preliminarily experiments. The performance of the two units was validated by comparing results from tests with fuel beds of uniform fuel particles operated at different airflows with similar tests in other furnaces reported in the literature [8] and [33], Figure 26 and Figure 27. The comparison shows that the units built provide the expected behaviour and can be used for more detailed investigations in future work.

Conclusions
Paper I. Fluidized bed combustors are usually considered to be practically isothermal due to the good mixing provided by the fluidization of inert material, which have a high thermal capacity. However, in the present work, measurements of surface temperature in a large scale fluidized bed combustor show that the fuel particles can have excess temperatures of 500 to 600 K in relation to the bed temperature. Model calculations show that these excess temperatures only can be present at higher oxygen concentrations than the mean concentration measured in the combustor. This indicates that the mixing is not as good and the temperature is not as even as expected.

19

Paper II. A general model of fuel loading and size distribution in a fluidized bed combustor (FBC) has been developed. For the circulating fluidized bed combustor (CFBC) the fuel concentration and size distribution along the riser are also modelled. The fragmentation is to some extent treated in a new way. The fragmentation rate constants obtained from the evaluation of experimental data show a good agreement with the literature data. The model accounts for the most influencing parameters and is computationally efficient. Paper III. Modelling of a great number of large solid fuel particles in a combustion system needs a model that account for the most essential features of the conversion process with as few equations and as small computational effort as possible. The model presented here treats particles that change in size during conversion and are realistically shaped. The model describes the conversion of the fuel particles in a one-dimension formulation, which makes it suitable for Eulerian calculations. A comparison of the model with experiments shows that the simplifications made do not significantly influence the overall agreement of the model. Paper IV. When modelling combustion systems, especially of high volatile fuels, a representation is needed of the composition of the volatile gases. In the present work, a sub model for the composition of the volatile gases leaving particles that are non isothermal during conversion and fulfils the elemental species and energy balances of the fuel is derived. The model is validated for wood and the physical properties required for the modelling are summarised. Paper V. When modelling combustion of large wood particles, the thermal conductivity is one of the most important properties, since internal heat transfer in most cases is decisive for the conversion rate of the solid fuels. Previously there have been suggestions on how to model the thermal conductivity of wet and dry wood, but the models differ slightly. Therefore a more detailed investigation, based on modelling, has been carried out. The result of the investigation is a more general model and suggestions on how to model the thermal conductivity of char. Paper VI. A simplified bed model is derived for the combustion behaviour of a cross-current (reciprocating) grate furnace. The model accounts for the most essential physical processes of the conversion of the fuel bed along the grate that are modelled and based on the generally accepted combustion concept. The concept presumes that a reaction front propagates from the surface of the bed down to the surface of the grate. The bed model is successfully connected to a CFD-calculation of the gaseous combustion above the grate by an iterative procedure. For model validation, a survey is made of the different gas concentrations in the furnace by measurements in a 31 MW grate furnace. Comparison with measurements from a 31 MWth reciprocating grate furnace shows a good agreement in the upper part of the furnace, but the lower part indicates that there are some fundamental controversies in the model assumptions for the bed (To be treated in Paper VII). Paper VII. The generally expected combustion behaviour of cross-current furnace is that the bed is ignited on the top of the bed by radiation from the flames and ignition aches (assumption made in Paper VI). However, the present paper shows that this is not the case, especially for

20

wet biofuels. The measurements performed both in the large scale furnace and in laboratory furnaces show that the ignition has to start close to the grate, which means that it starts on the opposite side related to what is generally expected. This conclusion explains why the bed model derived in Paper VI did not properly work for the investigated grate furnace. Paper VIII. In order to further investigate the propagation rate of a reaction front moving in the same direction as the one generally expected in cross-current combustion, two furnaces were constructed and their functionality is proven by comparing measurement data with similar units found in the literature.

Future Work
In the papers assembled in the present thesis a number of models are derived. They can be combined in different ways and be complemented for further use and analysis of measurements performed. There are also test facilities built up for future investigation, especially for the combustion in a fixed bed. The work that will hopefully be done in the near future is to introduce the transient model for single particles into a general bed model, for supplementary evaluation of measurement data obtained from the small grate furnaces built and the reciprocating grate. Further, there are still a large part of the measurements carried out during this work that are not analysed; both from the reciprocating grate furnace and the CFB combustor. For example, a third measurement campaign was carried out in the reciprocating grate in Trollhttan. Other interesting work for the future is to formulate a transient model for the fuel loading in the CFB combustor, and to extract more information on the oxygen fluctuations in the CFB combustor, based on the surface temperature measurements.

References
1. Salatino, P, Massimilla, L., A predictive model of Carbon Attrition in Fluidized Bed Combustion and Gasification of a Graphite, Chemical Engineering Science, 1989, 44, 10911099. 2. Brown R.C., Ahrens, J., Christofides, N., The Contribution of Attrition and Fragmentation to Char Elutriation from Fluidized Beds, Combustion and Flame, 1992, 89,95-102. 3. Arena, U., Cammarota, A.,, Chirone, R., Primary and secondary fragmentation of coals in a circulating fluidized bed combustor, Twenty-Fifth Symposium (International) on Combustion, The Combustion Institute, Pittsburgh, 1994, 219-226. 4. Stubington, J.F., Moss, B., On the Timing of Primary Fragmentation During Bituminous Coal Particle Devolatilisation in a Fluidized Bed Combustor The Canadian Journal of Chemical Engineering, 1995, 73, 505-509. 5. Winter. F., Single Fuel particle and NOx/N2O-emissions characteristics under (circulating) fluidized bed combustor conditions, Academic Dissertation, University of Technology, Vienna, 1995. 6. Ross, D.P., Heidenreich, C.A., Zhang, D.K., Devolatilisation times of coal particles in a fluidized-bed, FUEL, 2000, 79, 873-883

21

7. Ford N.W.J., Cook, M.J., Sage, P.W., Modelling of fixed bed combustion, Fuel Processing Technology, 1993, 36, 55-63. 8. Gort, R., On the propagation of a reaction front in a packed bed; thermal conversion of municipal waste and biomass, Academic Dissertation, University of Twente, ISBN 909008751-6, 1995. 9. Shin, D., Chio, S., The Combustion of Simulated Waste Particles in a Fixed Bed, Combustion and Flame, 2000, 121, 167-180. 10. Saastamoinen, J.J, Hortanainen, M, Taipale, R., Sakomaa, P., Propagation of ignition front in fuel beds of particles, Combustion and Flame 2000, 123, 214-216. 11. Marskell W.G., Miller J.M., Mode of combustion of coal on a chain grate stoker, Article series in Fuel Science and Practice, vol. 25, 1946, p. 4-11, 50-62, 78-85, 109-113, 159-162. 12. Palchonock, G., Heat and Mass Transfer to Single Particle in Fluidized Bed, Academic Dissertation, Chalmers University of Technology, ISBN 91-7197-712, 1998. 13. Britholtz, C., Heat Transfer in Circulating Fluidized Bed Boilers, Academic Dissertation, Chalmers University of Technology, ISBN 91-7197-891-7, 2000. 14. Evans R.J., Milne, T.A., Molecular Characterization of the Pyrolysis of Biomass. 1. Fundamentals, Energy&Fuels, 1987, 1,123-137. 15. Roberts, A. F., A Review of Kinetics Data for the Pyrolysis of Wood and Related Substances, Combustion and Flame, 1970, 14, 261-272. 16. Kanury, A. M., Thermal Decomposition Kinetics of Wood Pyrolysis, Combustion and Flame, 1972, 18, 75-83. 17. Beamont, O, Schwob, Y., Influence of Physical and Chemical Parameters on Wood Pyrolysis,Industrial & Engineering Chemistry Process Design and Development, 1984, 23, 637-641. 18. Chan, W-C. R., Kelbon, M., Krieger-Brockett. B., Modelling and experimental verification of physical and chemical processes during pyrolysis of a large biomass particle, Fuel, 1985, 64, 1505-1513. 19. Nunn, T.R., Howard, J.B., Longwell, J.P., Peters, W.A., Product Composition and Kinetics in the Rapid Pyrolysis of Sweet Gum Hardwood, Industrial & Engineering Chemistry Process Design and Development, 1985, 24, 836-844. 20. Scott, D.S., Piskorz, J., Bergeougnou, M.A., Graham, R., The Role of Temperature in the fast Pyrolysis of Cellulose and wood, Industrial & engineering chemistry research, 1988, 27, 8-15. 21. Figueiredo, J. L., Valenzuela, C., Bernalte, A., Encinar, J. M., Pyrolysis of Holm-Oak Wood: Influence of Temperature and Particle Size, Fuel, 1989, 68, 1012-1016.

22

22. Horne, P. A., Williams, P. T., Influence of Temperature on the Products from Flash Pyrolysis of Biomass, Fuel, 1996, 75, 1051-1059. 23. Reina, J., Vole, E., Puigjaner, L., Kinetic Study of the Pyrolysis of Waste, Industrial Engineering in Chemical Research, American Chemical Society, 1998, 37, 4290-4295. 24. Di Blasi, C., Signorelli, G., Di Russo, C., Rea, G., Product Distribution from Pyrolysis of Wood and Agriculture Residues, Industrial & Engineering Chemical Research, 1999, 38, 2216-2224. 25. Di Blasi, C., Branca, C., Santoro, A., Gonzalez Hernadez, E., Pyrolytic behaviour and Products of Some Wood Varieties, Combustion and Flame, 2001, 124, 165-177. 26. Siau, J. F., Transport Processes in Wood, Springler-Verlag, Berlin, ISBN 3-540-125744, 1984. 27. Saastamoinen, J. J., Richard, J-R., Simultaneous Drying and Pyrolysis of Solid Fuel Particles, Combustion and Flame, 1996, 106, 288-300. 28. MacLean, J. D., Thermal Conductivity of Wood, Transactions American Society of Heating and Ventilating Engineers, 1941, 47, 323-354. 29. Grnli, M, A Theoretical and Experimental Study of the Thermal Degradation of Biomass, Academic Dissertation, The Norwegian University of Science and Technology, ISBN 82-471-0009-6, 1996. 30. Palchonok, G.I., Dikalenko, V.I., Stanchits, L.K. Borodulya, V.A., Werther, J. and Leckner, B. Kinetics of the Main Stages of Fluidized Bed Combustion of a wet biomass particle, Proc. of the 14th Int. on Fluidized Bed Combustion, Vancouver, Canada, F.D.S Preto ed, ASME, New York, 1997, 125-134. 31. Lamb, B.W., Bilger, R.W., The combustion of wet cellulosic fuel bed, Second Australasian Conference on Heat and Mass Transfer, The University of Sidney, 1977, 501508. 32. Weissinger, A., Fleck, T., Obernberger, I, Investigation on the Release of Nitrogen and Volatile Compounds from the Fuel Bed in Grate Furnaces, 1st World Conference on Biomass, Seville, Spain, 2000. 33. Horttanainen, M.V.A., Saastamoinen, J.J., Sarkomaa, P.J., Ignition front propargation in packed beds of wood particles, IFRF Combustion Journal, www.ifrf.net, Article Number 200003, ISSN 1562-479X, May 2000.

23

Errata list for the included Papers


Paper I Page 871 col. 2 line 5: U-V/ should be U/-V Page 871, a notation is missing: max largest diameter of a inert or active particle Page 873 Eq (1);
2 2 should be 1 + ma / mi 1 + mi / ma

Page 873 col. 1 line 33: and the inert (bed) particle of density b,c. The expression 2/(1+ma/mi) should be and the suspension of density b,c. The expression 2/(1+mi/ma) Page 873 col. 2 line 21: times an amount of heat absorbed should be times an amount of heat absorbed per unit temperature difference Page 875 in Figure 2b: di, mm should be da, mm Page 875 caption to Figure 3 line 7: the solid line of 9.4 mm should be The dashed line of 9.4 mm Page 877, an acknowledgment is missing: The authors would like to thank Timo Joutsenoja, Plasma technology laboratory, Tampere University of Technology, Finland, for help with pyrometer measurements. Paper II Page 5 Eq (26): part (1) CD Aext g v fuel v fuel should be CD Aext g v fuel v fuel / 8 , in part (3) ( R + Rinert ) should be ( R + Rinert )
2

Page 12 2nd reference Siciliano, L., should be Salatino, P., Page 12 a reference is missing: Stubington, J.F., Moss, B., On the Timing of Primary Fragmentation During Bituminous Coal Particle Devolatilisation in a Fluidized Bed Combustor The Canadian journal of chemical engineering, 1995, 73, 505-509. Appendix Page XII Eq (A48): ... + F0,vol ( y ) + ... should be ... + Ff ,vol ( y ) + ...

Appendix Page XVII reference missed for Brown et al. [1992]: should be Brown R.C., Ahrens, J., Christofides, N., The Contribution of Attrition and Fragmentation to Char Elutriation from Fluidized Beds, Combustion and Flame, 1992, 89,95-102 Appendix Page XIX definition missed for Ff2,char: should be Ff 2,char ( R ) =
Rmax

h ( R , R) M
* n R

char char

(R )k
*

f , char

( R ) dR
*

24

Populrvetenskaplig sammanfattning ( In Swedish )


Bakgrund

Under alla tider har olika typer av fasta brnslen ssom ved, torv och kol anvnts fr uppvrmning och under de senaste seklerna fr vrme och kraftproduktion. Fr det senare r kol den idag, och i alla fall under detta sekel, det enskilt viktigaste brnslet. Frbrnningen av kol skapar dock bde regionala och globala problem. Regionala problem r frmst utslpp av svaveloxider, kvveoxider samt tungmetaller. Globala problem r frmst utslpp av vxthusgasen koldioxid. Under de senaste rtiondena har intresset fr att skapa ett uthlligt energisystem och att reducera produktionen av vxthusgaser kat. Ett steg i denna riktning r att erstta kol med frnyelsebara alternativ. Fr fasta brnslen bestr dessa i huvudsak av rester frn skogs- och jordbruksindustrin. Mlet med detta arbete och med all forskning inom detta omrde r att ka frstelsen fr frbrnningen i syfte att minimera pverkan p miljn och samtidigt producera el och vrme till en s lg kostnad som mjligt.
Generella frlopp

Oberoende om fasta brnslen frbrnns i en liten eldstad eller i en stor panna fr vrmeoch kraftproduktion r frbrnningsfrloppet detsamma. Brnslet torkar, avgasar och slutligen frbrnns, eller frgasas, den terstende koksen. Under torkningen avgr fukten frn brnslet vid en lokal temperatur omkring 100 C. Den efterfljande avgasningen r nr brnslet bryts ner pga. uppvrmning och avger sina flyktiga bestndsdelarna, vilket sker i temperaturintervallet 300 till 600 C. Avgasningen kan illustreras av de gaser som ger flammorna runt brinnande grillkol. Den slutliga koksfrbrnningen eller frgasningen r nr koksen reagerar med luftens syre eller med t.ex. vattennga och koldioxid. Detta kan skdliggras med gldande grillkol. Dessa frlopp sker antingen var fr sig eller samtidigt inne i brnslet beroende p storlek och med vilken hastighet brnslet vrms upp.
Avhandlingens innehll

Denna doktorsavhandling bestr av tta arbeten dr olika aspekter av omvandlingen eller fysikaliska storheter fr att beskriva omvandlingen av fasta brnslen i en rost eller en fluidiserad bddpanna undersks. Det som utreds i avhandlingen r fljande: Brnslepartiklar med en mycket hgre temperatur n omgivningen i en fluidiserad bdd Brnslemngden i en fluidiserad bdd. Transient beskrivning av omvandlingen av icke-sfriska brnslebitar med hg fuktoch flykthalt. Termokemiska data och sammansttningen av gaser som lmnar ett trbrnsle under avgasning. Vrmeledning i tr under olika stadier av frbrnningen. Modellering av en hel rostpanna. Faststllande av det generella frbrnningsfrloppet i en brnsle bdd p en rrlig rost. Konstruktion och byggande av experimentanlggningar fr underskning av frbrnningsfrloppen i en fast bdd.

25

Resultat

De huvudsakliga resultaten av arbetet r olika modeller, vilka r validerade mot genomfrda mtningar. Hrledningen av modellerna baseras p olika fysikaliska modeller och kan inkluderas i vergripande modeller av frbrnningen av fasta brnslen i fasta eller fluidiserade bddar, samt fr att ka frstelsen av frbrnning av fasta brnslen. Utver de framtagna modellerna r det fastlagt att den allmnna uppfattningen om frbrnningsfrloppet i en rrlig rostpanna eldad med vta biobrnslen inte stmmer. Istllet fr att frbrnningen av bdden tnds med hjlp av strlning frn flammor ovanfr bdden, visar mtningar och en matematisk analys, att bdden istllet tnds underifrn, nra den rrliga rosten. Detta kan f en stor betydelse fr konstruktionen av nya rostpannor, d detta arbete visar att de reflekterande ytorna ovanfr bdden som r avsedda fr att underltta tndningen inte har denna funktion.

26

Paper I

14th International Conference on Fluidized Bed Combustion, Vancouver, May 11-14, 1997

,03$&7 2) +($7 $1' 0$66 75$16)(5 21 &20%867,21 2) $ )8(/ 3$57,&/( ,1 &)% %2,/(56
* , 3DOFKRQRN & %UHLWKROW] + 7KXQPDQDQG % /HFNQHU
1

A.V.Luikov Heat and Mass Transfer Institute, Academy of Sciences of Belarus P.Brovka 15, 220072 Minsk, Belarus Phone:+375 17 268 4059, Fax:+375 17 232 1325
2

Chalmers University of Technology, Department of Energy Conversion S-412 96 Gteborg, Sweden Phone:+46 31 772 1431, Fax:+46 31 772 3592

$%675$&7 High excess temperatures of burning coal particles, up to 600 K, have been measured with a two- color pyrometer in the transport zone of a CFB boiler at a rather low average oxygen concentration of about 6 %. To understand this phenomenon, a model of heat and mass transfer between a burning char particle and its surrounding has been developed, based on measured heat transfer coefficients and the estimated slip velocity of a char particle. The gas-convective and radiative mechanisms of heat transfer were found to dominate in the core of the transport zone of a CFB furnace. The gasconvective transfer rate was 1.5 times as high as in a singlephase flow. Model calculations show that particles between 0.3 and 3mm could have as high a temperature as the measured ones, provided that there is a highly non-uniform oxygen distribution over the furnace cross-section. 120(1&/$785( a,b,c,d stoichiometric coefficients in reaction (12), cp specific heat, J/(kg*K) D molecular gas diffusivity, m2/s Da Damkhler number, kc/, d particle diameter, m E activation energy, J/mol F function of particle size ratio in Eq. (7) g gravity acceleration, m/s2 kapp apparent combustion rate, m/s kc chemical reaction rate, m/s ko frequency factor, m/s m particle mass, kg, and factor in Eqs (3), (4) N number of collisions, 1/(m2 s) Nu active particle Nusselt number, da/, p split factor in Eqs (13), Q heating value, J/K q heat transferred during one particle collision, J R universal gas constant, 8.314 J/mole K Ret particle Reynolds number, Ut,a d/, r reaction rate kg O2/m2 s S larger-to-finer particle size ratio, Sc Schmidt number, /D, -

Sh T t U Ut V x YO

Sherwood number, da/D, temperature, K time, s superficial gas velocity, m/s particle terminal or slip velocity, U-V/, m/s particle velocity, m/s distance between particle and probe tip, m oxygen mass fraction, -

*UHHN 6\PEROV heat transfer coefficient, W/(m2 K) mass transfer coefficient, m/s eff effective thickness of a gas lens, m voidage, r emissivity, split factor in Eqs (13), projected area of inert particles in a unit volume, 1/m gas thermal conductivity, W/(mK) gas kinematic viscosity, m2/s density, kg/m3 Stefan-Boltzmann constant, 5.6710-8 W/(m2 K4) angle of impact, rad 6XEVFULSWV a active particle b bed c core region e end g gas gc gas-convective i inert particle minsmallest diameter of inert or active particle o initial pc particle-convective r radiative s solid t terminal or slip velocity

871

,1752'8&7,21 Fluidized bed combustors of stationary (SFB) or circulating (CFB) type are usually regarded as isothermal systems. The bed temperature is kept at about 1123 K to optimize sulfur removal, to prevent oxidation of air nitrogen and ash melting. However, the particle phase is a mixture of a few active fuel particles, around 1 % by mass, and numerous inert ones: sand, ash and limestone. This makes the fluidized bed locally non-isothermal. During the drying and, to some extent, during devolatilization, the fuel is colder than the inert matrix of particles, but during char combustion the fuel is warmer than the surrounding bed. Numerous measurements, e.g. Borodulya et al. (1988), have shown that the excess temperature of large char particles burning in an SFB at an oxygen concentration of 21%, reaches up to 500 K and decreases nearly linearly with oxygen concentration. High local temperatures have to be considered when analyzing formation of pollutants, combustion efficiency and agglomeration in a fluidized bed combustor. The temperature of a burning particle is the result of the relationship between the chemical reaction rate and the heat and mass transfer rates to or from the particle. The heat and mass transfer coefficients of an active particle in a SFB combustor can be reasonably accurately calculated with models available in the literature, e.g. Leckner (1992). The same models can also be applied to the dense bottom bed of a CFB combustor, but there are no reliable data on the heat and mass transfer to an active particle in the upper zone of a CFB combustor. A Froessling type of equation, e.g. used by Basu and Nag (1993), has not been experimentally proven under CFB conditions. Such a calculation is based on the slip velocity of an active particle which can be affected by the motion of the inert particles. Besides, the turbulence of the external gas flow can significantly enhance the heat and mass transfer rate. Estimates of turbulence in the transport zone of a CFB combustor, caused by bubble eruption in the bottom bed (Palchonok et al., 1996), have shown that the turbulence intensity in the gas phase is around 20 %. According to Kutateladze (1990), this leads to an increase in the gasconvective heat and mass transfer rate with approximately 50 %. An analogy can be made with the gas-convective heat transfer in the freeboard of a bubbling bed, which has been found to be 50 to 70 % as high as in a non-disturbed single phase flow (Hassan and Palchonok, 1992). There is also a lack of experimental data on the temperature of the burning particles in a CFB combustor. The data available are limited to a few measurements with large fixed fuel particles in small-scale reactors. The excess temperatures of the fuel particles have been found to vary from 150 to 200 K (Bis et al., 1994), to 300 K (Krobarth et al., 1993). The transport zone of CFB combustors is characterized by a non-uniform lateral distribution of solids concentration and bed temperature (Leckner and Andersson, 1992). Close to the walls, a layer of descending particles is formed, in which strong gradients of the above parameters are observed, whereas the distribution is rather uniform in the core of the bed. The falling velocity of 0.17 and 0.26 mm bed particles (silica sand) was found to be, respectively, 1.3 m/s (Wu et al., 1991) and

2.2 m/s (Golriz and Leckner, 1992). These values are close to the terminal velocity of a bed particle, which implies that the gas velocity is small within the wall layer and that the combustion occurs mostly in the core of the bed. The gas velocity in the core is therefore higher than the superficial velocity and can be assumed to be uniform there. The present paper aims at improving the knowledge of the heat and mass transfer and temperature of burning char particles in the core of a CFB, based on measurements in quite a large scale CFB combustor. (;3(5,0(17$/ Measurements were carried out in the 12 MWth CFB boiler at Chalmers University of Technology described, e.g. by Palchonok et al. (1995). The combustion chamber has a 1.72 m by 1.44 m cross-section and is 13.5 m tall. Two of the four water-cooled membrane walls are refractory-lined. The test holes used in the measurements are situated in the center of the cool walls at heights 3.8, 8 and 11 m above the air distributor. The bottom bed temperature, 1123 K, was kept constant. Heat transfer coefficients to fixed metal spheres of 5, 10 and 15 mm in diameter were measured with the transient technique described by Palchonok et al. (1995). The instantaneous overall heat transfer coefficient was evaluated from the natural heating rate of the sphere measured by an embedded thermocouple. A method of dark (oxidized steel sphere, r=0.8) and light (ground silver sphere, r =0.35) calorimeters was used to evaluate the radiative component of the heat transfer coefficient. The time-average local bed temperature was measured with a K-type thermocouple protected with a double radiation shield. The probes were positioned outside the boundary layer whose thickness was around 0.1 m on all the three measurement levels. Two silica sand fractions with Sauter mean diameters of 0.35 and 0.22 mm, and density 2600 kg/m3, were used as bottom bed materials. The superficial gas velocity varied from 1.8 to 4.6 m/s during the measurements with the larger inert particles and from 1.5 to 2.7 m/s with the smaller ones. The cross-sectional average bed density at the measurement level varied from 1 to 35 kg/m3. The temperature of burning particles of bituminous coal was measured with a two-color optical pyrometer developed by Hernberg et al. (1993). The temperatures of the bed and of the active particle were calculated from the measured ratio of the incident radiation intensities at two wavelengths. The measurement volume was a diverging cone with an aperture of around 12 degrees. The model assumptions used in the pyrometry were: (i) the inert solids are gray bodies of the same emissivity; (ii) only one active particle is contained at the moment in the measurement volume; (iii) the wavelengths used are selected to avoid an influence of the flue gas. The optical pyrometer was contained in a water-cooled probe of 43 mm diameter positioned at a height of 3.8 m with the tip placed close to the axis of the furnace. The superficial gas velocity was 4 m/s in the bottom bed and 6 m/s at the measurement level, because of the added secondary air. The bed material was silica sand of 2600 kg/m3 density and with a Sauter mean particle diameter of 0.27 mm in the bottom bed and 0.25 mm at the measurement level. The cross-sectional average bed density

872

was 25 kg/m3 and the oxygen concentration was 5.7 % at the measurement level. The fuel was a bituminous coal. Simultaneous measurements of the instantaneous oxygen concentration with a zirconia cell showed fluctuations between oxidizing and reducing conditions. 02'(/ +HDW DQG PDVV WUDQVIHU The transfer rate is closely related to the flow pattern around the burning particle. An active char particle generally has different size and density from that of inert bed material. This results in a velocity difference between the active and the inert particles. A simplified situation, where spherical active particles with the parameters da, a and Va are surrounded by inert particles with the uniform parameters di, i and Vi , is considered for application in the model of Nowak et al. (1996). This model describes a quasi-steady state motion of a coarse active particle (da>>di, Va<Vi) during the upward flow of a dilute, uniform suspension of gas and inert particles. Because of the higher velocity, the inert particles collide with the stern of the active one. The collisions are assumed to be elastic, i.e. the restitution coefficient is equal to 1. As a result of the additional momentum transfer, the velocity of the active particle increases compared to the case of a single-size particle flow. In the present analysis, analogous to that of Nowak et al. (1996), an arbitrary active-to-inert particle size ratio is assumed, which results in a modified equation for the slip velocity, 8 W , D 8 W ,L 8 W , D 8 W ,L (G D + G L )2 2 E ,F + 1 + P D PL 4 2 + & ', J 8 W ,D 8 W , D G D J JPD = 0 4 2
2

component is calculated with a Froessling-type equation based on the slip velocity, 1X JF = 2 + P Re W , D 0.5 Pr 0.33 (3)

The parameter m was evaluated from the measurements with the fixed metal spheres. Equation (3) can be used for calculation of the mass transfer coefficient, provided that the analogy holds between the gas-convective heat transfer and mass transfer, 6KJF = 2 + P Re W , D 0.5 6F 0.33 (4)

The Sherwood and Schmidt numbers in Eq. (4) are based on the molecular gas diffusivity of oxygen in the flue-gas. This diffusivity was approximated with the one for a binary mixture of oxygen and nitrogen, and calculated according to Pomerantsev et al. (1986), & 2.5 (& + 273)( 2 + 273) ' = 18.4 10 6 (7 273) 1 (& + 7 )(& + 7 ) 1 2
0.5

(5)

where C1=138 and C2=107. This correlation is in agreement with the Chapman-Enskog theory (Bird et al., 1960). The particle-convective component is estimated, consistent with the flow pattern described above, as a number of collisions per unit time per unit surface of the active particle, N, times an amount of heat absorbed by a single inert particle during a short impact, q. According to Nowak et al. (1996)

1 = (3 2)8 W ,D 8 W ,L (1 )(G D + G L )2
(1)

(G

GL3 )

(6)

The first - collision - term of Eq.(1) can be interpreted as a drag force, acting on the active particle having an effective collision diameter (da+di), caused by the difference in velocity between the active particle (fuel) and the inert (bed) particle of density b,c. The expression 2/(1+ma/mi) is an effective drag coefficient of the solid phase. The second term describes the drag of the gas, where the drag coefficient is calculated as (Kovensky, 1996) & ', J = 24 Re W ,D + 0.44 F 4.75

The model of Zabrodsky (1966) for an elementary act of heat transfer between a surface and a single particle is used to calculate q; a derivation is given in Appendix A. The model assumes that the heat is transiently transferred through the gas lens between the relatively flat surface of the larger particle and the smaller spherical particle. A quasi-steady temperature profile in the film is assumed. The minimum effective thickness of the lens was estimated as d/6. An analogous estimate for two contacting spheres of an arbitrary size ratio, S = dmax/dmin, reads

HII G min = 1 6 + 1 2 6) 1 6 2 ) ( ) 3) +1 / 2 1

(7)

(2)

) = 1 cos(arcsin (1 6 ))
Provided a short duration of impact, which can be estimated as t=di/(Ut,a-Ut,i), the final expression for the particle-convective heat transfer coefficient reads

The local bed density in the core, b,c, is approximately two times as low as the cross-sectional average bed density, b (Zhang et al., 1994). The latter was evaluated from pressure drop measurements. It follows from Eq.(1) that the inert particles can either promote the movement of a slow (coarse) char particle or hinder a fast (fine, low-density) one. Equation (1) was obtained under a number of simplifying assumptions, and the prediction from this equation should be considered as a rough estimate. The overall heat transfer coefficient is assumed to be the sum of gas-convective, particle-convective and radiative components, = gc + pc + r. The gas- convective

SF = (3 8)( HII )( F )(G D + G L )2 G D 2 1

(8)

If the temperature of the inert particles does not change significantly during the short impact time, the radiative heat transfer coefficient is calculated as

U = 7D 2 + 7E 2 (7D + 7E ) ( U , D + 1 U ,E 1) 1

(9)

Emissivities r,a=0.85 of char (Ross et al., 1981), and r,i=0.6 of inert (sand) particles (Borodulya et al., 1982) were used in the

873

Gas conv. Nusselt number, Nu gc

calculations. The effective emissivity of the dilute isothermal suspension was calculated as (Borodulya and Kovensky, 1983)

60 40
15 10 5

Fe Ag

U ,E = U ,L 0.31
7HPSHUDWXUH RI WKH EXUQLQJ SDUWLFOHV

(10)

A simplified heat balance of a burning spherical char particle yields

(F S D D G D 6)G7D
,

GW = 4 . DSS <22 J (7D 7E )

(11)

20

The particle Biot number is assumed to be low enough to neglect the internal temperature gradient. A first-order chemical reaction is assumed to occur on the external surface of the particle,

10

D& + E22 = F&2 + G &22

(12)

The molar ratios of the reactants and the primary reaction products are assumed to depend on the particle size and temperature according to Arthur (1951) and Field et al. (1967)

200

400

600

1000

2000

Reynolds number, Re t,a


Figure 1: Nusselt number for gas convection between a sphere and the bulk bed in the transport-zone. The Reynolds number of the particle is based on the local gas velocity. The symbols are for measured values, with 5, 10 and 15 mm spheres of steel, (filled), and silver (open) (Palchonok et al., 1995). The large symbols are measured with di=0.35 mm and the small ones with di=0.20 mm. The line is calculated from Eq. (18). log[N0 ] = 2 + 0.2 10 4 ( the apparent activation energy, E, being 115 to 135 kJ/mole for bituminous coals. 5(68/76 $1' ',6&866,21 The gas-convective heat transfer coefficients, evaluated from the measurements as gc = - (pc + r), are presented in non-dimensional form in Fig. 1. The particle-convective constituent was found to be negligibly small, 0.5 to 1.7 % of the total heat transfer coefficient, at the actual bed densities in the core region, 0.5 to 17.5 kg/m3. The radiative constituent was 20 to 30 % for the silver probes and 40 to 60 % for the steel ones. A least square fit yields a value of the factor m in Eqs. (3) and (4) of 1.02, with Ret based on the superficial velocity. Provided that no gas flows through the 0.1 m thick wall layer, (Zhang et al., 1995), the gas velocity is around 1.3*U/, which leads to the following expressions 1X JF = 2 + 0.89 Re W , D 0.5 Pr 0.33 6KJF = 2 + 0.89 Re W , D 0.5 6F 0.33 (18) (19)

S F G = 2500 exp[ 6240 / 7D ]


D E = (2 S + 2 ) (S + 2 );G S 0.05 10 3 P

(13)

(2 S + 2)

S G S 0.05 10 3

)
;0.05 10 3 < G S < 1 10 3 P
3

(S + 2 )

0.95 10

= 1;G S 1 10 P
The above kinetics are questionable under SFB conditions (Prins, 1987) and (Hayhurst, 1996), but can be accepted for the dilute conditions in the CFB transport zone. The heat release of the chemical reaction per kg O2 is, derivation in Appendix B, 4 = 4&22 (2 ) + 4&2 ( 1) 4&22 = 12.34 0- NJ 22 ; 4&2 = 6.84 0- NJ 22 The apparent combustion rate constant (related to the consumption of oxygen) is N DSS = 1 ( N F + 1 ) 1 (15) (14)

The chemical reaction rate constant (related to oxygen consumption) is assumed to follow the kinetics of Field et al. (1967),

N F = (1 )(32 12)2247D exp[ 149200 57D ]

(16)

Equation (16) implies the same reactivity for different coals, which differs from recommendations of Pomerantsev et al. (1986). An alternative kinetics of Pomerantsev et al. (1986) was also used, and this predicts a somewhat lower reaction rate at high temperatures,

NF = N0 exp[ ( 57D ]

(17)

Equations (18) and (19) predict approximately 1.5 times higher transfer rate than the Froessling equation (m=0.6) for single phase flow. This agrees with the above mentioned recommendation of Kutateladze (1990) and implies that the turbulence intensity is higher than 10 %, in the same order of magnitude as estimates of Palchonok et al. (1996). The

874

800

1500

Excess temperature, K

600

400

Particle temperature, K

1000

500

200

0 -4 10
0

10

-3

10

-2

10

-1

10

100

200

300

400

X-factor, 10
di, mm 5 2

Time, s
Figure 3: Experimentally determined temperature histories of burning char particles from bituminous coal in an oxygen concentration of 21 %. The solid line is for a char particle with an initial diameter of 13 mm in a bed with a temperature of 805 to 825 K consisting of fire-clay particles with a diameter of 2.15 3 mm and a density of 2300 kg/m . The solid line is for a char particle with an initial diameter of 9.4 mm in a bed of fire-clay particles having a diameter of 0.83 mm and a temperature of 990 to 1015 K. radiation from the fuel particle due to inert particles, and the emissivites of the active particle and the bed. Assuming the emissivities to be close to unity, the X-factor can be represented simply as a function of the configuration factor and the extinction. The configuration factor was estimated as the ratio of the projected area of the active particle and the area of the field of view. The extinction depends on the distance between the particle and the probe tip, x, and on the projected area of inert particles in a unit volume, , and so the transmission becomes exp[-x]. The field of view increases proportionally to the distance from the probe tip, whereas the transmission decreases exponentially. Estimates of the X-factor versus the distance form the probe tip for different active particle diameters are plotted in Fig.2b. It is seen from Fig. 2a that excess temperatures up to around 600 K have been observed, the highest values corresponding to low X-factors. A low X-factor corresponds to either a small particle close to the probe tip or a larger particle further away, but when no high temperatures are observed at higher X-values the conclusion is that only small particles have high excess temperatures. The particle temperature is measuerd at a fixed location in the furnace (an Eularian approach), but for the calculation of the particle temperature from Eq(11) the particle is assumed to follow its environment (a Lagrangian approch). This problem can hardly be solved, because of the variety of the motion patterns of the particles and of combustion conditions. For instance, in a situation with strongly non-uniform oxygen distribution, Lyngfelt et al. (1996), a fine particle, easily carried away with the gas, can follow either oxygen-rich or

10

-1

1 0.5

X-factor, -

10

-2

0.2 0.1

10

-3

10

-4

10

-3

10

-2

10

-1

Distance from the probe tip, m


Figure 2: a) Temperature measured with two-color pyrometry as a function of the X-factor. b) The Xfactor as a function of the distance from the probe-tip, for different particle diameters. reference temperature in Eqs. (18) and (19) is defined as T=0.5(Ta+Tb). The standard deviation from Eq. (18) in Fig. 1 is around 30 %. The results of the particle temperature measurements are presented in Fig. 2a. The majority of the measured particle excess temperatures were below 200K, but excess temperatures up to 600K were recorded. The dependence of the measured excess temperature on the active particle diameter can be evaluated with a statistical routine. This statistical routine demands a higher measurement accuracy than that obtained and therefore, the excess temperature versus an X-factor dependence is plotted directly in Fig. 2a. The X-factor is an expression of the configuration factor for radiative exchange between the active particle and the probe, the extiction of the

875

600
1210
5

Da=5

10

Heat release/transport, W/m

1010 5

Excess temperature, K

500
2

810

Y =0.15 O
2

400
1

6105 410 210


5

0.11

300
0.5 16 14

0.07

200 100 0

12

0.2 0.1

Y ,% 10 O2 9 8 7 6 4

0.03

200

400

600

800

10

-4

10

-3

10

-2

Excess temperature, K
Figure 4: Heat supply by combustion (solid lines) for different oxygen concentrationsYO2 and heat removal by heat transfer (dashed line) for a 0.5 mm char particle as a function of the excess temperature. oxygen-poor gas streams, while a coarse particle most likely will face a fluctuating oxygen concentration. Therefore the measurements have been compared with the quasi-stationary form of Eq. (11), when dTa/dt = 0. The complete solution of Eq. (11) should look like Fig. 3, showing some coarse char particle temperature histories measured at 21 % O2 in a laboratory-scale stationary fluidized bed. The experimental curves have two plateaus, the first one corresponding to a quasi- steady state, a kinetically controlled combustion regime. A small permanent increase of the bed temperature, which was observed during the measurements, interrupted the thermal equilibrium and the second plateau was reached, corresponding to a quasi-steady state, a diffusion-controlled regime. Some stationary solutions of Eq. (11) for a 0.5 mm char particle are shown graphically in Fig. 4, where the heat release, QkappYO g, (solid lines)and heat loss from the particle, (Ta2 Tb) (dashed line), are plotted versus the excess temperature. The kinetics of Field et al. (1967) were used. Depending on the oxygen concentration, one or three steady state solutions can be found, the intermediate one in the last case being unstable, as can be shown in terms of the Semenov (1935) ignition theory. No char particle larger than around 4 mm has been found at the measurement level, and this size was chosen as the maximum diameter of the char particles in the calculations. All the solutions in the field of interest are presented in Fig. 5, which shows the excess temperature versus particle diameter for different oxygen concentrations and Damkhler numbers, Da. The Damkhler number is defined as the ratio of the reaction rate coefficient, kc, and the mass transfer coeffcient, . If Da < 1, the combustion is controlled by kinetics, and if Da>1 it is diffusion-controlled. A comparison between Fig. 5 and Fig. 2a shows that the high measured excess-temperatures

Particle diameter, m
Figure 5: Excess temperature versus particle diameter for different oxygen concentrations (solid lines), and Damkhler numbers (dashed lines). could not be reached at the measured average oxygen concentration YO =5.7%. Provided that the measurements and 2 the evaluation are correct, the highest temperatures can be attributed to fine char particles, e.g. products of secondary fragmentation, following oxygen-rich gas streams. Assuming local plug-flow conditions, the maximum excess temperature of 500 to 600 K is reached by particles of 0.3 to 3 mm in diameter. This agrees with the conclusion drawn from the comparison of Figs. 2a and 2b. Somewhat lower excess temperatures, but still consistent with the experiment, were obtained with the kinetics of Pomerantsev et al. (1986). Other reasons for the detected high temperatures, such as homogeneous volatile combustion, can not be excluded. The bulk of the excess temperatures lower than 200 K in Fig. 2a can be attributed to relatively large char particles, which did not follow the gas and which therefore, on the average, meet rather low effective oxygen concentrations. &21&/86,216 A model of heat and mass transfer to and from a burning char particle in a CFB furnace is presented, based on the difference between the velocities of fuel particles, bed particles and gas. The gas-convective and radiative mechanisms of heat transfer are the dominant ones under the conditions prevailing in the core region of the transport zone. The gas-convective transfer rate is 50 % as high as in a single-phase flow, implying a significant turbulence intensity in the gas phase. Extremely high maximum excess temperatures of the burning particles, up to 600 K, were measured at a rather low average oxygen concentration. This agrees with theoretical predictions, provided that oxygen is non-uniformly distributed over the furnace cross-section, and that fine char particles follow the gas flow and may burn in regions of high oxygen concentrations.

876

$&.12:/('*(0(17 The work was financed by the Swedish National Board for Industrial and Technical Development (NUTEK) and by the INTAS grant No. 94-4313. 5()(5(1&(6 Arthur, J.R., Reactions between Carbon and Oxygen, Trans.Faraday Soc., 47, 164, (1951). Basu, P., Yan, J., Characterization of the Fine Char Particle Combustion in Circulating Fluidized Beds, 12th Int. Conf. on FBC, Eds Rubow L. and Commonwealth G., ASME, (1993), 283. Bird, R.B., Steward, W.E., Lightfoot, E.N., Transport Phenomena, John Wiley and Sons, New York, (1960). Bis, Z., Gajevski, W., Nowak, W., CFB Combustion and Hydrodynamic Modelling, Proc. 2nd Int. Interfluid Symp. on Fluidized Bed Combustion, Nagoya, (1994), 141. Borodulya, V.A., Ganzha, V.L., Kovensky, V.I., Hydrodynamics and Heat Transfer in a Pressurized Fluidized Bed, Minsk, Nauka I Tehnica, (1982) (in Russian). Borodulya, V.A., Kovensky, V.I., Radiative Heat Transfer between a Fluidized Bed and a Surface, Int.J.Heat Mass Transfer, 26, 277, (1983). Borodulya, V.A., Palchonok, G.I., Vasiljev G.G., Dryabin, V.A. Galerstein, D.M., Heat and Mass Transfer and Combustion Kinetics of Solid Fuel in Fluidized Bed, Int.School-Seminar Heat and Mass Transfer Problems in Advanced Solid Fuel Combustion and Gasification Technologies, Minsk, (1988), 2, 3 (in Russian). Field, A.M., Gill, D.V., Morgan,B.B., Hawksley, P.G.W., Combustion of Pulverized Coal, BCURA, Leatherhead, (1967). Golriz, M.R., Leckner, B., Experimental Studies of Heat Transfer in a Circulatig Fluidized Bed Boiler, 1st Int. Conf. Engineering Application of Mechanics, Tehran, (1992). Hassan, A.F., Palchonok, G.I., Heat Transfer in the Freeboard of Fluidized Bed, Izvestija Vuzov: Energetika, No.4, 73-81, (1991) (in Russian). Hayhurst, A.N., Advances in Coal Combustion, Joint Meeting of the Portuguese, British, Spanish and Swedish Sections of the Combustion Institute, The Combustion Institute Madeira, (1996), P. 13.1. Hernberg, R., Stenberg, J. and Zethrus, B, Simultaneous In Situ Measurements of Temperature and Particle Size of Burning Char Particles in a Fluidized Bed Furnace by Means of Fibreoptic Pyrometry, Combustion and Flame, 95, 191-205 (1993) Kovensky, V.I., Account for the Constraining of Particle Motion in the Freeboard, Presented at the Int. Workshop on the INTAS project No.94-4313, Chalmers University of Technology Gothenburg, (1996). Krobarth, P., Winter, F., Hofbauer, H., Reactivity of Large Coal Particles under Fast Fluidized Bed Conditions, Nordic Seminar on Solid Fuel Reactivity, Gothenburg, (1993), P. 6. Kutateladze, S.S., Heat Transfer and Hydrodynamic Resistance: a Reference Book, Energoatomizdat, Moscow, (1990).

Leckner, B., Andersson, B.-., Characteristic Features of Heat Transfer in Circulating Fluidized Bed Boilers, Powder Technology, 70, 303, (1992). Leckner, B., Palchonok, G.I., Andersson, B.-., Representation of Heat and Mass Transfer of Active Particles, Presented at IEA Mathematical Modelling Meeting, International Energy Agency, Turku, (1992). Lyngfelt, A., mand, L.-E., Leckner, B., Progress of Combustion in the Furnace of a Circulating Fluidized Bed Boiler, 26th Symp.(Int.) on Combustion, Combustion Institute, Naples, (1996). Nowak, W., Bis, Z., Gajewski, W., Matsuda, H., Hasatani, M., Carryover of Coarse Particles from a Dense Bed in a Multy-Solid Fluidized Bed, Preprints of the 5th Int.Conf. on CFB, Beijing, (1996), P. DGS11. Palchonok, G.I., Breitholtz, C., Andersson, B.-A., Leckner, B., Heat Transfer in the Boundary Layer of a Circulating Fluidized Bed Boiler, Preprints of the 7th Int.Conf. on Fluidization, Engineering Foundation, Tours, (1995), 177. Palchonok, G.I., Johnsson, F., Leckner, B.,Estimates of Turbulence Effects in CFB Boilers, Preprints of the 5th Int.Conf. on CFB, Beijing, (1996), P. MSD7. Pomerantsev, V.V., Arefjev, K.M., Ahmedov, D.B., Fundamentals of Practical Combustion Theory, ed V.V.Pomerantsev, Energoatomizdat, Leningrad, (1986). Prins, W., Fluidized Bed Combustion of a Single Carbon Particle, Ph.D.Thesis, Twente University, Netherlands, (1987). Ross, I.B., Patel, M.S., Davidson, J.F., The Temperature of Burning Carbon Particle in Fluidized Beds, Trans.IChemE, 59, 83, (1981). Semenov, N.N., Chemical Kinetics and Chain Reactions, Oxford University Press, London, (1935). Wu, R.L., Lim, C.J., Grace, J.R., Brereton, C.M.H., Instantaneous Local Heat Transfer and Hydrodynamics in a Circulating Fluidized Bed, Int. J. Heat Mass Transfer, 34, 2119, (1991). Zabrodsky, S.S., Hydrodynamics and Heat Transfer in Fluidized Beds, M.I.T. Press, Cambridge, Mass., (1966). Zhang, W., Johnsson, F., Leckner, B., Characteristics of the Lateral Particle Distribution in Circulating Fluidized Bed Boilers, 4th Int. Conf. on CFB Technology, Ed A.A.Avidan, AIChE, Hidden Valley, (1994), 266. $33(1',; $ 3$57,&/( &219(&7,9( &2()),&,(17

+($7

75$16)(5

The particle convective heat transfer coefficient is calculated as the product of the heat transfered during one collision, q, and the number of collisions, N, divided by the temperature difference between the active particle and the bulk.

SF = T 1 (7D 7E )
T = & S ,L L G L 3 6 7L,H 7L , 0

(20)

)(

(21)

Ti,0 = Tb is the temperature of the inert particles before impact and Ti,e is the end temperature after the impact and Ti,e-Ti,0=T. Zabrodsky assumed a quasi-steady temperature profile in the

877

effective gas lens, which is considered as a flat disc of eff thickness, eff; this leads to: & S ,L L G L 3 6 G7L = HII G L 2 4 (7D 7L )GW Integrating from t=0 to the end of impact gives 7L ,0 7L ,H = 7D 7L , 0 [ exp( N W )] 1 where N = 3 2 HII L & S,L G L .

= (U1 + 2U2 ) (U1 + U2 )


which leads to the following set of equations.
U1

(32)

)(

(22)

( 1) + U2 ( 2) = 0
U1 + U2 = N F & V

(33)

(23)

(24)

where Cs is the concentration at the surface and the reaction rate coefficient, kc, is related to the consumption of oxygen. The solution for the reaction rates is:

For a short impact time (kt<0.03) the exponential expression, [1-exp(-kt)]kt, which leads to 4 = G L 2 4 HII (7D 7E )W

U1 = (2 )N F & V U2 = ( 1)N F & V

(34)

)(

(25)

The impact duration can be estimated as t=dmin/(Vi-Va). dmin is the smallest of the diameters of the inert and active particle. The number of impacts

In the last equation, dimensions are optional: either (mole O2) or (kgO2/m2s). Finally, the heat production becomes

4 = 4max (2 ) + 4min ( 1) .

(35)

1=
Finally,

L (1 F )(9L 9 D ) (G D + G L )2 4 L G L 3 6 G D 2

(26)

SF = (3 8)( HII )(1 F )(G min G L )(G D + G L )2 G D 2

(27)

The minimum effective thickness of the lens formed during a perfect contact between an inert particle and an infinitely large particle (a flat surface), related to the projected area of the inert particle, is eff,min =(dmin/6). The real curvature of the larger particle adds something to the minimum thickness dmin/6. Moreover, the particles are in relative motion during the impact, which should further increase the time average thickness of the lens. Actually, the present estimate of the collision time t is a twice the time for the smaller particle to pass a distance dmin/2 to and from the surface. For a direct ( = 0) elastic collision, such an estimate implies the time average lens thickness to be

HII = SHUIHFW + G min 4

(28)

For an arbitrary angle, eff=perfect+dmincos()/4. Averaging over = 0 to /2 leads to

HII = SHUIHFW + G min 2


$33(1',; % 5($&7,21 5$7(6

(29)

When carbon is oxidized, carbon-dioxide is formed with the reaction rate r1,

& + 22 &22 ,
but also carbon-monoxide with the reaction rate r2,
2& + 22 2&2 .

(30)

(31)

The split factor, , between coal and oxygen consumption is then,

878

Paper II

FUEL LOADING OF A FLUIDIZED BED COMBUSTOR


Henrik Thunman and Bo Leckner Department of Energy Conversion Chalmers University of Technology

Abstract
This is a study of the inuence of operating conditions on the loading and size distribution of fuel in a uidized bed combustor and on the vertical fuel concentration in a circulating uidized bed combustor (CFBC). For this purpose a model has been developed including the most important parameters and having a short calculation time, 30 sec on a Sun Ultra. The fuel loading depends on type of fuel, fragmentation and pressure; all these parameters have a great inuence on the combustion. A rise of pressure and/or supercial velocity increases the power output from the combustor, and also the fuel loading. The size distribution of fuel in the bed is mostly dependent on the fragmentation behavior of the fuel. For CFBC, the variation of fuel concentration along the riser is mainly affected by the supercial velocity.

an understanding of formation and destruction of emissions, especially NO and N2 O. The purpose of this work is to investigate how different operation conditions inuence fuel loading and size distribution of fuel. The model developed for this purpose shall have a short calculation time and shall include the most important inuencing parameters.

2 Theory
Most models for fuel loading in FBCs are developed for coal combustion, where the time of devolatilization is short compared with the time of char combustion, and the devolatilization can be assumed to be immediate e.g. Arena et al. [1995]. This assumption is not suitable for a general model, because the devolatilization time for many fuels, especially biofuels, is a large part of the total combustion time. A special case of the FBC is the circulating uidized bed combustor (CFBC), where material is transported up through the furnace and a certain amount of the fuel is above the inlet of the secondary air. This has not a great effect on the fuel loading, since most fuel is still in the bottom part of the furnace, but it can be of interest for formation and destruction of pollutants.

Keywords
Fluidized bed, fuel loading, fragmentation, fuel size distribution, parametric study

1 Introduction
The heat release of a ame combustor responds instantaneously upon a change in the fuel feed rate. A bed combustor red with solid fuels, and especially a uidized bed combustor (FBC), shows a certain time-delay between a change in the fuel feed rate and the corresponding change in the heat release rate. Similar time-delays occur for other changes in conditions, such as in air supply or bed temperature. The reason is, of course, that the fuel loading of the bed reacts with a certain delay depending on its size and time of combustion. The time-delay may cause additional difculties for a control system, since the fuel loading of a bed is not necessarily constant, it depends on a number of factors in a certain boiler: type of fuel, bed temperature, air supply (secondary/primary) etc., and it depends on type of boiler: uidization velocity and pressure. The fuel load has additional effects: the fuel distribution in the furnace and the surface temperatures of the fuel particles are important parameters needed for

2.1 General models for the FBC


The assumptions made in the present model are: The bed is fed with a fuel having a known Rosin-Ramler size distribution. Devolatilization and fragmentation start immediately when the particles enter the combustor. The devolatilization follows a shrinking core model and the fragmentation is proportional to the mass of the fuel particles. The char combustion follows a shrinking sphere model. During char combustion, the mass loss due to fragmentation and attrition are treated like a surface reaction rate constant. The loss of efciency from the combustor is due to small char particles carried away by the ue gases. The particles have a known

Feed Fuel

Fragmentation Volatiles Char

Devilatilization

Char FragmentaFlyCombustion ash tion

Feed Char

Particle size

Fuel transformation
Figure 1. Principle scheme of the transformation of fuel in a uidized bed combustor. The bars represent the mass of char or volatiles entering a particle size fraction, the boxes represent models for devolatilization and char combustion and the arrows represent mass ow from feed fuel entering the system to char leaving the system with the ue gases, Fash size distribution taken from Chirone et al. [1991] The fuel load of the bed is calculated from a mass balance over the combustion of the fuel. A principle calculation scheme for the mass balance is shown in Fig. 1. (3) F is the mass ow of fuel at a time and Fchar,0 is the initial mass ow of the char in the fuel having the particle size R. Vdev is the fraction of volatiles in the fuel. Assuming fragmentation of fuel particles to be proportional to the mass, which, under steady state condition, is proportional to the mass ow, the mass ow of the fuel particles at time becomes for a certain size element around R,
F (R;  ) = Fchar;0 (R)(1 Vdev (1 x(; R))) exp( kf;dev  (x0 ))(1 Vdev )exp(kf;dev  )

2.1.1 Devolatilization
The time of devolatilization, td , is represented by the empirical relationship, e.g. Zhang [1987], Stubington and Moss [1995], Winter [1995], (1) where R is the particles radius and a and b are constants related to the fuel. By introducing the assumption of a shrinking core, the part of the volatiles remaining in the fuel particle becomes,

(4)

kf,dev is the fragmentation rate constant and x0 is initial relative volatile content. The probability, Pf,dev , that a particle of certain size R fragments during devolatilization is, (5) This expression for the probability has the same form and gives nearly the same result as more complicated models, e.g. Chirone et al. [1991], where the fragmentation is related to models of the internal stress. During devolatilization the particles fragment into smaller particle sizes with a certain size distribution,

(2) and x is the ratio of mass of volatiles at time initial mass of volatiles. is the time passed from the beginning of devolatilization. Under steady state conditions if there were no fragmentation the mass ow of the devolatilizing fuel particles at time is,

and volatiles and char are considered separately (Fig. 1). The contribution of fragments to a certain particle size element centred at R of the initial mass ow, F0 (R), makes the average volatile content of the total mass ow of fuel, x0 , particle size element less than the volatile content of the initial fuel,

fragmentation and attrition are not treated separately. Instead, here the different size distributions of the fragments from attrition and secondary fragmentation are represented by a single size distribution. The rate of fragmentation is treated as a surface reaction rate constant, similar to the attrition models in the literature, e.g. Arena [1991], Chirone [1991]. The mass balance during char combustion is,

(6) Ff1,dev is the mass ow of the volatiles in the fragments and Ff1,char is the mass ow of the char in the fragments. The initial mass ow of char with a certain size is,

(7) By integrating Eq. 4 over all radii and over devolatilization time, the total mass of solid fuel during devolatilization, Mdev is obtained, (12) with the boundary condition, (13) (8) Fash is char leaving with the ue gases, Mchar total mass of char during char combustion, hchar size distribution of the char, Ff2,char mass ow of the fragments from larger particle sizes, and kf,char is fragmentation rate constant. The terms in Eq. 12 correspond to 1. Feed char from devolatilization 2. Char leaving with the ue gases 3. Char entering or exiting the interval due to char combustion 4. Mass reduction in the size interval due to combustion 5. Mass entering the interval from previous fragmentation 6. Mass leaving the interval due to fragmentation The char combustion is assumed to follow a shrinking sphere model. The shrinking rate is, (14) Mc is molecular mass of carbon, CO2 average oxygen concentration in the bulk, c density of the carbon in the char, kc surface reaction rate constant, split factor for CO and CO2 and is mass transfer coefcient. The split factor is calculated according to Arthur [1951] and the mass transfer coefcient from Sherwood number,

with the boundary conditions, (9) Rmax is the largest particle size in the size distribution of the particles fed. The size distribution, hdev , related to the total mass of solid fuel accumulated during devolatilization is

(10) The fuel leaving the devolatilization stage for char combustion is, (11)

2.1.2 Char Combustion


The model expressing the total mass during char combustion and the size distribution related to the mass of char particles during char combustion is in principle the same as the one presented by Salatino et al. [1989], which is an extension of the mass balance presented by Kunii and Levenspiel [1969]. The difference between the model presented here and the one of Salatino et al. is that secondary

(15) DO2 is molecular diffusivity of O2 . Sherwoods number is calculated according Halder et al. [1985] (16)
b is bed voidage, Re Reynolds number based on slip velocity and Sc is Schmidt number.

The particle temperature is obtained from the heat balance, (17)

During devolatilization the particles fall apart into a few large particles, Chirone et al. [1991] and Chern and Hayhurst [1996]. The relative radius is therefore be large during devolatilization, and the distribution follows Eq. (20). During char combustion, the fragments come from the nes produced by attrition and secondary fragmentation, Chirone et al. [1991]. If attrition is dominant, the relative radius is small ( <0.05), and if secondary fragmentation is dominant, the relative radius is large ( >0.5). The error caused by the size distribution chosen, i.e. Eq(18), related to the true size distribution, depends on the size distribution of the fuel fed; a narrow size distribution gives a larger error than a wide size distribution, because of the overlapping of the size distributions of fragments.

Hc is heating value for carbon oxidation, tot total heat transfer coefcient, T particle temperature, Tb bulk bed temperature, r radiative, and c convective heat transfer coefcient.

2.1.4 Fragmentation rate constant


The fragmentation behavior during devolatilization and char combustion differs, Chirone et al. [1991]. During devolatilization, fragmentation is caused by thermal shock and gas expansion inside the particle. This relates the fragmentation rate constant to the volume of the particle, (21) Af,dev is the fragmentation constant during devolatilization. Values of the fragmentation constant can be estimated from data on the probability to fragment, the number of fragments produced from a mother particle, and the mean diameter of fragments. Such data are found in the literature, e.g. Chirone et al. [1991]. The timing of fragmentation is not well known, but some work has been carried out, e.g. Stubington [1996], which shows a great variety of timing depending on type of coal. During char combustion, fragmentation is caused by irregular effects of combustion on the surface of the particle or by collision with other solid particles, which relates the fragmentation rate constant during char combustion to the particle surface. (22)

2.1.3 Distribution of fragments


The fragments are assumed to be spherical, their number is distributed according to a rst order gamma distribution and their masses according to a fourth order gamma distribution. The fragments must be smaller than the mother particle, a requirement that makes it necessary to normalize the distribution. The size distribution on a number basis, hf,n and on a mass basis, hf , is

(18)

y is the radius of the fragments and is a scale factor controlling the shape of the distribution. The mean size of the fragments is assumed to have the same relative radius, , for all initial particle sizes, (19) y mean is the mean size of the fourth order gamma distribution (not normalized). If the relative radius is large (>40) the normalized distributions can be simplied, (20)
*

The fragmentation constant during char combustion, Af,char is assumed to be much larger than the secondary fragmentation and analogous to the attrition rate constant found in the literature., which means that the attrition rate constants found in the literature, e.g. Arena et al. [1990] Chirone et al. [1991], can be used for the fragmentation constant during char combustion.

2.2 Special models for CFB


In a stationary FBC all the fuel is in the bed in the bottom part of the combustion chamber, but in a CFBC fuel and bed material are distributed in the entire combustion chamber. The size distribution and the concentration of fuel along the riser in CFBC is modeled here. These quatities have only a minor effect on the fuel loading, but they affect formation and destruction of emissions.

The gas velocity in the core is calculated from the supercial gas velocity, U and the bed voidage, b , (25) K is the ratio of the cross-sectional area of the riser and the core. The gas velocity in the boundary layer is assumed to be zero. The slip-velocity is calculated from a model of Nowak et al. [1996], which is modied to be valid for all particle sizes and not just for large particles,

2.2.1 Distribution of fuel along the riser


The ratio of the mass of fuel and inert material, the fuel concentration, decreases with height in the CFBC for large particles and increases for small particles. The velocity ratio of fuel and inert particles behaves like the mass ratio. Large fuel particles are carried upwards by collision with small inert particles. When the bed voidage becomes larger, the force acting on the large particles decreases, and so does the velocity of the large particles. The small fuel particles, having a lower slip velocity than the inert particles, experience the opposite situation; they are held down by collisions with the inert particles. In this case the collision force decreases with larger bed voidage, and the velocity of the small particles increases. Assuming that, along the riser, the ratio of the mass distribution of fuel and inert material is proportional to the ratio of the velocity distribution to the power of two, an assumption which is based on the kinetic energy of a single particle in vacuum, (23) hfuel is the vertical distribution of the fuel, hinert is the vertical distribution of the inert material. Ufuel is the velocity of the fuel particles, and Uinert is the velocity of the inert particles at the height z in the combustor. The vertical solids concentration can be estimated using a correlation by Johnsson and Leckner [1995], and since the concentration of fuel in the particle suspension is low, the solids concentration in the furnace is assumed to be that of the inert material. The inert material is assumed be mono-sized and uniformly distributed over the cross-section of the risers core. The velocity of the inert material and of the fuel is calculated from the slip-velocity, vinert or vfuel and the gas velocity, Ug .

(26) Aext is external surface area of fuel particles, g gas density, mfuel mass of the fuel particle, g gravity force, Rinert radius of an inert particle, minert mass of an inert particle and inert density of an inert particle. The drag coefcient CD is calculated according to Kovensky [1996] (27) The Reynolds number is based on the slip velocity of the fuel particle. The slip velocity of the inert particles is calculated for freely moving particles.

2.2.2 Average oxygen concentration and bed voidage


The mass balance of char combustion is solved for an average oxygen concentration, CO2 , and an average bed voidage, b . Both concentration and bed voidage are averaged over the mass of fuel,

(28)

(24)

CO2m is cross-sectional time-averaged oxygen concentration and bm voidage in the core. If the vertical distribution of the oxygen concentration is not known, it can be represented by the average oxygen concentration below the secondary air inlet, since nearly all fuel is located in the bottom

part. This underestimates the shrinking rate with approximately 15% for the small particle sizes, because some of the small fuel particles is in the oxygen richer areas above the secondary air inlet. However, since the large particles include nearly all the mass, this underestimation has only a small effect on the fuel loading.

3.2 Measurement compared with model calculation


The fragmentation constants and the relative size of the char fragments can be obtained by tting to measurements, knowing devolatilization and char combustion constants (a,b in Eq. 1 and corresponding for char combustion) from empirical data. The best t between measured and calculated fuel size distributions at 0.56m, 1.5m and 7.9m, and also the fuel concentration, is shown in Fig. 2 yielding a fragmentation constant during devolatilization Af,dev =0.11, a fragmentation constant during char combustion Af,char =3.2510-6 and a relative radius char =0.13. These values can be compared with fragmentation data for bituminous coal from Chirone et al. [1991], which can be converted to fragmentation constants varying between no fragmentation and 0.15 [1/s] during devolatilization, and between 210-6 and 710-6 [m/s] during char combustion, in a CFBC with a supercial velocity of 6 m/s, Arena et al. [1990]. The validity of the fragmentation constants obtained can be discussed comparing the measured size distributions and concentration of fuel, with the corresponding calculated ones if no fragmentation had occured. Fragmentation during devolatilization and char combustion leads to lower fuel concentration in the bed, but affects the size distribution of the fuel in different ways. Fragmentation during devolatilization controls the size distribution of the large particles; an increase of the fragmentation constant decreases the number of large particles in the bed. If no fragmentation took place during devolatilization, there would be a much higher concentration of large particles than shown by the measurement. This leads to the conclusion that there must have been fragmentation during the devolatilization. The number of particles following with the ue gases is much larger than the number of particles fed to the furnace, but also much larger than the number of particles which could be produced from fragmentation during devolatilization. This leads to the conclusion that the particles undergo fragmentation also during char combustion. The great increase of the number of particles below 1 mm can only be explained by fragmentation during char combustion. The number of small particles also gives an indication on the size distribution of the fragments produced during the char combustion, controlled in the model by the relative radius.

3 Experimental
The measurements were carried out in Chalmers 12MWth , CFBC, which is 13.5 m high and has a cross-section area of 2.5 m2 . The ratio K of the crosssectional area and the core area is approximately 1.3, Zhang et al. [1995].

3.1 Measurement
The fuel was bituminous coal with a known composition. The measured parameters were: 1. solids vertical distribution, by pressure difference measurement. 2. cross-sectional time-average oxygen concentration prole, by suction probe measurements at several positions in the cross-section and at several heights. 3. size distribution of feed fuel; samples taken from the fuel feeder were analyzed by sieving. 4. size distribution of char following the ue gases; samples were taken from the secondary cyclone and the bag-house lter and analysed by sieving. 5. size distribution and concentration of fuel in the furnace; samples were taken with a suction probe at the center position of the furnace at several heights, were sieved and the combustible content was determind on the entire sample and on every size fraction sieved. The analyzes of the samples taken with the suction probes were rough for the large fuel particles, for the fuel particles of the size of the inert material, and for the smallest particle sizes. There were only a few large particles present, and it was difcult to collect a representative amount of particles. The mass of fuel particles of the same size as the inert material was very small compared to that of the inert particles and close to what the equipment can analyze. The smallest particles can be attached to larger particles, and they can be produced during the sieving procedure.

Fuel mass p.d.f. [1/mm]

250

Fuel mass p.d.f. [1/mm] 0.1 1 10 Particle diameter [mm] Fuel concentration []

500

500

250

0 0.01 500

0 0.01

0.1 1 10 Particle diameter [mm]

Fuel mass p.d.f. [1/mm]

0.06

0.04

250

0.02

0 0.01

0.1 1 10 Particle diameter [mm]

0 0

5 Height [m]

10

Figure 2. Comparison between model (solid lines) and measurement (circles and stars) of the fuel mass probability density functions at 0.56m (top-left g.), at 3.7m (top-right g.) and at 7.9m (bottom-left g.) in the riser, and the fuel concentration along the riser (bottom-right g.).

4 Result
Fuel loading and overall size distribution in an FBC, and distribution at different heights in a CFBC are affected by the operating conditions.The following parameters have been investigated: 1. type of fuel 2. size distribution of fuel fed and fragmentation behavior 3. bed temperature 4. supercial velocity 5. air to fuel ratio 6. pressure To compare different operating conditions a number of parameters has to be held constant. If nothing else is said, the following parameters are held constant: 1. 2. 3. 4. 5. cross-sectional area power output average oxygen concentration average bed voidage supercial velocity

6. fragmentation constants and the relative radius are those determined above 7. rate constants for devolatilization and char combustion 8. size distribution of the fuel fed 9. size distribution and amount of fuel, entrained by the ue gases The input data are the same as in the case analyzed above. The fuel loading is related to this case. The fuel loading is the sum of the mass during devolatilization and tha mass during char combustion calculated from Eq 8 and the mass balans Eq 12.

4.1 Inuence of different fuels


The fuels considered are bituminous coal, woodchips and peat. Fuel specic data are time of devolatilization, kinetics during char combustion, fuel composition and fragmentation constants. The mass of fuel fed to the combustor has to be adjusted according to the different heating values of the fuels. Fragmentation rate constants of peat and woodchips has been estimated from visual observation

Mass p.d.f. [1/m] Coal Wood Peat

Fuel Loading []

200 100 0

0.5

0.1 1 10 Particle diameter [mm]

Figure 3. Fuel loading relative the fuel loading of coal (left g.) Size distribution of fuel in the bed (right g.), bituminous coal solid line, wood chips dashed line and peat dashed dotted line. N.B. the logarithmic scale. of the combustion in a CFBC and investigation of bed samples. The fragmentation constant during devolatilization is highest for peat and lowest for wood chips, the fragmentation constant during char combustion is lowest for coal and highest for peat and wood-chips. The relative radius is assumed to be the same for all fuels. The fuel loading and the size distribution of fuel in the boiler are shown in Fig. 3. The fuel loading is much higher for coal than for biofuels and lowest for peat. assumed to change proportional to bed temperature, since the driving force for fragmentation, the heating rate, increases with temperature. In Fig. 5 the inuence of bed temperature on loading and size distribution of fuel can be seen. An increase of the bed temperature decreases the fuel loading rapidly, and the number of small particles becomes smaller, a concequence of the constant fragmentation constant during char combustion. The reduction of fuel loading with increasing bed temperature is mainly due to the increase of the surface temperature of the particles, which is strongly connected to the bed temperature.

4.2 Inuence of size distribution of fuel fed to the combustor and fragmentation
The size distribution of the inlet coal in the case analysed above is a mean value of several samples. To show the inuence the size distribution of the inlet fuel and the fragmentation parameters on the fuel loading, the fuel loading has been calculated for the average, the largest and the smallest particle size distributions of the fuel fed, Fig. 4. Fig. 4 also shows the inuence of different fragmentation parameters on the size distribution of fuel in the combustor. The fuel loading decreases rapidly with an increase in fragmentation constant during devolatilization, and this mitigates the inuence of the size distribution of the feed fuel. The fragmentation constant during char combustion has a more moderate inuence on the fuel loading, and it does not have the same stabilizing effect for changes in the feed size distribution as that during devolatilization. The relative radius has nearly no infuence on the fuel loading but controls the distribution of the small particles (<1mm).

4.4 Inuence of supercial velocity


The amount of air supplied to the combustor is proportional to the supercial velocity. If the average oxygen concentration is held constant, also the fuel feed to the combustor and the power output become proportional to the supercial velocity. The combustion rate is only slightly affected by the supercial velocity in the special case of the CFBC, where the mass of fuel reaching the oxygen richer area above the secondary inlet increases with velocity. The fragmentation rate constant during char combustion is usually considered proportional to the velocity difference between the supercial velocity and the minimum uidization velocity. Then, the shrinking rate of fuel particles during char combustion increases with the supercial velocity, but the reduction of fuel loading created is much lower than the increase of the fuel loading caused by the higher input of fuel to the combustor. Fig. 6 shows the inuence of the supercial velocity on the fuel loading and the fuel concentration vs height in a CFB.

4.3 Inuence of bed temperature


The time of devolatilization is approximately inversly proportional to the bed temperature, Winter [1995]. The fragmentation constant during devolatilization is

Devolatilization 4 Fuel Loading [] Fuel Loading [] 3 2 1


Smallest size of fuel fed Largest size of fuel fed

Char Combustion 4 3 2
Largest size of fuel fed

Char Combustion 4 Fuel Loading [] 3 2


Largest size of fuel fed

1
Smallest size of fuel fed

1
Smallest size of fuel fed

0 0 0.05 0.1 0.15 Fragmentation Constant [1/s] 300 Mass p.d.f. [1/mm] Mass p.d.f. [1/mm]

0 0 0 3 6 9 0 0.25 0.5 0.75 1 Fragmentation Constant [m/s] Relative Radius [] 300 Mass p.d.f. [1/mm] 300

200

200
kfchar=9

200

100

kfdev=0.15

100

100

=0.02

0 0.01

kfdev=0

0 100 0.01

kfchar=0

0 100 0.01

=1

0.1 1 10 Diameter [mm]

0.1 1 10 Diameter [mm]

0.1 1 10 Diameter [mm]

100

Figure 4. The top gures show the fuel loading for the three size distributions of the fuel fed and the inuence of the fragmentation constant during devolatilization (left g.), fragmentation constant during char combustion (middle g.) and relative radius (right g.). The bottom gures show the inuence of the fragmentation constant during devolatilization (left g.), the fragmentation constant during char combustion (middle g.) and the relative radius (right g.). on the size distribution of the fuel in the combustor.

1.5 Mass p.d.f. [1/m] Fuel Loading [] 200


T=1023K

1 0.5 0 1000

100 0 0.01

T=1223K

1100 1200 Temperature [K]

0.1 1 10 Diameter [mm]

100

Figure 5. Fuel loading vs bed temperature (left g.) Size distribution of fuel in the bed for different bed temperatures (right g.)

4.5 Inuence of the air to fuel ratio


The average oxygen concentration is proportional to the air to fuel ratio. The inuence of the stoichiometric air to fuel ratio on the fuel loading and

the fuel size distribution is shown in Fig. 7. A rise in this ratio increases the shrinking rate during char combustion somewhat more than the proportional increase of the average oxygen concentration. The amount of small particles will be reduced accordingly

1.5 Fuel Loading [] 1 0.5 0 0

Fuel Concentration []

0.04

10 m/s 6 m/s

0.02 3.5 m/s 0 0 5 Height [m] 10

5 10 Superficial Velocity [m/s]

Figure 6. Fuel loading vs supercial velocity (left g.). Fuel concentration along the combustor (right g.).

2 Mass p.d.f. [1/m] Fuel Loading [] 1.5 1 0.5 0 1 1.2 1.4 Air to Fuel Ratio [] 200

100

Air/Fuel 1.05

0 0.01

Air/Fuel 1.5

0.1 1 10 Diameter [mm]

100

Figure 7. Fuel loading vs air to fuel ratio (left g.). Size distribution of fuel in the bed for different air to fuel ratios (right g.) and for the same reason as during an increase of bed temperature. The attrition rate constant increases with a rise in the oxygen concentration to some extent, Chirone et al. [1991], but here it is kept constant. Most signicant is that no attrition occurs if the oxygen concentration is close to zero. particles and inversely proportional to the square root of the pressure for large particles. If the air to fuel ratio and the supercial velocity are kept constant, then the mass of the fuel fed to the combustor, the power output and the average oxygen concentration are proportional to the total pressure. For kinetially controlled combustion, Eq. (14) shows that the shrinking rate (dR/dt) increases linearly with pressure. When diffusion dominates, the shrinking rate becomes independent of pressure. Only for large particles there is a square root dependence caused by the Reynolds number. For the size distribution considered these relationships give an increase of the shrinking rate, but due to the condition of constant air to fuel ratio and supercial velocity, the fuel loading increases as a consequence of the fuel feed rate being proportional to pressure. The resulting fuel loading and size distribution of fuel at different pressures are shown in Fig. 8. With the present size distribution the fuel loading increases with pressure to the power of 0.7, and the overall shrinking rate becomes proportional to the pressure to the power of 0.3, the shrinking rate being controlled by diffusion. The number of small particles becomes smaller with the rise of pressure, a consequence of the constant fragmentation constant.

4.6 Inuence of pressure


If a rst order reaction is assumed, the surface reaction rate constant, kc , is independent of pressure. This assumption can be questioned, and in recent works of Essenhigh [1996] and Croiset et al. [1996] there is some pressure dependence on the surface reaction rate constant. The pressure dependence on the mass transfer coefcient is different for small and large particles, caused by the correlation of Sherwoods number, Eq. (15). For small particles Sherwoods number is nearly constant, and for large particles it is proportional to the square root of Reynolds number. The kinematic viscosity and the molecular diffusivity are inversely proportional to pressure, and the slip velocity is only slightly affected by the pressure, due to a change in the drag force. This makes the mass transfer coefcient, , inversely proportional to pressure for small

10 Mass p.d.f. [1/m] Fuel Loading [] 200

100

1 bar 20 bar

0 0

10 Pressure [bar]

20

0 0.01

0.1 1 10 Diameter [mm]

100

Figure 8. Fuel loading vs pressure (left g.) Size distribution of fuel in the bed at different pressures (right g.)

5 Conclusions
A general model of fuel loading and size distribution of fuel in an FBC has been developed. For CFBC the fuel concentration and the size distribution along the riser are also modeled. The fragmentation is to some extent treated in a new way, and the fragmentation rate constants obtained from evaluation of experimental data show a good agreement with rate constants estimated from literature data. The model includes the most important inuencing parameters and has a short calculation time. (30 sec on a SUN Ultra). The fuel loading depends on type and size of fuel, fragmentation and pressure; all these parameters have a great inuence on the combustion. A rise of pressure and/or supercial velocity increases the power output from the combustor, and also the fuel loading. Bed temperature and air to fuel ratio have also an inuence on the fuel loading, but not as great as the other parameters investigated. The size distribution of fuel in the bed is mostly dependent on the fragmentation behavior of the fuel. For CFBC, the variation of fuel concentration along the riser is mainly affected by the supercial velocity.

Aext a,b CO2 CD DO2 F g Hc h K k M Mc m P R Re Sc Sh td U x Vdev v y

[m2 ] [mol/m3 ] [m2 /s] [kg/s] [m/s2 ] [J/kg]

[m/s] or [1/s] [kg] [kg/mol] [kg] [m]

Surface area of a fuel particle Fuel dependent constants Oxygen concentration Drag force coefcient Molecular diffusivity of O2 Mass ow Gravity Heat value Size distribution Area ratio of cross-section and core rate constant Total mass Molecular mass of carbon Mass of a single particle Fragmentation probability Particle radius Reynolds number Schmidt number Sherwood number Time for devolatilization Supercial velocity or velocity Relative volatile content Initial volatile content Slip velocity Radius of fragment

[s] [m/s]

[m/s] [m]

Greek letters [W/m2 K] [m/s]


b

Heat transfer coefcient Mass transfer coefcient Relative fragment size Bed voidage Density Scale factor Time for which a particle been subjected to devolatilization Split factor of CO and CO2

Acknowledgments
This work was supported by a scholarship from the Nordic Energy Research Program for Combustion of Solid Fuels and the Swedish National Board for Industrial and Technical Development (NUTEK).

[kg/m3 ] [s]

7 Nomenclature
Af,dev Af,char [1/s] [m/s] Fragmentation constant during devolatilization Fragmentation constant during char combastion

Index ash c char Char leaving with ue gases Carbon Char or during char combustion

dev f fuel g inert max n r tot 0 1 2

Volatiles or during devolatilization Fragments Fuel Gas Inert material Largest Number bases Radiation Total Initial During devolatilization During char combustion

Kunii D., Levenspiel O., Fluidization Engineering, John Wiley & Sons, New York, (1969) Kovensky, V.I., Account for the Constraining of Particle Motion in the Freeboard, Presented at the Int. Workshop on the INTAS project No.94-4313, Chalmers University of Technology Gothenburg, (1996). Nowak, W., Bis, Z., Gajewski, W., Matsuda, H., Hasatani, M., Carryover of Coarse Particles from a Dense Bed in a Multy-Solid Fluidized Bed, Preprints of the 5th Int.Conf. on CFB, Beijing, P. DGS11, (1996). Salatino, P., Massimilla, L., A Predictive Model of Carbon Attrition in Fluidized Bed Combustion and Gasication of a Graphite, Chem. Eng. Sci., 44, 1091-1099, (1989). Sasongko, D., Stubington, J.F., Signicant Factors Affecting Decolatilization of non Fragmenting, non Swelling Coals in Fluidized Bed Combustion, Chem. Eng. Science, 51, 3909-3918, (1996). Stanmore B.R., Brillard, A., Gilot, P., Delfosse, L., Fragmentation of Small Particles under Fluidized Bed Combustor Conditions, 26th Symp.(Int.) on Combustion, Combustion Institute, Pittsburgh, (1996). Winter. F., Single Fuel Particle and NOx/N2OEmission Characteristics under (Circulating) Fluidized Bed Combustor Conditions, Ph.D. Thesis, University of Technology, Vienna, (1995). Zhang, J.Q., Devolatilization an Combustion of Large Coal Particles in Fluidized Sand Beds, Technical Report QFBC.TR.87.2, Queens Fluidized Bed Combustion Laboratory, (1987). Zhang, W., Johnsson, F., Leckner, B., Characteristics of the Lateral Particle Distribution in Circulating Fluidized Bed Boilers, 4th Int.Conf. on CFB Technology, Ed. A.A.Avidan, AIChE, Hidden Valley, (1994), 266.

8 References
Arena, U., Cammarota, A., Chirone, R., Massimilla, L., Siciliano, L., Basu, P., Carbon Attrition During the Combustion of a Char in a Circulating Fluidized Bed, Combust. Sci. and Tech., 73, 383-394, (1990). Arena, U., Chirone, R., DAmore , M., Miccio M., Siciliano, L., Some Issues in Modelling Bubbling and Circulating Fluidized Bed Coal Combustors, Powder Technology, 82, 301-316, (1995) Arthur, J.R., Reactions between Carbon and Oxygen, Trans.Faraday Soc., 47, 164, (1951). Chern, J-S., Hayhurst, A.N., The Extent of Fragmentation of Various Coals during their Pyrolysis in a Hot Fluidized Bed, Joint Meeting of the Portuguese, British and Swedish Sections of the The Combustion Institute, Funchal, Madeira, April, (1996). Chirone, R., Salatino, P., Massimilla, L., Comminution of Carbons in Fluidized Bed Combustion, Prog. Energy Combust. Sci, 17, 297-326 (1991). Croiset, E., Chantal, M,. Rouan, J.P., Richard, J-R., The Inuence of Pressure on Char Combustion Kinetics, 26th Symp.(Int.) on Combustion, Combustion Institute, Pittsburgh, (1996). Essenhigh, R.H., Inuence of Pressure on the Combustion Rate of Carbon, 26th Symp.(Int.) on Combustion, Combustion Institute, Pittsburgh (1996). Halder P.K., Basu P., Mass Transfer from a Coarse Particle to a Fast Bed of Fine Solids, A.I.Ch.E. Symp. Ser No. 262., Vol. 84, pp. 58-64, (1988) Johnsson F., Leckner B,. Vertical Distribution of Solids in a CFB Furnace, Proceedings of the Thirteenth International Conferance on Fluidized Bed Combustion, Ed. K.J., Heinschel, ASME, New York, pp 671-679, (1995)

Appendix A: Fragmentation
A.1 Fragmentation during devolatilization of particles of a single size
During devolatilization fuel particles fall apart due to fragmentation. The fragmentation appears at different times for the feed particles, the principle behavior is shown in Fig. A1. A number of particles of a certain size is feed to the devolatilization process. After a time t1 one of the particles fragments, and the total mass decreases with the mass of one particle, at t2 another particle fragments, and the total mass of the particle size drops again, and so on. If the number of particles of a single size is large, the mass of the particle size can be treated as a continuous mass, and the mass loss can be described as:

dM (t) dt

f (t) M (t)

f (t) =

dM (t)=dt)=M (t)

(A 4)

M is the mass of a feed particle size, t is the time which a particle has been subjected to devolatilization and f(t) is a fragmentation rate function. The same example as in Fig. A1 is shown in Fig. A2 for a continuous mass instead of single particles. Assuming steady state conditions, the mass during devolatilization is given by the mass ow, F:

M=t

(A 5)

Rewriting Eq. A4 for the mass ow of a single particle size:

t1

t2

t3

td t

0 t1 t2 t3

td

Figure A 1. Principle fragmentation behavior of a group of particles having during devolatilization a certain size. M is total mass of the particle size, t is time and td is time of devolatilization. In this gure the mass loss due to the volatile release is not considered.
I

dF (R; t) dt

f (R; t)F (R; t)

F (R; t) = F 0 (R) exp(

f (R; t)dt)

(A 6)

F0 is the initial mass ow. The fragmentation rate function f(R,t) can be estimated from measurements, or it can be estimated from assumptions of the fragmentation behavior. The simplest assumption is that the fragmentation rate is independent of particle size and time:

f (R; t) = kf;dev

(A 7)

The fragmentation on a number basis is then proportional to the particle volume. For instance, for a spherical particle of the size R1 the probability to fragment at a given time compared with a smaller particle of size R2 becomes equal to (R1/R2)3 . Another simple assumption is that the fragmentation rate is independent of time, but dependent of size. Assuming that the pressure built up inside the particle can be modeled by a gas dome, a raise of temperature increases the pressure inside the dome and increases the stresses in the dome walls. The stresses in the wall of a gas dome is proportional to the ratio of the volume and the surface area, which means that the stresses in the dome wall is proportional

M(0)

M(t)
M (t)
frag

M(t )
d

0 M

0>t>t f(t)

Figur A 2. Principle fragmentation behavior of the mass of a single particle size during devolatilization. In this gure the mass loss due to the volatile release is not considered.
II

to the dome radius. The fragmentation rate assumed to be proportional to the surface stresses:

f (R; t) = kf;dev R

(A 8)

The fragmentation rate can also be assumed to be time dependent, for example: it may take some time before the fragmentation starts, and after a certain time the fragmentation stops. This can be modelled by spliting up the fragmentation rate function in steps:

f (R; t) =

kf;dev 1 f (R) 0 < t t1 kf;dev 2 f (R) t1 < t t2 kf;dev n f (R) tn < t td

(A 9)

The model assumes that a separation of devolatilization and char combustion is possible. The particle size is related to the feed particle size and not to the actual size during devolatilization, which makes it possible to model fuels which change size during devolatilization, due to shrinking or swelling. In the literature there are data on the probability to fragment for different fuels and particle sizes. These data can be used to estimate the fragmentation rate function f(R,t). The probability to fragment is the ratio of the mass of the fragments and the mass of fuel after devolatilization if no fragmentation had occurred:
t

F0 (R)exp(0) Pf;dev (R) =

F0 (R)exp F0 (R)exp(0)

f (R; t)dt
= 1

exp
0

f (R; t)dt
(A 10)

A.2 Fragmentation during char combustion


Fragmentation during char combustion differs fundamentally from fragmentation during devolatilization; during devolatilization the fragmentation is caused mainly by gas expansion and thermal decomposition inside the fuel particle, but during char combustion the fragmentation is caused mainly by particle collision and irregular combustion on the particles surface, Fig. A3. The different sizes of the fragments during char combustion make it convenient to split up the fragments into different categories and relate them to the origin. The ne fragments caused by the collisions between bed particles and fuel particles are produced by attrition, and the coarser fragments are caused by irregular combustion on the fuel particles surface, making the particle fall apart into larger pieces. The latter type is called secondary fragmentation. Other suggestions are also present in the literature. In this work the mass loss due to fragmentation is treated with
III

a single reaction rate, related to the surface area, Aext , and the slip velocity between bed particles and fuel particle, Us :

dM dt

M kf;char (Aext ; Us )

(A 11)

The fragment sizes are estimated by an size distribution.

A.3 Distribution of fragments


It is assumed that the fragment of a particle also can fragment, and the distribution of particle sizes can have been formed during several steps of fragmentation. Here the size distribution of a single fragmentation step is treated, the primary fragmentation. The primary distribution of fragments is not known, but experimental results on the size distribution after devolatilization and during char combustion are available in the literature for some fuels. If there is a large number of particles with different sizes, the primary distribution can be modelled by a ctitious distribution, which describes the size distribution after devolatilization and during char combustion. The primary distribution of fragments on a number basis is assumed to follow a rst order gamma distribution:

g (y ) =

exp (

y= ) 

(A 12)

where y is the radius of the fragments and  is a scale factor. The boundary condition of a gamma distribution is that g(y) has a value between zero and innity, but the maximum size of the fragments is limited to the size of the fragmenting particle, ymax . This makes it necessary to normalize the gamma distribution:

Bed Particles

Secondary Fragmentation

Fuel Particle

Bed Particle

Attrition products

Irregular combustion

Figur A 3. Fragmentation during char combustion


IV

gn (y ) =

exp( y=)
y

max

exp( y=)dy

= (1

exp( y=) exp( ymax = ))

(A 13)

The scale factor gives the shape of the distribution and the location of the mean radius. It is assumed that the relative location of the mean radius is at the same relative radius, , for all particle sizes. Dening the relative radius as the ratio of the mean radius of the non-normalized distribution, ymean , and the fragmenting particle size, yields:

ymean=ymax ymean =   = ymax

(A 14)

By changing the value of the relative radius, the distribution gn (y) is changed, and different fragmentation behaviors can be modelled, see Fig. A4. A small relative radius produces many small fragments and few large fragments. If the relative radius is large the distribution becomes uniform, which can be shown by an expansion of the rst order gamma distribution to a power series

exp (

y=) 

1 + (

y= ) +

y=)2
2!

y=)3
3!

y=)n n!

(A 15) The ratio y/ is much smaller than 1, if the relative radius is large, and this make the distribution uniform:
y

max
1

Const
0

dy = Const

ymax 

(A 16)

This gives

Const =

 ymax

g (y ) =

ymax

(A 17)

The model is not based on single particles. It is based on the mass of a large number of particles and the distribution on a mass basis is more interesting than the distribution on a number basis. If the fragments are assumed to be spherical, the mass can be related to a particle radius:

m=

y 3%
3

(A 18)

If all fragments have the same density, the mass distribution is given by the mass multiplied with the number of fragments of a given size:
V

hm (y ) = m N g (y ) =

y 3%
3

exp( y=) 

Const

y 3exp( y=) 

(A 19)

N is the average number of fragments produced from a fragmenting particle. For the probability density function, p.d.f, of the mass distribution of fragments the constant Const becomes:

1
Const
0

y 3 exp ( y=) dy = Const 


=

3

2y

y 2

y3

exp (

y=)

Const 6 3

= 1

(A 20)

This gives

Const =

1 6

3

(A 21)

The Const inserted into Eq. A19 converts the p.d.f of the mass of fragments into a fourth order gamma distribution:

h(y ) =

y 3 exp ( y= ) 4 6

(A 22)

In the same way as the distribution of the number of fragments the mass distribution is limited for particle sizes between zero and ymax , which makes it necessary to normalize also the mass distribution:

gn Small

Large

ymax
Figur A 4. The inuence of the relative radius, , on the distribution of the fragments
VI

hn

Small

Large

ymax y
Figur A 5. The inuence of the relative radius on the mass distribution of fragments

hn (y ) =

y 3 exp( y= )
y

max

(A 23)

y 3exp(

y=)dy

the scale factor of the mass distribution is:

ymean = 4  = ymax=4
and for large relative radii the mass distribution becomes:

(A 24)

hn (y ) =

4 ymax

y3

(A 25)

The inuence of the relative radius on the mass distribution can be seen in Fig. A5

Appendix B: Devolatilization
The devolatilization is modeled by the assumption of a shrinking core, see Fig. A6, and the time of devolatilization is modeled by the correlation:

td

aRb

(A 26)

Assume that the remaining devolatilization time t for a particle with the core radius r can be modeled with the same correlation:

t = arb

(A 27)

VII

Char residue

Devolatilization front
Figur A 6. Shrinking core assumption A relative volatile content x is dened as the ratio of the mass of the remaining volatiles and the initial volatiles:
Remaining mass of volatiles

x=

m0 (1

m0 Vdev

Vdev )

r R

t td

3=b

r
(A 28)
Initial mass of volatiles

The time during which a particle has been subjected to devolatilization,  , is:

= td

R
(A 29)

and the relative volatile content expressed as a function of the time  is:

x=

td

td

3=b

 td

3=b

(A 30)

The mass of a single particle mass at time  is:

m = m0 (1

Vdev ) + m0 Vdev x( ) = m0 (1


V olatiles

Vdev (1

x( )))

(A 31)

Char

For a steady state ow of fuel particles, feed to the devolatilization process the mass ow at a given time  is:

F0 (1

Vdev (1

x( )))

(A 32)

This expresses the mass ow of a fuel which does not fragment during devolatilization.

Appendix C: Simultaneous fragmentation and volatile release during devolatilization


VIII

C.1 A single particle size


Fragmentation means that particles fall apart into smaller particles, but during the devolatilization the mass decreases also due to the reduction of fuel density caused by the volatile release. This makes it necessary to compensate the initial mass ow in Eq. A6 in Appendix A for the density reduction Eq. A32 in Appendix B:

F (R;  ) = F0 (R) (1

Vdev (1

x(R;  ))) exp(

f (R;  )d )

(A 33)

V olatile release

F ragmentation

The total mass of a particle size is given by integration of the mass ow over the time of devolatilization:
td

Mdev (R) =
0

F (R;  ) d

(A 34)

The mass of char of the fragments is the difference between the char feed to the devolatilization process and the char of size R leaving the devolatilization for char combustion:
td
0

Ff p;char (R) = F0 (R)(1 F0 (R)(1

Vdev ) F0 (R)(1 Vdev ) exp

Vdev )exp
td
0

f (R;  )d
(A 35)

F eed char

Char af ter devlotilization

f (R;  )d

The mass of volatiles in the fragments is the difference between the volatiles fed to the devolatilization process and the volatiles leaving due to volatile release. The change of the mass ow is due to the mass reduction due to fragmentation and volatile release. When a derivation of Eq. A33 is done, the fragmentation and the volatile release are separated from each-other:

F (R;  ) = F0 (R) g (R;  )h(R;  ) g (R;  ) = 1 Vdev (1 x(R;  )) f (R;  )d ) h(R;  ) = exp(
and the derivative becomes:

(A 36)

F 0 (R;  ) = F0 (R) g 0 (R;  )h(R;  ) + g (R;  )h0 (R;  )


V olatile release F ragmentation

(A 37)

IX

where the primes denote derivatives. By integration of the volatile release over the time for devolatilization, the mass ow of volatiles leaving the particle size can be calculated. If the initial mass ow of volatiles is subtracted from the mass ow of volatiles leaving the particle size during devolatilization, the mass ow of volatiles in the fragments is achieved:

Ff p;vol (R) = F0 (R) Vdev + g 0(R;  )h(R;  )d


0

td

(A 38)

The plus sign is caused by the integration of the volatile release over time which turns out to be negative.

C.2 The size distribution of fuel particles


In a size distribution of fuel particles the model formulated for a single particle size corresponds to the largest particle size in the distribution; all other particle sizes will get a contribution of particles in form of fragments. If all fragments have the same relative volatile content as the fragmented particle, the fragments enter the corresponding particle size element having different relative volatile contents. For example, in Fig. A7, a particle of size R1 starts being devolatiles. After a time t1 the particle disintegrates into two pieces, one of the size R2 and one of the size R5. Both pieces are assumed to be spherical and have the same relative volatile content x1. After a time t2 the particle of size R2 falls into two pieces of the sizes R3 and R4, and both are assumed to have the same relative volatile content x2. The time t1 represents the time for a particle of size R1 to release volatiles corresponding to the relative volatile decrease 1x1. t2 is the time during which a particle of size R2 releases volatiles corresponding to the relative volatile decrease x1x2, and so on. With the assumption made above, the devolatilization process will be faster for a particle which fragments compared with a particle which does not fragment. This time reduction is strongly connected to the fragmentation behavior, e.g. if the particle falls apart into a large number of particles of approximately the same size in the beginning of the devolatilization, the time reduction will be signicant, but if the particles fall apart in one large particle and a few small particles in the beginning of the devolatilization or in a large number of particles in the end of the devolatilization, then the reduction of the devolatilization time will be small. To calculate the mass of a particle size, which is continuously fed with particles of different relative volatile content, it is necessary to build up a two-dimensional distribution of fragments, where the mass of the fragments is a function of relative volatile content and particle size. This is possible, but the calculation becomes both memory and time consuming. A faster way is to average the relative volatile content of all particles fed to one particle size. As shown above the total mass of volatiles and char in the fragments from one particle size are calculated in a rather simple way, and if the distribution of the fragments hn (y,R) is known, the mass of volatile and char fed to a particle size y from a larger particle size R becomes:
X

Ff;char (R; y) = Ff p;char (R)hn (y; R) Ff;vol (R; y) = Ff p;vol (R)hn (y; R)
and the mass of volatiles and char from all larger particle sizes R>y are:
Rmax

(A 39)

Ff;char (y ) =
y+

Ff p;char (R)hn (y; R) dR


Rmax

(A 40)

Ff;vol (y ) =
y+

Ff p;vol (R)hn (y; R) dR

The average relative volatile content in the fuel fed to a particle size, x0 , is then:
M ass of volatiles f ed to a particle size

x0 (y ) =

F0 (y )Vdev + Ff;vol (y ) F0 (y )(1 Vdev ) + Ff;char (y )


1

Vdev Vdev

(A 41)

Initial volatile mass related to the mass of char f ed to a particle size

When the average relative volatile content in the fuel fed to a particle size is less than one, a compensation of the initial mass in Eq. A33 must be done. This compensation is done by the introduction of a ctitious initial mass ow based on the initial mass ow of char fed to a particle size, F0,char :

F0;char (y ) = F0 (y )(1

Vdev ) + Ff;char (y )

(A 42)

The mass of char is only affected by fragmentation, and at  (x0 ) the mass ow of char is the same as the initial mass ow of char. The ctitious char mass ow, F* 0,char , is then the corresponding mass ow at time zero:
3 F0;char (y ) = F0;char (y )=exp
 (x0 (y ))
0

f (; y)d

(A 43)

and the ctitious initial mass ow, F0 * of fuel at time zero is:
3 3 F0 (y ) = F0;char (y )=(1

Vdev )

(A 44)

The equations for a single particle size, Eq. A33A35 and A38, can then be rewritten for a size distribution and the mass ow becomes:
3 F (y;  ) = F0 (y ) (1

Vdev (1

x(y;  ))) exp(

f (y;  )d )

(A 45)

V olatile release

F ragmentation

XI

The corresponding total mass is:


0

td

Mdev

=
Rmax  (x0 )

F (y;  ) ddy

(A 46)

The mass of char in the fragments:

Ff p;char (y ) = F0;char (y )

exp

td  (x0 )

f (y;  )d

(A 47)

and the mass of volatiles in the fragments:


3 Ff p;vol (y ) = F0 (y )Vdev + F0;vol (y ) + F0 (y )g 0 (y;  )h(y;  )d
0

td

(A 48)

The averaging of the relative volatile content produces an error, which is small (05 %) in most cases. When the fragmentation rate function is assumed to follow Eq. A7 the error is less than 2%. When a time dependence is assumed on the fragmentation rate function, e.g. Eq. A9, this error can become unacceptably large. The worst case if the fragmentation is takes place only at the very end of the devolatilization period, when the error can be around 30%. To keep this error at an acceptable level, the devolatilization can be split up into several steps related to the relative volatile content. For example, if all fragmentation occurs between a relative volatile content of x1 and 0, then the devolatilization can be split up into two steps, one for the relative volatile content between 1 and x1, and one between x1 and 0. The relative volatile content in the fragments can be modelled by two extreme cases: one is the one described above Fig. A7, where all fragments have the same volatile content, and the other is that the char layer falls of the shrinking core as illustrated in Fig. 8. A particle of size R1 starts devolatilization. After a

x R1

1 t1 R2

x1 t0 t2 R3 R4 R5

x2

R t1 t2 t0

t3 t4 t5 t5

t3 t4

Figur A 7. Time for devolatilization of a fragmenting particle


XII

x R1 R2 R3

1 t1

x1 t0 t2 t3

x2

R t1 t0 t2 t3

R4 R5
Figur A 8. Time for devolatilization of a particle who the char layer falls of the core time t1 the particle disintegrates into two pieces, one piece of size R2 originates from the core and the other piece of size R5 originates from the char layer. The particle which consists of char goes directly to char combustion. After a time t2 the same procedure is repeated and the core forms a particle of size R3 and the char layer leaves with a particle of size R4. In this extreme case the devolatilization time is independent of the fragmentation. To model this, two mass distributions of the fragments have to be assumed, one for the distribution of the cores, hcore and a second for the distribution of the char fragments, hf,char . The mass of the fragments, including the core, which is feed to a particle size can be calculated from the volatiles in the fragments:

Fcore (y ) =

Rmax

Vdev

hcore (R)Ff p;vol (y; R)dR


y+

(A 49)

and the mass of the fragments including the char of a particle of size R leaving the devolatilization for the char combustion is:
Rmax

Ff;char (y ) =
y+

hf;char (R)Ff p;char (y; R)dR

Fcore (y )(1

Vdev )

(A 50)

The devolatilization can then be calculated as for a single particle, with the initial mass ow compensated with the mass ow of the core fragments, and the char leaving for char combustion compensated with the mass ow of char fragments.

XIII

Appendix D: Char burning as shrinking spheres


The mass reduction of a single particle during char combustion can be expressed as a function of external surface area, Aext , a mass transfer coefcient, , a reaction rate constant, kc , an oxygen concentration CO2 and a split factor, , the molar ratio of C and O2 :

dmcarbon dt

f (Aext ; ; kc ; CO2 ; )

(A 51)

Assuming spherical particles, the mass of the particles can be related to particle radius

mchar

4%char R3 3

(A 52)

By dening the density of the carbon in the char as

%carbon = %char

%ash

(A 53)

the mass of carbon in the particle can be expressed:

mcarbon =

4%carbon R3 3

(A 54)

If the density is constant during char combustion, the mass reduction is only caused by a change of the radius:

dmcarbon dt
dR3 /dt can be expressed as dR/dt:

4%carbon 3

dR3 dt

(A 55)

dR3 dt
where dR3 /dR is:

dR3 dR dR dt
3 (R + dR)

(A 56)

dR3 dR

= lim dR 0

R3 ! R

(R + dR)

= 3R

(A 57)

Putting Eq. A56 and A57 into Eq. A55 yields:

dmcarbon dt

= 4%carbon R

2 dR

dt

(A 58)

This corresponds to the mass loss of a single particle, but in a combustor there is a mass loss of a collection of particles of the same size:

XIV

dM dt

dmchar dt
=

%char dmcarbon M %char 2 dR = 4%carbon R dt %carbon dt mchar %carbon M %char dR 3M dR 2 4%carbon R = 4%char R3 % dt R dt carbon 3
=

(A 59)

M is the mass and N is the number of char particles of a certain particle size. The mass reduction for a small particle is controlled by the kinetics of the surface reaction, which can be calculated from a reaction rate constant, kc . The reaction rate constant is dened in different ways in the literature. The most common and simple denition is that the reaction rate constant is related to the partial pressure, PO2inf or to the concentration, CO2inf , of oxygen in the bulk and the reaction is assumed to be of rst order. For a shrinking sphere model the mass decrease is related to surface area:

dm dt

3 3 3 Aext PO2inf kc1 = Aext CO2inf R3 T kc1 = Aext CO2inf kc

(A 60)

T is the surface temperature and R* is the gas constant. This gives the reaction rate constant k* c the unit [m kgcarbon / mol(O2) s]. In order to relate the reaction rate to the mass transfer coefcient of oxygen into the particle, the reaction rate constant has to have the unit [m/s]. This can achieved by separation of the relation between mass of carbon and mol of oxygen from the reaction rate:
m kgcarbon molO2 s
3 kc

= =

molcarbon molO2

MC kgcarbon molcarbon

kc
m s

(A 61)

Putting Eq. A61 into Eq. A60:

dmcarbon dt

Aext MC CO2inf kc

(A 62)

The mass reduction of a large particle is controlled by the diffusion of oxygen into the particles surface which can be calculated from the mass transfer coefcient, . When the combustion follows a shrinking sphere model and is controlled by diffusion, the oxygen concentration at the surface is much smaller than in the bulk, and the difference in the oxygen concentration between bulk and particle surface is approximately the same as the oxygen concentration in the bulk.

dmcarbon dt

Aext Mc CO2inf

CO2s

Aext Mc CO2 inf

(A 63)

The simplest and most usual way to relate the mass transfer and the reaction rate to each-other, is to treat them as two parallel oxygen resistances and call the resulting resistance apparent reaction rate, kapp :
XV

kapp = 01 0 + kc 1
The mass reduction expressed with the apparent reaction rate is:

(A 64)

dmcarbon dt

Aext Mc CO2inf kapp

(A 65)

By inserting Eq. A58 into Eq. A65, the mass reduction can be related to the shrinking rate of the particle, :
Aext
= 4R
2

2 dR 4%carbon R dt

Mc CO2inf kapp

dR dt

Mc CO2inf kapp %carbon

(A 66)

Appendix E: Determination of fragmentation rates from data in literature or from experiments


E.1 Fragmentation during devolatilization
Measurements of the fragmentation during devolatilization is presented in the literature in many different ways, dependent on the type of experiment performed. The most common way is to estimate the average number of particles produced from a fragmenting particle of a given size, often together with a mean size of the fragments. For some fuels these data are complemented with data on probability to fragment or/and size distribution of the fragments after devolatilization. In some rare cases an estimation on the timing of the fragmentation is made. Most experiments are carried out for coals, which show large variety in fragmentation behavior. Common for nearly all coals is that they produce a countable amount of fragments, which means that the particle falls apart in rather few large pieces. This corresponds to a large value of the relative radius and makes the primary fragmentation follow Eq. A25. The reaction rate function, f(R,t), has to be assumed. If there are data available on the timing of the fragmentation these data can support the choice of a fragmentation rate function f(R,t). The fragmentation rate, kf,dev , can be estimated by tting the calculated, Eq. A10, to the measured probability to fragment. The fragmentation rate can also be estimated from the number of particles produced and the mean size of fragments after devolatilization. The number of particles produced from one particle of the original size R, Nf , can be calculated:

XVI

Nf

mo F0
1=Number

R
0

3 4%char y 3

F (y; td )

dy

of particles feed to devolatilization per second

mass per second of particles Mass of one leaving devolatilization char particle
Number of particles per second leaving devolatilization

(A 67)

Eq. A67 reveals the weakness in the use of the number of fragments; even if the mass of the small particles is negligible, the number of these particles can be in majority. This means that the average number of particles produced from a particle of a given size R presented in the literature must be related to the possibility to collect particles of the smallest particle sizes in the experiments performed. For example: Chern et al. [1996] reported that for coal particles of an initial mass of 2g, corresponding to a particle radius of about 7.5 mm, less then 5% of the mass was found in fragments having a particle radius smaller than 1.5mm. For most fuels the average number of fragments after devolatilization was between 1 and 10. Assume that the fuel does not swell and that an average of 10 particles with a particle radius larger than 1.5mm is produced during devolatilization. A calculation with the fragmentation rate function following Eq. A7 and with a fragmentation rate, kf,dev , equal to 0.035 1/s, gives 10 particles with a particle radius larger than 1.5mm. These 10 particles corresponds to 97% of the mass, but the total number of particles given by Eq. A67 is 25. So 60% of the particles corresponds to 3% of the mass. Fig A9 shows the calculated cumulative distribution of mass and number of particles after devolatilization in the example above. The conclusion is that the number of particles produced from a particle of size R, reported from experiments, does not give sufcient information, if it is not complemented with the mass loss and the smallest size of the collected particles.

E.2 Fragmentation during char combustion


During char combustion fragmentation is caused by attrition and secondary fragmentation. Values on the attrition rate during char combustion in a CFBC are given by Arena et al. [1990]. Attrition rates for BFBC conditions have been reported in several works, e.g. Chirone [1991], Brown et al. [1992]. All these works use a dimensionless attrition rate constant, katt , which gives the fragmentation rate during char combustion, kf,char , according to:
XVII

1 Cumulative fraction 0.8 0.6 0.4 mass 0.2 num. of part. 0 0 3 6 9 12 Diameter [mm] 15

Figur A 9. Illustration of the cumulated mass and number of particle fraction of particles leaving the devolatilization for char combustion

dMf (R) dt

=k =

att

kf;char

katt U

2R

Umf Umf

M (R)

(A 68)

2R

U is the supercial velocity and Umf is the minimum uidization velocity. For a CFBC katt is in the range of 0.110- 6 to 0.710- 6 . The mass distribution of the fragments is controlled by the relative radius. The value of the relative radius is small (0 < < 0.05) if only attrition occur and large (0.5 < < 1) if only secondary fragmentation occur. In the mass balance given by Salatino [1989], which is an extension of the mass balance presented by Kunii and Levenspiel [1969], the fragments leaving a particle size element centered at R are only related to the secondary fragmentation:

Mchar

d 0 = F (R)dR dR dR (R)M h (R) dt h (R) 3 dR (R)dR + F (R)dR M


char char char char

Fash (R)dR hchar (R)kf;char (R) dR

R dt

2;char

char

dM f =dt

(A 69)

If there is only attrition, the size reduction, which makes the particle go from one size to another, will be affected and the shrinking rate due to char combustion has to be complemented with a shrinking rate due to attrition:

XVIII

0 = F (R)dR
char

d dR (R) + dRf (R) Mchar hchar (R) Fash(R)dR dR dt dt 3 dR (R)dR + F Mchar hchar (R) Mchar hchar (R)kf;char (R) dR f 2;char (R)dR R dt

dMf =dt

(A 70)

Index f indicates the size reduction due to attrition. The size reduction due to attrition can be calculated from the fragmentation rate:

dMf (R) dt

= 3MR(R) dR = k dt dR =k R dt 3
f f f f;char

f;char

Mf (R)

(A 71)

In order to handle the attrition and secondary fragmentation simultaneously, the two extreme cases have to be combined in some way. In the calculation made, the extra size reduction rate due to attrition was assumed to have only a minor effect on the total mass and on the mass distribution, and Eq. A69 was therefore used. This assumption overestimates the total mass. In the cases calculated the overestimation was less than 5 %. A better way is to connect a split factor, , to the relative radius

0 = F (R)dR
char

d dR (R) +  dRf (R) Mchar hchar (R) Fash(R)dR dR dt dt 3 dR (R)dR + F Mchar hchar (R) Mchar hchar (R)kf;char (R) dR f 2;char (R)dR R dt

dMf =dt

(A 72)

A large relative radius gives larger fragments, which by denition are produced by secondary fragmentation, and a small relative radius gives ne fragments, which by denition are produced by attrition. The relative radius can be related to the split factor

=

1 2 0 < < 0:5 0 0:5

(A 73)

XIX

E.3 Estimation of fragmentation rates from large scale experiments


The fragmentation rates can be estimated from experiments in a large scale boiler. The necessary measurements for this estimation are mass ow and mass distribution of feed fuel and y char, together with the mass distribution and concentration of fuel particles in the bottom part of the boiler, total mass of material in the combustor. The mass distribution fuel concentration has only to be measured in the bottom part of the combustor, since nearly all fuel is located in this part. The measured mass distribution in the combustor is assumed to be representative for the true mass distribution in the combustor. To make a reliable estimation of the fragmentation data the accuracy of the measurement should be evaluated, and a maximum and a minimum mass distribution of the fuel in the combustor had to be determined. The range of the fragmentation data can then be estimated from the maximum and the minimum mass distributions of fuel in the combustor. The estimation of the fragmentation rates during devolatilization and char combustion and of the relative radius during char combustion can be carried out as follows: 1. Determine the devolatilization constants, a and b in Eq. A26 and the reaction rate constant, kc during char combustion (literature data). 2. Assume the fragmentation rate function during devolatilization, e.g. Eq. A7-A9. 3. Estimate the minimum and maximum fragmentation rates, kf,dev during devolatilization. 4. Estimate the fragmentation rate, kf,char and relative radius, , during char combustion. The devolatilization constants and the reaction rate constant for char combustion can be taken from the literature or can be determined by experiments. Comparison between the calculated mass if no fragmentation occurs and the measured mass of fuel in the combustor gives an estimation of the accuracy of the chosen reaction rate constant and the devolatilization constants. If the measured mass is smaller than the calculated one, then the reaction rate and/or the time for devolatilization are underestimated, because some fragmentation and attrition of the fuel particles takes place in a FBC, and the total mass of the fuel will be reduced. This reduction of mass of fuel is most likely larger than the measurement error. The reaction rate constant, kc , has the greatest inuence on the total mass of fuel in the combustor and is fuel specic. Data on the reaction rate constant are available in the literature for different fuels, but the values are scattered and it is difcult to pick the correct reaction rate

XX

constant for the fuel used in a particular experiment. Only a range of possible values can be determined. The fragmentation rate function, f(R,t) is not known and has therefore to be assumed. Suggestion on assumptions is given by Eq. A7-A9. The minimum and maximum fragmentation rate, kf,dev , during devolatilization can be estimated by comparison between calculated and measured mass distribution of fuel in the combustor. Estimation of the minimum fragmentation during devolatilization is illustrated in Fig. A10: 1. Assume that no fragmentation occurs and calculate the mass distribution of the fuel during devolatilization, kf,dev =0 2. Compare the mass distribution with the measured one in the combustor 3. Increase the fragmentation rate until the calculated mass distribution of the fuel during devolatilization coincides with the measured mass distribution

Mass distribution [kg/m]

kf,dev

Particle radius [m]


Figur A 10. Determination of the minimum fragmentation rate. Full line is measured mass distribution of fuel in the combustor. The dashed lines are calculated mass distributions of fuel during devolatilization in the combustor for different fragmentation rates, kf,dev
XXI

The minimum fragmentation rate is then the smallest fragmentation rate, which fulls the condition that the mass distribution of the fuel during devolatilization coincides with the measured mass distribution. Estimation of the maximum fragmentation rate during devolatilization is illustrated in g A11: 1. Assume that no fragmentation occurs during char combustion, kf,char =0, and calculate the mass distribution 2. Compare the mass distribution with the measured one in the combustor 3. Increase the fragmentation rate, kf,dev until the calculated mass distribution coincides with the measured mass distribution The maximum fragmentation rate is then the largest fragmentation rate, which fulls the condition that the calculated mass distribution of the fuel, if no fragmentation occurs during the char combustion, and which coincides with the measured mass distribution. The fragmentation rate and the relative radius during char combustion are estimated from the measured normalized mass distribution. It can be concluded that there had been fragmentation during the char combustion, since the number of char particles leaving with the y char is much larger than the number of particles feed. The particles produced during devolatilization are limited and even if the largest possible fragmentation rate during devolatilization is used, the number of particles after devolatilization is nowhere near the number of particles which are leaving with the y char. Fragmentation during char combustion is then the only possible explanation for this dramatic increase of particles of smaller

Mass distribution [kg/m]

kf,dev

Particle radius [m]


Figur A 11. Determination of the maximum fragmentation rate. Full line is measured and dashed lines are calculated mass distributions of fuel in the combustor, with different fragmentation rates, kf,dev .
XXII

sizes. The fragmentation rate and the relative radius during char combustion can be estimated as follow, from Fig. A12: 1. The calculated and the measured normalized mass distributions of the large particles are tted to each other, by changing the value of the fragmentation rate during devolatilization between its minimum and maximum, and the value on the fragmentation rate during char combustion. An assumed value is used for the relative radius. 2. The relative radius is adjusted so that the small sizes are tted. Finer adjustment of the fragmentation rates during devolatilization and char combustion can be necessary. The fragmentation rates and the relative radius are then taken for the case of the best t between the model and the measured normalized mass distribution. Compare the calculated total mass with the measured one and compare the data on the fragmentation rates with data from the literature. If the calculated total mass is outside the possible range of the measured total mass in can this in most cases it can be adjusted by changing the reaction rate constant. Other parameters which have an inuence and have to be investigated in a full analysis is time for devolatilization, fragmentation rate function during devolatilization and primary distributions of fragments during devolatilization and char combustion.

Mass distribution [1/m]

kf,char char

Particle radius [m]

Figur A 12. Determination of the fragmentation rates and the relative radius. Full line is measured and dashed lines are calculated normalized mass distributions of fuel in the combustor, with different fragmentation rates kf,dev and relative radius
XXIII

Appendix F: Fuel distribution along the riser


In an FBC the bed particles have a rather narrow size distribution, making it possible to handle the bed particles as a single size. The fuel particles on the other hand, have a wide size distribution, but because the mass of the fuel particles only corresponds to some percent of the total mass the fuel particles can be treated as individual particles in a suspension of bed particles. When a single particle of a different size and density is moving in a suspension it will be affected by collisions with bed particles. The collision force makes it possible for large particles to be transported up through the furnace, even if the drag force from the gas is not large enough. The collision force is a direct function of the density of the suspension, where a high density corresponds to a large collision force and a low density corresponds to a small collision force. Also the small particles are affected by this collision force but in an opposite way; the collision force acts in the opposite direction and instead caring the particles upwards it holds them down. In a CFBC the suspension density is high in the bottom and low in the top. This makes the movement of the fuel particles in the bottom region controlled by collisions and in the top controlled by the drag force from the gas. When the movement of the fuel particle is controlled by collisions, all fuel particles attain approximately the same velocity as the bed particles. When the movement of the fuel particle is controlled by the drag force, the slip velocity increases with particle size. At a certain particle size the slip velocity will be larger than the gas velocity, and consequentially the particle will not be further transported up through the riser. Fig. A13 shows the velocity of the gas up through the furnace, the bed particles, and large and small fuel particles. The change of suspension density with height in a CFBC makes it necessary to assume how this inuences the vertical distribution of a single fuel particle size. From knowledge of the action of the collision force on different particles sizes the following behavior can be expected: 1. particles with a larger slip velocity than the bed particles have a smaller possibility to be transported up through the riser than the bed particles. 2. particles with a smaller slip velocity than the bed particles have a greater possibility than the bed particles to be

U
Gas Small fuel particle Bed particle Large fuel particle

zmax z

zmax z

Figur A 13. Particle and gas velocity versus height in the riser.
XXIV

transported up through the riser. Fig. A14 shows the mass ratio of bed and fuel particles of different sizes, and also the suspension density. The bed density decreases very fast in the riser. For a large particle the density is too low at a certain position to carry the particle any further. But the decrease of the mass ratio between large fuel particles and bed particles is not immediate, it changes successively with the bed density. The small particles experience the opposite, the mass ratio of bed particles and fuel particles increases with decreasing bed density. The conclusion is that particle velocity and mass ratio of fuel particles of a given size and bed particles behave in the same way. The fuel and bed particles can be correlated with each other:

Mf uel (R; z ) Mbed(z )

=K

Uf uel (R; z ) Ubed(z )

(A 74)

Assuming that the mass ratio is proportional to the possible height, according to the theory of kinetic energy a particle attain the height:

mgh =
and

mU 2

h=

U2 2g

(A 75)

Mf uel (R; z ) Mbed(z )

=K

hf uel (R; z ) hbed(z )

2g U 2 (R; z ) = K 2g U 2 (z) = K
f uel bed

Uf uel (R; z ) Ubed(z )


2 2

(A 76)

The vertical mass distribution of a fuel particle size can then be calculated:

hf uel (R; z ) =

hbed(z ) Uf uel (R; z )=Ubed(z )


H

hbed (z ) Uf uel (R; z )=Ubed(z )

(A 77)

dz

H is the height of the combustor.

/max 1 0

M(R)/Mb
Small particles

zmax

1 z 00

Large particles

zmax z

Figur A 14. Relative bed density vs. height in the riser and relative mass ratio between fuel particles of a given size and bed particles related to the ratio in the bottom of the riser.
XXV

Paper III

Combustion of wood particles a particle model for Eulerian calculations


Henrik Thunman, Bo Leckner, Fredrik Niklasson, Filip Johnsson
Department of Energy Conversion,Chalmers University of Technology,S-412 96 Gteborg, Sweden +46 31 7721430, +46 31 7723592 (fax), email energy.conversion@entek.chalmers.se

(Submitted for publication)

Abstract
A simplified thermochemical particle conversion model, independent of particle size and shape, is derived. The model operates with a small number of variables and treats the most essential features of conversion of solid fuel particles, such as temperature gradients inside the particle, release of volatiles, shrinkage and swelling, considering also typical shapes (spheres, finite cylinders and parallelepipeds). The model treats the particle in one dimension, and the conversion can be described by the heat and mass transport to the surface of the particle. When modelling a large combustion system this is a great advantage, as the model does not have to be limited to just a single particle. In fact, it can handle the conversion of a solid phase in a computational cell, where the conversion is related to surface area per unit volume instead of surface area of a single particle. The model divides the particle into four layers: moist (virgin) wood, dry wood, char residue and ash. The development of these layers is computed as function of time. The model shows satisfactory agreement with measurements performed on more than 60 samples of particles of different sizes, wood species and moisture contents. Comparison with the experiments shows that the simplifications made do not significantly influence the overall accuracy of the model. The model also demonstrates the great influence of shrinkage on the time of devolatilisation and char combustion.

Introduction
A fuel particle entering a combustor dries and devolatalises. Subsequently, or even simultaneously if the particle is large enough, the residual char layer is oxidised, and in the end only ash remains. During all these processes the particle shrinks. The conversion of single solid fuel particles is usually described by two categories of models. The first concerns combustion in a boiler containing a great number of particles, for example, [1], [2] [3], [4], [5], [6] and [8], where the principal assumption is that a single particle in the combustor can be treated as thermally thin and that the stages of conversion, drying, devolatilisation and char combustion, occur in sequence as the temperature of the particle raises. Alternatively, the stages of conversion, starting at the surface of the particle and moving towards the centre, are treated by an empirical consideration, as a result of heat and mass transport to the surface of the particle. The other category of models deals with modelling of the single particle, exposed to a well-defined surrounding. In this type of models the particle is divided into a great number of computational cells, often more than hundred, in order to catch drying, devolatilisation and char combustion while these processes move from the surface to the centre of the particle, for example, [8], [9], [10], [11], [12] and [13]. Even with this large number of cells, the models often only treat one-dimensional particles, such as infinite plates, infinite cylinders and spheres. The weakness of the first category of models is obvious, as the combustion, especially of large particles, is simplified so much that the combustion behaviour of the particle is missed. The second category of models becomes computationally too heavy and cannot be included in a large computational system describing a large number of particles

2 converted at the same time, for example in a bed of particles, and furthermore, typical particle shapes, such as parallelepipeds and finite cylinders, cannot be treated. There have been attempts to model different shapes by relating a shape factor, based on the initial particle shape, to heat and mass transfer to the surface of the differently shaped particles, for example in [14]. However, the conversion phenomena that are to be studied here take place to a large extent inside the particle, and the particle changes in size during the conversion. Therefore such shape factors are not suitable, and a more general formulation of the shape factor is needed. For this purpose a computationally efficient model should be formulated that operates with a constant, small number of variables and that is suitable for all particle sizes and shapes. The model should treat essential features of the fuel particle: shape (spheres, finite cylinders and parallelepipeds), temperature gradient inside the particle, shrinkage and swelling of the particle during conversion. Here, a generalised description of such a particle model will be derived and compared with measurements. The outline of the discretisation of the particle in such a model is shown for a generalised particle in Figure 1. The particle is divided into n layers from the centre to the surface. The layers inside the particle are denoted by p, followed by an index that specifies the layer. Here, the particle is divided into four layers: moist (virgin) wood (1), dry, devolatilising fuel (2), char residue (3) and ash (4). The boundaries between these layers are denoted by b followed by an index specifying the stage of combustion related to the boundary: drying (1), devolatilisation (2), char combustion (3) and particle surface (4). The amount of fuel at a certain stage of combustion inside a particle determines the size of the layer. As the conversion of the fuel particle proceeds, solid matter will leave layer j and enter layer j+1. The drying, devolatilisation and char combustion fronts moving from the surface to the centre of the particle coincide with the boundaries between the layers. In order to validate the model, more than 60 wood samples were dried and devolatalised in an inert atmosphere, followed by burnout of the remaining char in a nitrogen-oxygen atmosphere.
b,4 b,3 b,2 b,1 p,1 p,2 p,3 p,4

Figure 1. Fuel particle consisting of p,1..4 solid layers and b,1..4 boundaries between the layers. The dashed line indicates the initial shape of the particle before shrinkage.

Theory
In a generalised particle, such as in Figure 1, all heat and gas produced pass a control surface, whose position depends on the radius r, as expressed in the conservation equations. This control surface is related to a control volume V as
dr = dV

(1)

Applied to a spherical volume the control surface, that is, the shape factor, becomes =4r2, and for the well-known cases of infinite cylinder and infinite plate the shape factor is 2r and 1, respectively. The shape factor can be derived for a finite fuel particle of any shape, implicitly assuming an isotropic behaviour of the particle or that representative mean values over the surface can be formulated. The derivation based on Eq (1) assumes that the position of the control surface is determined by the requirement of equal distance from the surface of the particle. Hence, for a finite cylinder of initial length l and initial diameter d we obtain
= 2 3r 2 + r ( l d )

(2)

and for a parallelepiped with the initial dimensions l1 l2 l3, where l1 represents the shortest end
= 24r 2 + 8r ( ( l2 l1 ) + ( l3 l1 ) ) + 2 ( l2 l1 )( l3 l1 )

(3)

With this definition the starting location of the radius (r0), which for a sphere is r0 =0, becomes a line for a long cylinder (l>d) or a plate for a short cylinder (l<d) and a parallelepiped. This is illustrated for a finite cylinder in Figure 2. For all shapes, the starting location of the radius is r0 =0, except for a short cylinder, where the starting location is r0=(d-l)/2. In the one-dimensional model the effects of anisotropic material and corners (finite cylinders and parallelepipeds) are averaged over the surface , and therefore it is important to also give physical and thermo-chemical data in the form of surface averaged values.
l l

l/2 l/2

r0=(d-l)/2
d

d/2 d/2 r0=0

r d r

r r

Figure 2. Definition of the starting position of the radius r0 and the control surface (solid) at radius r inside a finite cylinder (dashed) having a length l, longer or shorter than the diameter d.

4 In [15] a one-dimensional model is presented for simultaneous drying and devolatilisation of a single particle having the shape of a sphere, an infinite cylinder or an infinite plate. This model is expanded here to include char combustion, shrinkage and the two particle shapes treated above, finite cylinder and parallelepiped. The model [15] was derived under the assumptions of thermal equilibrium between gas and solid phases inside the particle, no gas phase reactions inside the particle, immediate release to the surrounding of gases produced, homogeneous heating of the particles surface, and evaporation of water in an infinitely narrow region inside the fuel particle. By adding the assumption that also devolatilisation and char combustion take place in infinitly narrow regions inside the particle (covered by the ash layer), the conversion of the single particle can be described by the following energy conservation equation in one dimension,
I p t T 1 1 ( BpTp ) = r k p rp + b,4 r * rb, j Rb, j is, j r j b, j

(4)

Here I is enthalpy, B is heat capacity flow per unit area, T temperature, R reaction rate, delta function, i specific enthalpy and index s indicates a component in the solid. The heat capacity flow per unit area, originating from the conversion of the solid fuel to gas, is given by

r ( r * rb , j ) Rb , j B p = b ,4 dr * X i , j cg ,i j r0 i

(5)

where c is specific heat and X is molar stoichiometric coefficient for species i in reaction j. The continuous formulation of conversion of a particle, Eq (4), is discretisized into four layers by averaging temperature, physical and thermo-chemical data across each layer. The boundaries of the layers inside the particle coincide with the combustion stages and are given the same index (b) as shown in Figure 1. The source terms are now placed in the boundaries between the layers, as seen in Figure 3. By doing so, the source terms can be removed from the energy equation (Eq. (4)) for the single layer. Instead they will appear in the boundary conditions, and an additional convective flow is introduced (the third term in Eq. (6)) to
b,j

j q j+1

A)

B)

Figure 3. Section of fuel particle. Centre line dashed-dotted. A) A heat source/sink, q, located in a layer, according to Eq. (4). B) Same as A) but with the heat source/sink located at the boundary b,j. When the boundary moves during conversion of the particle, as indicated by an arrow, mass leaves layer j for layer j+1.

5 represent the solids that leave layer j for layer j+1 when the boundaries move toward the centre of the particle. The energy conservation equation of a single layer j then becomes, b ,4 Rb, k X k ,i k =1 i I , j p + Vp, j t
j 1

cg ,i dT Tb , j 1
Tb , j

4 4 0,k 0, k k= j b,4 Rb , j 1i p , j {Tb , j 1} k = j Rb, j i p , j {Tb, j } 0, j 1 0, j = Vp, j

(6)

(r

2k p , j

b, j

rb , j 1 )

( (T
b, j

b, j

Tp , j ) b , j 1 (Tp , j Tb, j 1 ) Vp, j

Tp,j is the mean temperature of layer j, assumed to be equal to the temperature at the centre of the layer at, (rb,j+rb,j1)/2. The amount of the components is expressed by a mean molar fraction in the particle, , related to the molar concentration C0, of the virgin fuel,

Y / (1 Ym ) Yv Y Y C0 = d m + + c + a Mm Mv Mc Ma

(7)

where is density, Y mass fraction related to initial fuel properties, and M is molar mass. The mass fraction of moisture Y1 related to dry solid, is equal to Ym/(1-Ym). The initial mean molar fraction of components i in the fuel is

0,i =

d Yi / M i
C0

(8)

where subscript i is 1 (moisture), 2 (volatiles), 3 (char) and 4 (ash). The specific enthalpy of solids in a layer at a given temperature inside the particle results from the specific enthalpies of the remaining solid components and the initial mean molar fractions in the layer as,

i p , j = 0,k is ,k {T }
k= j


k= j

0, k

(9)

The volume of a single layer j is obtained from the definition of the shape factor, Eq (1).
rb , j

Vp, j =

rb , j 1

dr

(10)

Eq. (6) is simplified by using enthalpy of each layer instead of enthalpy per unit volume and by expressing the integration over the heat capacity with an enthalpy difference,

6
I p , j
j 1 + b,4 Rb ,k X k ,i ig ,i {Tb , j } ig ,i {Tb , j 1} t k =1 i 4 4 0,k 0,k k= j k= j Rb, j 1i p , j {Tb , j 1} Rb , j i p , j {Tb, j } = b ,4 0, j 0, j 1 2k p , j (T T ) (T T ) ( rb, j rb, j 1 ) b, j b, j p, j b, j 1 p, j b, j 1

(11)

In Eq. (11) the shrinkage of the particle can be easily handled, as only position and area of the boundaries have to be calculated. This is done by means of the volume of each particle layer, which is given by the initial volume of the particle, Vp0, the volume ratio of a layer inside the particle, p,j, and an empirical shrinkage factor of the particle layer, p,j.
V p , j = V p 0 p , j p , j

(12)

The initial volume is defined as,


rb 0,4

Vp0 =

r0

dr

(13)

The volume ratio of a layer inside the particle, p,j, is related to a non-shrinking particle and is expressed by the present and initial molar fractions of the compounds in the solid phase,
j j 1 p, j = 0, j 0, j 1

(14)

The empirical shrinkage factor of the volume element is defined as the ratio of the volume at combustion stage j to the initial volume. The mean temperature in a layer is given by the enthalpy of the layer, I p , j = Vp 0 p , j Cs 0 0,k i p , j {Tp , j }
k= j 4

(15)

Initially the particle consists of moist virgin wood, whose enthalpy is,
I p 0,1 = V p 0Cs 0i p ,1 I p 0,2 = I p 0,3 = I p 0,4 = 0

(16) (17)

For calculation purposes the layers need to have some extension in space, and therefore the minimum volume of a layer is restricted to a volume corresponding to i/0,i=10-4. The boundary conditions of the entire particle are: At the calculation centre (r = r0)

7
T =0 r

(18)

and at the surface of the particle (r = rb,4)


k p ,4 T = qrad + qconv + qcond r

(19)

Subscript conv indicates external convection and cond external conduction. The radiative heat transfer to the surface of the particle is,
qrad = rad (T4 Tb4,4 )

(20)

In the present calculation convective and conductive heat transfer to the surface of the particle are treated together, estimated by a single heat transfer coefficient, h,
qconv + qcond = h (T Tb ,4 )

(21)

The boundary conditions for the individual layers inside the particle, which give the temperatures at the boundaries, are: At the calculation centre (r = r0),
Tb ,0 = Tp ,1

(22)

At the drying boundary (r = rb,1),


2k p ,2 rb ,2 rb ,1

(T

p ,2

Tb ,1 )

2k p ,1 rb ,1

(T

b ,1

Tp ,1 ) = Rb ,1

Hm M s ,m

(23)

At the devolatilisation boundary (r = rb,2),


2k p ,3 rb ,3 rb ,2

(T

p ,3

Tb ,2 )

2k p ,2 rb ,2 rb ,1

(T

b ,2

Tp ,2 ) = Rb ,2

Hv M s ,v

(24)

At the char combustion boundary (r = rb,3),


2k p ,4 rb ,4 rb ,3

(T

p ,4

Tb ,3 )

2k p ,3 rb ,3 rb ,2

(T

b ,3

Tp ,3 ) = Rb ,3

Hc MC

(25)

At the surface of the particle (r = rb,4),


qrad + qconv + qcond 2k p ,4 rb ,4 rb ,3

(T

b ,4

* * Tp ,4 ) = Rb ,4 ( ig , H 2 + 1 ig ,O2 ig , H 2O ) 2

(26)

8 The ash layer is assumed to be inert and fixed on the surface of the particle during the entire conversion process. Despite this assumption there is a reaction rate given in Eq (26). This is the rate of combustion of the volatile gases flowing out, which meet the oxygen transported by diffusion from the bulk gas to the surface of the particle. In the model, the combustion of the volatile gases flowing out is limited to combustion of hydrogen. This formulation of the surface temperature, Tb,4, causes a severe instability in numerical computations, since convection, conduction and especially radiation are very sensitive to the surface temperature. Also heat generation at the boundary of the char layer can cause a stability problem if there is a rapid change in the oxygen concentration. To overcome these problems, the surface temperature is assumed to be equal to the temperature of the ash layer (Tb,4=Tp,4) and the heat sources at the char and ash boundaries are placed inside the ash layer. By doing so, the char boundary condition, Eq (25), becomes,
2k p ,4 rp ,4 rp ,3

(T

p ,4

Tb ,3 )

2k p ,3 rp ,3 rp ,2

(T

b ,3

Tp ,3 ) = 0

(27)

and the ash boundary condition, Eq. (26), disappears. Instead, a source term, qp,4, is added to Eq (11) to represent the ash layer. This source term includes the source terms at the boundaries of the char and ash layers,
* q p ,4 = qrad + qconv + qcond p ,4 R,4 ( ig , H 2 + 1 ig ,O2 ig , H 2O ) p ,3 R,3 p p 2

Hc MC

(28)

Phase change reactions


The phase change reactions are related to vaporization of water, devolatilisation of volatiles and gasification of char. The vaporisation of water at low heating rates is caused by a combination of heat and mass transfer inside the particle. If the heating rate is high, the vaporisation mostly depends on internal heat transfer and the influence of the mass transfer on vaporisation is neglected. At ambient pressure, the saturation temperature for vaporisation is 100C, but inside the particle it becomes somewhat higher due to the pressure increase caused by the volume expansion during gasification of water. No re-condensation of water is considered in the model. The temperature gradient in the particle is steep, and the exact temperature of vaporisation is of little interest. Therefore it is sufficient to know that the water is vaporised close to the saturation temperature. This can be expressed by modelling the rate of vaporisation by an Arrhenius expression with a steep gradient around the saturation temperature. Hence, a rate of vaporisation at ambient pressure has been derived for these conditions
C0V p 0 1 b ,1 Rb ,1 = Cs 0110 27 exp [ 25000 / T ] = b,4 t b,4

(29)

The rate of devolatilisation is expressed by three competing reactions in a one-step mechanism according to [9],

9
A1 = 1.3 108 A2 = 2 108 A3 = 1.110
7

E1 = 16875 E2 = 16009 E3 = 14602

(30)
C0V p 0 2 C0 2 Rb ,2 = = b,4 t b ,4
rb0,2 0 Ai exp [ Ei T ] dr i =1 3

r0

where superscript 0 indicate coordinates for the case of no shrinking. There are a great number of other reaction mechanisms for devolatilisation of wood, given for example in [10], [16], [17], [18]. The choice of the reaction mechanism in [9] is made only to get a reaction rate that releases volatiles at approximately the right temperature level, and if the char yield, (1-Yv), is not known, this reaction mechanism gives approximately the correct level of the char yield, given by the last term in the mechanism multiplied by 0.9. The char yield calculated from [9] is obtained from an iterative solution of
0 rb ,2 = 0.9 A3 exp [ E3 / T ] dr 0 V p 0 r0 t (1 Yv ) = + Ya ; t

(31)

where is the mass fraction of the combustible part of the fuel, related to its initial mass. Devolatilisation is a thermal decomposition of the fuel, and volatiles are released in the entire region between the centre of the particle and the char layer, but due to the steep temperature gradient inside the particle most of the volatiles are released close to the char layer where the temperature is highest. This justifies the assumption that the volatiles are released in a front that moves from the surface to the centre of the particle. To compensate for the volatiles that are released before the arrival of the devolatilisation front, the reaction rate in Eq (30) and (31), is expressed as the rate of devolatilisation integrated from the centre of the particle (r0) to the boundary of the dry fuel layer (rb,2), and divided by the boundary area of this layer. Between the centre of the particle and the boundary of the dry layer, the temperature is assumed to be piecewise linear from the mean temperature of the wet (virgin) layer to the boundary of the wet layer, (rb,1), to the mean temperature of the dry layer and to the boundary of the dry layer. The gases produced during devolatilisation of biofuel consist of CO, CO2, H2O, H2, light hydrocarbons and of small quantities of heavy organic compounds. The composition of the volatile gases can be determined as a result of energy and species balances, [19]. The reactions at the char boundary are modelled as a sum of four reactions

c C+O 2 2 ( c -1) CO+ ( 2- c ) CO 2 C+CO 2 2CO C+H 2 O CO+H 2 C+2H 2 CH 4 whose rate is

I II III IV

10
C0V p 0 3 b ,3 Rb ,3 = ( Rb,I + Rb,II + Rb,III + Rb,IV ) = b ,4 t b ,4

(32)

Char from wood produced under rapid devolatilisation always contains small fractions of oxygen and hydrogen as illustrated by the fuel used in the present experiments, see Table 1. For simplicity the fractions of oxygen and hydrogen are not included in the reaction scheme represented by reactions I to IV. The char is predominantly gasified by a reaction between char and oxygen (reaction I),
Rb , I = (1 ) c CO2 ( r , I ) + d1
1

(33)

where is the fraction of oxygen consumed by volatile gases to be explained below. If the oxygen is consumed, the char can be gasified by carbon dioxide, water vapour and/or hydrogen (reaction II, III, IV),
Rb ,i = Ci ( r ,i ) + d1
1

(34)

where i represents the reaction. The rates of gasification, Eq. (33) and (34), are functions of the rate of the diffusion of the reactant from the bulk into the reactive surface, d, and the reaction rate, at the reactive surface, r, compensated by an efficiency factor, ,

3 / 0,3 3 / 0,3 + c

(35)

The efficiency factor takes into account the extension of the reaction front, c in the particle. In the model c is assumed to be 0.01. The oxidation rate of char by oxygen is [20],

r , I = 1.715T exp [ 9000 / T ]

(36)

A correlation for the formation of carbon monoxide and carbon dioxide, c, is given by the following expression [20] obtained in conjunction with Eq (36), c = 2 (1 + 4.3exp [ 3390 / T ]) 2 + 4.3exp [ 3390 / T ] (37)

The reaction rate chosen for charcoal from biomass, Eq (33) is similar to reaction rates used for biomass char in Russian literature, [21]. Gasification rates for water vapour, carbon dioxide and hydrogen are those of lignite, [3],

r , II = r , III = 103 r , IV = 3.42T exp [ 15600 / T ]

(38)

Even if char from lignite is not a charcoal from wood, the reaction rate is close to that of charcoal, as seen from a comparison between the oxidation rate of charcoal from biomass [20] and char from lignite [3]. There are other reaction rates of the gasification reactions available

11 in the literature, for example in, [22] and [23]. In [23] other proposed rates are validated, by comparing model calculations of single particles with measurements, but the resulting reaction rates are complex, and since the gasification process is of secondary interest here, they are not included. The assumption made here, that also the char conversion takes place in an infinitely narrow region, can be motivated by the high heating rate of the particle during combustion, which creates a steep temperature gradient inside the particle, and also by the fact that the conversion of char is controlled by diffusion of oxygen from the bulk to the surface of the particle. The conversion of char is controlled by external diffusion in nearly all combustion situations in fixed or fluidised beds, where the char is converted by oxygen. On the other hand, modelling of gasification as taking place on reaction surfaces is a great simplification. For some gasification situations the reaction rate totally controls the conversion of the char, and the intrinsic surface area of the fuel particle becomes of most interest. Hence, with the reaction rate expressions selected in the present formulation the model is mainly intended for situations when char is converted by fast oxidation reactions. The consideration that diffusion of the oxidant from the bulk is the rate-determining step of char combustion is also a simplification during simultaneous devolatilisation and vaporisation. Then there is an outflow from the interior of the particle of the oxidants water vapour, carbon dioxide and hydrogen, which are assumed not to gasify the char on its way out of the particle. Furthermore, the oxygen that is transported to the surface of the particle by diffusion not only oxidises the char but also the volatiles, flowing out from the interior of the particle. This is taken into account only for hydrogen, which has a much higher reaction rate than char. The hydrogen flowing out as a part of the volatiles is assumed to consume the oxygen before the char, at the ash boundary
Rb ,4 = 2 CO2 d

(39)

The fraction of the oxygen that is consumed by the hydrogen is ,

1 =

s ,2 Rb ,2 X 2, H 2 ; s ,4 2CO2 d

; 1 1 = 1 1 ; 1 > 1

(40)

Table 1. Properties of fuel used in the measurements


Birch/Char Spruce/Char Ash (% mass, dry) 0.3/2.0(a) 0.4/2.0(a) 3 Density (kg/m ) 54040/42040/(b) H (MJ/kg) 18.4/33.6 18.8/33.2(b) Elemental Analysis(c) C 49.2/93.6 50.2/92.6 O (by difference) 44.3/5.1 43.5/6.4 H 6.3/0.7 6.3/0.8 N 0.16/0.45 0.10/0.23 S <0.01/ <0.01/ (a) (b) (c) calculated, ash free, % mass, dry ash free fuel

12

Measurements
In order to validate the model, experiments were carried out in a laboratory fluidised bed reactor. In this reactor more than 60 wood particles, having sizes as used in utility boilers, were dried and pyrolysed in an inert atmosphere, followed by burnout of the remaining char. The composition of gas components was continuously measured. To a large extent the fuel particles were floating on the surface of the fluidised bed during the thermochemical conversion. Especially larger birch particles fell apart at the end of the devolatilisation phase. The Biot number was above 6 for all particle sizes. Hence, internal heat transfer controls the heating of the particles and the choice of reactor for the experiments does not affect the results. The fuel particles were prepared from raw wood of birch (hardwood) and of spruce (softwood), with material properties as shown in Table 1. Parallelepipeds of eight sizes were tested. The particles were cut with the long side (varied between 10 and 40 mm) along the fibre direction of the wood and with the shortest end (varied between 3 and 10 mm) always in the perpendicular direction. The sizes were chosen to cover the range of specific areas (surface-area to volume ratios), 320-800 m-1, that is expected to occur in a batch of forest waste wood as fired in utility boilers. Each wood-particle size was tested for birch and for spruce, as received (birch, 35 to 45% and spruce, 60 to 65% moisture) and after drying for 24 hours in an oven at 377 K. To simplify the validation of the different combustion stages, drying, devolatilisation and burnout of char, the particles were pyrolysed in an inert atmosphere, followed by burnout of the remaining char in air. Further details on the experimental conditions and set-up are reported in [19].

Table 2. Physical properties, S is initial specific area, b,4/Vp0

Property
k [W/mK] ^ / || kw [W/mK] kc [W/mK] sd [kg/m ] w [kg/m3] c [kg/m3]
3

= a1+a2T

a1
0.52 / 0.73 0.278 1.47 1480 1000 1950 2.0310-10

a2
1.1110-3 1.110-3 2

Dg [m2/s] = a1T ^a2

Property
Y3+Y4 [-] spruce/birch

= a1+a2S

a1
0.245/0.194

a2 (10-3)
-0.094/-0.092

13

Physical data used in the model


The mass and heat transfer coefficients are estimated Appendix A [24], and the mass transfer coefficient is compensated for the out-flowing gases [25]. The thermal conductivity, kp,i, is an averaged value over the surface , where the effective conductivity for a surface element is given perpendicular or parallel to the fibre direction. The thermal conductivities of wood and char are estimated according to Appendix B, [26]. The diffusivity in the gas mixture is given in Table 2, by assuming a binary gas consisting of CO2-O2 at ambient pressure, [27]. The specific enthalpies at reference temperature and the temperature dependent specific heat for gas and solid species are taken from [19], according to which the composition of the volatile gases is calculated. For simplicity, since the ash content is small, the ash is assumed to have the same specific enthalpy and conductivity as the solid part of the char.

Comparison between model and measurement


Measured and calculated devolatilisation times are compared in Figure 4. The devolatilisation time, more or less starting from the time of introduction of the fuel particle into the reactor, is defined to last until 95% of the volatiles are released. The model predicts the time of devolatilisation within 20%, with an overestimation, especially for the larger particle sizes. There are mainly two reasons for the discrepancy: larger particles form cracks or fall apart during devolatilisation which enhances the area for heat transfer from the surrounding and lowers the thermal distance inside the particle; the second reason is the discretisation of the particle in only four layers. The overestimation caused by the discretisation increases with particle size and decreases with a growing difference between the surrounding temperature and the temperature of devolatilisation. The error is small in the outermost position and then grows with the distance from the surface of the particle. This is a result of the assumption of a linear temperature distribution inside the layer, and the theoretical error is largest for spherical particles. However, most of the mass of a spherical particle is located in the outer part of the particle, and this makes the error small for release of mass, but it may be non-negligible for time of mass release. For parallelepipeds and finite cylinders this effect to some extent is compensated for by the shape factor, which causes an underestimation, because the corner effect of the heat transport into the interior of the particle is not exactly represented.
200
Measured time [s]
400

20%
Measured time [s]

150
-20%

300 200 100 0

20% -20%

100

50

100 200 Calculated time [s]

200 Calculated time [s]

400

Figure 4. Comparison between calculated and measured time of devolatilisation for 61 samples of, wet (o) and dry ( ) spruce, and

Figure 5. Comparison between calculated and measured time of char combustion for 61 samples of, wet (o) and dry ( ) spruce,

wet () and dry (+) birch.

and wet () and dry (+) birch.

14
x 10
-3

400

Measured time [s]

Particle radius [m]

4 3 2 1
3 4 1 2

200

200 500

400 Time [s] 700

600

200 400 Calculated time [s]

T[K] 300

900 1100

Figure 6. Times of devolatilisation (o) and char combustion (). Arrows indicate the influence of increasing shrinkage.

Figure 7. Example of simulated temperature and positions of drying (1), devolatilisation (2), char combustion (3) and particle surface (4) inside a wet spruce particle as a function of time.

The largest underestimation caused by the shape factor is for cubic particles and cylinders, having the same length as diameter. As the particle shape approaches an infinite plate or infinite cylinder the error disappears. It is difficult to quantify this error, since it strongly depends on size, particle shape, heating rate and final temperature, but the influence of the discrepancy on the result is not great for the most common particle sizes used in boilers as seen in Figure 4. However, the model is not suitable for very large particles, for example wood logs, devolatalised at low temperature (less than 873K). Figure 5 compares measured and calculated times of final char burnout, defined as the time between the introduction of air into the reactor, or in the calculation, and 95% burnout. Here the scatter is much larger than for devolatilisation, but the agreement is still good. One of the main reasons for the larger scatter during char burnout than during devolatilisation is the accuracy of the correlation for char yield (Table 2), which is plus minus 10%. Other reasons are that some fuel particles fall apart in the end of devolatilisation, especially large birch particles, and finally, the accuracy of the mass transfer coefficient is plus minus 20%. The mass transfer coefficients are related to spheres, and other shapes are accounted for by an equivalent diameter based on the surface area of the particle. This may have some limitations, and can be the reason for the large scatter for the thinnest particles, which thermally more or less can be represented as infinite plates. The only parameter that has a great effect on the conversion, but cannot be satisfactory measured, is the shrinkage of the particles. Although the shrinkage is difficult to measure, attempts were made to do so. The difficulties in measurement of shrinkage during drying are, for example, that the particles were exposed to a much lower final temperature and heating rate than particles in the experiments, in order to separate drying from devolatilisation. Furthermore, measurement of shrinkage during devolatilisation becomes complicated, since the char particles produced are fragile and large cracks are formed inside the particles. Especially large birch particles may fall apart. The measured shrinkage during drying for both wood species was between 10 and 15%. During devolatilisation the measured shrinkage, related to a dry particle, was between 35 and 50% for birch, and for spruce more than 20%. In the calculation the shrinkage was assigned to 10% during drying, and to 39%, (2=0.55), for birch, 28% (2=0.65) for dry spruce, and 39% for wet spruce during devolatilisation. Dry and

15 wet spruce is given different shrinkages, because the time of char combustion is shorter for initially dry particles than for initially wet particles of the same size. Since char combustion is controlled by diffusion of oxygen from the bulk of the surrounding gas to the surface of the particle, the only variable that can cause this difference is a smaller surface area of the particles. The conclusion is that wet particles shrink more than dry particles. In Figure 6 times of devolatilisation and char combustion are calculated for three wet spruce particles of the size of 102540mm. The shrinkage was assumed to be 10% during drying and 6, 28 and 39% (2=0.85, 0.65 and 0.55) during devolatilisation. The figure shows that the time of devolatilisation decreases with increasing shrinkage, because the remaining wood becomes compact, and this results in a higher thermal conductivity and consequently in an increased heat transport inside the particle. In contrast, the time of char burnout increases with increased shrinkage as a result of smaller surface area.
x 10 1
Simulated [kg/s]
-4

x 10 1
Measured (Nornalised)

-4

1
Simulated [kg/s]

1 0.75 0.5 0.25 0 500 1


Measured (Nornalised)
-1

a.
0.75 0.5 0.25 0 0.75 0.5 0.25 0 500 1

a.
0.75 0.5 0.25 0

0 x 10

100
-4

200 300 Time [s]

400

0 x 10

100
-4

200 300 Time [s]

400

Measured (Nornalised)

1
Simulated [kg/s]

0.75 0.5 0.25 0

0.75 0.5 0.25 0 500

Simulated [kg/s]

b.

b.
0.75 0.5 0.25 0 0.75 0.5 0.25 0 500 x 10 1

0 x 10

100
-6

200 300 Time [s]

400

0 x 10

100
-6

x 10 1

-1

200 300 Time [s]

400

Measured (Nornalised)

c.
Simulated [kg/s]

c.
Simulated [kg/s]

6 4 2 0

0.75 0.5 0.25 0

6 4 2 0

0.75 0.5 0.25 0

400

500

600 700 Time [s]

800

400

500

600 700 Time [s]

800

Figure 8. Measured normalised (solid line) and simulated (dashed line) mass release of water, (water in volatiles + moisture, a.), volatiles (b.) and char (c) as a function of time for a spruce particle of the size of 102540 mm, with a moisture content of around 60%.

Figure 9. Measured normalised (solid line) and simulated (dashed line) mass release of water, (water in volatiles + moisture, a.), volatiles (b.) and char (c) as a function of time for a birch particle of the size of 102540 mm, with a moisture content of around 40%.

Measured (Nornalised)

Measured (Nornalised)

16 Figure 7 shows a simulation of a 102540 mm moist spruce particle: temperatures and widths of the four layers change with time, as the reaction fronts move from the surface to the centre of the particle. Figures 8 and 9 give representative examples of comparisons between measurements and simulations for spruce (Figure 8) with a moisture content around 60% and for birch (Figure 9) with a moisture content of around 40%, both of the size of 102540 mm. The measured and simulated release of water, volatiles and burnout of char are shown. The measured concentration of water originates from the moisture in the fuel and to a small extent from the volatile gases. The volatiles are assumed to have a constant composition during the whole devolatilisation, and the measured rate of the devolatilisation is related to the measured CO concentration. The char is predominantly converted to CO2, and the measured release of char is related to the measured CO2 concentration. The releases of water, volatiles and char are observed as gas concentrations, and are not directly measured as mass release. For comparison with the simulated releases, they are therefore normalised to the highest measured concentration of water. However, this value is not exactly represented by the simulation, which results in an almost instantaneous peak of water (Figure 8a, 9a) as the particle enters the reactor, followed by a peak of volatiles (Figure 8b, 9b). The instantaneous peak of water, released from the surface of the particle, was correctly obtained by the simulation (as confirmed by visual observations during measurements), but could not be captured by the gas analyser (the FTIR) because its response time was not sufficiently short. For a fair comparison only the main part of the simulated water release is shown in Figure 8a and 9a. After the initial peaks, water and volatiles decline until all moisture is released. Then, the temperature in the centre of the particle rises rapidly, because the heat sink caused by evaporation disappears. This results in a second peak of water and volatiles, as a consequence of enhancement of devolatilisation. After the second peak the levels of water and volatiles decline once more until the devolatilisation is finished. The trends in release of water and volatiles of both spruce and birch particles are reasonably well described by the model. However, in the case of spruce, the second peak of the measured volatile release is more extended in time than that obtained from the simulations, probably a result of cracks formed in the particle. For the birch the agreement is very good. After 350 seconds air is turned on and the final burnout of the char starts (as seen in Figures 8c, 9c). As the char burns out, the surface rate of the char conversion decreases nearly linearly until complete char burnout, because the reaction surface diminishes when the combustion front approaches the centre of the particle, which can be seen both from measurements and simulations. As already stated above, agreement between measured and simulated times for char burnout is not as good as the predicted times of devolatilisation, as seen from Figure 8 and 9. The measured time of burnout of spruce is much longer than simulated, whereas it is shorter for birch. The much shorter time for the birch, as already mentioned, results from the disintegration of the birch particles. For spruce there is an almost perfect agreement between measured and simulated time for the nearly linear decrease in char combustion. The simulations show a sudden end of the char burnout, whereas measured char burnout ends slowly. This is probably a result of tree-dimensional effects, which cannot be properly modelled with the one-dimensional model.

Conclusions
A simplified thermochemical particle conversion model is derived, for use in a comprehensive model of a fuel bed. The model is independent of size and shape of the particle and operates with a low number of variables. It includes the most essential features of conversion of a solid fuel particle, such as fuel shapes (spheres, finite cylinders and parallelepipeds), temperature

17 gradient inside the particles, shrinkage and swelling. The model treats the particle in one dimension independent of shape, and the conversion of the particle can be described by the heat and mass transport to the surface of the particle. When modelling a large combustion system this is a great advantage, since it extends the validity of the model beyond the treatment of just a single particle. In fact, it can handle the conversion of a solid phase in a computational cell, where the conversion is related to surface area per unit volume instead of surface area of a single particle. The model is validated numerically by a heat and mass balance, and it agrees well with measurements performed on more than 60 samples of particles of different sizes, wood species and moisture contents. A comparison with the experiments shows that the simplifications made do not significantly influence the overall agreement of the model. The model calculations predict the great influence from shrinkage on time of devolatilisation and char combustion in agreement with experiments. Larger shrinkage results in a more compact char layer, and consequently higher thermal conductivity and a shorter time for devolatilisation. On the other hand, for char combustion the trends are the opposite: increased shrinkage means a smaller external surface, as shown by the model.

Acknowledgment
The project was financed by grant from the Swedish National Energy Board

Nomenclature
A Ar B C D E H I M N P Pr R Re S Sc T V X Y b c d h i k l [1/s] [-] [W/m2K] [mole/m3] [m2/s] [K] [J/kg] [J] [kg/mole] [mole] [Pa] [-] [mole/s] [-] [m2/m3] [-] [K] [m3] [-] [-] [-] [J/kgK] [m] [W/m2K] [J/mole] [W/mK] [m] Pre-exponential factor Archimedes number Ar = gd3(s-g)/( g2g ) Heat capacity flow per unit area Molar concentration, related respective phase Molecular diffusion, or particle dispersion Activation energy divided by gas constant Lower heating value Enthalpy Molar mass Mole Pressure Prandtl number Pr = gcgg/kg Global reaction rate Reynolds number Re = ugdeq/g Specific area, b,4/Vp0 Schmidt number Sc = g/Dg Temperature Volume Molar stoichiometric coefficient for species i in reaction j Mass fraction, related to dry fuel properties Reduction factor, compensating for gas outflow Specific heat Particle diameter Heat transfer coefficient gas-particle Specific enthalpy Heat conductivity Initial length of particles

18 q r t u Greek mf rad [W/m2] [m] [s] [m/s] [m2] [-] [-] [-] [m/s] [-] [1/m] [-] [-] [-] [m2/s] [-] Heat flow Particle radius Time Local velocity Shape factor, equal to area at particle radius r Volume fraction Stoichiometric ratio or molar fraction between two species Fraction of oxygen consumed by volatile gases inside particle Mass transfer coefficient, or reaction rate Mass fraction of the combustible part in the fuel, related to its initial mass Delta function Bed voidage at minimum fluidisation Emissivity Efficiency factor Kinematic viscosity Shrinkage coefficient, volume of a finite volume related to its initial volume Density Stefan-Boltzmann constant, 5.6710-8 W/m2K4 Dimensionless length Molar ratio related to initial molar concentration Dimensionless width of the char reaction front Coordinates for the case of no shrinking Integration variable, or if combined with the specific enthalpy i at the surface of the particle Per unit area, cross-sectional area for two-phase system and particle surface area for discrete particles Per unit volume

[kg/m3] [W/m2K4] [-] [-] [-] superscript 0 * // /// subscript A, B O2, H2, a b c conv cond d eq g i in j

Auxiliary variable Oxygen, Hydrogen, Ash Boundary Char Convection Conduction Diffusion, or dry fuel Equivalent Gas Gas species (O2, N2, H2O, ....) Inert particles Fuel component (s,j) (moisture, volatiles, char, ash), layer in particle (p,j) (wet fuel, dry fuel, char residue, ash), combustion stage (b,j) (drying, devolatilization, char combustion, particle surface) Counter

19 m p r rad s sc sd v w 0 Moisture Particle Reaction Radiation Solid solid in fibre wall inside the char solid in fibre wall inside the wood Volatiles Water Initial condition, or starting position

Literature
1. 2. Adams, T. N., Combust. Flame, 39:225 (1980). Purnomo, Aerts, D. J., Ragland, K. W., Twenty-Third Symposium (International) on Combustion, The Combustion Institute, Pittsburgh, 1990, p. 1025. Hobbs, M. L., Radulovic, P. T., Smoot, L. D., AICHE Journal, 38:681 (1992). Bryden, K. M., Ragland, K. W., Energy & Fuel, 10:269 (1996). Thunman, H., Leckner, B., Fuel loading of a fluidised bed combustor, (W. Leuckel, J. Ward, R. Collin, and A Reis Eds.), INFUB, Portugal, 1997. Shin, D., Choi, S., Combust. Flame, 121:167 (2000). Di Blasi, C., Chem. Eng. Sci., 55:2931 (2000). Pyle, D.L., Zaror, C.A., Chem. Eng. Sci., 39:147 (1984). Chan, W-C. R., Kelbon, M., Krieger-Brocket B., Fuel, 64:1505 (1985).

3. 4. 5.

6. 7. 8. 9.

10. Alves, S. S., Figueiredo, J. L., Chem. Eng. Sci., 44:2861 (1989). 11. Melaaen, M.C., Num. Heat Transfer, 29:331 (1996). 12. Di Blasi, C, Chem. Eng. Sci., 51:1121 (1996). 13. Hansson, K-M, Pyrolysis of Large Particles of Biomass Experiments and Modelling, Thesis for the degree of licentiate of engineering, Department of Energy Conversion, Chalmers University of Technology, 2001. 14. Kita, T., Sugiyama, H., Kamiya, H., Horio, M., AIChE Symposium Series No. 303, 92:120 (1996). 15. Saastamoinen, J. J., Richard, J-R., Combust. Flame, 106:288 (1996). 16. Roberts, A. F., Combust. Flame, 14:261 (1970). 17. Kanury, A. M., Combust. Flame, 18:75 (1972). 18. Reina, J., Vole, E., Puigjaner, L., Ind. Eng. Chem. Res., 37:4290 (1998). 19. Thunman, H., Niklasson, F., Johnsson, F., Leckner, B., Composition of volatile gases and thermochemical properties of wood for models of combustion and high temperature gasification, Submitted for publication, March 2001.

20

20. Evans, D. D., Emmons, H. W., Fire Research, 1:57 (1977). 21. Pomerantsev, V. V. Ed., Fundamentals of applied theory of combustion, (In Russian), Energatomizdat Publishing House, Moscow, 1986. 22. Janse, A. M. J., de Jonge, H. G., Prins, W., van Swaaij, P. M., Ind. Chem. Res., 37:3909 (1998). 23. Dasappa, S., Experimental and modeling studies on the gasification of wood-char, Academic dissertation, Department of Mechanical Engineering, Indian Institute of Science, Bangalore India, July 1999. Parts of the work are published in the proceedings of the Twenty-Fifth and Twenty-Seventh Symposium (International) on Combustion, The Combustion Institute, Pittsburgh, 1994, p. 1619 and 1998, p. 1335. 24. Palchonok, G., Heat and Mass Transfer to a Single Particle in Flidized Bed, Academic dissertation, Department of Energy Conversion, Chalmers University of Technology, Gteborg Sweden, ISBN 917197-712, 1998. 25. Bird, R. B., Stewart, W. E. Lightfoot, E. N., Transport Phenomena, John Wiley & Sons Inc., New York, 1960. 26. Thunman, H., Leckner, B., Thermal conductivity of wood during different stages of combustion, Submitted for publication, February 2001. 27. Kanury, A. M., Introduction to Combustion Phenomena, Gordon and Breach Science Publisher, New York, 1977. 28. Incropera, F. P., DeWitt, D. P., Fundamentals of Heat and Mass Transfer, Fourth edition, John Wiley & Sons Inc., New York,1996.

Appendix A
The mass and heat transfer coefficients are estimated from the maximum values obtained in a fluidised bed [24]. In [24] a correlation is developed, which is valid over a large range of particle sizes, based on measurement data and empirical correlations from various authors. The Nusselt number, Nu, and the Sherwood number, Sh, are related to the inert spherical particles in a fluidised bed (index in) and are expressed for two extreme cases: the size of the fuel particles is equal to the inert particle (index A) or they are much larger than the inert particles (index B),
Nu in Nu in , B Nu in , A Nu in , B Sh in Sh in , B Sh in , A Sh in , B d = in d eq
0.67

(41)

d = in d eq

0.67

(42)

where index eq stands for equivalent. The mass and heat transfer coefficients are then calculated as
h = Nu k g / d in

(43) (44)

d,0 = Sh Dg / din

Nusselt and Sherwood numbers for fuel particles of the same size as the inert particles are,

21

0.39 Nu in , A = 6 + 0.117Arin Pr 0.33 0.39 Sh in , A = 2 mf + 0.117Arin Sc0.33

(45) (46)

and for fuel particles much larger than the inert particles
0.19 0.5 Nu in , B = 0.85Arin + 0.006Arin Pr 0.33 0.5 Sh in , B = 0.009Arin Sc0.33

(47) (48)

where Pr is Prandtl number, Ar Archimedes number, mf bed voidage at minimum fluidisation, and Sc Schmidt number. The Archimedes number, Ar, is based on the inert particles as
Arin =
3 gdin ( in g ) 2 g g

(49)

The equivalent diameter of the fuel particles deq, is given by the surface area of the particle, b,a, d eq = ( b, a / )
1/ 2

(50)

The mass transfer coefficients are compensated by a reduction factor, bd, for gas flowing out from the particles during drying and devolatilisation. No corresponding reduction of the heat transfer coefficient is made, as the major part of the heat transfer depends on particle convection (and radiation). However, the heat transfer coefficient is reduced by a factor because the particles tend to float on the surface of the bed during the experiments. In [24] a reduction factor for this combustion situation of 0.85 is recommended. The reduction of the mass transfer coefficient is derived for spheres, infinite cylinders and plates, according to [25]. Here, it is assumed that this reduction factor can be used also for finite cylinders and parallelepipeds. Thus, bd = ug = ug / d 0 d = d 0 exp ( ug / d 0 ) 1 (51)

8.314 (Ts + Tg ) Rb,k X k ,i P 2 k i

(52)

where ug is the velocity of the gas flowing out from the surface of the particle. The physical properties, thermal conductivity, k, and kinematic viscosity, , of the gas are assumed to be the same as for air. Empirical correlations, which are based on table values from [28], are given in Table 2.

Appendix B
The thermal conductivity, perpendicular and parallel to the fibre direction, is calculated according to [26], from the conductivities and densities of dry fuel, liquid water, fibre wall,

22 and the solid part of the char, using information regarding composition of the fuel, conversion stage and shrinkage. Knowing these properties, the effective thermal conductivity can be calculated from an equivalent fibre structure in the wood and the char. Quadratic tubes with normalised wall thickness and a moisture layer represent the fibres. The normalised thicknesses of wall fibre, s, moisture layer, w, and gas layer are defined as,

s ,1 = 1 1 ( p ,2 / p ,1 ) ( d / 2 )

0.5

4 s , j =1 1 ( p ,2 / p , j )( d / j ) Yn n= j

0.5

for

{ j > 1}

(53)

w,1 = (1 s ,1 ) (1 s ,1 ) ( p ,2 / p ,1 ) ( d / 2 ) Y1
2

w, j =0 for { j > 1} g , j = 1 s , j w, j

0.5

(54) (55)

where the second index on the normalized thickness indicates the combustion stage (virgin fuel, 1, dry fuel, 2, char, 3 and ash, 4), index d on the density, , indicates dry fuel and index 1 moisture, 2 dry wood fibre, 3 char and 4 ash. The effective thermal conductivity perpendicular to the fibre direction is calculated from the minimum effective thermal conductivity for each conversion stage derived for an equivalent fibre,
s , j (1 s , j ) g, j + g , j s , j + w, j + + keff , j , A1 = s , j ks , j + w, j ks, j ks , j kw kw ( k g + krad ) and the maximum effective thermal conductivity
1 1

(56)

s, j w, j g, j + + keff , j , B1 = ks , j s , j k s , j + (1 s , j ) kw s , j ks , j + w, j kw + g , j ( k g + krad )

(57)

The minimum Eq (56) and maximum Eq (57) thermal conductivities, (n=1), are combined in serial and parallel paths, keff , j , An +1 = 1 keff , j , An + 1 keff , j , Bn 2 2 keff , j , Bn +1 = ( 1 / keff , j , An + 1 / keff , j , Bn ) 2 2
1

(58)

which gives the effective thermal conductivity at n=2 after combination of those of n=1, where n is a counter. The values obtained by Eq (58) over- and underestimate, respectively, the serial and parallel conduction paths. As n increases, the difference between the two conductivities decreases. When n approaches infinity the two values become equal and a resulting effective thermal conductivity is obtained.
n ; keff , j = keff , j , A = keff , j , B

(59)

Along the fibre the thermal conductivity is simpler to estimate, and pipes placed next to each other represent the fibres. The effective thermal conductivity, can be modelled by a parallel path

Paper VI

Composition of volatile gases and thermo-chemical properties of wood for modeling of fixed or fluidized beds
Henrik Thunman, Fredrik Niklasson, Filip Johnsson, Bo Leckner* Department of Energy Conversion Chalmers University of Technology S-412 96 Gteborg, Sweden +46 31 7721430, +46 31 7723592 (fax) email energy.conversion@entek.chalmers.se (Submitted for publication)

Abstract
Modeling of conversion of solid fuel in combustion or gasification systems needs a description of the composition of the volatile gases that leave a fuel particle of typical size in a fixed or fluidized bed during devolatilisation. Much work has been published on releases of volatiles, but a general model of the composition of volatile gases and a comprehensive presentation of the related thermo-chemical properties of the fuel are still missing. Here, a simplified model is presented whose structure is valid for any solid fuel. The model consists of a heat and mass balance complemented by empirical data. The empirical coefficients have to be specified for certain classes of fuel: the fuels treated in this paper are one type of hardwood and one of softwood, having particle sizes in the range used in utility boilers. The species in the volatile gases are represented by the time mean mass fractions of CO2, CO, H2O, H2, light hydrocarbons, and heavy hydrocarbons. In addition, a comprehensive set of data is presented for such properties of wood that are needed for modeling of conversion in fixed and fluidized beds. The data are taken from the literature and from measurements carried out in the present work.

Introduction
Models of solid fuel combustion devices, such as fixed and fluidized beds, tend to be extremely complex and time consuming. Therefore simplifications have to be found. The present model aims at characterizing composition and quantities of the volatile gases leaving the surface of non-isothermal fuel particles by means of a relevant number of gas species. Relevant number of species means species for the thermal process and for related computations. These components are readily outlined according to experience: CO, CO2, H2, H2O, light (non-condensable) hydrocarbons, and remaining hydrocarbons, here called lumped hydrocarbons. The problem to be solved then consists in the determination of the quantities of these gases. Characterization of the volatile gases is most essential for fuels having a high volatile content, since in this case an error made regarding the composition may have a significant impact. A characterization can be formulated in a general way, but it needs some empirical correlations and fuel data. As an example, data for trunk wood will be employed here and the empirical parts of the model will be specialized for this fuel. Similar investigations could produce the empirical input for other fuels. The volatile gases from the moisture-free part of wood have already been represented in various ways in models. For example, they are assumed to consist of a single equivalent gas 1, by a number of gases given by measurements 2, 3, 4 (the composition in 2, 3 was estimated from measurements 5). There are numerous publications describing the rate of devolatilization, the char yield and the composition of volatile gases 6, 7, 8, 9, 10, 11, 12, 13, 14 15, 16. Most of the works
*

Corresponding author

2 cited focus on devolatilisation, where the fuel is isothermally heated to a rather low final temperature, lower than the temperature in a combustor. The experiments have a main focus on the production of charcoals and/or pyrolysis oils. In contrast, here the interest is on combustion-like conditions, which means that devolatilization is controlled by internal heat transfer and the medium surrounding the particle has a higher temperature than that needed for devolatilisation. Besides, if the particles are thermally thick, similar to typical fuels, and the devolatilization takes place in a temperature gradient inside the particle. The volatile gases may be transformed on their way out of the fuel particle. Due to measurement complications it is difficult to close the elemental species balance, and attempts have only been made in some work. Efforts in closing the energy balance have been made in even fewer cases, probably due to the difficulty in closing the species balance. There are exceptions, e.g. 10 where the closure of both the elemental species balance and the energy balance has been tried. However, the closure was not complete. When modeling a combustion or a gasification system the balance of the elemental species and the energy must be fulfilled in order to satisfy the conservation equations that describe the system. A closure model has to be established that is valid during all conversion stages of the fuel. For this propose it is convenient to express the energy conservation equation in terms of specific enthalpy. By doing so, the heat of reactions at temperatures that differ from the reference temperature, as well as the energy balance over the system, are easy to obtain. This is what is done in present work, where heat and mass balances are used to formulate a model for evaluating the quantities of the gases leaving a devolatilizing fuel particle. The model requires thermochemical data of the fuel (virgin fuel, char, tar, etc) as input. These data are collected from literature sources and completed for wood particles of the sizes used in utility boilers by results from measurements carried out in the present work.

Theory
The specific enthalpies of the fuel particle and of the gases leaving the particle, as well as the composition of the gas, are needed in each step of the conversion to conserve the energy according to the first law of thermodynamics (applied to chemical reacting systems). Wood, used as fuel, usually has high moisture content, and consequently a major part of the gases released from the wood is evaporated moisture. The other major part consists of volatiles released during devolatilization. Here, the main focus is on the gases released during devolatilisation, and all considerations below concern the moisture-free part of the wood. Wood consists mainly of carbon, hydrogen and oxygen. Other species can be neglected when the global system is considered. The specific enthalpy, h, of the wood at a reference temperature, superscript 0, can be calculated from the lower heating value, H, of the moisturefree wood and the elemental analysis, given as the mass fraction, X, of respective species and related to the moisture and ash-free wood as, X X H 2 , wood 0 0 0 hwood = H wood + C , wood M CO2 hCO2 + M H 2O hH 2O (1 Yash ) MC M H2 X X H 2 , wood X O2 , wood 0 C , wood + M O2 hO2 (1 Yash ) MC 2M H 2 M O2

(1)

M is molar mass and Y denotes mass fraction of volatiles, char and ash in the moisture-free wood. The lower heating value of the dry wood can be divided into two parts; one belonging

3 to the char and one to the volatiles including the heat of devolatilization (ash is considered inert), as
0 H wood = Ychar H char + (Yvol H vol H dev )

(2)

Under combustion-like conditions the ash-free content of char consists of nearly pure carbon. The specific heat, cp, of char is therefore assumed to be equal to that of graphite (given in Table 4). This is a common assumption 4, 17. To further simplify the calculations, since the ash content is small, the ash is assumed to have the same specific heat and specific enthalpy as the char. With these assumptions the specific enthalpy of the char at temperature T becomes,
0 0 0 hchar = hchar + c p ,C dT = hchar hC + hC Tref T

(3)

where the specific enthalpy at the reference temperature can be calculated from the lower heating value and the elemental analysis of the char, X X H 2 ,char X O2 ,char Ychar 0 0 H char = hchar + C ,char + + M O2 hO2 MC M O2 Ychar + Yash 2M H 2 X Ychar X H 2 ,char 0 0 M H 2O hH 2O + C ,char M CO2 hCO2 MC Ychar + Yash M H2

(4)

A common assumption, if the lower heating value and elemental analysis of char are not known, is that the ash-free char consists of pure carbon (that is, XH2,char and XO2,char0). Then the specific enthalpy of char becomes equal to that of graphite, and the heating value of char can be calculated from Eq. (4). The heat of devolatilization is defined by,
0 0 0 0 H dev = (Ychar + Yash ) hchar + Yvol hvol hwood

(5)

but the literature values are usually given at the devolatilization temperature. The major part of the devolatilization occurs in the temperature range 700 to 900K 10, and the heat of devolatilization reported in the literature has to be recalculated to the reference temperature as,
0 H dev = H dev Tdev

Tref

(c

p , wood

(Ychar + Yash ) c p ,char c p ,vol ) dT

(6)

The specific heat of wood has been measured. The heating value, specific enthalpy at the reference temperature and the specific heat of the volatile gases are estimated from the composition of the volatile gases as,

H vol = i H i
0 hvol = i hi0 i i

(7) (8)

c p ,vol = i c p ,i
i

(9)

where is the mass fraction of species i in the volatile gases. The volatile gases was assumed to consist of CO, CO2, H2O, H2 CiHj, and CnHmOk, and,

CO + CO + C H + C H
2 i j n

m Ok

+ H 2 + H 2O = 1

(10)

The light hydrocarbons, CiHj, mostly consist of methane and ethylene 10, 15, 16, 18. All hydrocarbons except methane and ethylene are lumped together and referred to as lumped hydrocarbons in the following. The elemental composition, lower heating value and specific heat of the lumped hydrocarbons, CnHmOk, can be roughly estimated from previous knowledge 19. The specific enthalpy of the lumped hydrocarbons at reference temperature can be estimated from the heating value,
0 0 hCn H mOk = H Cn H mOk + nM CO2 hCO2 +

m m k 0 0 M H 2O hH 2O n + M O2 hO2 2 4 2

(11)

There are five more unknowns in the system of equations formed by Eq. (1) to (11) than there are equations. The balances of the elemental species give three additional equations,
X C , wood (1 Yash ) = Yvol X C ,vol + Ychar X C ,char CH CO2 C H O = Yvol CO + + i j +n n m k M CO M CO M C H M Cn H mOk i j 2 X H 2 , wood (1 Yash ) = Yvol X H 2 ,vol + Ychar X H 2 ,char M C + Ychar X C ,char M H 2 + Ychar X H 2 ,char

(12)

CH H H O m Cn H mOk = Yvol 2 i j + + 2 + 2 MC H 2 M Cn H mOk M H 2 M H 2O i j X O2 , wood (1 Yash ) = Yvol X O2 ,vol + Ychar X O2 ,char

(13)

1 CO CO2 k Cn H mOk 1 H 2O = Yvol + + + 2 M CO M CO 2 M C H O 2 M H O 2 2 n m k

M O2 + Ychar X O2 ,char

(14)

The final two equations needed are empirical correlations based on the most commonly measured species in combustion or gasification systems, carbon monoxide, carbon dioxide and total hydrocarbons. The correlations are expressed as the ratios of carbon monoxide to carbon dioxide and light hydrocarbons (obtained from the measurements of the total hydrocarbons) to carbon dioxide,
1 = CO / CO2

(15) (16)

2 = Ci H j / CO2

The balances of the elemental species, Eq. (12) to (14) the conservation of energy, Eq. (6) inserted in Eq. (2)and put equal to Eq. (7), together with the empirical ratios, Eq. (15) and (16), form a system of equations that gives the composition of the volatile gases . The system of equations can be solved by a matrix operation as,

0 H 2O 0 H 2 1/ M H O 1/ M H CO 1 / M 0 H O = 2 CO2 H H O CH 0 i j 0 Cn H mOk 0 0
2 2 2 2

1/ M CO
2

1/ M CO 0
1 2

i / MC H
i

n / MC H
n

m Ok

0
1 2

j 2

/ MC H
i

m 2 k 2

/ MC H
n

m Ok

/ M CO

/ M CO

/ MC H
n

m Ok

CO
1 0

CO 2
1 2

C H
i

C H
n

m Ok

0 1

0 0

X C ,vol / M C X H ,vol / M H X O ,vol / M O 0 0


2 2

(17)

where and are equivalent heating values. is related to the lower heating value of the species in the volatile gases and is defined,

i = Hi +

Tdev

Tref

c p ,i dT

(18)

is related to the lower heating value of the volatile gases and is defined,

= i i = H wood Ychar H char + H dev (Ychar + Yash ) c p ,char c p , wood dT / Yvol T


Tdev
ref

(19)

Measurements
In order to get the wanted properties of wood particles having sizes as used in utility boilers, single wood particles of these sizes were pyrolysed in an inert atmosphere, followed by a burnout of the remaining char. The composition of gas components was continuously measured during the experiment. The fuel particles to be analyzed were prepared from raw wood of birch (hardwood) and of spruce (softwood) with material properties as shown in Table 1. Parallelepipeds of eight sizes were tested (the total number of samples was 57). The particles were cut with the long side, varied between 10 and 40 mm, along the fiber direction of the wood and with the shortest end, varied between 3 and 10 mm, always in the perpendicular direction. The sizes were chosen to cover the range of specific areas (surface-area to volume ratios), 320-800 m-1, that is expected to occur in a batch of forest waste wood as fired in utility boilers. Each wood-particle size was tested for birch and for spruce, as received (wet) and after drying for 24 hours in an oven at 377 K.

The experiments were performed in a laboratory-scale fluidized bed reactor, as outlined in Figure 1. The reactor consists of a vertical quartz tube with a height of 1400 mm and an inner diameter of 64 mm. The quarts glass tube extends 50 mm on each side of the insulated part of the reactor. The fluidized bed material (silica sand) rests on a distributor plate made of a porous quarts filter, placed nearly at half height of the quartz glass tube. Three separate electrical heaters control the reactor temperature, dividing the reactor into three zones, as shown in Figure 1. The entering fluidizing gas flow, controlled by mass flow regulators, is preheated in Zone I before passing the gas-distributor plate to the fluidized bed in Zone II and the freeboard in Zone III. The reactor is designed for a maximum temperature of 1473 K.

Table 1. Properties of fuel used in the measurements


Birch/Char Spruce/Char Ash (% mass, dry) 0.3/2.0(a) 0.4/2.0(a) Density (kg/m3) 54040/42040/H (MJ/kg) 18.4/33.6(b) 18.8/33.2(b) Elemental Analysis(c) C 49.2/93.6 50.2/92.6 O (by difference) 44.3/5.1 43.5/6.4 H 6.3/0.7 6.3/0.8 N 0.16/0.45 0.10/0.23 S <0.01/ <0.01/ (a) calculated, (b) ash free, (c) % mass, dry ash free fuel

Table 2. Operating conditions of the fluidized bed reactor


Parameter Reactor Temperature Density of bed material Mean diameter of bed material Bed height Superficial gas velocity Residence time of gas in reactor (bed to exit) Operating Condition 1123K 2600 kg/m3 0.30 mm 80 mm 0.60 m/s 1.1 s At rest At 1123 K At 1123 K Remark Constant along reactor Silica sand

The top and the bottom of the glass tube reactor are water-cooled to prevent the gas inlet and outlet connections to become overheated and to reduce the temperature of the exiting gas to around 460 K. The fuel pieces were dropped down onto the fluidized bed from the top of the reactor tube. During the heating of the particle, including drying and pyrolysis, the fluidizing gas in the reactor was pure nitrogen, allowing the study of the composition of the pyrolysis gas. When the pyrolysis had finished, the fluidizing gas was changed to air in order to burn out the char residue in the bed. The mass flow controllers used allow any ratio of oxygen to nitrogen during combustion of the test samples. The operating conditions are listed in Table 2. For the condition in the experiments the Biot number was above 6 for all particle sizes, which means that the heating of the particles is controlled by internal heat transfer and the choice of reactor for the experiments does not affect the results.

The experimental set-up, including the gas-analysis system, is shown in Figure 2. The gasanalyzers are listed in Table 3. Calibration of the analyzers was performed daily before the experiments. The residence time of the pyrolysis gases in the hot reactor was around 1 second. The gas was sucked out close to the reactor outlet, where two parallel fiberglass filters, kept at 423K, filtered the gas. About 4 liters per minute of the sample gas was lead to an FTIR instrument by heated tubing, passing a second filter and a heated pump into the test cell of the instrument. Another part of the sample gas was lead by heated tubes to a total hydrocarbon instrument, measuring the hydrocarbons as methane equivalents by a flame ionization detector (FID). The third line of sample gas was lead to a dryer and a pump before being distributed between two CO/CO2 instruments and one O2 instrument. One of the CO/CO2 instruments was fed by sample gas diluted with nitrogen to avoid that the gas concentration of CO out from the reactor exceeded the instruments maximum. The dilution factor was estimated by comparison with CO levels measured below maximum concentration. The levels of CO2 were then used to verify the dilution factor. CO was preferred for the estimation of the dilution factor, because the concentration of the diluted CO2 occasionally exceeded the lower range of the CO2 analyzer.

7 The integrated gas flow of each species is calculated to estimate the time average ratios of the different gaseous species during devolatilization. The flow of the single species is estimated from the measured concentration and the total gas leaving the reactor. The total gas is the sum of the gas flow of inert gas into the reactor, given by mass flow regulators, and the gas produced during drying and devolatilization. The gas produced in the reactor is estimated from the mass leaving the fuel particle and is correlated to the concentrations of the measured species. The fuel particles were to a large extent floating on the surface of the fluidized bed, and especially larger birch particles fell apart at the end of devolatilisation phase. There was a small amount of soot and tar on the outlet tube walls, but no condensed tar was observed to fall back into the reactor. No tar, but soot and small sand particles could be seen on the filters.

Property data for modeling


The empirical ratios of 1 = CO/CO2 and 2 = CiHj/CO2 in the volatile gases are estimated from the experiments and from literature data 3, 15, 16. Literature data 3 give a ratio of 1 between 0.4 and 0.95 at 1092 K, and a temperature dependent ratio 16 expressed as 1 = 1.94 106 T 1.87 which, in the temperature range of 665 to 990 K, is between 0.37 and 0.78. The present measurements, Figure 3, show a ratio that increases linearly from around 2.4 to 3.4, as the specific area of the fuel particle increases from 320 to 800. The difference is small between the tested hardwood and softwood or between dry and wet particles. The high ratio, around 3, of CO to CO2 agrees with what has been found for rapid pyrolysis (1000K/s) of small hardwood particles, (45-88m) 10. In the literature the ratio of the light hydrocarbons to CO2 is stated to be between 0.29 and 0.46 3. A temperature dependent ratio has been expressed by 16, 2 = 1.305 1011T 3.39 + 3.007 1014 T 4.07 (21) (20)

in which the first term represents the ratio of CH4 to CO2, and the second term the ratio of C2H4+C2H6 to CO2. This ratio of 2 is between 0.06 and 0.24 in the measured temperature range of 665 to 990 K. The present measurements give a ratio of THC to CO2 between 1.2 and 1.7 (the THC measured on a gas cooled to 423K), depending on the specific area of the fuel, as illustrated in Figure 3. The ratio increases linearly with the specific area, for small specific areas, but for larger the ratio tends to become constant. No significant difference between wet or dry fuel or between hardwood and softwood can be seen. THC is expressed as methane equivalents. The THC analyzer (FID-type) counts the carbon atoms in the hydrocarbons, related to the calibration gas. The actual concentration of the different hydrocarbons is calculated from effective carbon numbers. The number is unity for methane, 0.95 for ethylene and ethane, and 0.89 for benzene (information in the manual of the FID analyser). Assuming that the effective carbon number is the same for the lumped hydrocarbons as for benzene, the ratio of THC to CO2 can be expressed as,

(1) 50 (4) (3) III II 1400 (6) I 650 50 (2) III II (6) I

Stack

Nitrogen for dilution Filter Filter MFC


gas dryer

Fluidised Bed Reactor

CO/ CO2 CO/ CO2 Stack O2

(5)

Heated Tube

Heated Tube

THC FTIR

MFC

MFC

65

Figure 1. Outline of the reactor, (1) gas outlet, (2) gas inlet, (3) insulated casing, (4) quartz glass reactor tube, (5) distributor plate, (6) heaters in different zones I, II, III.

Nitrogen or air Oxygen

Filter

Figure 2. Experimental set-up and gasanalysis system, MFC stands for mass flow controller.

1.8
THC/CO2 (by mass)

4.0
CO/CO2 (by mass)

1.6 1.4 1.2 1.0 0.8 300 1.8

3.5 3.0 2.5 2.0 300

Table 4. Summary of thermo-chemical data for wood


Trunk wood Hwood (MJ/kg) CHiOj; I J cp,wood Range 17-1-20.8 (a) 1.32-1.61 (a) 0.56-0.64 Eq. (24) to (25) Reference
20 20 20 21, 22, 23, 24

600 900 Specific Area [1/m]

600 900 Specific Area [1/m]

H2O/CO2 (by mass)

Char Yield [%]

1.6 1.4 1.2 1.0 0.8 300 600 900 Specific Area [1/m]

22 18 14 10 300 600 900 Specific Area [1/m]

Figure 3. Ratios of THC to CO2, CO to CO2 and H2O to CO2, and char yield, related to the specific area of the wood particles, birch (wet, +, dry, *, solid trend lines), spruce (wet, o, dry, , dashed trend line). Table 3. Gas analyzers used
Analysed Gas Instrument CO CO2 THC O2 H2O Rosemount, BINOS 100 J.U.M. Engineering, FID 3-300A Leybold-Heraeus AG, Oxynos-1 FTIR Bomem MB 100 Range 0 3.0 % 0 20 % 0 10% (CH4) 0 25 %

Char Range Reference 33 HChar (MJ/kg) Eq. (26) (b) (Birch Ychar=14%) 32.9 (b) (Spruce Ychar=20%) 32.5 (b) CHiOj; i 0.07-0.17 (b) J 0.004-0.055 4, 17 cp,char (kJ/kgK) Graphite, Table 5 Devolatisation Range Reference 26 Hdev (MJ/kg) >623 K 0.20 to 0.25 Light hydrocarbons Range Reference (a) 37 HCiHj (MJ/kg) 49.2 - 49.4 (b) 2-3 CH4 /C2H4 CiHj (b) cp,CiHj (kJ/kgK) CH4, Table 5 Lumped hydrocarbons Range Reference (b) HCnHmOk (MJ/kg) 26-40 12, 3 CnHmOk; n 6 2 3 m 6.2 or 8 or 2 3 k 0.2 or 1 or cp,CnHmOk 38, (b) C6H6, Table 5 (kJ/kgK) (a) (b) calculated from ref., obtained in present work

9
* THC X CH 4 ,vol + 1.90 X C2 H 4 ,vol M CH 4 / M C2 H 4 0.89n X Cn H mOk ,vol M CH 4 / M Cn H mOk + CO2 X CO2 ,vol X CO2 ,vol

(22)

where the non-condensable part of the lumped hydrocarbons at the temperature of the gas entering the analyzer is indicated with an asterisk. As stated above the light hydrocarbons in the ratio 2 are mainly methane and ethylene. This defines the ratio 2 as,
2 = X CH 4 + X C2 H 4 / X CO2

(23)

According to the definition Eq. (22) the ratio of THC to CO2 is always larger than the ratio of the light hydrocarbons to CO2, Eq. (23). To obtain the ratio of the light hydrocarbons to CO2 the mass fractions of methane and ethylene must be approximated from the ratio of THC to CO2, by Eq. (22). This is done by estimation of the ratios of methane to ethylene, and methane and ethylene to the non-condensable part lumped hydrocarbons in the THC. The mass ratio of methane to ethylene is estimated to be between 2 and 3, 10, 15, 16, 18. On a molar basis this corresponds to 78-84% methane. (If the mass ratio is 3, an equivalent hydrocarbon takes the form, C1.16H4). The ratio of methane and ethylene to lumped hydrocarbons in the measured THC is estimated to be between 4 and 9. This ratio is based a qualitative judgment of the concentrations of the non-condensable part of the lumped hydrocarbons at the temperature of the gas entering the THC-analyzer given by the FTIR-spectra that indicates that the concentration of these lumped hydrocarbons is much smaller than the one of methane and ethylene. With these estimates of the ratios of methane to ethylene and of methane and ethylene to lumped hydrocarbons in the THC, the ratio of the light hydrocarbons to CO2, can be approximated to be between 0.9 and 1.2 for the smallest specific areas and between 1.1 and 1.5 for the largest. The specific heat of the light hydrocarbons is assumed to be close to that of methane, given in Table 5. The ratios of CO to CO2 and of light hydrocarbons to CO2 obtained in the present measurements are much higher than those of the other results quoted. To get the same ratio of light hydrocarbons to CO2 as in the other quoted results, a major part of the hydrocarbons in the THC must be related to the lumped hydrocarbons. This can, however, be excluded judging from the qualitative measure given by the FTIR-spectra. The most likely reason for the discrepancies is the difference in residence time in the heated part of the reactor. In the measurements presented here, the residence time was around one second, whereas for the data 16 the residence time in the reactor was around six seconds. The residence time in the other set of measurements 3 cannot be estimated. Rapid pyrolysis of small particles 10, where the composition of the volatile gases was measured after a short residence time, shows on a ratio of CO to CO2 in the same region as the one reported here. For measurements where the light hydrocarbons is estimated by a FID-analyzer, in a system without FTIR, it would be better to cool the gases to ambient temperature before they enter the THC analyzer. In this way as much hydrocarbons as possible is condensed, and thereby a more defined composition of the THC is obtained. If a FTIR is used, and the major components are measured from the FTIR-spectra, then it is an advantage to measure the THC at the highest possible temperature, because then an additional control of the mass fractions given by Eq. (17) is achieved. This control involves the ratio of the THC to CO2 given by Eq. (22), using mass fraction of light and lumped hydrocarbons from Eq. (17), that must be larger than or equal to the measured ratio of the THC to CO2.

10 A ratio of the water vapor to carbon dioxide concentrations in the pyrolysis gases can be obtained from the FTIR measurements of water vapor formed during devolatilization of the dry wood particles. This ratio varied between 0.8 and 1.3, see Figure 3. The reliability of the measurements decreases with a decreasing particle size, due to the relatively slow exchange of gas in the measurement cell. The FTIR provided one measurement point every 7th second and, as the time of devolatilization for the smallest particles is around 30 seconds, only 4 to 5 measurement points were obtained for these particles. However, the ratio was rather constant, around 1.2 for spruce and 1.0 for birch. The property data of most interest are heating values, structure of the compounds CHiOj, and specific heats. These data are collected both from literature sources and from the present measurements. The results are summarized in Table 4. A database 20 of 25 types of wood allows determination of the heating value, Hwood and estimation of i and j in CHiOj, Table 4. For a particular kind of wood the range of variation of the heating value and the elemental composition is smaller than those given in Table 4. The specific heat of wood is temperature dependent and is represented by empirical correlations from the literature, e.g. for dry wood (i) 21, (ii) 22 (a mean value of six reported correlations), (iii) 23 (taken from 17), (iv) 24, c p , wood = 4.206T 37.7 c p , wood = 4.607T 132.8 c p , wood = 3.867T + 103.1 c p , wood = 2.45T + 531.2 (i ) (ii ) (iii ) (iv) (24)

The first three correlations (i) to (iii) are close to each other, whereas correlation (iv), results in same specific heat at ambient temperature, but has a temperature dependence which is about half of that of the others. Correlation (iv) is interesting, since data forming this correlation show a linear temperature dependence of the specific heat until devolatilization starts at 553K, whereas data forming the other correlation only do not exceed 450 K. Small differences between different woods can be expected, as they consist of different amounts of cellulose, hemicellulose and lignin. The heat capacity of pure cellulose is similar to the specific heat of wood, according to the three first correlations in Eq. (24), but has slightly higher temperature dependence 25. Moist wood has a greater specific heat than what would be expected from the simple law of mixtures 22, 23, as a result of the energy absorbed by the wood-water bounds. This is represented by a correction term 23,

c p , wet = ( c p , wood + 4190Ymoist / (1 Ymoist ) ) / (1 + Ymoist / (1 Ymoist ) ) + A A = ( 23.55T 1320Ymoist / (1 Ymoist ) 6191) Ymoist / (1 Ymoist )

(25)

Under combustion conditions the char yield becomes 10-25% of initially dry mass 9, 14, 26, 27, 28. However, flash pyrolysis of small particles can give a char yield below 10% for temperatures above 800 K 10, 29. In the present measurements the char yield was estimated from of CO2 concentration, elemental carbon content in char, and airflow through the reactor during the burnout of the char. For large, dry birch particles, with a specific area of 320, the char yield is around 16%, Figure 3, which is similar to a value reported in a previous investigation 30, where a char yield of around 14% for birch is reported at the same temperature as in the

11 present measurements for particles of a size of 10x10x70mm. The char yield declines nearly linearly with particle size, and for birch particles with a specific area of 800 it becomes around 11%. For wet, large particles the char yield is somewhat higher, but for smaller particles the difference disappears. The same trends are observed for birch and spruce, but the spruce shows 30% higher char yield than the birch. According to the elemental analysis of the present measurements, Table 1, and from the literature 12, 28, the content of carbon in the ashfree char is higher than 90% on dry mass, hydrogen is in the range of 0.6-1.5% and oxygen, determined by difference, in the range of 0.5-6.4%. This yields an approximate elemental composition given in Table 4. Empirical correlations have been published 31, 32, 33 for heating values of char related to the amount of fuel released as volatiles. The difference between the correlations is small 34. One of the correlations 33, for the range of the conversion (1-Yvol) between 1 and 0.17, is H char = 16.7 106 + 2.93 106 /(1 Yvol ) ; (26) (1 Yvol ) > 0.17 For a char yield of 20%, (1-Yvol)=0.2, the resulting heating value becomes 31.3 MJ/kg. If (1-Yvol)0.17 the char is nearly pure carbon, and the heating value of the char is assumed to be the same as for (1-Yvol)=0.17, which gives 34.0 MJ/kg. These data can be compared with those in Table 4. The composition of the lumped hydrocarbons is difficult to generalize, as it depends strongly on temperature and residence time 19. Here, the primary product of devolatilisation inside the particle is assumed to have a composition close to pyrolysis oils, produced in the temperature range of 700 to 900 K, where most of the devolatilization occurs under combustion-like conditions. These pyrolysis oils have an elemental composition close to that of wood 10, 35, and a heating value between 22 and 26 MJ/kg 13, 36. As the gas passes out through the particle, it meets higher temperature and active char surfaces, which crack the pyrolysis oils. The remaining lumped hydrocarbons that leaves the surface of the fuel particle contain more stable hydrocarbons than they did originally, for example, benzene, naphthalene and toluene 19. These hydrocarbons have a heating value around 40MJ/kg 37. The conclusion is that lumped hydrocarbons leaving the fuel particle, produced during combustion-like conditions, have a heating value between 26 and 40MJ/kg. In Table 4 the composition is assumed to be roughly the same as the molecule of tar. Judging from the different hydrocarbons in the lumped hydrocarbons in Table 4, the heating value is most likely higher than 40 MJ/kg for one proposed composition 2, and little less than 40 MJ/kg for another one 3.
Table 5. Specific heat, cp,i=a1+a2T+a3T2+ a4T3+ a5T4, and enthalpy for gas species and graphite

Species
(a) (a)

h0
0 0 -13.4106 0 -3.95106 -8.94106 -3.05106 Eq. (11) Eq. (4)

a1
811 939 1612 14400 982 508 1086 -9.31 -334

a2 (103) a3 (106) a4 (109) a5 (1012)


411 302 740 -369 139 1390 3820 2290 4410 -175 -81.0 -8.24 1620 138 -899 159 -763 -3160 37.5 8.23 -38.5 -467 -96.1 274 -682 112 1010 -2.97 -0.150 4.84 41.3 15.9 -31.5 141.2 -4.01 -119

O2 N2 (a) H 2O (a) H2 (b) CO (b) CO2 (b) C1.16H4 (cp for CH4) (b) CnHmOk (cp for C6H6) (b) Char (cp for graphite)

Correlations valid in the range (a) 273-4000K, (b) 273-3000K, error less than 2%

12
Table 6. Data for sensitivity analysis, dots indicate the same value as in reference case A.

Variable\Case

C
1.0
(b)

D
2

E
16

F
6

G
0.75

H
30

I
0.85

CO/CO2 2.4 2.5 CiHj/CO2 1.2 CH4/C2H4 3 Char (%) 15 n/k (CnHmOk) 30 n/m (CnHmOk) 0.97 (a) HCnHmOk 37 (MJ/kg) (b) H2O/CO2 0.95 (a) Calculates the ratio H2O/CO2

Calculates heating value of CnHmOk

The specific heat of the lumped hydrocarbons can be approximated with that of benzene, as the specific heat does not vary greatly for the hydrocarbons, which from experience are expected to be present in the lumped hydrocarbons, and the approximately equivalent molecule is close to that of benzene, see Table 4. This approximation has been made previously 38. The temperature dependent specific heats, in the form of polynomial fits (based on literature data 39, 37 ), and the reference enthalpies of the gas species used in the model, are summarized in Table 5.

Sensitivity analysis of measurements


As the system of equations, solved by Eq. (17), is rather stiff, and as the resulting composition of volatile gases easily can take unrealistic values, the influence of the uncertainty in the measurements on the resulting prediction of the composition was analyzed by a sensitivity analysis. The sensitivity analyzes was carried out for birch, having an elemental analysis and a heating value according to Table 1. The input data to the sensitivity analyzes, presented in Table 6, were varied in the range of uncertainties in the measurements on the largest particles, specific area of 320. The ratio of methane to ethylene was varied between 3 and 2 in the lumped hydrocarbons. The amount of light hydrocarbons in the THC was only approximately known, and therefore the ratio 2 was varied in a wider range than that resulting from the measurements of THC. The composition of the lumped hydrocarbons in the base case, case A, Table 6, is equal to a proposed composition 2, C6H6.2O0.2, and the variation of the atoms of hydrogen and oxygen follows another proposal 3, C6H8O. The heating value of the lumped hydrocarbons was varied between 30 and 37 MJ/kg. The measurements carried out here contain information on the ratio of H2O to CO2, and an alternative system of equations can be established, where the equation of conservation of energy is replaced by the ratio H2O/CO2 = H2O/CO2. The alternative system of equations can be expressed in the same way as Eq. (17),

13
0 H 2O 0 H 2 1/ M H O 1/ M H CO 1 / M 0 H O = 2 CO2 1 0 CH 0 i j 0 Cn H mOk 0 0
2 2

1/ M CO
2

1/ M CO 0
1 2

i / MC H
i

n / MC H
n

m Ok

0
1 2

j 2

/ MC H
i

m 2 k 2

/ MC H
n

m Ok

/ M CO 0 1 0

/ M CO 1 2

0 0 0 1

/ MC H
n

m Ok

H 2O / CO2

0 0 0

X C ,vol / M C X H ,vol / M H X O ,vol / M O 0 0 0


2 2

(27)

The result of the sensitivity analysis is presented in Figure 4 to 6. The sensitivity analysis was performed with either the heating value of the lumped hydrocarbons given, according Table 6 and Eq. (17), Figure 4, or with the ratio of H2O to CO2 given, according Table 6 and Eq. (27), Figures 5 and 6. In general, the sensitivity analysis shows that within the present measurement accuracy one obtains a stable solution both for the species concentrations in the volatile gases and for the heating value of the lumped hydrocarbons, when the composition of the volatiles is obtained from Eq. (27). The only parameter varied that leads to a pronounced deviation, is the atomic ratio of carbon to oxygen in the lumped hydrocarbon, case F. A reduction of this ratio raises the mass fraction of the light hydrocarbons and of the hydrogen and lowers the mass fraction of the lumped hydrocarbon, if the heating value is given according to Table 6. If instead the ratio of H2O to CO2 is given according to Table 6, the heating value is lowered. The heating value of the lumped hydrocarbons is expected to become lower as the atomic ratio of carbon to oxygen decreases, since in such a case there is relatively seen less carbon and hydrogen to oxidize. In general the sensitivity of the heating value of the lumped hydrocarbons is greater than that of the mass fractions. An increase of the amount of lumped hydrocarbons in the measured THC, (by a lower ratio of CiHj to CO2) shows a lower amount of light hydrocarbons (as expected), and a higher amount of hydrogen and in the lumped hydrocarbons, which tends to lower the heating value of the lumped hydrocarbons. Even case C shows some deviation, but it is mainly as expected due to the change of the ratio of CiHj to CO2.

x100

H Ox100

x1000

20

20

20

10

10

10

0
x100

0
20
COx50

0
20
CO x100
x100 C

H x1000

20

2O

H2

10

0
20

x50

20

CO

CO2

10

10

10

0
x100

0
C H x100

0
20

10

0
20

x100

CnHmOk

Ci Hj

10

10

10

nHmOk

20

20

i j

10

A B C D E F G H

A B C D E F G H

A B C D E F G I

A B C D E F G I

Figure 4. Resulting species distribution in the volatile gases from the sensitivity analysis according to Table 6, species distribution calculated by Eq. (17). (Observe the different scaling factors)

Figure 5. Resulting species distribution in the volatile gases from the sensitivity analysis according to Table 6, species distribution calculated by Eq. (27). (Observe the different scaling factors)

14

40
x100
20 10 0

H x1000 CO x100 x100 C

20 10 0 20 10 0 20 10 0 300 600 900 Specific area [1/m]

2O

HCnHmOk [MJ/kg]

CO x50

36

38

20 10 0

34
C H x100

20 10

i j

32

30 A

0 300 600 900 Specific area [1/m]

Figure 6. Resulting heating value from species distribution given by Eq. (27)from input data for the sensitivity analysis according to Table 6.

Figure 7. Estimated composition of volatile gases, expressed as mass fractions, versus specific area, for birch (* solid line) and spruce ( and dashed line). (Observe the different scaling factors)

Analysis of measurements
The measurements of the birch and spruce for specific fuel areas in the range of 320 to 780 were analyzed with Eq. (17) to obtain the most likely mass fractions of the components in the volatile gases. The empirical ratios and char yields from Figure 3, and the fuel properties according to Table 1 and 4 are input to Eq. (17). In order to get the most likely heating value and the elemental composition of the lumped hydrocarbons, the measured data for birch and spruce were analyzed, similar to that was previously done in the sensitivity analysis. This analysis results in a most likely heating value of the hydrocarbons of around 37 MJ/kg and a elemental composition of C6H6.2O0.2. The heating value of 37 MJ/kg for the lumped hydrocarbons is lower than that stated above for the given elemental composition (>40MJ/kg), but the heating value cannot have a higher value without producing an unrealistic composition of the volatile gases. This resulting heating value is closer to the one expected from the other proposed elemental composition, C6H8O, but if this elemental composition is inserted it would result in an unrealistically low heating value. The resulting quantities of in the volatile gases are presented in Figure 7 in the form of mass fractions of the volatile gases related to the specific area of the fuel. The most likely result obtained for each specific area is chosen. The analyzes shows nearly identical results for birch and spruce, except for the difference in moisture content. The most significant trend is the fall of the mass fraction of CO2 and the raise of that of the lumped hydrocarbons for increasing specific area of the fuel. Also, the mass fraction of the light hydrocarbons tends to decline as the specific area becomes larger. The data on Figure 7 indicate that the ratio of H2O to CO2 should increase with increasing specific area, but this contradicts the measurements, where the ratio was more or less constant. This is most probably due to the slow gas analyzer used for the measurement of H2O, making the measurement of H2O uncertain for the smallest particles. If there would be an uncertainty in the measurement of H2O also for the larger particles, this could be a reason for the low heating value of the lumped hydrocarbons.

nHmOk

15 In the quoted case 10, where the ratio of CO to CO2 was in the same order as the present one, although performed on thermally small particles, the amount of the lumped hydrocarbons was much higher and the amount light hydrocarbons much smaller. This indicates that cracking of the lumped hydrocarbons could have occurred in the present measurements, which produced a high quantity of light hydrocarbons. The cracking is to a large extent expected to take place inside the particles as the volatile gases flow out. Nevertheless, cracking of gases that already have left the particles cannot be excluded, but is most likely smaller, since the longer hydrocarbons that crack form more stable hydrocarbons on their way through the hot char surfaces in the outer part of the fuel particles. The absence of hot surfaces (except the reactor walls) in the reactor after the sand bed also limits the cracking. In the cases investigated the empirical data given here should be considered the best available, since the residence time is significantly shorter than in the other quoted measurements on thermally large particles.

Conclusions
The description of the volatile gases divided in a relevant number species, and leaving fuel particles of such sizes that are found in fixed or fluidized beds, is essential for sub models in comprehensive models of combustion devices. The main condition for the model is that matter and energy are conserved. Due to the complexity of the transformation of the fuel, it becomes excessively time-consuming to model the gas leaving the thermally large particles by means of a set of reaction rates. Because of time restrictions a more simplified description of volatile release is needed for comprehensive bed models. Here, a sub model is presented, whose structure is valid for any solid fuel. The model solves a system of six equations to obtain the six gas concentrations. The system of equations consists of three mass balances, one energy balance and two empirical ratios. In some cases, dependent on the available input data, the energy balance can be replaced by an additional empirical ratio, and the energy balance can be used for validation. The model includes empirical coefficients, and they have to be specified for certain classes of fuel. Gases from devolatilization of the dry part of highvolatile fuel are assumed to consist of CO2, CO, H2O, H2, light hydrocarbons (mainly methane and ethylene), and lumped hydrocarbons (that is, the remaining hydrocarbons). The method proposed is for estimation of the quantities of these gas components in a case when no data are available, or for a measured set of data needed to be checked. In addition to the empirical ratios, thermo-chemical properties of the fuel during its various phases of conversion are needed. Therefore, measurements have been carried out to provide data that can be compared with data from literature. This collection of data is especially selected to suit combustion and high temperature gasification. The data are specialized to wood. The measurements concern softwood (spruce) and hardwood (birch), but similar collection of data can be established for other fuels as well. The system of equations is analyzed for influence of uncertainties in input data. The result shows that the most important parameter is the C/O ratio in the lumped hydrocarbons. This parameter affects the heating value of the lumped hydrocarbons. An estimate gives 37 MJ/kg in the most likely case, to be compared with around 40 MJ/kg for tar and 20 to 26 for pyrolysis oils. Heating values and specific heats for wood, char, hydrocarbons and other gases are compiled. Measurements were carried out with two types of wood. 57 parallelepipeds of different size, chosen to cover the expected range occurring in a batch of forest waste as fired in utility boilers, were dried, pyrolysed and burned in a fluidized bed reactor operated at 1123 K under

16 internal heat transfer control. These results show a clear correlation between the size of the fuel, expressed as specific area, and several quantities such as char yield, ratios of CO to CO2, and light hydrocarbon to CO2. As the specific area increases (size decreases) a nearly linear decrease of the char yield and linear rise of the ratios of CO to CO2, and light hydrocarbon to CO2 can be seen. The lower char yield is accompanied by an increasing amount of carbon relative to hydrogen and oxygen in the volatile gases, which, together with the two ratios mentioned, affects the composition of the volatile gases, since it favors the lumped hydrocarbons on the behalf of the carbon dioxide. The two tested woods, hardwood (birch) and softwood (spruce), show the same trends. The only differences are in the level of char yield and in amount of H2O in the volatile gases. Further work could involve other fuels and a more detailed investigation on the influence of temperature. In the present work the most relevant temperature for the application was studied.

Acknowledgment
The Swedish National Energy Administration has supported this work financially. We thank Dr. Lars-Erik mand at this department for helpful discussions and the Swedish National Testing and Research Institute (SP), for the use of their experimental equipment and their kind assistance.

Nomenclature
H M T X Y cp h h0 i j n,m,k Greek [J/kg] [kg/mole] [K] [-] [-] [J/kgK] [J/kg] [J/kg] Heating value (lower), related to subscript Molar mass Temperature Mass fraction based on ash-free substance Mass fraction based on dry wood Specific heat, (related to subscript) Specific enthalpy, (related to subscript) Specific enthalpy, at reference temperature 298 K and ambient pressure (related to subscript) Atoms of carbon in light hydrocarbons, or index indicating specie Atoms of hydrogen in light hydrocarbons Atoms in equivalent lumped hydrocarbons, C, H and O Mass ratio of two gas species in the volatile gases Mass fraction in volatile gas Equivalent heating value (auxiliary variable) Equivalent heating value (auxiliary variable)

[-] [-] [J/kg] [J/kg]

Subscript C, H, O, etc ash char dev i n,m,k moist vol Carbon, Hydrogen, Oxygen, etc. Ash Char Devolatilisation Species Atoms in equivalent lumped hydrocarbon Moisture (mass fraction on wet fuel) Volatiles

17 wood Dry wood

References
(1) (2) (3) (4) (5) Shin, D., Choi, S. Combust. Flame, 2000, 121, 167-180. Adams, T. N. Combust. Flame, 1980, 39, 225-239. Bryden, K. M.; Ragland, K. W. Energy & Fuel, 1996, 10, 269-275. Di Blasi, C. Chem. Eng. Sci., 2000, 55, 2931-2944. Kaiser, E. R.; Friedman, S. B. in Incinerator and Solid Waste Technology, (P. E. Stephenson; R. K., Hampton; E. R. Kasier, and C. O. Veley Eds.), published by ASME, New York, 1975, 247-256. Roberts, A. F. Combust. Flame, 1970, 14, 261-272. Kanury, A. M. Combust. Flame, 1972, 18, 75-83. Beamont, O.; Schwob, Y. Ind. Eng. Chem. Process Des. Dev., 1984, 23, 637-641. Chan, W-C. R.; Kelbon, M.; Krieger-Brockett B. Fuel, 1985, 64, 1505-1513. Nunn, T.R.; Howord, J.B.; Longwell, J.P.; Peters, W.A. Ind. Eng. Chem. Process Des. Dev., 1985, 24, 836-844 Scott, D.S.; Piskorz, J.; Bergougnou, M.A.; Graham, R. Ind. Eng. Chem. Res., 1988, 27, 8-15. Figueiredo, J. L.; Valenzuela, C.; Bernalte, A.; Encinar, J. M. Fuel, 1989, 68, 10121016. Horne, P. A.; Williams, P. T. Fuel, 1996, 75, 1051-1059. Reina, J.; Vole, E.; Puigjaner, L. Ind. Eng. Chem. Res., 1998, 37, 4290-4295. Di Blasi, C.; Signorelli, G.; Di Russo, C.; Rea, G. Ind. Eng. Chem. Res., 1999, 38, 2216-2224. Di Blasi, C.; Branca, C.; Santoro, A.; Gonzalez Hernadez, E., Combust. Flame, 2001, 124, 165-177. Ragland, K. W.; Aerts, D. J.; Baker, A. J., Bioresour. Technol, 1991, 37, 161-168. mand, L-E. Personal communication, Department of Energy conversion, Chalmers University of Technology, Sweden, 2001. Evans, R.J.; Milne, T.A. Energy and Fuels, 1987, 1, 123-137. BIOBIB - A Database for biofuels; University of technology Vienna, Austria, http://www.vt.tuwien.ac.at/biobib, 2001. Koch, P. Wood Sci., 1969, 1, 203-214. Skaar, C., Wood-Water Relations, Springler-Verlag, Berlin, 1988. TenWolde, A.; McNatt, J.D.; Krahn, L. Thermal properties of wood and panel products for use in buildings, DOE/USDA-21697/1, Oak Ridge National Laboratory, Oak Ridge, TN, 1988. Harada, T.; Hata, T.; Ishihara, S. J. Wood Sci., 1998, 44, 425-431. Boutin, O.; Ferrer, M.; Ld, J. J. Anal. Appl. Pyrolysis, 1998, 47, 13-31. Roberts, A. F. 13th Symposium (International) on Combustion, The Combustion Institute, Pittsburgh, 1971; pp 893-903. Alves, S. S.; Figueiredo, J. L. Chem. Eng. Sci., 1989, 44, 2861-2869. Della Rocca, P. A.; Cerrella, E. G.; Bonelli, P. R.; Cukierman, A. L. Biomass & Energy, 1999, 16, 79-88. Scott, D.S.; Piskorz, J. Can. J. Chem. Eng., 1984, 62, 404-412. Hansson, K-M., Pyrolysis of large particles of biomass-Experiments and modelling, Thesis for the degree of Licentiate of Engineering, Department of Energy Conversion, Chalmers University of Technology, Gteborg, Sweden, 2001, forthcoming. Brenden, J. J. Combust. Flame, 1967, 11, 437-439.

(6) (7) (8) (9) (10) (11) (12) (13) (14) (15) (16) (17) (18) (19) (20) (21) (22) (23)

(24) (25) (26) (27) (28) (29) (30)

(31)

18 (32) (33) (34) (35) (36) (37) (38) Allan, F.J.; Cameron, D.E.; Lambie, D.A. Combust. Flame, 1966, 10, 394-396. Roberts, A.F. Combust. Flame, 1964, 8, 245-246. Roberts, A.F. Combust. Flame, 1967, 11, 439-441. Piskorz, J., Scott, D.S., Radlein, D. Pyrolysis Oils From Biomass, ACS symposium series 376, (J. Soltes and T.A. Milne Eds.), 1987. Nurul Islam, M.; Zailani, R.; Nasir Ani, F. Renewable Energy, 1999, 17, 73-84. TRC Thermodynamic tables - Hydrocarbons, (Frankel, M.; Hong, X.; Wilhoit, R.C. Eds.), NSRDS-NIST 75-120, U.S. Government printing office, Washington, 2000. Grnli, M, A Theoretical and Experimental Study of the Thermal Degradation of Biomass, Academic Dissertation, The Norwegian University of Science and Technology, ISBN 82-471-0009-6, 1996. Barin, I.; Platzki, G. Thermochemical data of pure substances, Weinheim, VCH, 1995.

(39)

Paper V

Thermal conductivity of wood during different stages of combustion


Henrik Thunman, Bo Leckner
Department of Energy Conversion, Chalmers University of Technology, S-412 96 Gteborg, Sweden +46 31 7721430 (tel), +46 31 7723592 (fax), email energy.conversion@entek.chalmers.se

(Submitted for publication)

Abstract
The effective thermal conductivity is one of the most important parameters for modelling of thermo-chemical conversion of wood. It changes both with temperature and with conversion of the wood. There have been suggestions on modelling of this problem, together with measurements, in earlier works, especially for wet and dry wood, but for char the knowledge is poor. Here, two principle models of effective thermal conductivity on the basis of the pore structure in wood are validated by a comparison with direct numerical simulation of the fibre structure. The validation leads to a more general model, both for conductivity in the perpendicular and in the parallel direction relative to the fibres in the wood. A secondary result is that the thermal conductivity of the solid phase in the fibre wall of the wood can be evaluated for dry and wet fuel, from measurement data on effective thermal conductivity. The effective thermal conductivity can be estimated from given values of temperature, density and moisture content of the wood. It can also be applied to pellets and chipboards. In addition, the general model expresses the effective thermal conductivity of char, since the wood material maintains its fibre structure during conversion. Keywords: Thermal conductivity, Wood, Modelling

Introduction
The effective thermal conductivity is one of the most important parameters for modelling of thermal conversion of wood. The effective thermal conductivity changes both with temperature and with conversion of the wood, and it is therefore of importance to describe this behaviour for modelling of the progress of combustion. A model is derived to calculate the effective thermal conductivity in parallel and perpendicular to the fibres in a fibrous material, and this model is applied here to wood. Wood particles have their largest surface area parallel to the fibres in the wood structure. This makes the thermal conductivity perpendicular to the fibres most important. However, the thermal conductivity along the fibres is also of interest in some cases, for example in pellets or chipboard, materials with random fibre orientation. Much measurement data, especially on conductivities of wet and dry wood, can be found in the literature. For example, the work of MacLean [1], is probably the largest single study ever made of thermal conductivity perpendicular to the fibres of different woods. Along the fibres measurement data are available, for example, in a review carried out by Grnli [2]. Thermal conductivity of chars from biofuels has been measured by, for example, Lee et al. [3], Evans and Emmons [4], Alves and Figueiredo [5]. As previously shown by Siau [6] and Saastamoinen [7], a model consisting of quadratic cells, where the conduction is evaluated along various paths in relation to the orientation of the cells, gives a good agreement with effective thermal conductivities measured perpendicular to the fibres of wood. Along the fibre the thermal conductivity is much simpler to estimate,

2 pipes placed next to each other can represent the fibres, and the effect of the solids that cut off the gas path in the ends of the fibres can be neglected, Siau [6].

Theory
As the derivation of the conduction paths perpendicular to the fibres differs to some extent between the models of Siau [6] and Saastamoinen [7], a numerical simulation was made, in which the energy equation was solved for a large number of quadratic cells located in a matrix between two thin plates having a very high conductivity. Figure 1 shows the principle set-up. The energy equation for this case,

( ki T ) = 0

(1)

where index i indicates the conductivities of solid (i=s), moisture (i=w) and gas (i=g) constituting the porous body. The boundary conditions of the simulation are: constant surface temperature, T2, on one end of the matrix and constant heat flow, q(T1), on the other end. The boundaries perpendicular to the matrix and to the plates were perfectly insulated. A quadratic computational grid with a smaller size than the width of the wall of the solid phase was applied. The temperature was calculated at the surface of the plate having the constant heat flow. The effective conductivity, keff, was then evaluated from the temperature difference and the heat flow according to the definition:
q = ( 2 y1 + y2 ) ( 2 y1 / k1 + y2 / keff )

(T1 T2 )

(2)

where y1 is the height of a bounding plate and y2 height of the matrix. k1 is the thermal conductivity of the plates, assumed to be large (500W/mK). Additional simulations were
q (T1)

y1

Plate
g

y2

w/2 s

y1

Plate
T2

Figure 1. Model of fibre structure of wet wood. s is width of solid, w width of moisture layer and g width of gas. For numerical validation the heat flux was calculated between boundaries having temperature T1 and T2. In the calculation not nine, but a large number of cells were placed in the matrix.

Figure 2. Fibre structure of hardwood, Maclean [8].

Figure 3. Fibre structure of softwood, Howard [9].

a)
1 2 s 3s w 2 s 1

Flux
s w g w s s w s s

b)

Flux

c)

Flux

Ai+2 Bi+2 Bi

s s w s 1 2

s w g w s 3

s w s s 2 1 s w g w s

Bi+1 Bi
1 2 3 2 1

Ai+1 Ai

Ai+1 Ai

Bi Ai

R1,s R2,s R2,w R2,s R1,s R2,w

R3,s R3,w R3,g Rs,1 Rw,2 Rg,3

Bi Ai Bi+1

R3,s R3,w R3,g

Figure 4. Paths of heat flux through the fibre represented by electrical circuits. a) and b) paths perpendicular to the fibre derived by Siau [6] and by Saastamoinen and Richards[7] respectively. c) paths parallel to the fibre, derived by Siau [6]. R stands for resistance, index 1,2,3 indicate the different paths, and s,w,g, indicate solid, water and gas.

Figure 5. The paths perpendicular to the fibre as derived of Siau [6] (A) and Saastamoinen and Richards [7] (B), combined in parallel and serial paths

4 made to consider the effects of size, shape and relative position of the perpendicular crosssection of the fibre. Rectangular shapes represented the cross-section of the fibres with different height to length ratios and different wall thicknesses. When uniform cross-sections of fibres with various height to length ratios and wall thicknesses, were placed in straight rows and columns in a matrix, the effective conductivity could vary between two extremes; the elements of the wood structure could either be modelled as placed parallel or serial to each other in the direction of the heat flux. The perpendicular fibre structure of wood, hardwood [8] Figure 2 and softwood [9] Figure 3, differs in size, but shape and wall thickness is rather constant. This was simulated with a matrix divided into sub-matrices, where each sub-matrix consisted of uniformed sized cross-sections of the simulated fibres, but the size, shape and wall thickness varied between the sub-matrices. A number of simulations were made and the result showed that the calculated effective thermal conductivities nearly coincided with the one obtained from the simulations using uniform quadratic cross-sections placed in straight rows and columns. This concludes that a uniform structure of quadratic fibres can represent a more complex one, as long as there is no systematic arrangement of non-quadratic crosssection of fibres in the wood. The result supports the modelling of thermal conductivity of wood as a system of serial and parallel conduction paths. In the derivation of conduction paths both Siau [6] and Saastamoinen and Richard [7] assumed that the water remains on the surfaces of the fibre wall in an even layer. The relative thickness of fibre walls and the thickness of the water layer are calculated from the volume occupied by the solid and the water. The relative volume is estimated by the apparent density of the wood divided by the real density of the solid and the moisture. Knowing the volume occupied by each component and assuming the type of fibre structure, the normalised thickness, , of each layer in the pore can be calculated from the densities, , of the fuel and the fibre walls and from the shrinkage of the particle. The shrinkage is the ratio of a volume element to its dry volume element. For a quadratic cell structure, the thickness of each layer will be (the details of the derivation are given in Appendix A),

s ,1 = 1 (1 ( 2 / 1 )( d / 2 ) )

0.5

s , j =1 1 ( 2 / j )( d / j ) Yn for
4 n= j

0.5

w,1 = (1 s ,1 ) (1 s ,1 ) ( 2 / 1 )( d / 2 ) Y1
2

w, j =0 for { j > 1} g , j = 1 s , j w, j

{ j > 1}

(3)

0.5

(4) (5)

where the first index is related to the solid (s), moisture (w) and gas (g), the second index j is related to the different stages of combustion: virgin fuel (1), dry fuel (2), char residue (3) and ash (4). For the density j represents: water (1), solid fibre wall in the dry and the virgin wood (2), solid fibre wall in the char (3) and solid fibre wall in the ash (4). Siau [6] derived the effective conductivity of the cell by three layers in parallel, Figure 4a, and therefore underestimated the conductivity, for conversion stage j,
s , j (1 s , j ) g, j + g , j s , j + w, j + (6) + keff , j , A1 = s , j ks , j + w, j ks , j ks , j kw k w ( k g + krad )
1 1

5 Saastamoinen [7] derived the cell in three layers in series, Figure 4b, and therefore overestimated the conductivity, for conversion stage j,
s, j w, j g, j (7) + + keff , j , B1 = k s , j s , j k s , j + (1 s , j ) k w s , j k s , j + w, j k w + g , j ( k g + krad )
1

The results from Eq (6) and (7), (n=1), are combined in serial and parallel paths, as illustrated in Figure 5,
keff , j , An +1 = 1 keff , j , An + 1 keff , j , Bn 2 2 keff , j , Bn +1 = ( 1 / keff , j , An + 1 / keff , j , Bn ) 2 2
1

(8)

which gives the effective thermal conductivity at n=2 after combination of those of n=1, n is a counter. The values obtained by Eq (8) over- and underestimate, respectively, the serial and parallel conduction paths. As n increases, the difference between the two conductivities decreases. When n goes towards infinity the two values become equal to each other and a resulting effective thermal conductivity is obtained. This resulting conductivity is nearly the same as the one obtained from the simulation.
n ; keff , j = keff , j , A = keff , j , B

(9)

Along the fibre the thermal conductivity is simpler to estimate, pipes placed next to each other represent the fibres. The effective thermal conductivity, can be modelled by a parallel path, as suggested by Siau [6], Figure 4c,
2 2 keff , j , = 1 (1 s , j ) k s , j + (1 s , j ) g , j k w + g , j ( k g + krad ) 2 2

(10)

The effective thermal conductivity in the different layers in the particle is of most interest during drying and devolatilisation, since the heat transfer in the particle drives both these phenomena. The temperatures during this period are moderate and the pore diameter is small, which makes the radiative contribution to the effective thermal conductivity negligible. During char combustion, when drying and devolatilisation are ended, a high temperature (above 1400C) can be attained, and then the radiative contribution has to be included. However, during this combustion stage the effective thermal conductivity is of minor interest, since diffusion of oxygen into the particle and reaction rate determine combustion of char and, together with the external heat transfer, the temperature of the particle. The radiative thermal conductivity can be approximated by, Hottel and Sarofim [10], krad = 16 rad T 3 d pore 3 (11)

where rad is the emissivity in the pores, Stefan-Boltzmann constant and dpore is an equivalent pore diameter in the wood. A partially converted solid should have an effective thermal conductivity somewhere between the values of the present (j) and the next (j+1) conversion stage. When modelling the intermediate stage it is important to consider all input parameters, such as shrinkage, density

6 and thermal conductivity of the solid fibre wall. The simplest approach is a linear approximation of the thermal conductivity between the two conversion stages j and j+1 as, keff = (1 X i ) keff , j +1 + X i keff , j (12)

where Xj is the mass ratio of the component being converted and the initial mass. In pellets and chipboards the fibres are nearly randomly oriented and the effective thermal conductivity becomes the same in all directions. From geometrical consideration, 2/3 of the fibres are perpendicular and 1/3 parallel to the direction of the heat flux, and the effective thermal conductivity can be estimated by, keff ,r = 2 keff + 1 keff , 3 3

Results
The thermal conductivity perpendicular to the dry fibre, ks,2, is estimated to 0.52 W/mK for the solid material in the dry wood by a least-square fit of Eq (9) to measurement data from MacLean [1] and other authors, reviewed by Grnli [2]. Parallel to the fibre the corresponding thermal conductivity, ks,2, is estimated to 0.73 W/mK from Eq (10) and measurements reviewed by Grnli [2]. The density of the solid in dry wood, 2, is chosen to 1480kg/m3, Siau [6]. The conductivity and density of the solid in dry wood are assumed to be independent of temperature. The density, 1=1000 kg/m3 and conductivity, kw0.487+5.88710-3T7.39106 2 T W/mK are used for the moisture in the wood, according to data for water, Incropera and DeWitt [11]. The thermal conductivity of gas is assumed to be the same as for air, which is motivated by the small difference between the major components produced during drying and devolatilisation: water vapour, CO, CO2, O2 and N2, Kanury [12]. However, during devolatilisation CH4 and especially H2 will be present in rather large quantities. These gases have a much higher conductivity and stronger temperature dependence than the other gases, and the effective thermal conductivity will be somewhat underestimated. The largest error occurs in the char layer formed during devolatilisation, as the volatile gases pass through. For air, the thermal conductivity can be calculated from an empirical correlation based on table values, Incropera and DeWitt [11], given in the temperature range 250 to 3000K, as follows,
k g = 7.494 103 + 1.709 10 4 T 2.377 107 T 2 + + 2.202 1010 T 3 9.463 1014 T 4 + 1.581 1017 T 5

(13)

There is good agreement between Eq (6) to (10) and the measurement data for dry wood, Figure 6, and for wet wood, Figure 7. The difference between the models, represented by Eq (6) to (7), is rather small. The difference becomes clear if one regards the heat conductivity of the solid in the fibre. Siau [6] has suggested this heat conductivity to be 0.43W/mK, but Saastamoinen and Richard [7] chose 0.6W/mK. If the values mentioned are inserted into their respective equations, there is an even better agreement than the one shown in Figure 6, but for high density fuels the error becomes noticeable. If these values are used, for example, to estimate the heat conductivity of pellets, which have a high density (1000-1300 kg/m3), the difference between Eqs (6) to (7) can be more than 15 %. When char is produced at a high heating rate and with a high final temperature, the resulting porous structure is nearly pure carbon, Figueiredo et al. [13], Della Rocca et al. [14]. It is assumed that carbon in the porous structure has the same thermal conductivity and density as

7 amorphous carbon, for which the conductivity, ks,3=1.47+0.0011T W/mK and density, 3= 1950 kg/m3, Incropera and DeWitt [11]. The thermal conductivity of the carbon in the porous structure is assumed to be the same in the perpendicular and parallel directions relative to the fibre. The solid in the small quantity of ash is given the same density and conductivity as the carbon in the char. For char it is not possible to estimate the thermal conductivity of the solid material in the cell wall from available measurement data, as the data of Lee et al. [3], Evans and Emmons [4] and Alves and Figueiredo [5] are given without sufficient information on the experimental conditions, but the reported values indicate that Eq (8) overestimates the conductivity, perhaps because the samples used for the measurements were not completely carbonised.

Conclusions
Comparison with measurements show that the model of effective thermal conductivity perpendicular, Eq (14), and parallel, Eq (10), to the fibres, Figure 6 and 7, can be used to calculate the effective thermal conductivity for all kind of wood species from density, moisture content and shrinkage, parameters that can be measured easily. Knowing from Grnli [2] that the wood maintains its fibre structure after devolatilisation the effective thermal conductivities for char can be calculated from the same models. For char there is no comprehensive work available in the literature on measurements of effective thermal conductivity, as it is for wet and dry wood, and this makes it necessary to assume the density and the corresponding thermal conductivity of the solid material in the char. In present work the solid in char is modelled by amorphous carbon, motivated by the high carbon content in the char. The radiative contribution to the effective thermal conductivity can be neglected for calculation of the thermal phases of conversion of wood, due to small pore diameters and moderate temperatures during drying and devolatilization, which are the phases of conversion when the effective thermal conductivity is most important.
0.5
Measured thermal conductivity [W/mK]

0.8 0.7

Heat conductivity [W/mK]

0.4

0.6 0.5 0.4 0.3 0.2 0.1 0 0 500 1000 Dry density [kg/m3] 1500

0.3

0.2

0.1

0.1 0.2 0.3 0.4 0.5 Calculated thermal conductivity [W/mK]

Figure 6. Heat conductivity of dry wood as a function of density. Conductivity parallel to fibre, Eq. (10) (solid line with dots), measured data, Grnli [2] (stars). Conductivity perpendical to fibre, Eq. (15) (solid line), Eq. (7) (dashed line), Eq. (6) (dotted line). Measured, MacLean [1] (circles), Grnli [2] (rhombs).

Figure 7. Comparison of measured, MacLean [1], and calculated effective thermal conductivity, Eq (16), perpendicular to the fibres in wood, with a moisture content of 7-45% (based on wet wood).

Acknowledgment
The Swedish National Energy Administration has supported this work financially. The authors would like to thank Gennadij Palchonok, Department of Energy Conversion, Chalmers Technical University, for valuable discussions.

Appendix A
Derivation of the thickness of the solid walls and the water layer in a fibre is made from geometrical considerations. If the fibres are represented as quadratic tubes, the ratio of the area of the solid in the perpendicular cross-section, As, to the area of the perpendicular crosssection of the fibre (solid + gas), Af, becomes equal to the volume ratio of the solid in the fibre wall, Vs and the fibre, Vf, which in its turn, is equal to the ratio between the apparent density of the fibre, f, , and the density of the solid in the fibre wall, s (assuming that the density related to the gas is negligible). As Vs f = = Af V f s The density of the dry fuel together with the shrinkage coefficient and the stage of conversion give the apparent density of the fibre,

f , j = ( 2 / j ) d Yn
n= j

where the shrinkage is the ratio of the volume of the fibre during conversion stage j to the volume of the dry fibre, conversion stage j=2. By normalizing the area of perpendicular crosssection, the normalized thickness of the solid wall becomes,

s = ( Af / Af ) g =

0.5

(( A )

As ) / Af = 1 s

0.5

= 1 1 ( f / s )

0.5

and for the gas volume,

(( A

As ) / Af

0.5

For the virgin fuel (j=1), the solid wall in the fibre is divided into a solid and a liquid part, and the apparent density of the fibre is then given by,

f ,1 = ( 2 / 1 ) d Y1 + ( 2 / 1 ) d Yn = ( 2 / 1 ) d Y1 + ( 2 / 1 ) d
n=2

In the virgin wood, conversion stage j=1, the moisture is assumed to stay on the surface of the solid in an even layer, and the thickness of the solid layer then becomes,

s ,1 = 1 (1 ( 2 / 1 )( d / 2 ) )

0.5

the thickness of the moisture layer,

w = (1 s ) (1 s ) ( 2 / 1 )( d / 1 ) Y1
2

0.5

and the gas volume, w = 1 s w

Nomenclature
A T V X Y k q y [m2] [K] [m3] [-] [-] [W m-1 K-1] [W m-2] [-] Area Temperature Volume Mass ratio mass Fuel composition related to dry fuel Thermal conductivity Heat flow Height

Greek rad [ - ] [-] [kg m-3] [W m-2 K-4] [-] Subscripts A B d eff f g i j Conductive paths, Siau Conductive paths, Saastamoinen dry Effective Fibre Gas Component, solid, water, gas 1,virgin fuel, 2,dry fuel, 3,char, 4,ash, special for density, water (1), solid fibre wall, in dry and virgin wood (2), in char (3) and in ash (4) n Counter pore Pore r Randomly oriented fibres, as in pellets or chipboard rad Radiation s Solid w Water Parallel to fibre Emissivity Shrinkage coefficient Density Stefan-Boltzmann constant Dimensionless length

References
1. MacLean, J. D., Thermal Conductivity of Wood, Transactions American Society of Heating and Ventilating Engineers, 1941, 47, 323-354. 2. Grnli, M, A Theoretical and Experimental Study of the Thermal Degradation of Biomass, Academic Dissertation, The Norwegian University of Science and Technology, ISBN 82-471-0009-6, 1996. 3. Lee, C. K., Chaiken, R. F., Singer, J. M., Charring Pyrolysis of Wood in Fires by Laser Simulation, The Sixteenth Symposium (International) on Combustion, The Combustion Institute, Pittsburgh, 1976, pp 1459-1470. 4. Evans, D. D., Emmons, H. W., Combustion of Wood Charcoal, Fire Research, 1977, 1, 57-66. 5. Alves, S. S., Figueiredo, J. L., A Model for Pyrolysis of Wet Wood, Chemical Engineering Science, 1989, 44, 2861-2869. 6. Siau, J. F., Transport Processes in Wood, Springler-Verlag, Berlin, ISBN 3-540-12574-4, 1984.

10
7. Saastamoinen, J. J., Richard, J-R., Simultaneous Drying and Pyrolysis of Solid Fuel Particles, Combustion and Flame, 1996, 106, 288-300. 8. Maclean, J.D., Preservative treatment of wood pressure mothods, US Dep Agr Handbook 40, 160 pp, 1952 9. Howord, E.T., Manwiller, F.G., Anatomical characteristics of southern pine steamwood, Wood Science, 1969, 2, 77-86. 10. Hottel, H. C., Sarofim, A. F., Radiative Transfer, McGraw-Hill Book Company, New York, 1967. 11. Incropera, F. P., DeWitt, D. P., Fundamentals of Heat and Mass Transfer, Fourth edition, John Wiley & Sons Inc., ISBN 0-471-30460-3, 1996. 12. Kanury, A. M., Introduction to Combustion Phenomena, Gordon and Breach Science Publisher, New York, ISBN 0-677-02690-0, 1977. 13. Della Rocca, P. A., Cerrella, E. G., Bonelli, P. R., Cukierman, A. L., Pyrolysis of Hardwoods Residue: On Kinetics and Chars Characterization, Biomass & Energy, 1999, 16, 79-88. 14. Figueiredo, J. L., Valenzuela, C., Bernalte, A., Encinar, J. M., Pyrolysis of Holm-Oak Wood: Influence of Temperature and Particle Size, Fuel, 1989, 68, 1012-1016.

Paper VI

Final report May 1999


Page 1 of 79

Activity 3.1.3 Modelling and verifying experiments on the whole furnace


H.Thunman, L-E. mand, F. Ghirelli, B. Leckner Department of Energy Conversion Chalmers University of Technology S-412 96 Gteborg, Sweden

March 1999 EU-contract: JOR 3CT96 0059

Final report May 1999


Page 2 of 79

Final report May 1999


Page 3 of 79

Summary A model has been developed for combustion of a fuel layer on a moving grate, which can be connected to a CFD-calculation for combustion and gas flow in the free-room of the furnace. Special attention has been paid to formation and reduction of nitrogen oxides. The work has not looked at the formation and destruction of nitrogen oxides by detailed chemistry, but on the conditions in the fuel layer and in the free-room. From these conditions the nitrogen chemistry has been modelled in a simplified way. To verify the calculations extensive measurements have been performed at a 31MW th industrial moving grate furnace burning wood chips, in co-operation with Kvaerner Pulping Power Division. The modelling of the whole furnace gives acceptable results for the gas flow and the reactions of the main species, such as oxygen, carbon monoxide, carbon dioxide and hydrocarbons. Especially good is the agreement in the upper part of the furnace. In the lower part of the furnace the agreement with measurements is not as good as in the upper part, which is a result of the difficulty to define the fuel layer on the grate. This is a problem of determining the empirical data needed in the fuel layer model, which could not be established by the measurements, due to practical measurement problems at the test facility. The modelling of the nitrogen chemistry showed a limitation in the use of reduced mechanisms available for the CFD-calculation, especially at the low temperature levels that are present in a grate furnace operating with biofuels. The nitrogen chemistry in the fuel layer model shows that the region of devolatilization is the most interesting and that the biofuels have a great potential in themselves to reduce the nitrogen oxides to low levels.

Final report May 1999


Page 4 of 79

Final report May 1999


Page 5 of 79

Table of content

Page 1. Introduction 2. The boiler and the conditions for modelling and measurements 3 Measurement preparation 4. Modelling 5. The NOx modelling 6. Experimental 7. Results 8. Discussion and conclusion 9. References Appendix A. Composition of the volatile B. Analytic solution for reactions controlled by mixing 77 78 7 9 11 15 38 42 43 73 76

Final report May 1999


Page 6 of 79

Final report May 1999


Page 7 of 79

1. Introduction Grate-firing has been used to burn bio-fuels for many decades, particularly in the pulp and paper industry, where the fuel is supplied together with the raw material for pulp production or from wastes, such as bark. Later it also became used in the furnaces of the district heating systems. The original design was the plane grate, but later sloping grates of various shapes and inclinations were employed. These grates are quite sensitive to the adaptation of the inclination of the grate to the type of fuel. If the fuel does not fit properly, it happens that too much fuel slides down the grate and ends up only partly burned in the lower end. The most modern development, the reciprocating sloping grate, solves this problem and allows a rather controlled movement of the fuel along the grate by means of mechanically controlled rods, constituting the surface of the grate, moving back and forth according to an adjustable scheme, thus allowing to control the speed of the fuel layer's movement along the sloping grate, Figure 1.1. This is the type of grate treated in the present work. Previous development work has been directed to achieving a complete and reliable combustion on the grate. During recent years, however, the emission limits have been lowered and a complete combustion is not a sufficient criterion of acceptable operation: now a low emission of nitrogen oxides also has to be achieved at the same time as the desire for complete combustion remains. These requirements have made it necessary to improve the understanding of the combustion processes in grate-fired furnaces, just as in other combustion devices. Tools to evaluate the operation of the grate and to predict the behaviour of combustion and pollutant formation and destruction are needed. Much is also gained simply by an increased qualitative understanding of the processes taking place, since the behaviour can be improved by simply adjusting the existing grate for a certain fuel. Several parameters of adjustment are available: initial bed height, air distribution along the grate, movement of the grate (speed of the fuel layer along the grate), secondary air jets, flue gas re-circulation etc. This is, indeed, quite a complex situation that has to be assessed in addition to the design information needed for design of the unit. The present work uses existing knowledge elements to develop a model of the combustion behaviour of the fuel layer on a grate and its connection to the gas space above the grate, the free-room with its secondary air jets and combustion in the gaseous phase. Moreover, special attention is paid to formation and destruction of nitrogen oxides. This item can be treated either by detailed nitrogen chemistry or by assuming that the conditions in the fuel layer and free-room control the nitrogen chemistry. The latter approach is followed here. It is quite obvious, however, that the existing knowledge on the behaviour of the fuel layer is not sufficient and that a considerable model development is needed. In the first place a fuel layer model should be developed to give input data at the lower boundary to a computerised fluid dynamic calculation of the processes in the gas space, including various reactions and heat transfer. In this part of the furnace, in contrast to the fuel layer, available commercial programmes can be applied with advantage. In the present case FLUENT is used. In order to validate the computational procedures and to obtain

Final report May 1999


Page 8 of 79

additional information about the combustion process, measurements have been carried out. The measurements intend to describe the overall combustion behaviour inside the furnace. Measurements are performed in several positions in the free-room and in the gas flow entering and leaving the furnace. For the fuel layer only the velocity of the fuel layer along the grate and the surface temperature of the grate were measured. These measurements together with the boiler design data give the input to the modelling and a base for validation of the simulation. In summary, the present work consists of three parts: modelling of the fuel layer, freeroom modelling and measurements predominantly carried out in the free-room and in the furnace exit. The purpose of the work is to provide tools that could be used for design predictions related to fuel layer and free-room.

Figure 1.1 Reciprocating sloping grate (Kvearner)

Final report May 1999


Page 9 of 79

2. The boiler and the conditions for modelling and measurements The simulation concerns an existing 31MW grate-fired boiler that produces hot water for the district heating network of Trollhttan, designed and built by Kvaerner Pulping, Power division. The fuel used is wood waste from the pulping industry, consisting of splinters, bark and sawdust. The boiler is operated in the load range of 25 to 100% and with fuel having moisture content of 35 to 55% on mass basis. Due to the highly moist fuel, the plant is equipped with a flue gas condensation unit to recover the latent heat of the water vapour. Of the 31MW produced 6MW are delivered by this flue gas condensation unit, 17MW by the convection area of the flue gas pass and 8MW by the tube walls of the furnace. The wide fuel and load range of operation requires a number of special design considerations. The fuel is supplied to the grate through two fuel chutes, covering the total boiler width. Raising or lowering ceramic beams in front of the first section of the grate controls the height of the fuel layer. The grate is a reciprocating one consisting of a steel support with intermediate rollers on which the grate sections are placed. The sections are put into operation by means of hydraulic pistons in the boiler front. All castings exposed to radiant heat from the furnace are composed of a chrome/nickel alloy, capable of withstanding high operation temperatures. The grate is built up of movable grate bars. A row of grate bars is alternating between fixed and moving every second. The velocity of the fuel layer along the grate is reduced significantly half way on the grate in order to burn out the char. The primary air supply is divided into 2 times 5 air zones (P), see Figure 2.1. Details of the grate are shown on Figure 2.2. The primary air-flow is preheated to 150 C and optimised for the combustion of the fuel layer above each wind-box. A relatively high pressure-drop across the grate has been chosen to avoid blowing of holes through the fuel layer. The combustion and pyrolysis gases evolving from the fuel layer are mixed by the recirculation gas jets (R in Figure 2.1) and the mixture passes through the narrow section of the combustion chamber (the neck) where it is mixed with the secondary air (S in Figure 2.1) supplied in three rows. Each row of secondary air has a separate air-box with control valves. The high moisture content of the fuel requires refractory lined walls and roofs below the neck in the furnace. A flue-gas recirculation system is installed to improve the mixing in the lower part of the furnace and to better control the combustion temperature, independent of the moisture content, which makes low NOx operation of the boiler possible. Downstream of the furnace, the flue gases are cooled down to a temperature of 170 C in the boiler back passes by the heat recovery surfaces. The fly ashes are separated from the gas flow by an electrofilter. Condensation equipment humidifies the air to further increase the heat production. Technical data of the boiler are given in Table 2.1.

Final report May 1999


Page 10 of 79

Figure 2.1 Scheme of test plant, Kvaerner Pulping, Power Division.

Figure 2.2. The reciprocating sloping grate of the test boiler in Trollhttan.

Final report May 1999


Page 11 of 79

Table 2.1. Technical data for the test facility in Trollhttan. Maximum capacity Maximum condensation capacity Moisture content of the fuel Boiler water temperatures Inlet district heating temperature Outlet district heating temperature Grate surface area Free-room volume Excess air Flue gas temperature NOx CO Dust (13% CO2 dry gas) 25 6 35-55 200-140 72 120 39.8 295 25 170 72 90 35 MW th MW th % C C C m2 m3 % C mg/MJ mg/MJ mg/Nm3

3 Measurement preparation It was necessary to prepare the test facility in Trollhttan for the measurements in the present project. In order to get access to the boiler furnace with probes, 29 measurement holes were taken up. Banisters were moved or taken away, termocouples and pressure transducers were installed as well as cooling-water and sewage systems in connection to the measurement holes. For collection of measurement data, signal cables were drawn from each of the measurement levels to the control room were the data collection took place. A great number of other practical arrangements were also necessary for the measurements. The measurement holes had to be designed in different ways, depending on the tube walls. The tube walls in the lower part of the furnace were prepared with large round holes, Figure 3.1. Narrow rectangular holes were made in the fin between the tubes in the upper part of the furnace, Figure 3.2, in order to reduce the costs and to install more holes. The rectangular holes restrict the measurement to probes having a simple cooling system. The positions of the measurement holes are presented in Figure 3.3.

Final report May 1999


Page 12 of 79

Figure 3.2. View of rectangular holes and equipment for guiding and fixing the with probe. .probe.

Figure 3.3. View of the round holes equipment for guiding and fixing the

Figure 3.1. The measurement holes in the test boiler in Trollhttan. Four kind of probes were specially designed and built for the measurements: A cold suction probe for local gas concentration and temperature measurements in the rectangular holes. With a cold probe is meant that the gas is cooled down to the temperature of the cooling water, Figure 3.5.

Final report May 1999


Page 13 of 79

A hot suction probe for local gas concentration measurements in the round holes. With a hot probe is meant that the gas is cooled down to a temperature of 200C, Figure 3.6. A zirconia cell probe for local oxygen fluctuation and temperature measurements to be used in the round holes, Figure 3.4. A flow direction probe, to be used in the upper part of the furnace, for examination of the direction of the gas flow.

The cold suction probe measures the concentration of CO, CO2, O2, NOx and total hydrocarbon concentration, THC (for analysis equipment, see Table 3.1). The hot suction probe measures NH3 with a FTIR (Fourier Transform Infra Red) analyser, together with the same gas species as those measured with the cold suction probe. For the ammonia measurements, a hot suction probe must be used to prevent absorption of ammonia in the condensate. The arrangement to keep a rather high gas temperature is space consuming, and this restricts the ammonia measurement to the round holes in the furnace. Table 3.1. Gas analysers Analyser O2 CO* CO2 NOx Principal of measurement Paramagnetic NDIR Manufacturer M&C PMA 10 Binos 100 Measurement range 0-30 % 0-30000 ppm 0-5000 ppm 0-20 % 0-1000 ppm Calibration gas 9.97 % 4040 ppm 12,0 % 91.7 ppm NOx 90.7 ppm NO 899 ppm CH4

chemiluminiscent Eco Physics

THC* FID FID-analyzer 0-30 000 ppm NH3 FTIR Bomem MB 100 *) When the concentration of CO and THC exceeded the range of the analyser, the gas concentrations were analysed by the FTIR.

Figure 3.4 The zirconia cell probe.

Final report May 1999


Page 14 of 79

Figure 3.5. The cold suction probe for gas concentration and temperature measurements.

ceramic filter

thermostated shell

heated sampling tube

Figure 3.6a. The front part of the hot gas extraction probe.
backflush tubes Teflon electric insulator

low voltage high current transformer heated sampling tube 220 VAC

Figure 3.6b. Rear part of the hot gas extraction probe.

Final report May 1999


Page 15 of 79

4. Modelling The present work uses existing knowledge to develop a model of the combustion behaviour of the fuel layer on the grate and its connection to the gas space above the grate, the free-room with its secondary air jets and combustion in the gaseous phase. Moreover, special attention is paid to formation and destruction of nitrogen oxides. This item can be treated either by detailed nitrogen chemistry or as controlled by the conditions in the fuel layer and free-room. The latter approach is chosen here, but it is quite obvious that the existing knowledge is not sufficient and that a considerable model development is needed. In the first place a fuel layer model should be developed that gives input at the lower boundary to a computational fluid dynamics (CFD) calculation of the gas space, including various reactions and heat transfer. In this part of the work, in contrast to that related to the fuel layer model, available commercial programmes can be applied with advantage. In the present case FLUENT/UNS 4.2.5. is used. This software package has built in the necessary physical models to simulate gas phase combustion (turbulent fluid flow, chemical species mixing and reaction, conductive, convective and radiant heat transfer). The package includes software for domain definition and discretization and has an integrated nitrogen oxide reaction module. The complexity of modelling of a whole furnace is obvious, and it is therefore important that already from the beginning a realistic accuracy level is defined on what should be required from the overall solution. In the present work the knowledge on the combustion of a moving fuel layer sets the limit. The understanding of the combustion of the fuel layer is the key for the nitrogen chemistry, because a modern boiler is designed to prevent the formation of thermal NOx, the main source of NOx is from the combustion of the fuel. In the fuel layer the solid fuel is converted to gas, and the precursors of the pollutants are formed and transported up to the free-room. Since the detailed knowledge of the combustion of the fuel layer is not previously available, the present modelling has been focused on the combustion of this fuel layer. Special efforts have been made to model the parts that are important for the NOx chemistry. The fuel layer is exposed to a rather complex combustion situation, both the primary airflow and the velocity of the fuel layer along the grate change several times while the fuel is transported along the grate. For the nitrogen chemistry the position where the fuel-bound nitrogen leaves, the local oxygen concentration and the height of the char layer are the main parameters. If the nitrogen is released at or close to the surface of the fuel layer, formation and destruction of the nitrogen pollutants will take place in the free-room above the fuel layer. If there is a char layer, both formation and destruction of nitrogen pollutants can take place in it, depending on the oxygen level. In the reaction front passing through the fuel layer, precursors of the nitrogen oxidation will be released. Dependent on the oxygen level, these precursors will form or reduce nitrogen oxides as they pass through the char layer. The models of the fuel layer and the free-room run independently of each-other; the fuel-layer model represents one of the boundary conditions for the free-room model and vice versa. This requires an iterative calculation procedure to reach a converged result for the combustion inside the entire furnace. Convergence is attained when the heat flux to the fuel layer, given by the free-room calculation, agrees with the heat flux to the fuel layer given in the fuel layer model. Since a commercial programme is

Final report May 1999


Page 16 of 79

used for the free-room, the fuel layer model has to be described in such a way that the output from the fuel layer and free-room model easily can be used for the iteration. The chemistry of combustion is divided into two parts called main chemistry and trace chemistry. The main chemistry includes the reactions and species relevant for the determination of velocity and temperature fields, while the trace chemistry includes species present at concentrations that may be neglected in the heat and mass balances. The first part, requiring most computational time and power, calculate the balance equations of the main variables: temperature, average velocity and concentrations of non-trace species. The trace chemistry is neglected in this part. Subsequently, during the second part of the calculation, the post processing, the solution from the first part is fixed, and the balance equations for the trace chemistry are solved. Therefore the post processing will not influence the main fields. This separation of the chemistry is useful, and saves considerable computational time and power, which makes it possible to test several cases with different emphasis on the trace chemistry on the same calculation.
H eatFl ux

xs

Bed height

xc xr
Re act

Figure 4.1. Illustration of the four layers; surface, char reaction front and unreacted fuel. The calculation time of the fuel layer on the grate is important to limit the calculation effort. Therefore the model of the fuel layer is simplified as much as possible. For this purpose the fuel layer is split up in four characteristic layers, as illustrated in Figure 4.1 surface layer, char layer, reaction front and unreacted fuel layer. The surface and the char layers consist of char. This representation of the fuel layer is supported by experiments in pot furnaces, e.g. Gort 1995, Saastamoinen 1999. The difference between them is that the surface layer, in contrast to the char layer, is influenced by the conditions in the free-room above the fuel layer. The reaction front is the layer where drying and devolatilization take place, and the unreacted fuel layer consists of the fuel not yet reached by the reaction front. The most complex part of a fuel layer model is the representation of the reaction front; a narrow layer where the fuel is dried and devolatilized simultaneously. In this layer the heat and mass transport is complex and a detailed model is required. Saastamoinen [1998], for instance, models the behaviour of a single particle through the reaction front, but in the present work a more simplified approach is made. The heat and mass transfer components in the

Unreac ted fue l layer

Su

rfa

ce

lay

er

ion fro nt

xU

Time or Length of grate

Ch ar la ye r

Primary air

Final report May 1999


Page 17 of 79

reaction layer are qualitatively modelled by the radiation across the reaction layer and adjusted by fitting parameters. The fitting parameters give a qualitative description of the heat flow through the reaction front, the heat flow through single fuel particles in the reaction front and the consumption of oxygen and gaseous fuel constituents. The char layer is modelled in more detail. The fuel layer is divided into a number of sublayers where the char particles are assumed to burn with a shrinking core. It is assumed that the reaction rate is controlled by diffusion of oxygen from the gas between the particles through the ash layer that builds up on the particle surface into the char core. The height of the fuel layer is modelled by the continuity equation in one dimension. The height of the fuel layer decreases due to particle shrinking during the conversion of the fuel particles and rises due to the reduction of the velocity of the fuel layer along the grate. The simplifications mentioned make it possible to describe a model, which produces the necessary input to the free-room model at a sufficiently short calculation time, and gives the local condition for the nitrogen chemistry. Due to the fitting parameters of the reaction front some knowledge is needed about the combustion behaviour of the fuel layer on the grate, especially estimation on the time when the reaction front reaches the surface of the grate. The fuel layer model is validated numerically by a molar and an enthalpy balance over the fuel layer. Input data to the fuel layer model are given by the operation of the boiler, by the incident heat flux from the free-room above the fuel layer and of the fitting parameters in the fuel layer model itself. Data given by the operating condition are: type of fuel, fuel feed rate, initial height of the fuel layer, primary airflow and velocity distribution of the fuel layer along the grate. The incident heat flux from the free-room is given to the fuel layer model as an equivalent temperature distribution of the radiation, given by the free-room calculation. Fitting parameters are: relation between the actual heat flux and the heat flux given by radiation over the reaction front, maximum heat flux into single particles, mixing rate of gas inside the fuel layer and minimum voidage in the fuel layer. The fuel layer model gives the flow rate, composition and temperature of the gas leaving the fuel layer. It also gives the height, mass loss, and temperatures inside and on the surface of the fuel layer and the position of the reaction front inside the fuel layer. In the Trollhttan boiler measurements of the temperature on the metal surface of the reciprocating grate were made and these measurement has been used to choose the fitting parameters in the fuel layer model. 4.1 Description of fuel layer model Four layers; the unreacted fuel layer, the reaction front, the char and surface layers characterise the bed, Figure 4.1. The unreacted fuel layer consists of the fuel between the grate and the reaction front. The reaction front is located as the area where devolatilization takes place. The char reaction layer is the layer downstream of the reaction front. This layer is not influenced by the conditions above the bed surface. The surface layer is the layer downstream of the reaction front and the char layer, and this surface layer is influenced by the conditions above the bed surface. By assuming that no heat transfer takes place in the unreacted fuel layer and that the reaction front is thin, the behaviour of the bed can be described by the velocity of the

Final report May 1999


Page 18 of 79

reaction front and by the heat balance over the char and the surface layers. To simplify the calculation, the velocity of the reaction front is calculated for a dimensionless height of the reaction front, xU. This height is unity at the surface of the bed and zero at the grate. The velocity of the reaction front is obtained by dividing the mass release per unit area in the reaction front, mr/t, with the amount of fuel per unit area. The amount of fuel per unit area is calculated from initial mass of fuel per unit area, given by the fuel density F, and initial packing of the bed 1-0, and the bed height under the condition that the fuel keeps its size and that the packing is given by the initial bed height xb0 and the ratio between the initial and current velocity of the bed along the grate ub0/ub. If the fuel maintains its size and the bed keeps its initial packing the solution of the continuity equation in one dimension gives that the height of the bed change proportionally to the ratio of the initial and actual velocity along the grate ub0/ub. The change in bed height becomes x' U 1 ub mr = t F 1 0 x b 0ub 0 t

(4.1)

The actual position of the reaction front xu, is the dimensionless height of the reaction front and the bed height under the condition that the fuel maintains its size but nots its packing, xb. xU = x' U xb (4.2)

The solution of the continuity equation in one dimension for the bed height under the condition that the fuel keeps its size but not its packing is, xb = x b 0 ub 0 1 0 ub 1

1 6 0 5

(4.3)

The energy balance over the char layer consists of two heat sources; heat from the reaction front, qr, and heat from the char combustion, qc.

01 5 c
c

pc

Tc qc + qr = x s + xc xrad t xc

(4.4)

T is temperature, density, cp heat. In the initial stage when the layer downstream the reaction front, surface plus char layer xs+xc, is thinner than the penetration depth of the radiation from above the bed into the bed, xrad, no char layer is defined. This initial stage is modelled by a Heaviside step function . The energy balance over the surface of the bed consists of: radiation from the surrounding, q0,rad, char combustion, qs, heat convection from the gas, qsg, and in the initial stage, when the char layer is not developed, heat from the reaction front, qr.

01 5 c
s

ps

Ts q0 ,rad + qs + qsg + qr x rad x s + x c = t xs

(4.5)

Final report May 1999


Page 19 of 79

4.1.1 Reaction front The temperature in the reaction layer is in the model defined by the temperature in the char layer and therefore is the heat produced or consumed in the thin reaction front in he reaction front is immediately transported to the char layer. The gas and the char are assumed to leave the reaction front at the temperature of the char layer (in the beginning when the layer downstream the reaction front is thinner than the penetration depth, of the radiation this temperature is equal to the surface temperature). The mass release per unit area in the reaction front is equal to the heat flux per unit area to the unreacted fuel layer, qUeff, divided by the heat per unit mass needed for heating up the fuel to the devolatilization temperature and to evaporate the moisture, QU: mr qUeff = t QU (4.6)

The reaction front can be divided into four zones see Figure 4.2. The first zone is where the surface of the particles is heated up to the evaporation temperature. Heat is transported to the particle surfaces in this zone by radiation and conduction from the warmer zones downstream and heat is transported away from the particle surfaces by convection to the gas. The convective heat transfer coefficient is high and the gas temperature rises to approximately the same temperature as that of the particle surface. The second zone is dominated by drying. In this zone the gas flow out from the particle reduces the influence of the convective flow and the radiative and the conductive heat flow transfer the heat. The third zone is defined by the devolatilization. Dependent on particle size, there will be a period of simultaneous drying and devolatilization. In this case the gas flow out from the particles is high and the convective heat transfer is small like in the second zone. In this third zone ignition of the gas takes place and the gas temperature rises fast. If there is a substoichiometric condition in the reaction front, the temperature of the gas becomes higher than that of the char layer. The radiative heat flux of the gas can increase the heat flux to the particle surfaces in all zones, and the surface temperature of the particles can reach a higher temperature than the temperature in the char layer. The fourth zone is at the very end of devolatilization. In this zone the gas flow from the particles is low, and if there is oxygen left, char combustion will start. The convection is once more important, and the temperature of the gas and the particles converge to the same temperature.

Final report May 1999


Page 20 of 79

Water Volatiles zone 1


Mass loss

zone 2

zone 3

zone 4

Length along reaction front

Figure 4.2. The mass loss of moisture and volatiles in the reaction front. The figure is reproduced from figure of Saastamoinen [1998] The maximum radiative and conductive heat flux which evaporates moisture and devolatilizes the particles is proportional to the difference between the highest temperature in the reaction front, Tr, and the temperature in the unreacted fuel layer, TFs, minus the heat needed to heat up the primary gas, qUg, to the evaporation temperature, Tdry . The true heat flow for evaporation and devolatilization is smaller. It is assumed to be equal to the maximum heat flux reduced by a coefficient k. The model does not calculate the temperature profile of the gas and the solid in the reaction front and the maximum heat flux is not calculated. The heat flux is therefore related to the radiative heat flux between the char layer and the unreacted fuel layer, and a fitting coefficient k1 is introduced. qU = k max Tr

 4 1 6 

4 TFs +

eff max Tr TFs x

2 1 6

7 q

Ug

4 = k1 Tc 4 TFs qUg

(4.7)

is emissivity, Stefan-Bolzmanns constant, heat conductivity and x is the distance between the position of the highest temperature in the reaction front and the unreacted fuel layer. The reaction front can only go in one direction and therefore has the heat flux be restricted to a positive value. qU =

%q &0 '

qU > 0 qU 0

(4.8)

Due to the low heat capacity of the gas and the high convective heat transfer coefficient in the bed the gas attains the same temperature as the fuel, and the heat needed to heat up the gas from this temperature to evaporation temperature is qUg = Ji Ii Tdry Ii TFs
i

(4.9)

Final report May 1999


Page 21 of 79

J is molar flow per unit surface area, I enthalpy, subscript i indicates the different species, and dry stands for drying.. The mass release per unit area given by Equation (4.6) assumes heat transfer control of devolatilization, but if the temperature decreases the devolatilization will be controlled by kinetics. It is also assumed that there exists a heat source by the oxidation of the volatiles leaving the particles. So, there is only a reaction front if the temperature is high enough for devolatilization. An empirical function is therefore introduced to describe the transition between kinetic and heat transfer controlled devolatilization. When the devolatilization is controlled by kinetics, the rate is low and fr 0, but when it is controlled by heat transfer fr1. The empirical function is based on the kinetics of devolatilization: fr Tc = exp E / RTc k2 + exp E / RTc

(4.10)

E is activation energy, R gas constant and k2 is an empirical constant which determines the position of the temperature range of devolatilization. At high temperatures internal heat transfer in the bed particles controls the mass release. The maximum internal heat transfer can be calculated using the Biot module, Bi. By limiting the maximum heat flow for the drying and devolatilization, qmax, the fitting constant k1 in Equation (4.7) can be optimised for the temperature range between the extremes. The effective heat flux for drying and devolatilization is restricted by the heat flux from Equation (4.7) qUeff =

%q &q '

f Tc ; qU fr Tc < qU max Bi qU fr Tc qU max Bi U max Bi ;


U r

(4.11)

The limiting maximum heat flow is at present adjusted as a fitting coefficient. The heat per unit mass needed to evaporate the moisture and to heat up the fuel to the devolatilization temperature is QU = YH2 O I H2O( g ) Tdry I H2O( l ) TUs / M H2O + Yvol Ivol ( s ) Tvol Ivol ( s ) TUs / Mvol ( s ) + Ychar Ichar Tvol Ichar TUs / Mchar + Yash
ash

2 2I

Tvol Iash TUs / Mash

(4.12)

Y is mass fraction of fuel, M molar mass and subscript vol stands for volatiles. The estimation of the composition and enthalpy of the volatiles are presented in Appendix A. In the reaction front the solid and the gas leaving the particles are heated up further to the temperature of the char layer. The heat per unit mass needed is:

Final report May 1999


Page 22 of 79

Qr =

YH2O I H2O( g ) Tc I H2O( g ) Tdry / M H2O + Yvol Ivol ( g ) Tc Ivol ( s ) Tvol / Mvol ( g) + Ychar Ichar Tc Ichar Tvol Yash
ash

2 2I

Tc Iash Tvol / Mash

/ Mchar +

(4.13)

The assumption that no heat transfer takes place in the unreacted layer gives the energy needed to rise the temperature of the primary air to the temperature of the char layer: qrg = Ji Ii Tc Ii TUg
i

(4.14)

The major heat source in the reaction front is oxidation of hydrocarbons and carbon monoxide. The entire set of reactions of the major species in the bed is modeled as follows: CHm + 1 1 + m O2 CO + m H2 O 2 2 CO + 1 O2 CO2 2

4. I 4. II

m is the number of hydrogen atoms per carbon atom in the hydrocarbons. It is assumed that the local mixing controls the reaction rate according to the concept of Magnussen and Hjertager (1977), which states that the reaction rate is controlled by the minimum concentration of oxidant, reactant or product. Reactions that produce the products, carbon dioxide and water, are already ongoing during the devolatilization process, and the products never restrict the reaction rate. The reaction rate is therefore controlled by the minimum concentration of oxygen or fuel. The gas leaving the particles changes in the reaction front, both in quantity and in composition, from water rich to volatile rich. To simplify the descriptions of gas flow and composition in the reaction layer, the quantity and the composition of the gas leaving the particles are assumed to be constant along the entire reaction front. The Magnussen and Hjertager concept is described by the mixing rate, b, and the minimum of the oxygen or the fuel. The mixing rate is assumed to be proportional to the turbulent kinetic energy divided by the dissipation of turbulence. Inside the bed the flow can not be turbulent due to the narrow channels between the particles. The local mixing is therefore not created by the turbulence but of the impulse force of the gas flowing out from the particles. For this reason the local mixing rate is assumed to be proportional to the gas flow out from the particles. b = jk3 (4.15)

j is gas flow per unit time from the particles and k3 is a coefficient to be determined. The total molar flow out from the particles, jtr, tr is the residence for the gas in the reaction front is given by the mass loss in the reaction front per unit area.

Final report May 1999


Page 23 of 79

jtr =

mr Yvap / MH2O + Yvol / Mvol ( g ) t

(4.16)

The residence time in the reaction front can be estimated by assuming the width of the reaction front xr, where the J0+1/2jtr is the mean molar flow in the reaction front. tr = x r P 1 J 0 + 2 jtr RT (4.17)

The change of the molar flow of oxygen, JO2, and fuel, JCH4, JCO, is the difference between the total molar flow out from the particles and what is consumed by reactions. JCH m t = jXCH m b min( JCH m , JO2 ) J0 + jt

JCO b min( JCO , JO2 ) min( J CHm , JO2 ) = jXCO t J0 + jt JO2 t = b J0 + jt

8 8

(4.18)

3 01 + m5 min( J
1 2

CH m

, JO2 ) + 1 min( JCO , JO2 ) 2

t is the local residence time in the reaction front, t=0 when the primary air enters the reaction front and t=tr when the gas leaves the reaction front, J0+jt is the total molar flow at time t and X is molar fraction. An analytical solution of the system of differential equations, Equation (4.18) is favourable in order to speed up the calculation. Such a solution can be found under the following conditions,

43 J 43 J 43 J

CH n

JO2 & JCO JO2 > JO2 > JO2


CO

CH n

CH n

8 3 8& 3 J 8& 3 J

JO2 > JO2

CO

89 89 89

(4.19)

These conditions are fulfilled in all situations of interest for combustion of volatile gases from wood. The only case when condition (4.19) is not fulfilled is when the fraction of carbon monoxide in the volatile gases is not much smaller than the fraction of hydrocarbons. The analytical solution is presented in Appendix B. The production of reaction products is: J H2O = J CO2 =
m 2 2 m

3J

CH m

0 + XCH m jtr JCH m tr

J H2O + JCO 0 + XCO jtr J CO tr

(4.20)

The energy release of reaction is: q R = Ji Ii Tc


i

(4.21)

Final report May 1999


Page 24 of 79

Subscript R stands for reaction. The heat flux from the reaction layer to the char layer is then obtained from the Equations (4.6),(4.12), (4.13), (4.14) and (4.21) qr = q R qrg QU + Qr

6 mt

(4.22)

4.1.2 Char and surface layer During the drying, devolatilization and char combustion, the size of the particle changes; wood chips shrink by volume with 10 to 20% of the initial fuel size during drying, with 40 to 60% of the dry fuel during devolatilization and with 95 to 99% of the devolatilized fuel during char combustion. The shrinkage factors of drying, Sdry, devolatilization, Sdev, and char combustion, Schar, describe the shrinkage. They are defined Sdry = 1 Vdry / V0 Sdev Schar
dev ash

2 = 21 V = 11 V

/ Vdry

7 /V 6
dev

(4.23)

Vdry is the volume of the dry fuel, Vdev volume of devolatilized fuel and Vash volume of ash-For spherical particles with the same shrinkage in all directions, the char particle diameter is a function of the moles of char divided by the initial moles of char Xc. d = d0 1 Sdry 1 Sdev 1 Schar 1 Xc

42

71

62

679

1/ 3

(4.24)

The thickness of the surface layer, xs, is the minimum of the penetration depth of the radiation coming from downstream the bed surface, xrad, and the thickness of the bed layer downstream the reaction front, xs+xc: xs = min xrad , x s + xc

6 71 6 I 21 S 11 X 67dx
xb char char xU

(4.25)

The thickness of the bed layer downstream the reaction front is: xs + xc = xb xU 1 Sdry 1 Sdev

62

(4.26)

Equations (4.4) and (4.5) re not defined initially when xs and/or xc are equal to zero. Therefore it is necessary to limit these quantities in the calculation to a small value. The penetration depth of the radiation flux, qrad, is defined as the length xrad where 95% of the radiation is attenuated,

Final report May 1999


Page 25 of 79

qrad = exp q0,rad

 

x rad

I
0

1 dx = 1 0.95 lm

 0 

(4.27)

q0,rad is the radiation fluxes incident on the bed surface, x=0, from above and lm is the mean path-length of radiation. The mean path of radiation is defined as lm = 1 n p Ap (4.28)

np is the number of particles per unit volume and Ap is the mean projected area of the particles. For spheres with the diameter d, the mean path-length is, Hernberg et al. [1993]: lm = 2 d 3 1 (4.29)

Downstream of the reaction front the char combustion starts. The char combustion is modelled by a single reaction: C + nO2 2 2n CO + 2n 1 CO2

4.III

n are the moles of oxygen consumed for one mole of carbon. If the reaction is controlled by the diffusion of oxygen to the particle, the reaction rate becomes, CO2 t = CO2 x CO2 = ug = CO2 eff As x xU t x x (4.30)

C is concentration, mass transfer coefficient to the external surface of the particle, As specific surface area of bed particles, ug gas velocity. The step function restricts the char combustion to downstream of the reaction front. Since the char combustion is assumed to be diffusion controlled, the model can not describe the extinction of the char combustion at a high flow of primary air. The introduction of a reaction rate for the char combustion does not solve this problem, because the temperature in the reaction front is not modelled. Setting a limiting temperature for the char combustion can make the error smaller. The effective mass transfer coefficient, eff, is the external mass transfer coefficient including also mass transfer trough the ash layer built up on the particle surface during char combustion,

eff CO2 = CO2 CO2 ,sur =

DO2eff x ash

3C

O2 , sur

0 eff =

DO eff 2 x ash

DO2eff x ash

(4.31)

DO2,eff is the effective diffusivity in the ash layer. Because the porosity in the ash layer is close to unity, the effective diffusivity is about the same as the diffusivity of the gas. The thickness of the ash layer xash is:

Final report May 1999


Page 26 of 79

x ash =

1 2

 d d 421 S 711 S 6X 9 
1/ 3 0 dry dev c 1 3

(4.32)

The Nusselt number, Nu, for the channels between the particles inside the with a hydraulic diameter, dh, and a channel length l becomes

 Pe  Nu = 1.615   
l dh

(4.33)

The Peclet number, Pe, is Pe = ug g dh c pg (4.34)

and the hydraulic diameter and channel length is dh = l= 2 d 3 1 (4.35) (4.36)

d 2

The mass transfer coefficient is calculated by assuming that Nusselts number is equal to Sherwoods number,
3 3 DO2 c pg ug g 1 d Nu = Sh = h = DO2 g d 2

 

0 5 

1/ 3

(4.37)

The specific surface area of bed particles is: 6 1 As = d

0 5d  d
c

(4.38)

dc is the char core diameter. dc = d0 1 Sdry 1 Sdev Xc

42

71

6 9

1/ 3

(4.39)

The gas velocity in each bed layer is given by the molar flow of each species: RT Ji
i

ug =

(4.40)

P is pressure. By dividing the bed layer into z sub-layers and assuming that gas velocity, mass transfer coefficient and specific area are constant in each layer, Equation (4.30) can solved for a height element xi,

Final report May 1999


Page 27 of 79

CO2 ,i = CO2 ,i 1 exp

 

eff i Asi xi f1 i ugi

 

(4.41)

f1[i] is the step function, , integrated over the element xi. f1=0 if the element is located upstream of the reaction front and f1=1 if the element is located downstream the reaction front, see Figure 3.
Reaction front 1

f, 1
0 x1 x2 x3 x4 x5 x6

Figure 4.3 delta function integrated over element xi If the reaction front is inside the element, f1 gives the relative position of the reaction front in the element. The position of the element i is given by the summation of the elements j=1 to j=i, where iz, this can be expressed

f1

% 0 K K    K i = &  x  x  / x K    K 1 K '
i j U j =1

x
j =1 i 1 i j =1

< xU
i

x j xU x j
j =1

(4.42)

xU < x j
j =1

i 1

The thickness of the active char layer elements xi is f1 i > 0 xi =

xb 1 f1 i Sdry 1 f1 i Svol 1 Schar 1 Xchar ,i 1 f1 i z f1 i = 0

71

64

3 2

67 / f i 89
1

(4.43)

xi = 0

Final report May 1999


Page 28 of 79

The consumption of oxygen in the active char layer element is: JO2 ,i = JO2 ,i 1 exp and of char: nC ,i 1 = JO2 ,i t n Xchar ,i 1 JO2 ,i = t n nC ,i 0 The initial moles of char in the active char layer element is: nC ,i 0 = xb Ychar F 1 f1 i z Mchar

   

i Asi x i 1 ugi

   

(4.44)

(4.45)

0 5

(4.46)

The bed temperature in the sublayer Ti is the same as for the char layer when x is smaller than xs and Ts when x is larger than xs. For the bed sub-layer that includes the border between the char and the surface layers an integrated mean temperature is used, Ti = Tc f2 i + Tc 1 f2 i

6
x
j =1 i i

(4.47)

f2 is a step function integrated over an element in the same way as Equation (4.42) , with the difference that f2 is 1 below the surface layer and 0 above.

f2

% 1 K K    K i = &1   x  x  / x K    K 0 K '
i j s j =1

< xs x s x j
j =1 i

x
j =1

i 1

(4.48)

x s < x j
j =1

i 1

The heat release during the char combustion is: qi = nC ,i IC Ti nIO2 Ti 2 2n ICO Ti 2n 1 ICO2 Ti t

(4.49)

The heat release in the char and the surface layers is:

Final report May 1999


Page 29 of 79

qc = qi f2 i
i =1 z

qs = qi 1 f2 i
i =1

6 7 0 5

(4.50)

The heat corresponding to the temperature change of the gas between the char layer and the surface layer is: qsg = Ji Ii Ts Ii Tc
i

(4.51)

The density and heat capacity of the fuel in the char and in the surface layer are:

i =
c pi =

Xc ,i Ychar char + Yash ash 1 Xc,i Ychar + Yash Xchar ,i Ychar c pchar Ti + Yash c pash Ti Xchar ,i Ychar + Yash

(4.52) (4.53)

Subscript i indicates the char and the surface layer. 4.1.3 Packing of bed The packing of the bed changes with time. For wood chips the bed voidage is highest in the beginning and decreases during the combustion process. A major reason for a change in the bed voidage is a change of velocity of the bed along the grate. While the velocity of the bed decrease along the grate and the pressure forces inside the bed is built up. These pressure forces decrease the bed voidage and increase the height of the bed according to the continuity equation. If it is assumed that the bed voidage is proportional to the change in velocity of the bed along the grate, and the initial and minimum bed voidage are known, the bed voidage at a certain position on the grate becomes: v = min + 0 min

6 11uu uu 66
b b min b0 b min

(4.54)

ubmin is lowest velocity of the bed along the grate, and the minimum bed voidage is reached at this velocity. 4.2 Description on free-room calculation and CFD Software used The software used for the calculations of the free-room is Fluent/UNS 4.2.5. This software package has built in the necessary physical models to simulate gas phase combustion (turbulent fluid flow, chemical species mixing and reaction, conductive, convective and radiant heat transfer). The package includes software for domain definition and discretization, and may be integrated with a nitrogen oxide reaction module. It also contains some bugs, which will be discussed later in this report.

Final report May 1999


Page 30 of 79

The software uses a finite volume discretization scheme for solving the equations that define the problem. The equations are: The continuity equation, The Navier-Stokes time averaged equations, The transport equations for the various chemical species, The transport equations for turbulent variables (k and ). The discretization transforms each partial differential equation in a set of linear equations (one for each cell), which is solved not inverting the set but using an iterative tool called Algebraic Multigrid Linear Equation Solver. The steps in the problem set-up and solution procedure are the following: definition of the geometry of the volume (domain) where the equations will be solved, discretization of the domain in a number of control volumes (cells), choice of the physical models to be applied, definition of the boundary conditions, choice of the numerical methods and parameters, iteration, refinement (adaption) of the cells structure (grid) in the critical areas, iteration, post processing, solution analysis.

4.2.1 Numerical scheme To illustrate the discretization of an equation a simple one-dimensional diffusion equation is used. The differential equation is written:

( u ) = + S x x x

(4.55)

After integration over a control volume (cell) and discretization the equation is written as:

J e e J w w = ( e

E P P W w ) A + S V x e x w

(4.56)

Where E, P and W are the values of the variable in the centroids of the cells, e and w are the values on the surfaces of the cell, Je is the mass flow through surface e, A is the cross-sectional area, and V is the volume of the cell, S is a source term and x is the distance between two cells, as shown in Figure 4.4.

Final report May 1999


Page 31 of 79

xw W w P

xe E e

Figure 4.4 One-dimensional discretization of equation The UNS code stores the value of the variables at centroids, and interpolates the surface values. The interpolation can be performed in three ways: assigning to the variable on the surface the value of in the upstream cell (first order upwind discretization scheme) computing the value as a second order Taylor expansion from the centroid of the upstream cell. The value of the gradient is computed satisfying the divergence theorem. (second order upwind discretization scheme) solving the Peclet equation, which is an interpolation between the cells centroids based on the features of the flow. It assigns to the surface the value of the upstream cell centroid if the flow is dominated by convection, it computes a linear interpolation if the flow is dominated by diffusion and uses an intermediate value otherwise (power law discretization scheme). The choice of the discretization scheme affects the solution in terms of the precision obtained, but it also affects the convergence speed of the iterative process. When it is possible, it is suggested to use the second order discretization scheme. For non-reacting flows the second order discretization scheme may be applied successfully, but in the present case, though, when fast chemical reactions such as combustion are included in the flow, the algorithm is not stable enough to converge, not even slowly. Even with the first order discretization scheme, reacting flows cause convergence problems, but they can be overcome using a stronger underrlaxation than the default. The pressure-velocity-coupling algorithm is the one suggested for transient flows (SIMPLEC). The equations are solved separately, beginning from the momentum equations, following with the continuity equation, and then the other scalar equations. The convergence criteria used are the default ones of the code, which are quite strict. The criteria are satisfied when the normalised residual for a variable has decreased to a fixed percentage of the initial normalised residual. The normalised residual of a variable is defined as the sum over the entire domain of the modulus of the residuals of the linearized equation, divided by the total flow of the variable through the domain. The computer used has a Sun Ultra processor and 500MB of RAM memory, and it takes approximately 7 min to perform one iteration in the complete case (about 300000 cells, 12 independent variables, and the radiation model). 4.2.2 Domain discretization The discretization of the domain is performed using the P-Cube software, trying to achieve grid independence, which is accomplished when a further refinement of the

Final report May 1999


Page 32 of 79

grid does not cause any significant change in the solution. The grid used is the hexahedral one, since the geometry of the boiler can be easily represented as a set of rectangular prisms and pyramids. Section of the grid is shown in Figure 4.5.

Figure 4.5 Section of grid An excessive refinement of the grid must be avoided in order to limit the computation time. The correct approach is therefore to refine the grid only in the areas where strong gradients of important variables are present (adaption). The gradients that that controls the reacting flow are the ones of velocity and of oxygen concentration, those have very high values on jets and on reaction fronts. By refining the grid in the volumes occupied by the jets, a more accurate description of the velocity field increases the penetration of the jet and improves the description of the flow around them. A typical procedure to obtain grid independence is to refine the grid in the jet area until the penetration of the jet does not change with a further refinement. Refining the grid in the reaction front area could increase significantly the quality of the description of combustion where NOx reactions take place, therefore improving the estimation of the NOx production. Adaption of the grid is also useful to obtain a grid that matches the size of the boundary layer, for reasons discussed further on. The adaption tool unfortunately doesnt work in simulations where the Discrete Transfer Radiation Model is used, due to a bug described in paragraph 4.2.4.2. Other criteria to be observed in the production of the grid are the alignment of the grid to the flow and the minimisation of cells skewness. Alignment of the grid to the flow can be obtained only where the flow field has a simple and obvious direction, i.e. where the air and gas jets are located. Extreme cell skewness can be avoided if the geometry is not too complex, otherwise it may be better to use a tetrahedral mesh.

Final report May 1999


Page 33 of 79

4.2.4 Physical models for fluid flow The program solves the momentum equations and the continuity equation for time averaged variables. The Reynolds stress tensor is considered proportional to the laminar stress tensor through a multiplying factor, which is a function of turbulent kinetic energy and its dissipation rate. Two more balance equations are solved for the turbulent kinetic energy k and its dissipation rate , the so-called k- closure of the turbulent problem. At the wall boundaries, the set of equations is not valid, because the hypothesis of homogeneity and anisotropy of the turbulence, which is at the basis of the model, is no longer valid. Therefore, empirical functions are used to describe the variables in the cells adjacent to the wall. In the present case the description of the flow in the vicinity of the walls is a matter of secondary importance, since the most important mechanism of heat transfer is radiation, due to the high temperature difference between the flame and the tube walls, and the production of k and at the walls is negligible compared to the production in the secondary air and the recirculation jets. For this reason the computational resources are not used to obtain a correct representation of the boundary layer. The code offers three options for modelling the flow near walls, the Standard Wall Function, the Non Equilibrium Wall Function and the Two Layers Zonal Model. The first two are functions that describe the profile of the variables on an empirical basis instead of solving a set of equations, the third option is a model that solves a modified set of equations inside the boundary layer. This last model requires many cell layers inside the near wall area and has to be excluded because of the large computation time requirements. The Non Equilibrium Wall Function is a model intended for flows with strong pressure gradients, which is not the case in the boiler studied. Therefore the simpler Standard Wall Function is employed. This approach assumes that the production of turbulent kinetic energy is equal to its dissipation rate (local equilibrium hypothesis). The transport equation for k is solved, while is computed using an empirical function. The velocity function is linear in the viscous sublayer (y*<11) and logarithmic outside (y*>11). The velocity function is valid in the boundary layer (y*<60), but it is applied to the whole adjacent-wall cell, so if this cell is larger than the layer, an error is introduced in the calculation of the velocity in this cell. Producing a grid that satisfies this condition on all walls is possible only for simple flows such as one in a duct, but it is practically impossible in a complex flow like the one in the furnace analysed. The size of the cells that would fit in the boundary layer is much smaller these of the rest of the domain, and to represent the boundary layer it would be necessary to decrease the size of the cells at the walls. Since the boundary layer is considered to be of secondary importance, this requirement is not fulfilled, therefore introducing an error. The error involves not only the velocity in the walladjacent cell, but also the temperature, since also the temperature field is not solved but determined by an empirical function. The temperature may be underestimated (if the cell is larger than the boundary layer) because in the laminar sublayer and in the transition zone the heat transfer is lower than in the turbulent region. However, the thermal boundary layer thickness is in general different from the velocity boundary layer, so the error would exist even if the grid was built with a thickness of y*=60. To verify whether the error affects significantly the solution, a pair of two-dimensional case was studied where all conditions are identical with the exception of the grid in the vicinity of the walls. In one case the dimension of the cells is such that the condition 30<y*<60 is satisfied almost everywhere, while in the other case these cells

Final report May 1999


Page 34 of 79

have a dimension similar to that of the neighbouring cells, which is much larger. The results from the two cases show that the difference between the heat powers transmitted to the walls is negligible, confirming that the main heat transfer mechanism is radiation, and that the thickness of the cell layer does not affect the heat power significantly. All the fields were compared in a qualitative way, with particular attention to the velocities and turbulent quantities, and no important differences appear. For the final tri-dimensional case it was therefore decided not to refine the grid near the walls. The boundary conditions are set as follows: the secondary air and recirculation gas inlets are defined as surfaces with constant velocity over the area. the turbulent energy and the dissipation rate of the jets entering the combustion chamber are computed using the formulas available in the Fluent manual for flows in a duct. the outlet is defined as a surface at the constant pressure of 1 bar (the rest of the boiler has a slightly higher pressure). on the upper surface of the bed (the inlet to the combustion chamber) the velocity is a function of the distance from the fuel inlet. (The function is an output of the bed model) the turbulence exiting the bed is quite difficult to estimate, but it is reasonable to say that the fuel layer has an effect similar to a porous medium. Therefore the flow should have the characteristics of weak grid turbulence. The importance of this boundary condition, however, is limited, since the turbulence of the recirculation jets reaches the upper surface of the bed, so that the contribution from the bed itself would be negligible. at the walls the boundary condition is set by the wall function.

4.2.5 Physical models for heat transfer The conservation of energy is satisfied solving the thermal energy equation:

ui T Dp + ( ik ) eff + Sh ( h) + ( ui h) = (k + kt ) h j Ji j + t xi xi xi xi x k Dt

(4.57)

Where h is the specific enthalpy and the terms on the right hand side of the equation represent: conduction (kt is the conductivity due to turbulence) enthalpy transfer due to diffusion of species (Jij is the diffusive flux of species j in direction i) the reversible work irreversible dissipation of kinetic energy due to friction (probably negligible if compared to the turbulent dissipation rate) a source term that contains the effect of radiation.

Final report May 1999


Page 35 of 79

The Fluent/UNS code includes two options to model radiation, which are called Discrete Transfer Radiation Model (DTRM) and P1 model. Both models use a weighted sum of grey gases to determine the emissivity of a layer of fluid as a function of the concentrations of CO2 and H2O and of the thickness of the layer. This function is a polynomial of temperature fitted to experimental values. The contribution of soot to the emissivity of the fluid cannot be included in Fluent/UNS. The role of soot in radiation is quite important and the neglecting of soot may produce an underestimation of the radiant power transmitted to the bed.

Figure 4.6 Rays and clusters, as used in the DTRM mosel. The DTRM describes the radiation exiting from a surface as a finite set of rays that transport the energy to the cells that they cross, see. Figure 4.6. The variation of power along a ray is expressed by the following equation:

I aT 4 + aI = s

(4.58)

where I is the power flowing along the ray, a is the absorption coefficient per unit length of the fluid and is the Stefan- Boltzmann constant. The volume discretization of the DTRM is coarser than the discretization for the solution of the other equations. One single DTRM cell (called a cluster) is actually a group of 30 cells of the grid. The clusters are defined before the iterative process starts, together with the rays. A problem that arises during the preparation of the rays and clusters is the lack of an instrument capable of evaluating the effectiveness of the representation. For example it is not possible to verify whether each cluster is actually crossed by at least one ray, and therefore whether it can actually exchange radiant heat with the neighbouring cells. Furthermore, when defining the rays (ray tracing phase) a bug appears. This bug causes the rays to cross the boundaries of the domain, ending in a not defined space. This error is limited to a few rays in the case of a basic grid but increases highly for adapted grids. The result is that the DTRM cannot be used on refined grids. This problem strongly limits the usefulness the code, because it should be possible to perform the grid adaption in order to obtain a solution that has a sufficient

Final report May 1999


Page 36 of 79

precision. An effort was made to produce a structured grid already refined in the critical areas, but the skewness increased, drastically slowing down the convergence. In the P1 model the radiant heat transfer is expressed by the following balance equation of incident radiation:

(G ) aG + 4aT 4 = 0

(4.59)

where G is the incident radiation flux and is a function of the absorption coefficient. The equation has the shape of a diffusion equation and therefore this kind of model works well for optically thick media, where radiation travels small distances. In the case treated the optical thickness of the medium is small, but it is interesting to apply the model anyway, since the DTRM does not work properly. However, the result is not satisfactory because the algorithm needs a very high number of iterations to reach convergence. For this reason it is not practical to use this model for tridimensional cases. Using the P1 model some information can be obtained on the effect of grid adaption. The main difference that appears when adapting the grid in the jet zones is a slight increase of the penetration of the jet. Adaption on the reaction front was not performed. The boundary conditions are set as follows: Temperature of the refractory walls: 900K Temperature of the tube walls: 500K Emissivity of all walls: 0.7 (using this value, the computed heat power transmitted to the walls equals the measured one) Emissivity of the bed: 0.9 Emissivity of the other boundaries: 1

Reflection and emission are considered to be diffuse for all surfaces, the values of temperatures are taken from measurements in another plant, and the variation of temperature along the tubes, due to heating of the water is neglected. Anyway the lower temperature plays a minor role in the determination of the radiant flux, since it depends on temperature to the power of four. 4.2.6 Physical models for reaction kinetics In the studied case the reactants enter the combustion chamber separately, so that a mixing is required before the reaction can take place. The mixing rate can be the limiting factor of the combustion process, and therefore it is necessary to estimate this rate in the various cells. To perform this estimation, the code uses a model built on the Eddy Dissipation Theory of Magnussen and Hiertager [1989]. This model assumes that the last step of the turbulent energy cascade (the dissipation of kinetic to thermal energy) takes place in fine structures where the mixing occurs at a molecular scale, making it possible for the various species to react. The quantity of fuel that may react per unit time is therefore a function of the flow of reactants that enters the fine structures. The model of Magnussen and Hjertager produces two expressions for the generic reaction B+C->D+F :

Final report May 1999


Page 37 of 79

k r ( B) = const 2 ([ D] + [ F ]) k
r ( B) = const 1 ([ B] + [ C ]) These expressions are compared to the Arrhenius rate:
1 1 1 1 r ( B) = A[ B] [C ] [ D] [ F ] exp(

(4.60) (4.61)

E ) RT

(4.62)

and the lower (limiting) one of the three is used by the code. The resulting production or consumption of chemical species is introduced in the transport equations of the various species as a source term. In cases where the chemical kinetics are very fast, it is not necessary to calculate the value of the Arrhenius expression, and some computational resources may be spared. In the present case, though, both the reaction rates of CO and CH4 are not fast enough. The second expression, which appears contradictory at a first look, is derived from a consideration that may be briefly summarised saying that the absence of products of reaction implies that a reaction has not occurred recently. Therefore radicals which are necessary for the activation of further reaction are not available. The concept cannot be applied to calculations were products of reaction are provided by other sources. In the present case for example, the products of reaction (H2O and CO2) are present in high concentrations already in the combustion air and in the recirculated flue gas entering the chamber, so the expression always assumes high values, preventing it from being the limiting factor. It may also be argued that such expression has a value of zero where there are no products of reaction, and hinders combustion even if there are reactants at a high temperature. In the numerical method anyway the residuals always provide a minimum presence of reactants, that produce the numerical ignition of the flow. Then the radical diffusion is modelled together with that of the reactants. 4.2.5 Simplified combustion schemes In the processes of combustion and pyrolysis is involved a large number of chemical species that interact in a complex system of reactions. Therefore it is not possible to describe the complete mechanism in a fluid dynamic simulation of a complex flow and simplified combustion schemes are used which include a limited number of species and reactions. The main chemistry has been modelled by two different reaction schemes, and the results were compared. The first scheme is the one available in the database of the Fluent code for methane combustion. It includes the following reactions: CH4+3/2O2->CO+2H2O CO+O2->CO2 4.IV 4.V

In the database the carbon dioxide dissociation reaction was also included, but the rate of dissociation is negligible at the temperatures present in the furnace. Hydrogen

Final report May 1999


Page 38 of 79

is not included because this pyrolysis gas is not measured in the boiler, furthermore it is not very relevant to any of the reactions involving nitrogen compounds (in fact, its concentration does not appear in the reaction rate expressions). The second reaction scheme for the main chemistry is the one formulated by Jones and Lindstedt [1988] for methane combustion. It includes the following reactions: CH4+1/2O2->CO+2H2 CH4+H2O->CO+3H2 H2+1/2O2->H2O CO+H2O->CO2+H2 4.VI 4.VII 4.VIII 4.IX

In this case hydrogen obviously plays an important role, being the product of both methane and carbon monoxide partial oxidation, so it has to be included. This reaction scheme, however, proved to be inadequate to the conditions existing in this boiler, i.e. very high moisture content at all stages of combustion (the fuel has a very high water content and the combustion air is moisturised). Using this scheme causes all the methane and the carbon monoxide to be split by the water just above the grate, producing a high hydrogen concentration and a complete absence of other combustible gases in the combustion chamber. This clearly contradicts the measurements that instead show the presence of both carbon monoxide and hydrocarbons below the secondary air injection. Both combustion schemes lack the soot formation reaction. The inclusion of soot modelling in the combustion simulation would probably modify significantly the solution for two reasons. First the presence of soot increases the emissivity (and absorption coefficient) of the gases, making the flame more luminous and therefore increasing the radiant heat transfer, second it affects the rate of oxidation of the carbon (the soot oxidation is different from that of CO or CH4), displacing the flame to another position in the furnace. 5. The NOx modelling Like the modelling of the main chemistry, the NO x modelling is calculated separately for the fuel layer and for the free-room, where the fuel layer also in this case acts as a boundary condition for the free-room. The NOx add-in model available from Fluent describes the NOx processes in the free-room. This model is developed for coal combustion, and HCN is used as a NOx precursor and not NH3, which would be more appropriate for biofuel. On the other hand the add-in model includes well known reaction rates for formation of thermal NOx and for the formation or destruction of NOx by NH3. In the present case NH3 is assumed to be formed instantaneously from the HCN. A reduced mechanism developed for biofuel would be preferred, but the implementation and validation of such a model element into Fluent is too large a work for this project. The nitrogen chemistry inside the fuel layer is qualitatively modelled on the basis of a general knowledge of the nitrogen chemistry under different conditions, such as: HCN and NH3 forms NO at high oxygen concentration and temperature HCN and NH3 reduce NO at low oxygen concentration and high temperature

Final report May 1999


Page 39 of 79

CO reduces NO in presence of char if the oxygen concentration is very low the char acts as a catalyst for the reduction of NO by HCN and NH3

From experiments on single particles following can be assumed fuel nitrogen bound to the volatiles forms HCN and NH3 fuel nitrogen bound to the char residue forms NO around half of the fuel nitrogen is bound in the char residue, Lepplahti 1995, Saastamoinen 1997. Reaction kinetics can be found for NO reduction over char surfaces for different biofuels, Zevenhoven (1998), and for coal char, Johnsson (1990), which makes it possible to model the NOx formation and destruction in the char layer. What enters the char layer from the reaction front is unknown. The situation inside the reaction front is very complex, the volatiles are released inside the particles, probably as HCN and NH3 and already inside the particles some of it is converted to NO or N2. This gas mixture leaves the surface of the particle into a gas stream, which has an oxygen concentration between 0 and 21% and a temperature between 600 and 1500K. The reaction front is not modelled in detail and therefore a model based on kinetic data of various reactions can not be established. Instead a model based on split factors is developed. The first split factor gives the part of the fuel nitrogen that is released together with the volatiles. The first part of the reaction front, where the oxygen still not is consumed, is defined as an oxygen-rich region and the subsequent region of the reaction front, where the oxygen more or less is consumed, is defined as an oxygen-lean region. In the oxygen-rich region HCN and NH3 most likely form NO, and in the oxygen-lean region this NO to some extent is reduced by the HCN and NH3 leaving the particles. The second split factor defines the fraction of fuel nitrogen that leaves the particles in the oxygen-rich region. The third split factor defines the degree of conversion of HCN and NH3 to NO in the oxygen-rich region. The fourth split factor, the last of the split factors, is defined as reduction of NO by HCN and NH3, where the split factor is based on the lowest concentrations of NO and the sum of the concentrations of HCN and NH3. 5.1 Description of nitrogen chemistry model in fuel layer The model for the nitrogen chemistry in the fuel layer assumes that all NOx is fuel nitrogen. The nitrogen chemistry in the fuel can be separated into the reaction front and the char layer. The reaction front is also separated in to two regions dependent on the oxygen concentration. In the first region, the oxygen rich region HCN and NH3 are assumed to produce NO: HCN , NH3 + O2 NO 5. I

and in the second region, the oxygen lean region HCN and NH3 are assumed to reduce NO: HCN , NH3 + NO N2 + H2O + CO 5. II

Final report May 1999


Page 40 of 79

If all oxygen is consumed in the reaction front the reduction of the nitrogen oxides continues in the char layer above the reaction front and is catalysed by the char surfaces: HCN , NH3 + NO char N 2 + H2 O + CO NO + CO N2 + CO2
char

5. III 5. IV

If the oxygen not is consumed oxidation of NH3 and HCN to NO and N2 will take place. After the drying and devolatilization is completed and only char combustion take place. During the char combustion all fuel-bound nitrogen is assumed to form NO. This assumption give that in this area of the grate no NH3 or HCN are present and the only mechanism for the reduction of NO is reaction 5.IV. The nitrogen chemistry in the reaction front is modelled in three steps and the reactions are modelled by split factors that are put manually. is used for the split factors and subscript 0 indicates particle surface,1 oxygen rich region, 2 oxygen lean region and 3 char layer. The oxygen rich region is: x1 J N F ,dev = x1 J0 , N 2 + J 0 , NO + J 0 , NH3 + HCN J1, N2 = x1 J 0, N2 J1, NO = x1 J 0, NO + 1 x1 J 0 , NH3 + HCN J1, NH3 + HCN = 1 1 x1 J0 , NH3 + HCN x is the fraction of volatile gases entering the oxygen rich region. The oxygen lean region is modelled:

8
(5.1)

11 x 6J
1

J2 , N 2 = J1, N 2 + 1 x1 J0 , N 2 + 2 J1, NH3 + HCN + 1 1 1 x1 J 0, NH3 + HCN J2 , NO


1 0 , NO 1, NO 2 1, NH 3 + HCN 1 1

J2 , NH3 + HCN

1 6 = 11 x 6 J +J = 11 63 J
2

N F ,dev

= 1 x1 J 0 , N2 + J 0, NO + J 0 , NH3 + HCN

63

1, NH3 + HCN

1 61 6 3J + 11 611 x 6 J + 11 611 x 6 J 8 3
1 1 0 , NH 3 + HCN

0 , NH 3 + HCN

8 8

(5.2)

The char layer is modelled: CNO = k R.VI + k R.V t CNH3 + HCN = k R.V t

(5.3)

where the reaction rate are assumed to be similar to the ones of coal given by Johnsson [1990].

Final report May 1999


Page 41 of 79

k R.V

k C k C +k = 1410 1 NO 2 CO 3 k1CNO + k2CCO + k3

0 .64 0 .64 k R.VI = 2.36 10 6 CNO CNH3 + HCN

% k = 67 exp06070 / T 5 ( Kk = 2.4 10 exp020400 / T 5K & K k = 8.9 10 exp031700 / T 5 ) K ' * exp0 10000 / T 5


1 5 2 5 3

(5.4)

Boundary condition J NO 0 = J 2, NO + J N F ,char J NH3 , HCN 0 = J 2, NH3 , HCN (5.5)

The residence time tc in for the nitrogen compounds in the char layer is calculated from the integrated mean time for the nitrogen leaving the char tcc and the time for the gas leaving the reaction front, tcr.
x mean

 J I  J
0 i

N F ,char

N F ,char

tcc = x mean tcc = tc =

dx = 0.5; 0 x x + x  J R1 x T + x T 6 / 1 x + x 6
/x
mean s i s s c c s c i

(5.6)

J R1 x T + x T 6
i s s c c

(5.7) (5.8) (5.9)

P J N F ,dev tcr + J N F ,char tcc J N F ,dev + J N F ,char

Nitrogen species leaving fuel bed J3, NO = J NO tc J3, NH3 + HCN = J NH3 + HCN tc (5.10)

J3, N2 = J2 , N 2 + 1 J2, NO + J F ,char J3, NO + J2 , NH3 + HCN + J3, NH3 + HCN 2

42

7 3

89

5.2. Description of nitrogen chemistry in free-room calculation The thermal oxidation of molecular nitrogen and the reburning reaction can be neglected in the plants where temperatures do not reach 1800K, and this condition is always verified in the furnace in exam, where the maximum computed temperatures dont exceed 1620K and the measured ones are even lower. The main contribution to NOx production is therefore the nitrogen contained in the fuel and an important fraction of the total NOx emission is produced inside the fuel layer itself. Another contribution comes from the prompt mechanism, which usually plays a minor role. As for the main chemistry, it is not possible to implement all the reactions in the calculation together with fluid dynamics, but simplified mechanisms are required. The models do not include the HCN->NH3 step, and represent all the NOx intermediates

Final report May 1999


Page 42 of 79

as HCN. The NOx formation and destruction mechanisms are summarised as in Figure 5.4. Nitrogen chemistry is not yet completely known, and some reaction rates are highly non-linear functions of temperature and of the reactant concentrations. For this reason it is difficult to predict exactly the production of nitrogen oxides, because a small fluctuation of the reaction conditions may cause an important increase of the reaction rates. Such fluctuations are present throughout the furnace, due both to turbulence and to small discontinuities in the fuel and air feed, and cannot be modelled in detail, impeding an exact estimation of the total NOx formation. To take into account the effect of the turbulent fluctuations a probabilistic approach is used. This consists of evaluating the mean production inside each cell, weighing the reaction rate function with the probability density function of the temperature. The temperature probability density function is a beta function, the main value is the one computed by the averaged equations, and the variance is expressed as a function of the turbulent variables.

+ CH i (prompt) +NO

N2
+ NH i

HCN

NH3
+ O (thermal) +OH + CH i (reburning)

NO

: Figure 5.4 Description of the model for the nitrogen chemistry in FLUENT. HCN, NO and N2 are used as in and out-put data. 6. Experimental The measurements described in the contract was planned before the plant was built, and the estimated cost and time for the measurements were based on experience from a boiler of the same scale as the one in Trollhttan. This estimation turned out to be too optimistic both concerning time and expenses, because of very high cost required to run the boiler independent of the heat demand from the district heating system. To optimise the activities, considering the funding given, characteristic operation conditions were defined and much more detailed measurements were performed during these test conditions than what was originally planed. By doing so the quality of the comparison between the simulations and the measurements was maintained in a satisfactory way.

Final report May 1999


Page 43 of 79

The first measurement campaign in Trollhttan the survey measurements, was the main measurement in this project. This survey included 49 permanent measurement positions and 81 positions for probe. During the test, fuel samples were taken continuously and analysed. The measurement period was 75 hours for the permanent positions, 5 minutes per position for the cold suction probe and zirconia cell probe, and 10 minutes per position for the hot suction probe. In all positions 10 seconds mean values were logged continuously, except for the FTIR, which gave two mean values during a 10 minutes period, and the zirconia cell, for which 40 values per second were collected. For the survey measurements, special, well prepared fuel (pure wood chips) was ordered and the test furnace had to be specially trimmed. Since the probe measurements only took place during short periods it was important to have stable operating conditions for the whole test period in order to achieve a representative picture of the combustion chamber. This was possible in almost all tests. Some disturbances did occur however, most probably due to a temporary change of the fuel size and moisture content. During these periods the measurements were paused. Apart from the survey measurements some additional measurements were carried out using the normal fuel, consisting of 40% bark, 40% wood chips and 20% sawdust. Ten water-cooled probes were inserted simultaneously in the upper part of the combustion chamber above the neck to achieve data on temperatures and gas concentrations of O2, CO2, CO, NO and THC for validation of the modelling. An additional cooled suction probe that allowed a much higher suction flow verified previous temperature measurements. The direction of the gas flow along one horizontal position above the neck was also measured. 7. Results The results consist of two parts: the results from the fuel layer, the free-room calculations and the measurements, and a comparison between modelling and measurements 7.1 Result of Bed model In Figure 7.1 an example of a simulation of the fuel layer can be seen. The primary gas flow of nitrogen, oxygen and water vapour is kept on a low level in the first part of the grate allowing the combustion to start. The primary gas flow is then raised in order to increase the combustion intensity in the fuel layer. When the reaction front reaches the surface of the grate, the gas flow and the velocity of the fuel layer along the grate are decreased to allow the char layer to build up and to increase the residence time of the gas. Dependent on the reduction of the content of char in the fuel layer, a further reduction of the primary air is necessary before the end of the grate to keep up the temperature and for the final burn out. In this the profile of the equivalent radiation temperature from the free-room is assumed. The output from the fuel layer model is the position of the reaction front and the height of the fuel layer. At half way along the grate, the height of the fuel layer rises. In Figure 7.1 this rise looks rather dramatic and unrealistic, but this is a result of the reduction of the velocity of the fuel layer along the grate and is emphasised by the vertical scale of the diagram. The major gas components in the part of the fuel layer where the

Final report May 1999


Page 44 of 79

reaction front develops are water vapour and nitrogen, whereas oxygen is consumed. In the part where only char combustion takes place, the oxygen concentration first falls as a consequence of the increased height of the fuel layer and the decreased primary gas flow. Then it rises, while the char is gradually consumed. When the primary gas flow once more is reduced the oxygen concentration again falls and then rises slowly as the char combustion continues. The gas is assumed to leave the fuel layer at the temperature of the surface of the fuel layer. The simulation shows that the surface temperature of the fuel layer is lower and in some parts much lower than the temperature inside the fuel layer, especially during char combustion. In Figure 7.2 the result of the post-processing of the nitrogen chemistry presented. With the reaction rates of the model, the reduction of NO in the char layer is small, somewhere between 0 and 20% dependent on the char layer surface and gas flow through the layer. When the reaction front is not present, the reduction is neiglible due to the low temperature. Obviously, the nitrogen compounds leaving the fuel layer are already determined in the reaction front, and given by the assumed split factors. Otherwise the model behaves as expected; the reduction of NO is related to the residence time in the char layer and the temperature. An increased residence time and temperature increases the reduction of NO.
Fuel-N converted to NOx, HCN or NH 3
1 0.8

NO and NH 3 +HCN [ppm]

450

NO NH3

300

0.6

0.4 0.2 0 0

150

4 6 Length along grate [m]

4 6 Length along grate [m]

Figure 7.2. Result of the nitrogen chemistry simulation. Left figure, concentration of nitrogen oxides, and NH3+HCN. Right figure, fraction of fuel nitrogen converted to nitrogen oxides or NH3+HCN 7.2 Result of the free-room model The aim of the simulation was to find a steady state solution. The algorithm, though, did not converge. Instead of a trend towards a single value, a non-decaying oscillation could be observed. In order to eliminate the fluctuations, many different options of the numerical solver were tried, but without success. This behaviour is due to the instability of the physical flow, which does not evolve to the exact equilibrium, but oscillates around it. Such instability is quite realistic in flows with gas jets, and oscillations of the flow can actually be observed by looking into the boiler. It is reasonable to say that this is an instability of the turbulent type, which may not be accounted for by the k and model, since it includes vortexes which are much larger than the control volumes, and that hinder the convergence to a steady state solution.

Final report May 1999


Page 45 of 79
2000
Temperature [K]

1600 1200 800 400 0 10


Outgoing Molar Flow [moles/s]

Char layer Surface layer Surrounding (Input)

CH 8

CO O2 CO 2 H2O N2

0
Relative mass loss [] Bed velocity [mm/s] (Input)

0.5
Bed height [m]

1 Bed surface Reaction front Mass loss 0.8 0.6 0.4 0.2 0 7.5 N2 O2 H2O (Input) (Input) (Input)

0.4 0.3 0.2 0.1 0 10

Ingoing Molar Flow [moles/s] (Input)

Bed vel. (Input)

4.5

1.5

4 Bed length [m]

Velocity change bed Velocity change gas

Figure 7.1 Input data and result of modelling. The grey areas indicate the given changes in the primary airflow and velocity of the fuel layer along the grate.

Final report May 1999


Page 46 of 79

The time variable must therefore be introduced in the calculation, in order to obtain convergence. This complicates dramatically the solution of the problem, that grows to a size n times larger than a steady state solution, where n is the number of time steps necessary to describe the full evolution of the flow in time. A solution of this size is not practically storable in memory, but this is not a significant limitation because it would be practically impossible to analyse such amount of data. The alternative is to monitor (i.e. store the value at every time step) some variable in a finite set of points. In order to understand the nature of the fluctuations it is also possible to store the values of a variable on a whole surface, and visualise it as a contour diagram. By placing the images for each time step in a sequence, it become possible to animate the result, which provides an easily interpretable representation of the unsteady combustion process.

Figure 7.3 Sections used to display result. Grey planes illustrate the sections The overall look of the various fields (velocity, temperature, concentrations) does not vary very much with time, so it is possible to make relevant considerations on the flow looking at a single frame of the animation even if it does not represent the whole solution. Two vertical sections of the boiler, shown in the Figure 7.3, are used for displaying the result from the free-room calculation. The velocity vectors on the vertical section parallel to the jets show the presence of a large vortex in the upper part of the furnace, see Figure 7.4. The presence of a downward flow inside the combustion chamber is seen, and has been reported from a similar furnace in Mjlby, Schuster 1994. Such a vortex causes an internal recirculation of a part of the flue gases, that therefore have a longer residence time inside the combustion chamber than the flue gas which goes straight to the exit.

Final report May 1999


Page 47 of 79
1.62e+03 1.49e+03 1.36e+03 1.23e+03 1.10e+03 9.69e+02 8.39e+02 7.10e+02 5.80e+02 4.50e+02 3.20e+02

Y Z X

Figure 7.4 Velocity vectors and temperature field on a vertical section parallel to the jets. The section is shown in Figure 7.3 The difference in residence times is also shown by the temperature field which includes a colder area in the middle of the vortex see Figure 7.4, where the gas has a longer time to radiate its energy to the tube walls. Such difference in residence should be avoided if possible, even if in this particular case it does not seem to have important consequences. In order to verify the existence of this downward flow a special hole in the furnace wall was opened and an appropriate probe was built to detect the direction of the flow. The probe is a 4.7m long steel tube, water cooled to resist the furnace temperatures, which emits a small acetylene jet from the tip. The jet ignites when introduced into the flue gas flow, due to the high temperature and follows the flow of the flue gas, if this is strong enough to overcome the buoyant effect of the hot flame. The flame is clearly visible from the window above the hole, as it is much more luminous than the surroundings, even if sometimes it extinguishes due to a lack of oxygen. The experiment was carried out twice, in different operating conditions, and both times it supported the calculation result of a downward flow. It appears to be a quite strong flow and it is located in a layer, which is approximately 0.5 meters thick in the vicinity of the rear wall of the furnace. Extracting the probe a little at a time the flame follows irregular directions first and it goes upwards later. The result of the experiment is therefore completely in agreement with the calculation. In order to avoid this vortex, or to decrease the size of the downward flow a rise of the flow from the front secondary air inlet and a decrease of the one from the rear wall secondary air inlet should be a sufficient remedy that would not imply any modifications of the plant itself. The velocity and temperature fields on the vertical section perpendicular to the jets show another feature of the flow that causes different residence times, see Figure 7.5.

Final report May 1999


Page 48 of 79

Figure 7.5 Velocity vectors and temperature field on a vertical section perpendicular to the jets. The section is shown in Figure 7.3 The pyrolysis gases are passing through the front recirculation jets along the symmetry plane of the boiler and beside the walls. This causes the combustibles to be more concentrated in these zones, and the temperature to be higher when the flow meets the secondary combustion air. The displacement of the combustibles to the sides of the furnace is clearly observed in the concentration measurements. An improvement of the homogeneity of combustion could be obtained using a more rational spacing of the jets. The temperature contours on the section parallel to the jets also show the presence of a high temperature area around the first part of the front recirculation jets. This may be due to the high oxygen concentration coming from the evaporation zone, that is blown in the hot pyrolysis gas flow, and reacts quickly thanks to the high turbulence. 7.3 Result measurement The results from the survey measurement can be summarised as follows: 1. Reducing conditions measured by the zirconia cell probe in holes 1 and 2 position a,b,c, Figure 7.6 corresponds well to a relatively low O2 concentration, Figure 7.7 and high concentrations of CO2, CO and THC, Figures 7.8-10. The measurement in holes 1 and 2 are in the flame zone above the grate, where the main part of fuel volatiles burns, Figure 1.1. Further down along the grate the char residue burns (burn-out zone) and the gas phase above the grate always show oxidising conditions, Figure 7.6. In this area the O2 concentration is high, Figure 7,7. The CO2, CO THC concentrations are low, Figures 7.8-10, in relation to what was measured in the flame zone.

Final report May 1999


Page 49 of 79

2. The temperatures at level 1 are shown in Figure 7.11. Higher temperatures were expected in the flame zone in comparison with the results obtained above the burn-out zone. This was also obtained with some exception, hole 1, position A, B and hole 2, position B, Figure 7.11. The reason for the low temperatures in these positions is that flue gas is re-circulated from a nozzle located above the grate. The effect of the flue-gas re-circulation can also be seen in the temperature measurement carried out at level 2, hole 7, position A, B and C, Figure 7.12 where a temperature level between 190-300 C was found. As a contrast to these low temperatures was the highest temperature measured in hole 8, position D (835 C), see Figure 7.12. It seems likely that oxygen and volatiles are pressed to the sides by the flue-gas re-circulation, mixed and burned up along the wall. 3. Combustion gases (of various mix between oxygen and unburned species, and with various extent of burnout) from the lower part of the combustion chamber is mixed with re-circulated flue gas and secondary air in the restriction (the neck, see Figure 3.1). The combustion gases have different content depending on their history whether they originate from the drying zone in the beginning of the grate, the flame zone or the burn-out zone. When the gases pass the neck an intensive combustion takes place leading to a large decrease of the levels of CO and total hydrocarbons (THC), down to levels between 0.1-0.4%, Figures 7.13 and 7.14. At the same time the CO2 concentration increases from levels between 11-14% up to 15-17% despite the fact that the total volume of combustion gas increases as a result of the secondary air injection. 4. By the measurement of the gas concentrations at level 4, Figure 7.16 and a comparison of the CO-levels at level 3, Figure 7.15 is it possible to follow the burnout of the gas phase in the upper part of the combustion chamber. The temperature is still high at this level, especially in the 10 measurement locations, holes 16-20, positions A and B, Figure 7.17. These positions form a core of the cross-section, where the temperature was measured to 839-909 C. The levels of CO and THC also decrease from a concentration of 1000-4000 ppm at level 3, Figure 7.15 to only 40-500 ppm at level 4, Figure 7.16. 5. With the nitrogen content measured in the fuel (0.1 % by weight measured on the combustibles matter) a theoretical level of 280 ppm NH3 and/or NO can be achieved at an excess air ratio of 1.2. If the fact that the gas flow through the grate is lower than that required for an excess air ratio of 1.2 is taken into account (despite the flow of re-circulated flue gas) then the sum of NO and NH3 in hole 1 and 2 ought to be in the range of 400 to 600 ppm. The highest level of the sum of NO and NH3 was measured to 100 to 200 ppm. An important question is what happened with rest of the fuel nitrogen. One hypothesis is that NH3 and/or NO decomposes already when the gas passes through the fuel layer on the grate. This should be possible to study separately in a laboratory scale reactor. 6. Above the burnout zone NO levels between 26-58 ppm were found, Figure 7.18. At the same time the CO2 level was measured to only 2.6-5.3 %, Figure 7.8. The calculated average of NO and CO2 in hole 3 and 4 is 40ppm and 3.76 % respectively. If one calculates the corresponding average in the flame zone, hole 1 and 2, 85.1 ppm NO and 14% CO2 is obtained. Normalising these levels of NO to a fictitious CO2 concentration of 15%, a concentration of 160 ppm NO is calculated for the measurement above the burnout zone to be compared to only 91 ppm NO above the flame zone, Figure 7.19. This difference is interpreted as an effect of a higher degree of oxidation to NO of the fuel nitrogen during the char

Final report May 1999


Page 50 of 79

burnout compared to the case when the volatile part of the fuel is oxidised in the flame zone. This interpretation is consistent with the hypothesis that NH3 and/or NO are decomposed already when the combustion gases pass the fuel bed, since this lowers the fuel nitrogen conversion to NO. During the char burnout different conditions prevail on the grate in comparison with the situation in the flame zone. Most of the fuel layer in the burnout zone of the grate consists of ash. In the flame zone the fuel concentration is much higher, and this increases the probability that the NO being formed in the gas phase between the fuel particles is reduced on the passage through the bed. 7. The passage of combustion gases through the neck has already been commented on. The highest temperatures were measured along the rear wall downstream of the neck. In these positions also the highest levels of NO were obtained. This means that the temperature levels in the flame zone are of importance for the formation of NO. Since the concentration of NH3 is low, it is likely that some thermal NO formation occurs in the flame zone. From this, it can be concluded that the flue gas re-circulation plays an important role in order to prevent too high temperatures of the gas phase during its passage through the neck. 8. The average concentration of NO measured on dry flue gas for the whole test period is calculated to 119 ppm. This corresponds to an emission of NO of 84 mg NO2/MJ fuel supplied, which is equal to 129 mg NO2/nm3 at 15% CO2, dry.

Final report May 1999


Page 51 of 79

1 .0

20 2 .8m fro m w a ll 1 .9m fro m w a ll

T im e fra c tion w ith re d. c on d ition s

O x yg e n c o n c e n tra tio n (vo l-% ,d ry)

0 .8

16

1 .1m fro m w a ll 0 .2m fro m w a ll

0 .6

12

0 .4

2 .8 m from w a ll 0 .2 1 .9 m from w a ll 1 .1 m from w a ll 0 .2 m from w a ll 0 .0 0 2 4 6 8

0 0

<_ fuel in

x (m )

as h out _>

< fu e l in

x (m )

ash out

>

Figure 7.6 Time fraction with reducing conditions at level 1 measured at four positions from the right wall: A=2.8m; B=1.9m; C=1.1m and D= 0.2m.
20

Figure 7.7 Oxygen concentrations at level 1 measured at four positions from the right wall: A=2.8m; B=1.9m; C=1.1m and D=0.2m.
12

C a rb o n d io xid e c o n c . (vo l-% ,d r y)

16

C a rb o n m o n o xid e co n c. (vo l-% ,d ry )

12

4 2 .8 m from w a ll 1 .9 m from w a ll 1 .1 m from w a ll 0 .2 m from w a ll 0 0 2 4 6 8

2 .8m fro m w all 4 1 .9m fro m w all 1 .1m fro m w all 0 .2m fro m w all 0 0 2 4 6 8

<_ fu e l in

x (m )

ash out

>

< fu e l in

x (m )

ash out

>

Figure 7.8 Carbon dioxide concentrations at level 1 measured at four positions from the right wall: A=2.8m; B=1.9m; C=1.1m and D=0.2m.

Figure 7.9 Carbon monoxide concentrations at level 1 measured at four positions from the right wall: A=2.8m; B=1.9m; C=1.1m and D=0.2m

Final report May 1999


Page 52 of 79
5 9 00

T e m p e ra tu re of th e c o m b u stio n g a s (C )

To ta l h yd roca rb o n co nc . (vo l-% ,d ry )

8 00

7 00

6 00

2 .8 m fr o m w a ll 1 1 .9 m fr o m w a ll 1 .1 m fr o m w a ll 0 .2 m fr o m w a ll 0 0

2 .8 m fro m w a ll 5 00 1 .9 m fro m w a ll 1 .1 m fro m w a ll 0 .2 m fro m w a ll 4 00

<

fu e l in

x (m )

ash o ut

>

< fu e l in

x (m )

a sh o u t

>

Figure 7.10 Total hydrocarbon concentrations at level 1 measured at four positions from the right wall: A=2.8m; B=1.9m; C=1.1m and D=0.2m.

Figure 7.11 Temperatures of the combustion gas at level 1 measured at four positions from the right wall: A=2.8m; B=1.9m; C=1.1m and D=0.2m.
120000 2 .8 m fro m w a ll 1 .9 m fro m w a ll

C a rb on m o no xide co nc. (v ol-pp m ,dry)

T em p era ture of th e com bustion g as (C)

8 00

1 .1 m fro m w a ll 0 .2 m fro m w a ll 80000

6 00

4 00 2 .8 m from w all 1 .9 m from w all 1 .1 m from w all 2 00 0 .2 m from w all

40000

0
0 _ 2 4 6 _ 8

<

fu e l in

x (m )

a sh o u t

>

<_ fue l in

x (m )

ash o ut _>

Figure 7.12 Temperatures of the combustion gas at level 2 measured at four positions from the right wall: A=2.8m; B=1.9m; C=1.1m and D=0.2m.

Figure 7.13 Carbon monoxide concentrations at level 2 measured at four positions from the right wall: A=2.8m; B=1.9m; C=1.1m and D=0.2m.

Final report May 1999


Page 53 of 79
5 00 0
120000 2 .9 m fro m w a ll 1 .4 m fro m w a ll 0 .1 m fro m w a ll

C a rb o n m o n o x id e co n c. (v o l-p p m ,d ry)

C a rb on m o no xide co nc. (v ol-pp m ,dry)

4 00 0

80000

3 00 0

2 00 0

40000

1 00 0

2 .9 m fro m w a ll 1 .4 m fro m w a ll

0 0

0 .1 m fro m w a ll 0

<_ fue l in

x (m )

ash o ut _>

< fu e l in

x (m )

a sh o u t

>

Figure 7.14 Carbon monoxide concentrations at level 3 measured at three positions from the right wall: A=2.9m; B=1.4m and C=0.1m.

Figure 7.15 Carbon monoxide concentrations at level 3 measured at three positions from the right wall: A=2.9m; B=1.4m and C=0.1m.
1 000 2 .9 m fro m w a ll 1 .4 m fro m w a ll

5 000

2 .9 m fro m w a ll 1 .4 m fro m w a ll

T em p e ra ture o f th e co m bu stio n g a s (C )

0 .1 m fro m w a ll

C a rbo n m o n ox ide co n c. (vo l-p p m , d ry))

4 000

0 .1 m fro m w a ll

9 00

3 000

2 000

8 00

1 000

7 00 0 0 0

< fu e l in

x (m )

as h ou t

< fu e l in

x (m )

as h ou t

>

>

Figure 7.16 Carbon monoxide concentrations at level 4 measured at three positions from the right wall: A=2.9m; B=1.4m and C=0.1m

Figure 7.17 Temperatures of the combustion gas at level 4 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m.

Final report May 1999


Page 54 of 79
2 00 2 00

N orm alised conc. of nitric oxid e (vol-ppm ,dry)

1 60

1 60

N itric o xid e con c. (vo l-pp m ,d ry)

1 20

1 20

80

80

2.8 m fro m w all 40 1 .9 m fro m w a ll 1.1 m fro m w a ll 0.2 m fro m w a ll 0 0 2 4 6 8

2 .8 m fro m w a ll 40 1 .9 m fro m w a ll 1 .1 m fro m w a ll 0 .2 m fro m w a ll 0 0 2 4 6 8

<

fu e l in

x (m )

ash out

>

<

fu e l in

x (m )

a sh o u t

>

Figure 7.18 Nitric oxide concentrations at level 1 measured at four positions from the right wall: A=2.8m; B=1.9m; C=1.1m and D=0.2m.

Figure 7.19 Normalised concentration of nitric oxide for the measurement positions at level 1.

The results from the additional measurements are compared with the survey measurements in Figures 7.20-41 and in Tables 7.1-3 below. Following observations can be made: 1. The operating conditions were much more stable during the additional measurements compared to the survey measurements despite the higher quality of the fuel (only wood chips) used the first time. This result in less spikes of CO in the stack during the additional measurements. See Table 7.1. 2. The nitrogen content in the fuel was much higher for the fuel mix compared to the wood chips fraction only, see Table 7.2. Despite the much higher fuel nitrogen content the NOx emission did not increase proportionally for the fuel mix, see Table 7.3. 3. The combustion gas temperatures below the "neck" at level 3 show different trends and levels in the two test cases, Figures 7.20 and 7.21. 4. The oxygen profiles in Figures 7.22 and 7.23 measured at the same level, (level 3) are more similar but somewhat lower during the additional measurements with the fuel mix. This is interpreted as somewhat earlier combustion during the second test, an observation also supported by the profiles of total hydrocarbon concentration in Figures 7.28 and 7.29. 5. At the level above the "neck", level 4 the difference in oxygen concentration between the two test cases is even more pronounced, Figures 7.34 and 7.35.

Final report May 1999


Page 55 of 79

Table 7.1. Emissions of CO. Survey measurement 16% of the time was the CO levlel above 500 ppm. Conclusion: Large variations in the CO level obtained during the measurement. Additional measurement No "spikes" of CO above 500 ppm. Conclusion: Much more stable operating conditions.

Table 7.2. Operating conditions during the measurements. Survey test Additional test Fuel High quality wood Mix of 40% wood chips, 40% bark and 20% chips saw dust N-cont. in fuel 0.1% on m.a.f. 0.4% on m.a.f. Load 100% 100% Excess air ratio 1.2-1.3 1.2 Primary air/ total 0.52 0.54 air Flue gas recirc/ 0.10 0.21 total air Table 7.3. Emissions of NO. Survey test 112 ppm (@ 6% O2, dry) 55 mg NO/MJ 84 mg NO2/MJ 146 mg NO/m3n (@15% CO2,dry) 224 mg NO2/ m3n (@15% CO2,dry) Fuel-N conv. to NO: 46% Additional test 131 ppm (@ 6%O2, dry) 69 mg NO/MJ 106 mg NO2/MJ 175 mg NO/ m3n (@15% CO2,dry) 269 mg NO2/ m3n (@15% CO2,dry) Fuel-N conv. to NO: 14%

Final report May 1999


Page 56 of 79

10 0 0

1 00 0 2 .9m fro m w a ll 1.4 m fro m w a ll 0 .1 m fro m w a ll

T e m p e ra tu re ( C )

800

T e m p e ra tu re ( C )
2 .9 m fro m w a ll 1 .4 m fro m w a ll 0 .1m fro m w a ll 0 1 2 3 4 5

900

900

800

700

700

< fu e l in

x (m )

ash out

>

< fu e l in

x (m )

ash out

>

Figure 7.20 Temperatures of the combustion gas at level 3 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. First measurement with wood chips
10

Figure 7.21 Temperatures of the combustion gas at level 3 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. Second measurement with a mix of biofuel.
10 2.9m from w all 1 .4 m from w a ll

O x yg e n co n ce n tra tio n (vo l-% ,d ry)

O x yg e n co n ce n tra tio n (vo l-% ,d ry)

0.1m fro m w all

2.9m fro m w all 1.4 m from w all 0 .1m from w a ll

0 0

< fu el in

x (m )

a sh o u t

>

< fu el in

x (m )

a sh o u t

>

Figure 7.22 Oxygen concentrations at level 3 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. First measurement with wood chips.

Figure 7.23 Oxygen concentrations at level 3 measured at different positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. Second measurement with a mix of biofuel.

Final report May 1999


Page 57 of 79
20 20

C arbo n dio xide co nc. (vol-% ,d ry)

12

C arb o n d io xid e co n c. (vo l-% ,d ry)

16

16

12

2.9m fro m w all 1.4m fro m w all 0.1m fro m w all

2.9m fro m w a ll 1.4m from w all 0 .1m from w a ll

0 0

<

fu e l in

x (m )

ash out

>

< fu el in

x (m )

a sh o u t

>

Figure 7.24 Carbon dioxide concentrations at level 3 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. First measurement with wood chips.
50 0 0

Figure 7.25 Carbon dioxide concentrations at level 3 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. Second measurement with a mix of biofuel.
5 000

C a rb o n m o n o xid e c o n c. (v o l-p p m ,d ry )

40 0 0

C a rb o n m o n o xid e co n c . (v ol-p p m ,dry )

4 000

30 0 0

3 000

20 0 0

2 000

10 0 0

2 .9 m fro m w a ll 1 .4 m fro m w a ll 0 .1m fro m w a ll

1 000

2.9m fro m w all 1 .4 m from w all 0.1m fro m w a ll

0 0

< fu e l in

x (m )

ash out

0 0

>

< fu e l in

x (m )

a sh ou t

>

Figure 7.26 Carbon monoxide concentrations at level 3 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. First measurement with wood chips.

Figure 7.27 Carbon monoxide concentrations at level 3 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. Second measurement with a mix of biofuel.

Final report May 1999


Page 58 of 79
2 .9 m fro m w a ll

T o ta l h y d ro c a rb o n c o n c . (vo l-p p m ,d ry )

6000

T o ta l h yd ro ca rb on co n c. (vo l-pp m ,dry )

6000

1 .4 m fro m w a ll 0 .1 m fro m w a ll

4000

4000

2000 2.9 m from w a ll 1 .4m fro m w a ll 0 .1 m fro m w a ll 0 0 1 2 3 4 5

2000

0 0

<

fu e l in

x (m )

ash out

>

<

fu e l in

x (m )

ash out

>

Figure 7.28 Total hydrocarbon concentrations at level 3 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. First measurement with wood chips.
2 00

Figure 7.29 Total hydrocarbon concentrations at level 3 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. Second measurement with a mix of biofuel.
2 00

Nitric oxide conc. (vol-ppm ,dry)

Nitric oxide conc. (vol-ppm ,dry)

1 60

1 60

1 20

1 20

80

80

40 40 2 .9 m fro m w a ll 1 .4 m fro m w a ll 0 .1 m fro m w a ll 0 0 0 1 2 3 4 5 0

2.9m fro m w a ll
1 .4m from w all

0.1 m fro m w a ll 1 2 3 4 5

< fue l in

x (m )

a sh ou t

<

fu e l in

x (m )

ash out

>

>

Figure 7.30 Nitric oxide concentrations at level 3 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. First measurement with wood chips.

Figure 7.31 Nitric oxide concentrations at level 3 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. Second measurement with a mix of biofuel.

Final report May 1999


Page 59 of 79
1 00 0 2.9 m fro m w a ll 1 .4m fro m w a ll 0.1 m fro m w a ll 100 0 2 .9 m fro m w a ll 1 .4 m fro m w a ll 0 .1 m fro m w a ll

T e m p e ra tu re ( C )

T e m p e ra tu re (C )

900

900

800

800

700 0

700 _ 1 2 3 4 _ 5 0

< fu e l in

x (m )

ash out

>

<

fu e l in

x (m )

as h o ut

>

Figure 7.32 Temperatures of the combustion gas at level 4 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. First measurement with wood chips.
10

Figure 7.33 Temperatures of the combustion gas at level 4 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. Second measurement with a mix of biofuel.
10 2 .9 m fro m w a ll 1.4 m fro m w a ll

O xy g e n c o n ce n tra tio n (v o l-% ,d r y)

O x yg e n co n ce n tra tio n (vo l-% ,d ry)

0 .1 m fro m w a ll

2.9m fro m w all 1.4 m from w all 0 .1m from w a ll

0 0

< fu el in

x (m )

a sh o u t

>

< fu e l in

x (m )

a sh o u t

>

Figure 7.34 Oxygen concentrations at level 4 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. First measurement with wood chips.

Figure 7.35 Oxygen concentrations at level 4 measured at three different from the right wall: A=2.9m; B=1.4m; and C=0.1m. Second measurement with a mix of biofuel.

Final report May 1999


Page 60 of 79
20 20

C arb o n d io xid e co n c. (vo l-% ,d ry)

12

C arb o n d io xid e co n c. (vo l-% ,d ry)

16

16

12

2.9m fro m w all 1.4 m from w all 0 .1m from w a ll

2.9m fro m w a ll 1.4m from w all 0 .1m from w a ll

0 0

< fu el in

x (m )

a sh o u t

>

< fu el in

x (m )

a sh o u t

>

Figure 7.36 Carbon dioxide concentrations at level 4 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. First measurement with wood chips.

Figure 7.38 Carbon dioxide concentrations at level 4 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. Second measurement with a mix of biofuel
50 0 0 2 .9 m fro m w a ll 1 .4 m fro m w a ll

5 0 00

2 .9 m fro m w a ll 1 .4m fro m w a ll

C a rb o n m o n o x id e c o n c . (v o l-p p m ,d ry )

0 .1 m fro m w a ll 4 0 00

C a rb o n m o n o x id e co n c . (v o l-p p m ,d ry )

40 0 0

0.1 m fro m w a ll

3 0 00

30 0 0

2 0 00

20 0 0

1 0 00

10 0 0

0 0

0 _ 1 2 3 4 _ 5 0

<

fu e l in

x (m )

a sh o ut

>

< fu e l in

x (m )

ash out

>

Figure 7.39 Carbon monoxide concentrations at level 4 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. First measurement with wood chips.

Figure 7.40 Carbon monoxide concentrations at level 4 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. Second measurement with a mix of biofuel

Final report May 1999


Page 61 of 79
2 0 00 2 .9 m fro m w a ll 1 .4 m fro m w a ll 2 00 0 2 .9 m fro m w a ll 1 .4 m fro m w a ll

T o ta l h y d ro c a rb o n c o n c . (v o l-p p m ,d ry )

1 6 00

0 .1m fro m w a ll

T o ta l h y d ro c a rb o n c o n c . (v o l-p p m ,d ry )

1 60 0

0 .1 m fro m w a ll

1 2 00

1 20 0

8 00

800

4 00

400

0 0

<

fu e l in

x (m )

a sh o ut

>

< fu e l in

x (m )

ash o ut

>

Figure 7.40 Total hydrocarbon concentrations at level 4 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. First measurement with wood chips.
200

Figure 7.41 Total hydrocarbon concentrations at level 4 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. Second measurement with a mix of biofuel.
200

N itric ox id e co nc. (vol-p pm ,dry)

N itric o xide co nc. (vol-pp m ,d ry)


2 .9 m fro m w a ll 1 .4 m fro m w a ll 0 .1 m fro m w a ll

160

160

120

120

80

80

40

40

2 .9 m fro m w a ll
1 .4 m fro m w a ll

0 .1 m fro m w a ll 0 5 0 1 2 3 4 5

0 0

< fu el in

x (m )

ash o ut

>

< fu el in

x (m )

a sh ou t

>

Figure 7.42 Nitric oxide concentrations at level 4 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. First measurement with wood chips.

Figure 7.43 Nitric oxide concentrations at level 4 measured at three positions from the right wall: A=2.9m; B=1.4m; and C=0.1m. Second measurement with a mix of biofuel.

Final report May 1999


Page 62 of 79

7.4 Comparison between model calculation and measurements Comparison between the survey measurements during operation with pure wood chips and the simulation is seen in Figures 7.44 to 7.50, for the temperature and gas species O2, CO2, CO, NO and total hydrocarbons at all measurement levels and NH3 at level 1 and 2. The measurement levels are presented in Figure 3.1. The following observations can be made: The simulation predicts the temperature to be approximately 300C higher than the measured temperature at level 1 and 2. This is due to unstable operation of the boiler during the time for these measurements. However, the calculation and the measurement predict the same trend: the temperature decreases towards the ash outlet see Figure 7.44. At level 3 and 4 the temperature measurements were made at stable operating condition and the agreement of between the measurement and calculation is good The oxygen concentration is lower in the simulation than in the measurements, especially towards the ash outlet. This indicates that the fuel layer model predicts the char combustion to be extended further along the grate than it was during the measurements. In the main part of the fuel layer the agreement between the simulation and the measurements is good except at the wall, where there probably is a by-pass of the primary gas that can not be described by the present fuel layer model, see Figure 7.45. At level 3 the agreement between measurement and calculation is good, but at level 4 the measurements shows twice as much oxygen as the calculation. In this case the measurements are probably wrong, because all fuel is finished and the oxygen level in the stack is just half of the level at the outlet from the furnace. The reason for the measurement error is most certainly air leakage into the measurement system. The air leakage flow is estimated to 30% of the flue gas flow out from the probes. As can be expected the carbon dioxide behaves opposite to the oxygen. This is seen both from the simulation and the calculation, Figure 7.46. A good agreement between calculation and measurement is achieved at level 3 and 4, if the measured concentrations on level 4 are compensated for the air leakage in the measurement system. The simulation predicts much lower concentrations of carbon monoxide and total hydrocarbons than what are measured, Figure 7.47 and 7.48. In the simulation are carbon monoxide and total hydrocarbons released closer to the fuel chutes then the measurements show. Compared with the carbon dioxide, the rate of oxidation seems to be too high. The burn-out along the furnace is the same both, for in calculation and in the measurements. The concentration of unburnt is high in the low levels. A large reduction is obtained between level 2 and 3. Between level 3 and 4 the concentration of unburnt is further reduced to very low levels The nitrogen oxides in the simulation are much higher than the measured values, Figure 7.49, but both give the same trend along the grate. A good agreement between simulation and measurement is obtained for the ammonia concentration,

Final report May 1999


Page 63 of 79

Figure 7.50. The over prediction of the simulated nitrogen oxides is a result of the assumed high conversion of fuel bound-nitrogen to nitrogen oxides in the fuel layer model. The measurement does not show any dilution effect between level 2 and 3, which is very clear in the simulation. Either the measured concentration is too low in the lower part of the furnace or nitrogen oxides are formed in the area between level 2 and 3.
1500

1 50 0

1200

1 20 0

T e m p e ra tu r e ( C )

900

T e m p e ra tu re ( C )

9 00

600

6 00

300

3 00

0 0

<
1500

fu e l in

x (m )

a sh ou t

>
1500

<

fu e l in

x (m )

a sh o ut

>

1200

1200

T e m p e ra tu r e ( C )

900

T e m p e ra tu r e ( C )

900

600

600

300

300

0 0

<

fu e l in

x (m )

a sh ou t

>

<

fu e l in

x (m )

a sh ou t

>

Figure 7.44. Calculated (dotted) and measured (solid) temperature profiles at level 1 (upper left figure) and level 2 (upper right figure), at 0.2m (red, star) 1.1m (green, diamonds), 1.9m (blue, squares) and 2.8m (black, triangulars) from the side-wall, and at level 3 (lower left figure) and level 4 (lower right figure), at 0.1m (red, diamonds) 1.4m (blue, squares) and 2.9m (green, triangulars) from the side-wall

Final report May 1999


Page 64 of 79

20

20

O x yg en c onc entration (vol-% ,dry )

16

O x yg en c onc entration (vol-% ,dry )


0 2 4 6 8

16

12

12

< fue l in
20

x (m )

as h out

0 20

>

< fue l in

x (m )

as h out

>

O x yg en c onc entration (vol-% ,dry )

16

O x yg en c onc entration (vol-% ,dry )


0 1 2 3 4

16

12

12

< fue l in

x (m )

as h out

>

< fue l in

x (m )

as h out

>

Figure 7.45. Calculated (dotted) and measured (solid) oxygen profiles at level 1 (upper left figure) and level 2 (upper right figure), at 0.2m (red, star) 1.1m (green, diamonds), 1.9m (blue, squares) and 2.8m (black, triangulars) from the side-wall, and at level 3 (lower left figure) and level 4 (lower right figure), at 0.1m (red, diamonds) 1.4m (blue, squares) and 2.9m (green, triangulars) from the side-wall

Final report May 1999


Page 65 of 79

20

24

C arb on diox ide conc . (vo l-% ,dry)

12

C arb on diox ide conc . (vo l-% ,dry)


0 2 4 6 8

16

20

16

12

< fue l in
20

x (m )

a sh out

0 20

>

< fue l in

x (m )

a sh out

>

C arbon d iox id e c onc . (v ol-% ,d ry )

12

C arb on diox ide conc . (vo l-% ,dry)


0 1 2 3 4

16

16

12

< fue l in

x (m )

as h out

>

< fue l in

x (m )

as h out

>

Figure 7.46. Calculated (dotted) and measured (solid) carbon dioxide profiles at level 1 (upper left figure) and level 2 (upper right figure), at 0.2m (red, star) 1.1m (green, diamonds), 1.9m (blue, squares) and 2.8m (black, triangulars) from the side-wall, and at level 3 (lower left figure) and level 4 (lower right figure), at 0.1m (red, diamonds) 1.4m (blue, squares) and 2.9m (green, triangulars) from the side-wall

Final report May 1999


Page 66 of 79

15 00 0 0

15 00 0 0

C a rb o n m o n o x id e co n c. (p p m ,d ry )

12 00 0 0

C a rb o n m o n o x id e co n c. (p p m ,d ry )

12 00 0 0

90 00 0

90 00 0

60 00 0

60 00 0

30 00 0

30 00 0

0 0

<
10000

fu e l in

x (m )

ash o ut

>
1000

<

fu e l in

x (m )

ash o ut

>

C a rb o n m o n o x id e c o n c . (v o l- p p m ,d r y)

C a rb o n m o n o x id e c o n c . ( v o l-p p m ,d r y )

8000

800

6000

600

4000

400

2000

200

0 0

<

fu e l in

x (m )

a sh out

>

<

fu e l in

x (m )

a sh ou t

>

Figure 7.47. Calculated (dotted) and measured (solid) carbon monoxide profiles at level 1 (left figure) and level 2 (right figure), at 0.2m (red, star) 1.1m (green, diamonds), 1.9m (blue, squares) and 2.8m (black, triangulars) from the side-wall, and at level 3 (lower left figure) and level 4 (lower right figure), at 0.1m (red, diamonds) 1.4m (blue, squares) and 2.9m (green, triangulars) from the side-wall

Final report May 1999


Page 67 of 79

50000

50000

T o ta l h yd ro ca rb o n c o n c e n tra tio n s ( p p m , d ry )

40000

T o ta l h yd ro ca rb o n c o n c e n tra tio n s ( p p m , d ry )

40000

30000

30000

20000

20000

10000

10000

0 0

<
10000

fu e l in

x (m )

ash o ut

>
3000

<

fu e l in

x (m )

ash o ut

>

T o ta l h y d ro c a rb o n c o n c . (v o l-p p m ,d ry )

T o ta l h y d ro c a rb o n c o n c . (v o l-p p m ,d ry )

2500

8000

2000

6000

1500

4000

1000

2000

500

0 0

<

fu e l in

x (m )

a sh out

>

<

fu e l in

x (m )

a sh ou t

>

Figure 7.48. Calculated (dotted) and measured (solid) total hydro carbon profiles at level 1 (left figure) and level 2 (right figure), at 0.2m (red, star) 1.1m (green, diamonds), 1.9m (blue, squares) and 2.8m (black, triangulars) from the side-wall, and at level 3 (lower left figure) and level 4 (lower right figure), at 0.1m (red, diamonds) 1.4m (blue, squares) and 2.9m (green, triangulars) from the side-wall

Final report May 1999


Page 68 of 79

50 0

50 0

40 0

40 0

N itric oxide c onc. (ppm , dry)

30 0

N itric oxide c onc. (ppm , dry)


0 2 4 6 8

30 0

20 0

20 0

10 0

10 0

< fuel in
30 0

x (m )

as h out

0 30 0

>

< fuel in

x (m )

as h out

>

25 0

25 0

N itric oxide c onc . (vol-ppm ,dry )

20 0

N itric oxide c onc . (vol-ppm ,dry )


0 1 2 3 4

20 0

15 0

15 0

10 0

10 0

50

50

< fuel in

x (m )

as h out

>

< fuel in

x (m )

as h out

>

Figure 7.49. Calculated (dotted) and measured (solid) nitrogen oxides profiles at level 1 (left figure) and level 2 (right figure), at 0.2m (red, star) 1.1m (green, diamonds), 1.9m (blue, squares) and 2.8m (black, triangulars) from the side-wall, and at level 3 (lower left figure) and level 4 (lower right figure), at 0.1m (red, diamonds) 1.4m (blue, squares) and 2.9m (green, triangulars) from the side-wall

Final report May 1999


Page 69 of 79

50

50

A m m onia/hy drogen cy anide c onc . (p pm , dry )

40

A m m onia/hy drogen cy anide c onc . (p pm , dry )


0 2 4 6 8

40

30

30

20

20

10

10

< fue l in
50

x (m )

as h out

0 50

> A m m onia /hyd rog en c yan ide con c. (v ol-ppm ,dry )

< fue l in

x (m )

as h out

>

A m m onia /hyd rog en c yan ide con c. (v ol-ppm ,dry )

40

40

30

30

20

20

10

10

0 0

< fue l in

x (m )

as h out

>

< fue l in

x (m )

as h out

>

Figure 7.50. Calculated (dotted) and measured (solid) ammonia profiles at level 1 (left figure) and level 2 (right figure), at 0.2m (red, star) 1.1m (green, diamonds), 1.9m (blue, squares) and 2.8m (black, triangulars) from the side-wall, and at level 3 (lower left figure) and level 4 (lower right figure), at 0.1m (red, diamonds) 1.4m (blue, squares) and 2.9m (green, triangulars) from the side-wall

Final report May 1999


Page 70 of 79

A comparison between the measurements during operation with a mix of biofuels and simulation is seen in Figures 7.51 to 7.55, for the temperature and gas species O2, CO2, CO, NO and total hydrocarbons at level 3 and 4. The following observations can be made: At level 3, the predicted temperatures are slightly higher than the measured ones (2.9 meters from the wall) as well as lower (1.4 meter from the wall). The shape of the temperature profile with the highest temperature in the third hole is correctly modelled in the positions located 2.9 meter from the wall. At level 4, the calculated temperatures are always slightly lower than the measured ones. Figure 7.51 At level 3 the calculated oxygen concentrations generally become higher in the model calculations in comparison with the measurements. The opposite is then expected for CO2 as confirmed. To a great extent this difference has disappeared at level 4, except for the locations close to the wall, Figure 7.52. The total hydrocarbons, which is modelled, as CH4 seem to follow the temperature predictions at level 3. Modelled hydrocarbon concentrations are lower than measured, in the positions 2.9 meters from the wall, where the higher temperatures were predicted. For the positions located 1.4 meter from the wall an opposite result is obtained for both temperature and hydrocarbons. At level 4, the hydrocarbon concentrations are low in measurements as well as in the model calculations. The calculated and measured profiles of hydrocarbons can be seen in 7.53 and the temperature profiles is Figure 7.51. The model predicts also in this case a much lower CO concentrations then the measurements at both levels, Figure 7.54. The predicted NO concentrations at level 3 and 4, Figure 7.55, are in the same range as the measured ones. This good agreement is obtained by assuming lower conversion of fuel-bound nitrogen to nitrogen oxides in the fuel layer model.

Final report May 1999


Page 71 of 79
1200 1200

1000

1000

T e m pe ra tu re (C )

600

T e m pe ra tu re (C )

800

800

600

400

400

200

200

0 0

<

fu el in

x (m )

ash o ut

>

<

fu el in

x (m )

ash o ut

>

Figure 7.51. Calculated (dotted) and measured (solid) temperature profiles at level 3 (left figure) and level 4 (right figure), at 0.1m (red, diamonds) 1.4m (blue, squares) and 2.9m (green, triangulars) from the side-wall
10 10

O xyge n con cen tra tion (vo l-% ,d ry)

O xyg en co ncen tra tio n (vol-% ,d ry)


0 1 2 3 4

<
20

fu e l in

x (m )

a sh o u t _ >

0 20

< _ fu e l in

x (m )

a sh o u t _ >

C a rb o n d io x id e c o n c. (v o l-% ,d ry )

16

C a rb o n d io x id e c o n c. (v o l-% ,d ry )
0 1 2 3 4

18

18

16

14

14

12

12

10

10

<

fu e l in

x (m )

a sh o u t

>

<

fu e l in

x (m )

a sh o u t

>

Figure 7.52. Calculated (dotted) and measured (solid) concentration profiles of oxygen (upper figures) and carbon dioxide (lower figures) at level 3 (left figures) and level 4 (right figures), at 0.1m (red, diamonds) 1.4m (blue, squares) and 2.9m (green, triangulars) from the side-wall

Final report May 1999


Page 72 of 79

2000

2000

T ota l hyd roca rb o n co nc. (vol-pp m ,dry)

1600

T ota l hyd roca rb o n co nc. (vol-pp m ,dry)

1600

1200

1200

800

800

400

400

0 0

<

fu el in

x (m )

ash o ut

>

<

fu el in

x (m )

ash o ut

>

Figure 7.53. Calculated (dotted) and measured (solid) concentrations profiles of total hydro-carbons at level 3 (left figure) and level 4 (right figure), at 0.1m (red, diamonds) 1.4m (blue, squares) and 2.9m (green, triangulars) from the side-wall
5000 5000

C a rb on m on oxide con c. (vo l-p pm ,dry)

C arbo n m o no xide co nc. (v ol-pp m ,dry)

4000

4000

3000

3000

2000

2000

1000

1000

0 0

< _ fu el in

x (m )

ash o ut _ >

<

fu el in

x (m )

ash o ut

>

Figure 7.54. Calculated (dotted) and measured (solid) concentrations profiles of carbon monoxide at level 3 (left figure) and level 4 (right figure), at 0.1m (red, diamonds) 1.4m (blue, squares) and 2.9m (green, triangulars) from the side-wall

Final report May 1999


Page 73 of 79
20 0 20 0

N itric o xid e co n c. (vo l-pp m ,d ry )

12 0

N itric o xid e co n c. (vo l-pp m ,d ry )


0 1 2 3 4

16 0

16 0

12 0

80

80

40

40

<
40

fu e l in

x (m )

as h o ut

>
40

<

fu e l in

x (m )

as h o ut

>

H y d ro g e n c ya n id e c o n c. (v o l-p p m ,d ry)

30

H y d ro g e n c ya n id e c o n c. (v o l-p p m ,d ry)
0 1 2 3 4

30

20

20

10

10

<

fu e l in

x (m )

a sh o u t

>

<

fu e l in

x (m )

a sh o u t

>

Figure 7.55. Calculated (dotted) and measured (solid) concentration profiles of nitrogen oxide (upper figures) and hydrogen cyanide (lower figures) at level 3 (left figures) and level 4 (right figures), at 0.1m (red, diamonds) 1.4m (blue, squares) and 2.9m (green, triangulars) from the side-wall 8. Discussion and conclusion The conclusion from present result is that the upper part of the furnace above the secondary air inlet is much easier to model than the lower part of the furnace. The model calculation in the upper part of the furnace agrees very well with the measurements in both cases examined. The secondary air jets determine the gas flow up to the upper part of the furnace, and the gas entering this part is well mixed. This creates a suitable situation for the CFD-calculation and this is also reflected in the modelling result. The lower part of the furnace is much more chaotic, which is a reflection of the mixing process going on. This part of the furnace is much more difficult to model, with high accuracy. In the lower part of the furnace there are jets with a high velocity and turbulence intensity, dissipating into a slow gas flow in connection with complex chemical reactions and flames. Both the turbulence and the chemistry is treated in a simplified way in the CFD-calculation, flames are not treated at all. The jets also create an unstable situation, which makes a time dependent solution necessary, and time mean values have to be calculated for comparison with measurements. The

Final report May 1999


Page 74 of 79

lower part of the furnace includes also the description of the fuel layer, which in itself is a simplified model. It is therefore quite reasonable that the agreement between measurements and model calculations in this part of the boiler is not as good as in the upper part. Despite all simplifications, gives the CFD calculation a great deal of information. The fuel layer has the greatest importance for the behaviour in the lower part of the boiler and the condition in the free-room above the fuel layer is a reflection of the fuel layer. To validate the fuel layer model measurements close to the surface along the grate surface would have been preferred. These kinds of measurements were not possible to perform in this project, due to the large surface and the limited possibility to enter the furnace with probes. Some information can be gained from measurements in an experimental pot, but it is not known how to implement this experimental result to a moving fuel layer model at present. There is a great difference between an industrial grate and a laboratory bed as is most clearly seen from the moisture content of the fuel that can be burned in the different facilities. Experiences show that a pot furnace can operate with fuel having moisture content of up to 40%, while the moving grate in Trollhttan operates with fuel having moisture content between 40 and 60%. The general conclusions of the modelling in the lower part of the furnace are that: the fuel layer model predicts a too early completion of the drying and devolatilization the overlap between drying devolatilization and char combustion is underestimated the oxidation of carbon monoxide and total hydrocarbons is to fast the reactions related to nitrogen oxides are too slow in the free-room model

The too rapid drying and devolatilization of the model is reflected by the much wider distribution of measured total hydrocarbons compared to the simulation. Also the distribution of the carbon monoxide indicates this. It can be seen from the distribution of oxygen, carbon monoxide and carbon dioxide that a greater overlap between drying, devolatilization and char combustion must take place in the fuel layer than what is predicted by the model. Close to the ash removal the levels of carbon monoxide, carbon dioxide are low, and the oxygen concentration is close to that of the primary gas. This means that the char combustion has already ended long before the ash falls into the ash removal pit. It can be seen from the analysis of the ash that the fuel is burned out. On the other hand gives the fuel layer model that the char combustion continues nearly up to the ash removal. In the central part of the grate the concentrations of carbon monoxide and total hydrocarbons are high. This together with a low oxygen concentration indicates that much of the oxygen is consumed by the char combustion. The fitting parameters that are estimated from the temperature profile on the metal surface of the grate could be misleading. The reciprocating grate may cause local mixing of the solid fuel that may course the temperature on the surface of the grate to rise before the reaction front, as defined in the model, reaches the grate surface. This is probably the reason, why the drying and the devolatilization are too fast in the model calculation. The calculated NO concentrations below the neck show that there is no active nitrogen chemistry modelled by the free-room model. This is confirmed by the

Final report May 1999


Page 75 of 79

calculated concentration profiles of HCN. This means that the nitrogen species evolving from the grate in the form of NO, NH3 and HCN, modelled as NO and HCN, do not undergo any further reactions in the free-room. The reason for this could be the relatively low temperatures given by the simulation, much lower than is normally obtained in modelling of flame combustion, an application where the nitrogen chemistry model probably works better. On the other hand the nitrogen chemistry linked to CFD-codes is generally of a low quality. This is further discussed by Kilpinen et al. (1997) where (MS20 of this EU-project) two common nitrogen mechanisms used in CFD calculations were compared with a detailed elementary reaction mechanism at 850 and 1100C. The study shows that the simplified nitrogen mechanisms do not work at the low temperature, while the detailed nitrogen mechanism does predict a conversion of NH3 into NO. The nitrogen compounds in the present model calculation are merely diluted by the flue gas recirculation and by the secondary air. The dilution caused by the secondary air and flue gas recirculation in the calculation is not seen in the measurements. As a result nearly 40% of the nitrogen oxides is formed between measurement level 2 and 3, in the area for the secondary air injection, if the dilution is the same in the real case as in the calculation. The remaining 60% of the nitrogen oxides are formed in the fuel layer. Both the calculation and the measurement just above the secondary air inlet show that the temperatures in the region of the secondary air inlet is lower than the ones where formation of thermal NOx may take place. Nevertheless, the temperature in the flames that are visually observed, and which are not possible to model with the present CFD code or measured by thermocouples, can lead to the formation of thermal NOx. Another reasons for the formation of NOx, could be that fuel-bound nitrogen leaving the fuel layer as ammonia or in entrained char particles is oxidised to NO when meeting the secondary air between measurement level 2 and 3. The measurement can not support any of these hypotheses, because nearly no ammonia was found close to the fuel layer, and char particles in the free-room was only indicated by visual observations (video film). The conclusion is that the nitrogen oxides can be reduced further by an improvement of the arrangement of the jets of flue gas recirculation and air for final combustion in connection with the design of the furnace. This can theoretically decrease the emissions of nitrogen oxides with up to 40%. The conclusion from the NOx modelling inside the fuel layer is that the remaining 60% of the nitrogen oxides formed, can be reduced, but better understanding on the formation of the nitrogen oxides and its precursors is needed. Special effort has to be made to understand the reaction front. Theoretically a fuel layer of biofuel offers an excellent opportunity to reduce the nitrogen oxides to extremely low levels. During the devolatilization ammonia and hydrogen cyanide are formed and these species can contribute to the reduction of nitrogen oxides, if the conditions are favourable. By arranging the combustion of the fuel layer in an optimal way, one should be able to create the same situation in the fuel layer and in the lower free-room as recommended for secondary measures, such as selective not catalytic reduction (SNCR). Also the formation of a char layer above the reaction front could improve the NOx reduction performance but this needs to be studied in more detail in the future both by theoretically and experimental works.

Final report May 1999


Page 76 of 79

9. References Fluent Inc. Fluent/UNS 4.2.5 users guide,1997 Gort R. On the propagation of a reaction front in a packed bed, PhD thesis University of Twente, ISBN 90-9008751-6, 1995. Johnsson, J.E. A new NOx module for the IEA-model. 21th IEA-AFBC Meeting Beograd, Department of Chemical Engineering, Technical University of Denmark, 1990. Jones, W.P., Lindstedt, R.P., Global reaction schemes for hydrocarbon combustion. Combustion and Flame, vol. 73, pp. 233-249, 1988. Kilpinen, P.; Brink, A.; Bostrm, S.; Mueller, C.; Nordstrm, T.; Hupa, M.; Simplified ammonia oxidation mechanism for modelling NOx emission from wood firing tests at ideal plug flow conditions, Report 97-5, bo Akademi, Faculty of Chemical Engineering, Combustion Chemistry Research Group, 1997. Lepplahti, J., Formation of NH3 and HCN in slow-heating-rate inert pyrolysis of peat, coal and bark, Fuel, vol.74, pp. 1363-1368, 1995. Schuster, R. Matematisk modellering, CFD; Stiftelsen fr Vrmeteknisk Forskning, Report no. 516, ISSN 0282-3772, 1994. Magnussen, B.F. Hjertager, B.H. On mathematical modeling of turbulent combustion with special emphasis on soot formation and combustion. 16th Symp. Int. on Combustion, The Combustion Institute, Pittsburgh, pp. 719-743, 1977. Magnussen, B.F., The Eddy dissipation concept, Division of Thermodynamics, Norwegian Institute of Technology, 1989 Saastamoinen, J.J.; Haukka, P.; Simultaneous drying and pyrolysis in fixed bed combustion of wet biofuel. Proceedings of the 11th International Drying Symposium (IDS 98), Halkidki, Greece August 19-22, pp.1975-1982, 1998. Saastamoinen J.J., Horttanainen, M., Taipale, R., Sarkomaa, P., Propagation of ignition front in fuel beds of wood particles, VTT Energy, 1999. Saastamoinen J.J. Hmlinen J.P. and Kilpinen P. Release of nitrogen compounds from wood particles during pyrolysis. Proceedings of the Fourth International Conference on Technologies and Combustion for a Clean Environment, Clean Air IV, Lisbon, Portugal, pp. 3.1:1-11, 1997. Zevenhoven, R., Hupa, M., The reactivity of chars from coal, peat and wood towards NO, with and without CO, Fuel vol.77, pp 1169-1176, 1998.

Final report May 1999


Page 77 of 79

Appendix A Composition of the volatile The composition of the volatile gases released from the fuel is unknown. By assuming that the hydrocarbon in the volatile gases is methane and that the char is pure carbon, the composition of the volatile gases can be estimated. The molar balance is based on dry and ash-free substance. The ash-free char content of the fuel is: Ychar = Ychar 1 Yash

The volatile gases are assumed to include carbon dioxide, carbon monoxide, methane and water. CHn Om n1CO2 + n2 CO + n3CH4 + n4 H2O n and m is n= 12YH YC Ychar m= 0.75YO YC Ychar

h1 and h2 are normalised heating values of carbon monoxide and methane. The heating value of the volatile part of the fuel and the normalised heat of devolatilization is h3 and h4. Normalisation is made so that all values in the matrix are in the same order of magnitude,

3 8 h = 3I I 2I 8 / I h = 4 H M Y 3 I I 89 / I h = H M 11 Y 6 / I
h1 = ICO ICO2 / ICO2
2 CH 4 CO2 H2 O CO2 3 u C char C CO2 4 dev C char CO2

CO2

Hu is the lower heating value of the fuel and Hdev is the heat of devolatilization. The composition of the volatile gases is obtained by solving the following system of equations:

1 !

1 0 0 2 1 0 h1

1 4 0 h2

0 2 1 0

1Y Y "#n "# ## n ## = Y 1Y 012YY n 6 .75 #$!n #$ 1 ! h h


1 2 C H 3 4 C char 3

char

O 4

6"# ## #$

The enthalpy, heat capacity and molar mass of the volatile gases are:

Final report May 1999


Page 78 of 79

Ivol g =
 #

n1 ICO2 + n2 ICO + n3 ICH 4 + n4 I H2O n1 + n2 + n3 + n4 n1c pCO2 + n2 c pCO + n3c pCH4 + n4 c pH2O n1 + n2 + n3 + n4 n1 MCO2 + n2 MCO + n3 MCH4 + n4 MH2 O n1 + n2 + n3 + n4

c pvol g =
 #

Mvol g =
 #

The enthalpy, and molar mass of the volatile solid are: Ivol s = Ivol g Tref
 #  #  #

H n1 + n2 + n3 + n4 + dev + c pvol s dT Mvol s Tref


 #  #

I
T

Mvol s = MC + nM H + mMO Appendix B Analytic solution for reactions controlled by mixing The analytic solution of the system of differential equations: JCH m t = jXCH m b min( JCHm , J O2 ) J0 + jt

JCO b min( JCO , JO2 ) min( JCH m , JO2 ) = jXCO t J0 + jt JO2 t = b J0 + jt

8 8

3 01 + m5 min( J
1 2

CH m

, JO2 ) + 1 min( J CO , JO2 ) 2

for the condition


=

43 J

CH n

JO2 & JCO JO2 ,


b/ j

8 3

89

JCHm JCO

jXCHm 1 J 0 + jt 6 b+ j
b/ j

+ C1 1 J 0 + jt 6 +

= C2 1 J 0 + jt 6

1 b/ j 2 2bj J 0 1 J 0 + jt 6b / j XCHm + 2 41 J 0 + jt 6 j 0b + j 5
b/ j

+ bj 3t 1 J0 + jt 6 + bj 3t 1 J0 + jt 6

b/ j b/ j

XCHm + bj 2 J0 1 J0 + jt 6 XCO + j 4 t 1 J 0 + jt 6
b/ j

XCO + j 3 J 0 1 J 0 + jt 6

b/ j

XCO +

XCO + b 3C1 ln1 J0 + jt 6 +


  

+2b 2 jC1 ln1 J0 + jt 6 + bj 2 C1 ln1 J 0 + jt 6 =

JO2

bjt 3 2bXCHm jXCHm + bmXCHm + jmXCHm bXCO jXCO 8 20b + j 5


b/ j  1 2 2

+ C3 +

+1 J 0 + jt 6

1C2

+ 2C1 mC1 6 +

bC1 ln1 J0 + jt 6   2j 

Final report May 1999


Page 79 of 79

for the condition = =

43 J

CH n

> JO2 & JCO JO2 ,


b/ j

8 3
0

89

JCH m JCO

jXCH m J0 + jt

0b + j 50b + 2 j 5 + C + 2C 1 J + jt 6 C m1 J + jt 6 0m 151 J + jt 6 4bjX J 1 J + jt6 + bj X t1 J + jt6 J = 0b + j 50b + 2 j 5 + C 1 J + jt 6 and for the condition 43 J > J 8& 3 J > J 89 .
b/2 j b/ j 3 2 0 1 0 b/ j b/ j 2 0 CH m 0 0 CH m 0 O2 b/2 j 2 0 CH n O2 CO O2

jt 2b 2 XCHm + 2bjXCH m b 2 mXCH m + b 2 XCO + 3bjXCO + 2 j 2 XCO

b+ j

6 + C 1 J + jt 6
1

8+
C1 b 2 3bj 2 j 2

b/ j

79 +

JCHm JCO JO2

= jtXCHm = jtXCHm

2C3  J0 + jt $ + C1 m2 + C2
b m2#/ 2 j

b m 2 # / 2 j

= C3  J 0 + jt $

Paper VII

Fuel 80 (2001) 473481 www.elsevier.com/locate/fuel

Ignition and propagation of a reaction front in cross-current bed combustion of wet biofuels
H. Thunman, B. Leckner*
Department of Energy Conversion, Chalmers University of Technology, 412 96 Gothenburg, Sweden Received 12 May 2000; revised 9 August 2000; accepted 12 August 2000

Abstract Grate ring is the most common way to burn bio-fuels in small-scale units. Different combustion modes are achieved depending on how fuel and primary air are introduced. In continuous systems fuel and air are usually fed in cross-current and counter-current ow. Here, combustion of wet biofuels is studied in a 31 MW reciprocating grate furnace (a cross-current ow combustor), and additional experiments have been made in batch-red pot furnaces. The fuel was forest waste with moisture content of approximately 50%. The combustion in a cross-current ow furnace is generally assumed to start by ignition on the surface of the bed, followed by a reaction front propagating from the surface down to the grate. Measurements and visual observations presented in this paper show, however, that in the case of wet fuels the ignition takes place close to the grate, followed by a reaction front propagating from the grate up to the surface of the bed. Hence, the progress of combustion in the bed is opposite to the expected one. 2001 Elsevier Science Ltd. All rights reserved.
Keywords: Grate combustion; Biofuels; Reaction front

1. Introduction Grate ring is used to burn bio-fuels in small-scale units. Different combustion behaviours can be achieved depending on how fuel and primary air are introduced. The most common congurations of fuel and air supply in continuous systems are cross-current and counter-current ow. An example of units where fuel and air are introduced according to the two modes is shown in Fig. 1. In cross-current ow the fuel is fed at one end of the grate and transported along the grate while burning to completion. Primary air is introduced perpendicular to the grate from below. The airow can be controlled separately in each part of the grate. In counter-current ow the fuel is thrown by a spreader or dropped onto the surface of the bed, and the air is evenly introduced through the grate from below. Combustion in cross-current ow units has been described by several authors [15] as taking place in a bed, which is heated by radiation from ames and refractory surfaces above the bed until it ignites on the upper surface. This generally accepted combustion process can be described as follows. After ignition, a reaction front (Fig. 1(a)) propagates from the surface of the bed down to the
* Corresponding author. Tel.: 46-31-772-1431; fax: 46-31-772-3593. E-mail address: energy.conversion@entek.chalmers.se (B. Leckner). 0016-2361/01/$ - see front matter PII: S0016-236 1(00)00127-7

grate against the direction of the combustion air. The heat, generated in the reaction front, is transported against the ow of combustion air and dries and devolatilises the raw fuel. This allows the reaction front to propagate. Due to the different directions of heat and airow, the heat is not transported downwards far from the position where it is released, and the reaction front is narrow. The heat generated in the reaction front originates from oxidation of fuel and, if not all oxygen is consumed in the narrow reaction front, a char layer will be formed above the reaction front. When the reaction front reaches the surface of the grate, a secondary reaction front, propagating up towards the surface of the bed, burns the char layer previously formed. The counter-current ow has been investigated more than cross-current ow due to its use in gasiers. (In Ref. [6] not less than 37 references on modelling of counter-current ow combustors and gasiers are presented.) In a counter-current ow combustor the air is supplied from the bottom and the fuel from the top (Fig. 1(b)). The combustion stages from the grate and up through the bed are char combustion followed by devolatilisation and drying. When char combustion takes place upstream of devolatilisation, the heat will be produced over a larger region due to the rather slow diffusion of oxygen into the char. The hot product gases provide heat for devolatilisation and drying of the fuel during their way up through the bed. In this system

2001 Elsevier Science Ltd. All rights reserved.

474

H. Thunman, B. Leckner / Fuel 80 (2001) 473481

Nomenclature a b Cp d H h M n k T t u x X Y Greek l n r s constant in correlation of Cp (J/kg K) constant in correlation of Cp (J/kg K 2) specic heat (J/kg K) diameter (m) heating value (J/kg) enthalpy at reference temperature (J/mol) molar mass (kg/mol) number of moles () heat conductivity (W/m K) temperature (K) time (s) velocity of reaction front (m/s) width of reaction front (m) mass fraction in wet fuel () mass fraction in dry fuel ()

wet fuels, such as forest waste. In such cases it can be suspected that ignition takes place close to the grate and, in contrast to the above conventional description, the ignition front propagates through the bed in the same direction as the airow, as shown in Fig. 2(b). The behaviour of a large-scale fuel bed is investigated by measurements in a 31 MW reciprocating grate furnace as well as in pot furnaces. 2. Theory In order to evaluate the limits of the velocity of the reaction front in a fuel batch ignited from one side and with air supplied from the other, a maximum reaction front velocity is estimated by a heat balance across the reaction front. Heat is transported in the reaction front by thermal conduction, radiation and convection. Due to the opposite directions of the reaction front velocity and the airow, the convective heat transfer reduces the heat ux. Here, the interest is in the maximum possible velocity, and therefore the convection is neglected. The radiation and heat conduction are modelled by an effective thermal conductivity of the fuel layer, keff, consisting of a radiative and a conductive constituent [7]: keff 4nsd s T 3 1

air-to-fuel ratio () bed voidage () density (kg/m 3) StefanBoltzmann constant, 5:67 10 8 W=m2 K4 Subscript (capital letters indicate species) ad adiabatic b boiling c char vap vaporisation i species m moisture, moist max maximum n normal (used to indicate normal cubic meter) eff effective s solid u lower v volatiles w water 0 initial

nks

where n is the bed voidage, s the StefanBoltzmann constant, ds the particle diameter, T the temperature and ks the thermal conductivity of the fuel, estimated according to MacLean [8]. The maximum heat transport is achieved if the temperature rises from the initial temperature, T0, to the adiabatic temperature, Tad, in the reaction front. The effective thermal conductivity is temperature dependent. In this evaluation of the maximum velocity the effective thermal conductivity is estimated at the adiabatic temperature. The heat ux across the reaction front must be equal to or smaller than the energy contained in the fuel entering the reaction front. If the width of the reaction front is x, this gives: k eff;max Tad x T0 us H rs 1

nearly all volatiles leave the bed unreacted. The supply of air and fuel creates an even, steady state reaction front, extending over the cross-section of the bed. In the present work the cross-current ow unit is studied by combustion of a fuel batch in a pot furnace. The combustion time for a fuel batch in a pot furnace corresponds to a certain transport distance on the reciprocating grate, Fig. 2, and the progress of combustion is similar to that of a fuel batch in the pot furnace. The concept of the travelling pot furnace is a classical approach [5], which is suitable when the horizontal velocity of the bed along the grate is faster than the vertical velocity of the propagating reaction front inside the bed, making the two-dimensional effects negligible. The purpose of the present work is to investigate the combustion behaviour of a cross-current fuel bed burning

where us is the velocity of the reaction front, H the lower heating value of wet fuel per kilogram dry fuel and r s the density of dry fuel. H and Tad are derived in Appendix A. The maximum velocity of the reaction front is attained when the width of the reaction front is minimum. This width is determined by the drying time t of the particles: x us t: 3

The maximum velocity of the reaction front is obtained by inserting Eq. (3) into Eq. (2): s keff;max Tad T0 : 4 us H rs 1 nt In some instances the propagation rate of the reaction front (kg/m 2 s) is used instead of the velocity of the reaction front

H. Thunman, B. Leckner / Fuel 80 (2001) 473481

475

heat balance of the particle becomes:


4 sT ad

Ts4

2 ddm =2 Xm 1 rs Hvap 1 Xm MH2 O dt

2ks T ds dm s

Tb Ts
T0

3 Cp;H2 O dT : 5

The drying time is then: t 0


d

2 Hvap

Xm rs Xm MH2 O Ts
T0

Cp;H2 O dT

ds dm dd 4ks Ts Tb m

Fig. 1. (a) Grate with cross-current ow. (b) Grate with counter-current ow.

(m/s). The propagation rate is velocity multiplied by apparent density, rs 1 n; of the packed fuel bed. The drying time of a particle is estimated in a simplied way for a onedimensional slab, exposed to a black surrounding, radiating with the adiabatic temperature. The vaporisation is modelled as a drying front moving from the surface of the particle to its centre. The radiative heat ux, received by the surface of the particle, is transported to the drying front within the particle by thermal diffusion. Heat accumulation in the particle is neglected. Under the given conditions the

Fig. 2. Principle combustion behaviours in a cross in a cross-current ow unit: (a) if the bed is ignited from the surface of the bed, or (b) from the surface of the grate. A indicates position of ignition, B vertical section, C reaction front, I unreacted fuel, II drying, III devolatilisation, IV char combustion or gasication, and V ash.

where Ts is the temperature of the particle surface, Tb the boiling temperature, ds the particle diameter, dm the moist core diameter, Xm the moisture content, MH2 O the molar mass of water, Hvap the heat of vaporisation, and Cp;H2 O the specic heat of water (gas). The velocity (or propagation rate) of the reaction front is given by the adiabatic temperature, which is a function of air-to-fuel ratio and moisture content, as shown in Appendix A. Once the velocity (or propagation rate) is known, the corresponding airow can be calculated from the air-to-fuel ratio. In contrast to the above, a fuel batch that is ignited and supplied with air at the same side of the bed will not have a stable reaction front, unless the batch is very long in the direction of the airow. When the bed is ignited, a narrow char layer is formed. The combustion of this layer produces heat that is transported with the gas ow and devolatilises and dries new fuel. This expands the layer of char combustion and more heat will be produced, resulting in an accelerating drying and devolatilisation front. A reaction front with stable velocity does not develop until the heat produced is equal to the heat needed for drying, devolatilisation and, dependent on moisture content, gasication of some of the char. The velocity and the length of the stable reaction front are directly correlated to the airow. Under the condition of no external heat losses, combustion can be maintained until the theoretical limit of moisture content in the feed fuel is such that the lower heating value of wet fuel is equal to zero. For a fuel batch of moderate height the combustion behaviour differs signicantly depending on where ignition takes place in relation to the location of primary air introduction. Not only the order of the combustion stages will differ, as described in Section 1, but also the behaviour of the reaction front velocity. A batch ignited from the surface opposite to where the air is supplied attains a nearly stable reaction front velocity, whereas a batch ignited from the same surface as the air is supplied has a transient reaction front velocity. The latter case is the most favourable combustion situation, not only because the heat ows in the same direction as the gas, but also because the oxygen is consumed in the front, precluding the propagation of a reaction front from the

476

H. Thunman, B. Leckner / Fuel 80 (2001) 473481

Fig. 3. Position of measurements in the reciprocating grate furnace. Gas sampling (A), thermocouples (P) (31 MW boiler, Kvaerner Pulping AB).

opposite side of the bed. Consequently, for ignition on the air supply side, the reaction front propagates with the gas ow through the bed towards the opposite surface of the bed. Except for the top particle layers, which are exposed to radiation from ames and refractories in the freeboard, the combustion is independent of the conditions at the opposite side of the bed. 3. Measurements Measurements in the reciprocating grate furnace were made for three operating conditions using two types of fuel. One fuel was chips of trunk wood with a moisture content of 40%, henceforth referred to as pure wood

chips. The other fuel was forest waste (wood chips 40%, bark 40% and saw dust 20%) with a moisture content of about 50%. Thermocouples were mounted in drilled holes in the rods of the grate to measure the surface temperature of the grate, in various positions along the grate and gas concentrations were measured above the surface of the bed. The total hydrocarbon (THC) concentration was determined by a FID (Flame Ionisation Detection) analyser, but concentrations that exceeded the measurement range of the FID-analyser were analysed by a FTIR (Fourier Transform Infrared) analyser. The furnace and the positions of the measurement equipment are shown in Fig. 3. The grate is 5 m wide and 8 m long. The velocity of the grate was approximately 6 mm/s on the rst 4 m and then 1 mm/s, resulting in a fuel residence time of 1 h and 15 min in the combustor. The residence time on the rst 4 m was 12 min and the corresponding primary airow was between 0.1 and 3 0.2 mn /s. The primary air was saturated by water at 45 C and pre-heated to 150 C before entering the grate. No recirculated ue gas was added to the primary air. The initial height of the fuel layer was approximately 0.5 m. The results presented here were obtained during a larger project reported in Refs. [911]. Experiments in pot furnaces using the same forest waste as in the reciprocating grate furnace were carried out at VTT (Technical Research Centre of Finland, VTT-Energy, Jyvaskyla) and at SP (Swedish National Testing and Research Institute). The measurement set-up at VTT is described in Ref. [12]. The pot furnace has a circular cross-section with a diameter of 244 mm and a height of 300 mm. Three thermocouples are introduced into the bed at different heights. The pot furnace at SP is 480 520 mm 2 by cross-section and 700 mm by height and is provided with nine thermocouples introduced into the bed at different heights. The two furnaces differ mostly in the direction of the airow and the movement of the reaction front; the air moves upward and the reaction front moves downward at VTT and in the opposite directions at SP. Also, the ue gas measurement was carried out upstream of the secondary air injection at VTT and downstream of this location at SP. An average velocity of the reaction front through the bed can be calculated from these data. The intention was that both units should be operated at the same primary gas ow as used in the reciprocating grate.

4. Calculation results The theoretical reaction front velocity, in a bed with air supply from one side and ignition from the other, is estimated by Eq. (4) for 10-mm particles with a dry density of 300 kg/m 3 (forest waste) and moisture contents varying between 10 and 70%. The bed voidage is assumed to be 0.5. This value is a typical average for various particle beds having voidages between 0.4 and 0.6, depending on type and shape of the particles [13]. The resulting reaction

Fig. 4. Calculated maximum propagation rate of the reaction front for different supercial air velocities. Both propagation rate and air velocity are given by the adiabatic temperature, which is a function of air-to-fuel ratio, l , and moisture content. Calculation was made for 10-mm wood particles with different moisture contents: (a) 10, (b) 30, (c) 50 and (d) 70%.

H. Thunman, B. Leckner / Fuel 80 (2001) 473481

477

transfer mainly comes from radiation, a lower temperature reduces greatly the reaction front velocity. A maximum reaction front velocity is attained close to stoichiometric conditions. Since the same adiabatic temperature is reached for both over and under-stochiometric conditions, the velocity of the reaction front is the same at two different airows, as seen in Fig. 4.

5. Results of measurements The surface temperature of the reciprocating grate is presented in Fig. 5, and to give an impression of the combustion activity, the concentration of total hydrocarbons (THC) above the fuel bed is shown in Fig. 6. In the case of pure wood chips the temperature rises at a location between 1.4 and 1.8 m from the entrance of the fuel. For forest waste the temperature was high already in the rst measurement position at 1.4 m. These positions correspond to residence times on the grate of 45 min. The gas concentrations show that hydrocarbons are released early and that devolatilisation goes on during the rst 4 5 m. Computational uid dynamics calculations including combustion verify that the high concentration of hydrocarbons measured above the front part of the grate is a result of gas brought there from the bed and not by the stirring of the gas space above the grate [9,10]. The rst part of the bed, 1 1.5 m, is visible through a view-glass located in the wall of the furnace. In this region there were no ames, but a large amount of smoke left the bed. There was also no stirring of the bed during its transport along the grate in this rst visible part of the bed. The measurements at SP are presented in Fig. 7. The height of the fuel batch in the pot furnace was 0.5 m. The same primary airow as in the bed of the reciprocating grate, 0.16 and 0.27 kg/m 2 s, was wanted, but the airow had to be reduced to about 0.03 kg/m 2 s to maintain a sufciently high temperature inside the bed for propagation of the reaction front. After ignition, the reaction front moved down through the bed layer with a velocity of 0.07 mm/s corresponding to a fuel consumption of 0.012 kg/m 2 s (based on dry substance). The measurements at VTT are presented in Fig. 8. In this case the height of the batch was 0.3 m. To ignite the fuel at the desired airows, the fuel was pre-dried from the moisture content of 56.6 to 40.3%. Both airows, 0.16 and 0.28 kg/m 2 s, gave the same velocity and the same behaviour of the reaction front. The reaction front moved down through the bed layer with the velocity of 0.05 mm/s corresponding to a fuel consumption of 0.008 kg/m 2 s (based on dry substance). When 20% of the mass was gone, the reaction rate increased to 0.16 mm/s corresponding to 0.025 kg/m 2 s (based on dry substance). Results from pot furnace experiments with biofuels [2,4,14], show a similar inuence of air velocity. The velocity of the reaction front (ignition front) increases rst with

Fig. 5. Measured time-average temperature on the reciprocating grate. Filled circles indicate measurement with pure wood chips and unlled circles are measurement with forest waste.

front velocities can be seen in Fig. 4. The calculation shows that the reaction front velocity and the airows for which a reaction front exists are very sensitive to moisture content. 3 For the airows in the reciprocating grate, 0.1 and 0.2 mn /s, the moisture content must be less than 45% for a reaction front to exist. As expected, the lower the moisture content, the higher the velocities of the reaction front can be before extinction takes place. The reason is that a raise of the moisture content increases the energy needed for the vaporisation and decreases the adiabatic temperature. As the heat

Fig. 6. Total hydrocarbon concentration above the bed of the reciprocating grate. Filled circles indicate the second measurement row and unlled circles the rst measurement row (cf. Fig. 3). Measurements from operation with pure wood chips.

478

H. Thunman, B. Leckner / Fuel 80 (2001) 473481

Fig. 7. Bed mass in pot furnace experiment at SP. Moisture content 56.5% (wet basis). Primary airow 0.03 kg/m 2 s. Solid line is measurement, dashed line tted mass reduction during the propagation of the reaction front. The velocity of the reaction front is estimated to 0.012 kg/m 2 s (dry fuel).

increasing air velocity, reaches a maximum, and then drops slightly until the reaction extinguishes, Fig. 9. Higher moisture content in the fuel decreases the velocity of the reaction front [2,4]. In Fig. 9 the reaction velocities quoted above are compared with measurements from Ref. [2]. It is obvious that the higher moisture content reduces the velocity of the reaction front. This agrees with the observations reported in Ref. [4].

Fig. 9. Propagation rate of the reaction front as a function of air velocity. Experimental data from Gort [2]: 30-mm wood cubes, 10% moisture (A), 10-mm wood cubes, 10% moisture ( ), and 10-mm wood cubes, 30% moisture (K). Experimental data from forest waste SP, 56.6% moisture (X), and VTT, 40.3% moisture (W). The ignition rate is based on dry substance.

6. Comparison between calculation and measurement In Fig. 10 the data presented in Fig. 9 are compared with calculations of the maximum velocity according to Eq. (4) for 10-mm particles. The data of Gort [2] for 10-mm wood cubes with moisture contents of 10 and 30% (curves A and

Fig. 8. Bed mass in pot furnace experiment at VTT. Moisture content 40.3% (wet basis). Primary airow: (a) 0.16 and (b) 0.28 kg/m 2 s. Solid lines are measurements, dashed lines represent tted mass reduction during the propagation of the reaction front. The velocity of the reaction front is estimated to (I) 0.008 and (II) 0.025 kg/m 2 s (dry fuel).

Fig. 10. Calculated maximum propagation rate of the reaction front for 10mm wood particles, dotted lines. Solid lines are curve ts of the measurements presented in Fig. 9. The measurements are indicated by capital letters and calculation with small letters, the different moisture contents are indicated by: A,a 10% [2], B,b 30% [2], C,c 40.3% (VTT) and D,d 56.6% (SP).

H. Thunman, B. Leckner / Fuel 80 (2001) 473481

479

B), measured at several airows, t inside the calculated maximum velocities of the reaction front. The model is qualitative, based on several simplifying assumptions, but it provides a measure of the theoretical maximum velocities of the reaction front for given airows. At low airows the situation in the reaction front agrees with the assumptions of the model and there is good agreement between calculation and measurement. For low airows an insulating char layer develops downstream of the reaction front, and all oxygen is consumed inside the bed, resulting in a low air-to-fuel ratio and consequently a low adiabatic temperature. As the airow increases the air-to-fuel ratio raises and the char layer becomes less signicant, which enhances the external heat losses. As the airow raises further the gas residence time inside the reaction front is no longer sufcient for the reaction of the volatiles and all heat generation will not take place inside the bed. Furthermore, an increasing airow enhances the convective heat ow between the gas and the particles, reduces the heat ux across the reaction front and in the end quenches the reaction front. Since the convective heat transfer was neglected in the calculations the reaction front velocity diverges more and more from the calculated maximum velocity with an increasing airow. In the experiments at SP and VTT the forest waste with high moisture content was used. The measured data t the calculated values quite well. At the highest moisture content (curve D, 56.6%), the velocity of the reaction front is higher than the calculated one, but it is inside the limits of accuracy of the model. In this case no reaction front can propagate at the airows present on the reciprocating grate, 0.1 and 3 0.2 mn /s (0.12 and 0.23 m/s at 298 K). At the second highest moisture content, 40.3% (curve C), the reaction front propagates slightly outside of the calculated limit of the airow and the measurements themselves show a certain scatter. However, calculation and measurements agree sufciently well to explain the inuence of moisture; the fuel had to be dried to allow the desired air velocities. The calculation shows that it is possible to achieve two stable velocities of the reaction front at the same airow. The change of velocity observed, the scatter (C in Figs. 8 and 10) could be a consequence of a transition of the velocity from one of the stable velocities to the other.

This observation is supported by the model calculations. 2. Even the highest velocity of the reaction front, 0.35 mm/ s, measured for wood cubes with a moisture content of 10%, is not sufcient for the front to reach the surface of the reciprocating grate within the rst 4 m, if the bed is ignited from the top. For a bed height of 500 mm, it would take 1430 s for the reaction front to propagate from the surface of the bed to the grate. With the velocity of the grate of 6 mm/s, the reaction front would reach the grate 8.6 m after ignition on the surface. The corresponding distance at velocities measured for wet fuels is 18 60 m. In contrast, in the boiler case, according to Figs. 5 and 6, the high temperatures clearly indicate that combustion takes place at or close to the surface of the grate after less than 1.8 m. 3. The temperature on the grate, Fig. 5, shows that heat was generated in the bed before ignition was visually observed on the surface of the bed. 4. There was a heavy evolution of smoke from the bed before the position of ignition on the surface of the bed and before the visible ame zone, according to visual observations in the reciprocating grate furnace. 5. Volatiles were found far from the fuel inlet, Fig. 6. A moist wood-chip particle (50% moisture) dries and devolatilises completely in two minutes at a temperature of 973 K [15]. The temperature inside the bed is around 1300 K [2,14] during char combustion, and the actual time for complete drying and devolatilisation should be less than 2 min. The fuel is transported 0.7 m along the grate in 2 min. Ignition is measured within the rst 1.8 m for pure wood chips on the reciprocating grate, Fig. 5, and all moisture and volatiles should have left the bed before 2.5 m (1.8 0.7 m) distance from the feed point if the reaction had propagated from the top of the bed to the grate. However, Fig. 6 shows that volatiles leave the bed up to 45 m from the beginning of the grate. The conclusion from these facts is that the ignition has not taken place on the surface of the bed, but in the bed, most likely on the surface of the grate. The arguments supporting this observation and that reaction propagates up through the bed, Fig. 2(b), are: 1. In such a case, the theoretical maximum moisture content is attained when the lower heating value for moist fuel is equal to zero. This occurs at a moisture content of 85 90%, excluding heat losses. The reaction takes place inside the bed in a location which is quite well insulated from the surrounding. A moisture content of up to 70 or 80% should therefore generate sufcient heat for drying and devolatilisation. The heat is generated at the bottom of the bed by char combustion and is transported up through the bed by the gas, which dries and devolatilises fresh fuel. For a counter-current bed, which is ignited from the top, on the other hand, this moisture content is too high, since in this case, devolatilisation

7. Conclusions from measurements As mentioned above, it is generally assumed that ignition takes place on the surface of the bed and that the reaction front propagates down through the bed in a cross-ow situation, such as in Fig. 2(a). The measurements show, however, that for the boiler concerned this cannot be the case: 1. In the pot furnaces the reaction front could not propagate through the bed for the same type of fuel and airows as in the boiler, where the fuel burned without problems.

480

H. Thunman, B. Leckner / Fuel 80 (2001) 473481

and combustion takes place in a narrow front. The calculation shows that the maximum moisture content in the range of airows considered is between 35 and 45%. It is also clear from the calculation that the velocity of the reaction front becomes very slow for these moisture contents. 2. Combustion and heat release in the bottom part of the bed generates gas, which transfers its heat to the fresh fuel and exits the bed at a temperature not much higher than that of entering fuel particles. This explains the heavy evolution of smoke, observed to leave the rst 11.5 m of the bed. 3. The reaction front propagates up through the bed and the devolatilisation continues until the devolatilisation front reaches the surface of the bed. This explains the volatile release far down on the grate, see Fig. 2. 4. Ignition of the fuel on the surface of the grate takes place as soon as the fuel has reached the ignition temperature. Larger particles need longer time to heat up and ignite, and consequently there is a later ignition of the pure wood chips than of the forest waste that includes a large portion of ne saw-dust, although the moisture content was 10% higher, Fig. 5. Large scale mixing of the fuel bed could be another reason for the temperature raise before 1.8 m from the fuel inlet, but ignition or stirring caused by the transport of the bed along the grate were not observed on the surface during the rst 1 1.5 m from the fuel inlet. Another fact that is against large scale mixing is the tendency of wood chips to stick together.

9. Conclusions The traditional description of the combustion behaviour, with ignition on the top of the bed and a reaction front propagating down through the bed, Fig. 2(a), was not observed on the reciprocating grate studied. As the fuel moisture content raises, the reaction front will have greater difculties to propagate from the surface of the bed down to the surface of the grate. The possibility for the reaction front to propagate depends on the airow and the fuels moisture content. In the reciprocating grate studied the airows were 3 0.1 and 0.2 mn /s, and for these airows the limiting moisture content is between 35 and 45% according to calculation. This conclusion is also supported by pot furnace experiments. Both calculation and measurements show that the velocity of the reaction front is very slow for fuel having high moisture content. A combustion behaviour, that explains the measurements and allows fuel with a high moisture content to burn, takes place if the reaction front moves in the same direction as the primary airow, Fig. 2(b). In such a case the bed has to be ignited close to the surface of the grate. The char combustion produces hot gases, providing heat for drying and devolatilisation on their way up through the bed. This combustion situation should not be sensitive to disturbances in airow and moisture content of the fuel due to the stability of the char combustion. It becomes more similar to a counter-current bed than to the traditional description of a cross-current unit. The cross-current combustion of wet fuel should be described by a fuel batch ignited from the side where the primary air is introduced and not from opposite side as is traditionally assumed. If this is the actual combustion behaviour, it will have a great inuence on the principles for boiler design. The ignition refractory arch is not needed for ignition. Instead, shape, position and movement of the metal rods of the grate control where and when ignition takes place. Further experiments in reciprocating grate furnaces and in pot furnaces are needed to nally conclude if this is the typical combustion behaviour of wet biofuel, but the circumstantial evidence presented gives a strong indication that this is the case. A general conclusion from this work is that, if it is possible to create a steady state ignition at or close to the surface of the grate, the reaction front always propagates from the grate up to the surface of the bed. There are locations for burning char particles that can ignite the bed at or close to the grate surface in all designs of travelling grates and the reciprocating grate is not a special case.

8. How can the fuel ignite on the grate? There are at least two mechanisms that can explain a steady state ignition of the bed at the surface of the grate: 1. burning char particles that are not transported along the grate by the reciprocating rods; 2. conduction of heat through the metal rods. It is nearly impossible to extinguish a burning char particle by an airow containing oxygen. The most probable reason for ignition of the bed is by such burning particles. The initiation of combustion on the grate is achieved by the ignition of a pile of wood chips soaked in diesel oil. The burning pile heats up the furnace and the metal rods. The fuel is then slowly fed and is ignited by the burning pile. More fuel is introduced until a stable operating condition is attained. The steps of the reciprocating grate provide places for small burning char particles to stay, and they are always supplied by oxygen from the primary air. These particles ignite the fuel above them and heat up the metal rods and new fuel particles are heated by backward transportation of heat through the metal rods. New char particles replace the old ones as they burn out or are transported away, and so a steady state ignition procedure is maintained.

Acknowledgements This work was supported by a grant from the Small Scale Combustion Programme of the Swedish National Energy Administration. Part of the work was carried out under the European Union contract JOR3-CT96-0059.

H. Thunman, B. Leckner / Fuel 80 (2001) 473481

481

Appendix A The adiabatic temperature for different air ratios is estimated by Tad ni Cp;i dT {Cp;i ai bi T} H T0 i 2 3 2 3 2 ni bi =2 Tad ni ai Tad A i i 2 3 2 2 3 3 2 H ni bi =2 T0 ni ai T0 0 A1
i i

system of equation: YC Ychar =MC nCH4 ;v YO2 =MO2 nCO2 ;v YH2 =MH2 2nCH4 ;v Hu nC;c hC hO2 nCH4 ;v hCH4 nCO;v hCO

nCO;v

nCO2 ;v ;

0:5nCO;v nH2 O;v ; hCO2 hO2 hO 2

nH2 O;v ; A6

hCO2 hCO2

2hH2 O

where h is the enthalpy at reference temperature. References


[1] Ford NWJ, Cooke MJ, Sage PW. Modelling of xed bed combustion. Fuel Processing Technology 1993;36:5563. [2] Gort R. On the propagation of a reaction front in a packed bed; thermal conversion of municipal waste and biomass. Academic Dissertation, University of Twente. ISBN 90-9008751-6, 1995. [3] Ryu C, Shin D, Choi S. Simulation of waste bed combustion in the municipal solid waste incinerator. Second International Symposium on Incineration and Flue Gas Treatment Technologies. IChemE, The University of Shefeld, 46 July 1999. [4] Saastamoinen JJ, Horttanainen M, Taipale R, Sakomaa P. Propagation of ignition front in fuel beds of wood particles. Combustion and Flame 2000;123:21426. [5] Marskell WG, Miller JM. Mode of combustion of coal on a Chain grate stoker. Article series in Fuel in Science and Practice, vol. 25, 1946. p. 411, 5062, 7885, 10913, 15962. [6] Smoot LD. Fundamentals of coal combustion. Amsterdam: Elsevier, 1993 (ISBN 0-444-89643-0). [7] Siegel R, Howell JR. Thermal radiation heat transfer. 3rd ed. Philadelphia, PA: Taylor & Francis, 1992 (ISBN 0-89116-271-2). [8] MacLean JD. Thermal conductivity of wood. Transactions of the American Society of Heating and Ventilation Engineers 1941;47:32354. [9] Ghirelli F, Thunman H, Amand L-E, Leckner B. Comparison between CFD calculation and measurement on a grate boiler burning biofuel. ToTEM14 and SwedishFinnish ame-day 1999. IFRF-net, the Swedish and the Finnish Flame Research Committees of the International Flame Research Foundation in co-operation with the Swedish Bioenergy Association, Vaxjo University, 2930 September 1999. [10] Thunman H, Amand L-E, Ghirelli F, Leckner B. Modelling and verifying experiments on the whole furnace. Report to the European Commission JOR3-CT96-0059, Department of Energy Conversion, Chalmers University of Technology, Goteborg, Sweden, 1999. [11] Padinger R. et al. Reduction of nitrogen oxide emissions from wood chip grate furnaces. Final report to the European Commission JOR3CT96-0059, 1999. [12] Horttanainen MVA, Saastamoinen JJ, Sarkomaa PJ. Ignition front propagation in packed beds of wood particles. IFRF Combustion Journal, www.ifrf.net, Article Number 200003, ISSN 1562-479X, May 2000. [13] Kaviany M. Principles of heat transfer in porous media. Berlin: Springer, 1991 (ISBN 0-387-97593-4). [14] Ronnback M, Axell M, Gustavsson L, Thunman H, Leckner B. Combustion process in a biomass fuel bed experimental results. Progress in Thermochemical Biomass Conversion, Tyrol, Austria, 1722 September 2000. [15] Palchonok GI, Dikalenko VI, Stanchits LK, Borodulya VA, Werther J, Leckner B. Kinetics of the main stages of uidized bed combustion of a wet biomass particle. In: Preto FDS, editor. Proceedings of the 14th International Conference on Fluidized Bed Combustion, Vancouver, Canada. New York: ASME, 1997. p. 12534.

where n is the number of moles of species i, Cp the specic heat, T the temperature and H the heating value of wet fuel per kilogram dry fuel: H Hu Hvap Xm =1 Xm A2

where Hu is the lower heating value of dry fuel, Hvap the heat of vaporisation and Xm =1 Xm the moisture content based on dry fuel. The number of moles of each species is calculated from the assumption that the char is pure carbon and the volatiles consist of methane, water, carbon monoxide, carbon dioxide and that the devolatilisation is thermally neutral. The number of moles of each specie as a function of air-to-fuel ratio l can be expressed as: nN2 l3:77nC;c 2nCH4 ;v 2nCH4 ;v nCH4 ;v 0:5nCO;v 0:5nCO;v

1 1nC;c

nO2 l nC 0

nCO2 nCO2 ;v nCO 0 nCH4 0 nH2 O nH2 O;w

nC;v

nCO;v

nH2 O;v

2nCH4 ;v

A3

nO2 0 nC 1 nCO 1 nCH4 1

lnC;c lnC
nCH4 ;v

nCO2 nCO2 ;v

lnCO;v lnCH4 ;v
nH2 O;v 2lnCH4 ;v A4 A5

nH2 O nH2 O;w nH2 O;w Xm =1 nC;c Ychar =MC

Number of moles in the moisture is: Xm MH2 O :

Number of moles in the char is:

where Y is the mass fraction in dry fuel and M the molar mass. The molar composition of the volatiles is given by the

Paper VIII

Progress in Thermochemical Biomass Conversion 17-22 September, 2000, Tyrol, Austria

Combustion processes in a biomass fuel bed Experimental results


Marie Rnnbck, Monica Axell, Lennart Gustavsson SP Swedish National Testing and Research Institute, BOX 857, SE-501 15 Bors, Sweden Henrik Thunman, Bo Leckner Chalmers University of Technology, Energy Conversion, SE-412 96 Gteborg, Sweden

ABSTRACT: Combustion processes in a biomass bed are investigated experimentally. Special attention is paid to the influence of primary airflow and particle properties on the ignition front, its temperature and on the composition of the gas leaving the front. Two test rigs have been built: a large rig in the same size as a boiler for domestic use and a small laboratory test rig. In both rigs the ignition front moves in opposite direction to the primary airflow. Three combustion regimes are identified: a sub-stoichiometric regime with incomplete consumption of oxygen, a sub-stoichiometric regime with complete consumption of oxygen and an over-stoichiometric regime. The results show that a fuel with higher density and thermal conductivity (but in other respects similar to other fuels) has a wider sub-stoichiometric regime where oxygen is completely consumed. If the particle size is increased (for the same fuel quality) the airflow range of this regime becomes shorter and starts at higher airflow.

INTRODUCTION Fixed or moving bed combustion is the most common technology for biofuels. Incineration of wastes and gasification of biomass are also often performed in a bed. The design of the grate (stationary, vibrating, reciprocating or moving) on which the bed rests and the method of fuel supply (from underneath, from the side or from above) depend on fuel characteristics and on the size of the plant. Devices with bed combustion have as common features drying, devolatilization, gasification and combustion occurring in particular zones in the bed. The extension of these zones depends on fuel and air supply and on the initial ignition of the fuel bed. The bed may be operated as a gasifier producing a combustible gas, but it may also be operated with excess air. The purpose of a boiler is to attain complete combustion, although this is achieved downstream of the bed in the form of gas phase combustion controlled by secondary air. For fuels with a high content of volatile matter, the gas combustion downstream of the bed is crucial for emission control. The fuel bed is the first stage in the combustion process and generates the conditions for the latter part. A review of available literature on the experimental simulation of solid fuels confirms that the knowledge of coal combustion is more detailed than that of biomass and municipal solid waste. The knowledge of biomass gasification devices today is extensive; however, see for example La

Progress in Thermochemical Biomass Conversion 17-22 September, 2000, Tyrol, Austria

Fontaine and Reed (1) and this knowledge also suits the combustion bed. Also, recently several publications have described experiments on high volatile fuels in onedimensional batch-type reactors. The fuel beds are ignited on the top and the ignition front propagates against the primary airflow. The results from such experiments can be extended to steady-state combustion on a stationary or moving grate. Gort (2) did a thorough investigation on the effects of moisture and volatile content, as well as of particle size. He burned wood cubes, coke and municipal solid wastes in a batch-type laboratory grate furnace. He distinguished three global reaction regimes, depending on the ratio of ignition rate and superficial air velocity. Shin and Choi (3) studied the effects of air supply rate, fuel particle size and calorific value on the combustion of wood cubes in a similar combustor. Depending on the availability of oxygen, distinct reaction zones were identified. Fatehi and Kaviany (4) performed similar experiments in a 7x7-cm furnace on wood spheres and described two combustion zones. Furthermore, Kuo et al. (5) pointed out the importance of the arrangement of air supply for combustion efficiency. In their experimental device primary air could be supplied through the grate and through the walls of the fuel chamber. They found that the CO-emissions were related to the oxygen content of the flue gas, depending on the mixing of gas and air in the bed. Horttanainen et al. (6) presented results from experiments on biomass particles of various sizes and moisture content. They focused on the speed of the ignition front as an important factor determining the release of volatiles and that affects the combustion power and the stability of combustion. The aim of the present work is to further analyse the influence of primary airflow and fuel particle parameters on the combustion process. The work ranges from substoichiometric to over-stoichiometric supply of primary air to the bed, to cover the influence of the parameters studied. Experiments were performed with the combustion front moving both upward (as in the studies referred to) and downward (as in a modern boiler with down-draught combustion). The work aims at forming a basis for modelling of wood burning, as well as for design of small boilers. The experiments were conducted so that the results can be used for both batch and continuously fired plants. EXPERIMENTAL CONDITIONS The experiments were carried out in two test rigs, here called the large and the small rig. The small rig is cylindrical with a diameter of 0.2 m and a height of 0.6 m. The inner wall is made of 3-mm stainless steel to minimise the heat capacity of the wall. It is insulated with a 50-mm glass-wool cover. Primary air is supplied through the steel grate at the bottom of the cylinder and exhaust gases leave through the top, where secondary air is supplied before the gases enter the chimney. The fuel is put on the grate and ignited by a torch from the top. Then, the ignition front moves against the primary airflow towards the grate, see Figure 1. This is a onedimensional representation of a continuously burning fuel bed on a grate, where fuel and air are supplied co-currently and the fuel is ignited on the top of the bed, see e.g. Thunman and Leckner (7). Two thermocouples (type K) located 150 mm and 300 mm above the grate measure the temperatures inside and above the fuel bed. Gas analysis is carried out in the gases leaving the bed, before they reach the secondary air supply. The large rig is of the same size as a domestic boiler. The design is chosen to ensure well-defined start- and boundary conditions for the fuel bed. The rig is equipped to measure airflow, weight loss and bed height. Temperatures can be measured upstream, in and downstream of the fuel bed and in the grate by shielded 1-mm thermocouples (type K), mounted both from the side (orthogonal to the movement of the ignition

Progress in Thermochemical Biomass Conversion 17-22 September, 2000, Tyrol, Austria

front) and from the air supply. A comparison with shielded 2- and 3-mm thermocouples showed that the mounting of the thermocouples and their sizes gave no significant differences in results. The fuel bed can be observed through sight glasses. The rig works with down-draught combustion, i.e. the primary air is supplied from the top of the fuel bed and the fuel is ignited from the bottom (in this case by electrical spirals in the grate). The ignition front moves against the primary airflow towards the top of the fuel bed, see Figure 2. The fresh fuel moves downwards into the ignition front and influences the ignition front with its weight.

Figure 1 Ignition front moving against the Figure 2 Ignition front moving against the airflow in the small experimental rig. airflow in the large experimental rig. The fuels were pellets and wood cylinders (pine). The pellets were made of compressed sawdust and had a diameter of 8 mm. The wood cylinders had three diameters, 8, 12 and 34 mm. The proximate analyses and elemental composition of the fuels were almost identical, as seen in Table 1. The density and the thermal conductivity of the pellets are about twice those of the wood. The pellets were burned both in the large and in the small rig, while the wood cylinders were burned only in the small one. Table 1 Fuel characteristics. Pellet C (weight-%, dry basis) H (weight-%, dry basis) O (weight-%, dry basis) N (weight-%, dry basis) S (weight-%, dry basis) Moisture (weight-%) Hi (MJ/kg, dry basis) Thermal conductivity (J/mK) (8) Diameter (mm) Fuel density (kg/m3) Bed density (kg/m3) Bed void Air mass flow (kg/ms) 50.5 6.2 43.1 0.15 0 8.3 20.2 0.32 8 1259 680 0.46 0.035-0.41 Wood 51.4 6.7 41.7 0.047 0 9.1 19.3 0.16 12 585 305 0.48 0.07-0.53

8 579 307 0.47 0.07-0.53

34 581 279 0.52 0.07-0.53

Progress in Thermochemical Biomass Conversion 17-22 September, 2000, Tyrol, Austria

Emissions of CO, CO2, O2 and THC (Total Hydro Carbon) were measured on-line. The gas was sampled as close to the fuel bed as practically possible. In the large rig, gas was extracted 50 mm downstream of the grate. The distance to the ignition front is then 100-350 mm, increasing as the ignition front moves upward in the fuel bed. In the small rig, gas was extracted in a position 450 mm above the grate. The distance to the ignition front is then up to 450 mm. The gas probe is cooled to quickly stop any combustion reaction, and the temperature of the extracted gas falls to 100C in less than 200 mm. Because of the high levels of CO downstream of the fuel bed, a dilution devise was used to dilute the sampled gases about 10 times. A conventional instrument could then measure the CO-concentration. The dilution factor was continuously calculated by comparing the CO2-concentration before and after dilution. RESULTS AND DISCUSSION The ignition rate of the fuel bed is defined as kg ignited fuel per grate area and time (kg/ms). In the small rig this quantity was determined by the times when two thermocouples at a distance of 150 mm reached 500C, multiplied by the (original) bed density of the fuel. The exact level of temperature chosen (500C) is not critical, because the temperature change in a fuel layer is quite fast, see Figure 3. The primary air mass flow through the grate is presented as kg air per grate area and time (kg/ms). Figure 3 shows the temperature at 150 mm and at 300 mm above the grate. The short peak in temperature is the front temperature, most likely influenced by the surface temperature of the nearby particles. In this example the air mass flow is low, and all fuel is not converted but accumulates upstream of the ignition front. The temperature in the partly converted layer is about 100C lower than in the front. As the bed shrinks the upper thermocouple finds itself above the bed. Because of radiative cooling by the walls the uncorrected temperature is about 200C lower than the bed temperature.
1300 1200 1100 Temperature (C) 1000 900 800 700 600 500 400 300 200 100 0 0
300 mm 150 mm

65 60 55 50 45 40 35 30 25 20 15 10 5 0 60 Time (min) Air flow (m3/h)

Flow

20

40

Figure 3 Temperatures at 150 mm and 300 mm above the grate and airflow in the small rig. The fuel was 8-mm wood cylinders and the air mass flow 0.11 kg/ms. The temperature curves show how the width of the accumulated layer broadens as the front moves downwards. When the ignition front reaches the grate, there is a peak in temperature (at about 43 minutes) caused by combustion of the accumulated char.

Progress in Thermochemical Biomass Conversion 17-22 September, 2000, Tyrol, Austria

As can be seen from the gas analysis in Figure 4 there is also a peak in CO- and a dip in CO2 emissions followed by a peak in CO2 while CO goes to zero. The peak in CO coincides with a dip in temperature. When char combustion starts the production of water vapour from drying and combustion falls and the CO-emission increases because of lack of water vapour for CO-oxidation as illustrated by the expression: -d[CO]/dt = ko[CO][O2]1/2[H2O] 1/2exp(-E/RT) from Howard (9). The final increase in CO2 follows from an increase in O2-concentration, enhancing combustion, when the bed is almost finished and very shallow. This dip-and-peak behaviour is not seen at high air mass flows (above 0.17 kg/ms for wood cylinders and above 0.40 kg/ms for pellets).
30 25 O2, CO2, CO (%) 20 15 10 5 0 30000 25000
CO

THC

20000 15000 10000 5000 0 60 Time (min) THC (ppm)

CO2

O2

20

40

Figure 4 Emissions of O2, C O2, CO and THC downstream of the grate in the small rig. Data as in Figure 3. The peaks in O2 with corresponding dips in CO2, CO and THC originate from back-blowing the analysis system. In the large rig electrical heating of the grate ignites the fuel. The ignition front moves upward, opposite to the direction of the airflow. The temperature profiles are steep and parallel during the devolatilization. Figure 5 shows temperatures at different levels in the centre of the bed between 10 mm and 400 mm above the grate. Two phases, devolatilization (ca 125-180 min) and char combustion (after 180 min) are indicated by the temperature curves and by the gas analysis downstream of the grate. Figure 6 shows the gas analysis from the same experiment. After bed break-through the bed consists mainly of char and the fuel burns from the top downwards. The temperature in the top layer, where the air meets the fuel, shows peaks to maximum 1100C. The temperature increases continuously in the remaining bed during char combustion. In this rig, the ignition rate of fresh fuel cannot be determined in the same manner as in the small rig, because the bed shrinks during conversion of the fuel. Instead, the time between the moment when the gas downstream of the bed reaches 500C and when the ignition front breaks through the top of the fuel bed was measured. The ignition rate in kg/ms is calculated by dividing the mass of the fuel in the bed by the grate area and the corresponding time. The break-through of the ignition front at the surface of the bed can be determined in several ways: visual observation of flames above the bed, a rise in temperature in the topmost fuel layer and changes in the gas analysis. In the ignition front there is a balance between heat generated by chemical reactions and heat transfer into the fuel particles, to the combustion air and to new layers of fresh

Progress in Thermochemical Biomass Conversion 17-22 September, 2000, Tyrol, Austria

1200

1000

10 mm

800

Temperature (C)

30 mm 600 50 mm 100 mm 150 mm

400 175 mm 200 400 mm

120

140

160

180

200 220 Time (min)

Figure 5 Temperatures in C at levels above the grate from 10 mm to 400 mm in the large rig. The air mass flow was 0.12 kg/ms and the fuel 8-mm pellets.
30 25 O2, CO2, CO (%) 20 15
CO2
CO

10 5
O2

0 120

140

160

180

200 220 Time (min)

Figure 6 Emissions of O2, CO2 and CO downstream of the grate in the large rig. Data as in Figure 5. The peaks in O2 with corresponding dips in CO2, CO and THC originate from back-blowing the gas analysis system. fuel. The heat transport to new fuel is dominated by radiation, because the direction is opposite to the airflow. The heat transport into the fuel particles depends on the thermal conductivity of the material. The particles used were all thermally thick, i.e. the Biot number = hDp /k > 1, and a temperature gradient was present in the particles during devolatilization and char combustion. As a particle is heated, it dries and devolatilizes. Water vapour and combustible gases are produced and transported to the surface of the particle and out into the gas between the particles, where the gases ignite if the conditions are suitable. Devolatilization of biomass starts already at temperatures about 200C and spontaneous ignition of

Progress in Thermochemical Biomass Conversion 17-22 September, 2000, Tyrol, Austria

wood at 500-600C. If a pilot flame is present, such as the flame front in a fuel bed, ignition can take place already at 300-400C, Saastamoinen et al. (10). When the volatile matter has left the particles, heterogeneous reactions, such as gasification with CO2 and H2O and char combustion, begin. Char combustion is a slower process than gas combustion and demands a higher temperature ( 800C) to be complete. As long as oxygen is present combustion dominates, since gasification is slow. COMBUSTION REGIMES Figure 7 shows the maximum temperatures and Figure 8 the ignition rates in the small rig with 8-mm wood cylinders and 8-mm pellets as fuel. The lines in Figure 8 represent the theoretical stoichiometric combustion rate that would occur if the fuel was exposed immediately to oxygen, i.e. if the fuel particles were thermally infinitely thin.
1500

1300 Temperature (C)

1100

sub-stoich. O20, wood sub-stoich. O20, pellet

900

700
Tmax wood 8 mm Tmax pellet 8 mm

500 0.00

0.10

0.20

0.30 0.40 0.50 0.60 Air mass flow (kg/ms)

Figure 7 Maximum front temperature for 8-mm wood cylinders and 8-mm pellets as function of air mass flow rate. The sub-stoichiometric regimes with complete oxygen consumption for the different fuels are marked in the figure. Based on the results, the combustion process has been divided into three regimes. (1) Sub-stoichiometric combustion with incomplete consumption of oxygen. This regime is found at low primary airflow and is characterised by a clear division in time in a devolatilization phase followed by char combustion. During the devolatilization, partly converted fuel accumulates downstream of the ignition front. Despite the sub-stoichiometric condition, oxygen is not fully consumed in the bed, either because of slow kinetics or of insufficient mixing between the devolatilized combustible gases and the primary air. As a result, just after the bed where the gas sampling was made, unburned gases as well as oxygen can appear in the sampling gas. The ignition rate and the ignition front temperature are strongly influenced by the primary airflow. Figure 3 and Figure 4 show combustion in this regime from 7 to 11 minutes. (2) Sub-stoichiometric combustion with a complete consumption of oxygen. This regime occurs at higher primary airflow, see Figure 7 and Figure 9, and is characterised by complete consumption of oxygen in the bed. A layer of partly converted

Progress in Thermochemical Biomass Conversion 17-22 September, 2000, Tyrol, Austria

fuel forms downstream of the ignition front, but the layer does not grow in thickness as in Regime 1. The devolatilization phase is followed by a short char combustion phase. The influence of primary airflow on ignition rate and ignition front temperature is not so pronounced. The maximum ignition rate and front temperature are found in this regime. Figure 5 and Figure 6 show an example.
0.12 0.10 Ignition rate (kg/ms) 0.08 0.06 0.04 0.02 0.00 0.0
Ign. rate wood 8 mm Ign. rate pellet 8 mm Stoichiometry wood 8 mm Stoichiometry pellet 8 mm

0.1

0.2

0.3 0.4 0.5 0.6 Air mass flow (kg/ms)

Figure 8 Ignition rate for 8-mm wood cylinders and 8-mm pellets as function of air mass flow rate. The lines represent stoichiometric combustion. (3) If the primary airflow is increased even more the combustion moves into the overstoichiometric regime and excess oxygen leaves the fuel bed. As the excess air is heated, the bed cools. With increasing airflow the ignition rate slows down and the front temperature decreases. Finally, the combustion is extinguished. The final char combustion phase is negligibly short. INFLUENCE OF DENSITY AND THERMAL CONDUCTIVITY The density and the thermal conductivity of the pellets are about twice those of the wood cylinders, but the fuels used differ very little in elemental composition and moisture content. In general, the differences in front temperatures and ignition rates are small for the fuels, as seen in Figures 7 and 8. The ignition rate of pellets is less influenced by airflow in a wider range, and the temperature in the front is higher for pellets at higher airflows when for wood cylinders. The interesting difference lies in the combustion regimes. Pellets have a wider sub-stoichiometric, full oxygen consumption regime, starting at a lower airflow. At the temperatures and residence times during the change from Regime 1 to Regime 2, the conversion of CO, according to the rate expression of Howard (9), should be complete and no oxygen should be present. The reason for finding CO and O2 downstream of the bed is then most likely bad mixing. The pellets are thermally thinner than wood, and their higher thermal conductivity leads to a higher devolatilization rate at the same airflow. This may have contributed to extending the sub-stoichiometric regime with full oxygen consumption to higher airflows in the case of pellets. The higher temperature in this regime (compared to wood) can be explained by a higher burning rate (kg mass loss/area and second) for pellets.

Progress in Thermochemical Biomass Conversion 17-22 September, 2000, Tyrol, Austria

INFLUENCE OF PARTICLE SIZE Figure 9 shows the maximum temperature and Figure 10 the ignition rate in the small rig for 8-, 12- and 34-mm wood cylinders. No gas analysis was carried out on the 12and 34-mm wood cylinder experiments and the temperature curves were used to separate the combustion regimes.
1500

1300 Temperature (C)

1100
sub-stoich, 8 mm, O2 0

900

sub-stoich, 12 mm, O2 0 sub-stoich, 34 mm, O2 0

700
T max wood 8 mm T max wood 12 mm T max wood 34 mm

500 0.0

0.1

0.2

0.3 0.4 0.5 0.6 Air mass flow (kg/ms)

Figure 9 Maximum front temperature for 8-, 12- and 34-mm wood cylinders as function of air mass flow rate. The sub-stoichiometric combustion regimes with complete oxygen consumption for the different sizes are marked in the figure.
0.12 Ignition rate (kg/ms) 0.10 0.08 0.06 0.04 0.02 0.00 0.0
Ign. rate wood 8 mm Ign. rate wood 12 mm Ign. rate wood 34 mm Stoichometry wood

0.1

0.2

0.3 0.4 0.5 0.6 Air mass flow (kg/ms)

Figure10 Ignition rate for 8-, 12- and 34-mm wood cylinders as function of air mass flow rate. The line represents stoichiometric combustion. At low airflow the front temperature is lower and the ignition rate appears to be somewhat higher for the largest fuel particles. However, the variations are generally

Progress in Thermochemical Biomass Conversion 17-22 September, 2000, Tyrol, Austria

small and have to be compared with the accuracy of the experiments, especially for large particles whose size reduces the precision in the calculation of the ignition rate. The sub-stoichiometric regime with complete oxygen consumption starts at a higher airflow for the larger particles. This can be explained by the larger particle being thermally thicker. The resistance towards heat flow into the particles increases with the diameter. More of the heat produced is transported to new layers of fresh fuels and not towards the centre of the fuel particle. This is reflected in a higher ignition rate and a lower front temperature at low airflow. Less combustible gases reach the surface of the large particles, the power produced is lower and the burning rate also becomes lower. When the particle size increases the transition from Regime 2 to Regime 3 is extended to higher airflow. The reason for this has to be further investigated. Parameters of interest are devolatilization rate compared to ignition rate, heat transfer between particles and primary airflow. COMPARISON BETWEEN THE TWO RIGS A comparison between ignition rates from 8-mm pellets burned in the small and in the large rig shows a reasonable agreement as seen in Figure 11. The front temperatures agree at low air mass flows, but at higher flows the temperatures in the large rig are lower, see Figure12. Four differences between the rigs can be noted: (1) The small rig has a smaller grate area and channelling inside the fuel bed is less probable than in a large rig. In the large rig, horizontal temperatures were measured in the bed and in the gas phase downstream of the bed. At several occasions an uneven start phase in the large rig resulted in an uneven ignition profile. An uneven ignition front could lead to a faster ignition of the bed, but the measured ignition rates are consistent and similar in the rigs (Figure 11). The method of measuring the ignition rate over the batch in the large rig is possibly removing any unevenness in the ignition profile. (2) In the small rig there is a risk for air passing between the outer edge of the fuel bed and the walls without taking part in the combustion process. In the large rig there is a seal between the grate and the walls, and any air and gas channelling between the fuel bed and the walls is forced to move over the surface of the grate and leave through holes in the grate. This difference in design may lead to higher O2 levels downstream of the bed in the small rig. (3) The large rig has a larger heat loss during batch firing, caused by the heavier design. The ignition front, however, is well insulated by fresh fuel on one side and by partly converted fuel or char on the other, and there is only a heat loss at the walls. (4) The most noticeable difference between the results from the two rigs is in the temperature of the reaction front (Figure12) at airflows higher than 0.3 kg/ms when the temperature is 200C higher in the small rig. Also the burning rate in the small rig is higher at these airflows. There are two explanations for the higher temperatures. The first is that in the small rig secondary air is supplied a few decimetres downstream of the bed, and the combustible gases leaving the bed burn with a gas flame that radiates upstream and heats the bed. At lower airflow the ignition front is covered by partly converted material and is not influenced by the heat from the flame. At higher airflow the layer of partly converted material is very thin and the front temperature increases by the heat from the flame. The second explanation is that in the combustion front of the small rig, particles with reduced size may start

Progress in Thermochemical Biomass Conversion 17-22 September, 2000, Tyrol, Austria

to fluidise at high airflow, as shown in experiments of Fatehi and Kaviany (4), and that would increase the conversion rate and the temperature in the front.
0.12 Ignition rate (kg/ms) 0.10 0.08 0.06 0.04 0.02 0.00 0.0

Large test rig, pellet 8mm Small test rig, pellet 8mm

0.1

0.2

0.3 0.4 0.5 0.6 Air mass flow (kg/ms)

Figure 11 Ignition rate for 8-mm pellets as function of air mass flow rate in the large and in the small test rigs.
1500

Temperature (C)

1300

1100

900

700
Large test rig, pellet 8mm Small test rig, pellet 8mm

500 0.0

0.1

0.2

0.3 0.4 0.5 0.6 Air mass flow (kg/ms)

Figure12 Maximum front temperature for 8-mm pellets as function of air mass flow rate in the large and in the small test rigs. COMPARISON WITH OTHER AUTHORS Figure13 shows the maximum temperature in the small rig using 8-mm wood cylinders with a moisture content of 8.3 % compared to 10-mm wood cubes with a moisture content of 10 % from Gort (2). Figure14 shows the ignition rate of the same cylinders and cubes and of 5-20 mm wood chips with a moisture content of 10.8 % from Horttanainen et al. (6). The results coincide quite well, except for the higher ignition rate of

Progress in Thermochemical Biomass Conversion 17-22 September, 2000, Tyrol, Austria

the wood chips. Horttanainen et al. suggest that smaller chips act as pilot flames for larger chips, and this enhances the ignition rate.
1500

1300 Temperature (C)

1100

900

700 wood 8 mm wood 10 mm, Gort 500 0.0 0.1 0.2 0.3 0.4 0.5 0.6 Air mass flow (kg/ms)

Figure13 Maximum front temperature for 8-wood cylinders and 10-mm wood cubes as function of air mass flow rate.
0.12 0.10 0.08 0.06 0.04 0.02 0.00 0.0 wood 8 mm wood 10 mm, Gort wood 5-20 mm, Horttanainen 0.1 0.2 0.3 0.4 0.5 0.6 Air mass flow (kg/ms)

Figure14 Ignition rate for 8-mm wood cylinders, 10-mm wood cubes and 5-20-mm wood chips as function of air mass flow rate. Figure15 and Figure16 show the maximum temperature and the ignition rate in the small rig with 34-mm wood cylinders having a moisture content of 8.3 % compared to 30-mm wood cubes with a moisture content of 10 % from Gort (2). The results from the 34-mm wood cylinders are more scattered, probably because the cylinders form a less homogenous bed, which decreases the accuracy of the calculation of the ignition rates. Despite the scatter, the temperatures coincide well. For the larger particles (the cylinders) the maximum ignition rate is found at a higher air mass flow.

Ignition rate (kg/ms)

Progress in Thermochemical Biomass Conversion 17-22 September, 2000, Tyrol, Austria

1500

1300 Temperature (C)

1100

900

700 wood 34 mm wood 30 mm, Gort 500 0.0 0.1 0.2 0.3 0.4 0.5 0.6 Air mass flow (kg/ms)

Figure15 Maximum front temperature for 34-wood cylinders and 30-mm wood cubes as function of air mass flow rate.
0.12 0.10 Ignition rate (kg/ms) 0.08 0.06 0.04 0.02 0.00 0.0

wood 34 mm wood 30 mm, Gort 0.1 0.2 0.3 0.4 0.5 0.6 Air mass flow (kg/ms)

Figure16 Ignition rate for 34-wood cylinders and 30-mm wood cubes as function of air mass flow rate. Other authors have identified combustion regimes similar to this work. Gort (2) divides the combustion process into 1) a partial gasification regime characterised by accumulation of partly converted fuel and full oxygen consumption, 2) a complete gasification regime characterised by a layer of partly converted fuel with constant thickness and full oxygen consumption and 3) a complete combustion regime characterised by a layer of partly converted fuel with constant thickness and incomplete oxygen consumption. These regimes coincide in principle with the ones defined in the present paper with the exception that Gort found no excess oxygen after the bed in the first regime. Shin and Choi (3) identified three combustion modes depending on the air supply rate. When the air supply is low, the reaction rates are controlled by the oxygen supply

Progress in Thermochemical Biomass Conversion 17-22 September, 2000, Tyrol, Austria

(oxygen-limited combustion). When the air supply increases, the flame propagation speed increases. However, the flame propagation speed is limited by the reaction rate of the fuel (reaction-limited combustion). When the air supply further increases, excess air cools the bed and puts an end to the flame (extinction by convection). Fatehi and Kaviany (4) identified in a similar way two modes and called them an oxygen-limited and a fuel-limited mode. None of these authors related the modes or zones to a description of bed events. Nevertheless, the oxygen-limited mode should correspond to the sub-stoichiometric regimes, and the reaction- or fuel-limited one to the overstoichiometric regime. CONCLUSIONS Two test rigs have been built for investigation of the combustion processes in a bed of solid fuels and particularly the influence of primary airflow and of particle properties (size, density and thermal conductivity) on the rate and temperature of the ignition front. Also the gas composition downstream of the bed has been investigated. The following conclusions can be drawn from the results: (1) The two test rigs show the same ignition rates, but the temperatures in the ignition front diverge at higher airflow. This divergence may be caused by a difference in design of the secondary air supply, giving a higher heat transfer to the bed from the gas flame in the small rig. The higher temperatures may also be influenced by fluidisation of particles in the front in the small rig, where the air is supplied from beneath. (2) Three combustion regimes were identified. At low airflow a sub-stoichiometric combustion regime with incomplete consumption of oxygen was found. This regime is characterised by a clear division in a devolatilization phase followed by a phase dominated by char combustion. During the devolatilization, partly converted fuel accumulates downstream of the ignition front. Although the combustion is sub-stoichiometric, oxygen is not fully consumed in the bed. At higher airflow there is a sub-stoichiometric combustion regime with a complete consumption of oxygen in the bed. A layer of partly converted fuel forms downstream of the ignition front, but it is not growing in thickness as in Regime 1. The devolatilization phase is followed by a short char combustion phase. When the primary airflow is increased even more, the combustion moves into the over-stoichiometric regime with excess oxygen leaving the fuel bed. The excess air cools the bed. At higher airflow the ignition rate slows down and the front temperature falls. Finally, the combustion is extinguished. In this regime, the final char combustion phase is negligibly short. (3) The two fuels compared show differences with respect to the combustion regimes. The thermally thinner fuel has a wider sub-stoichiometric, full oxygen consumption regime, starting at a lower airflow. Because of the higher devolatilization rate, the thermally thinner fuel has a sub-stoichiometric regime with full oxygen consumption sustained at higher airflow. (4) Three sizes of the same fuel have been compared. At low airflow the ignition rate is higher and front temperature lower with the larger fuel. The resistance towards heat flow into the particles increases with diameter, and more of the heat produced is transported to new layers of fresh fuels and not towards the centre of the fuel particles. The larger particles are thermally thicker, and the devolatilization rate is lower. The sub-stoichiometric regime with complete oxygen consumption starts at

Progress in Thermochemical Biomass Conversion 17-22 September, 2000, Tyrol, Austria

a higher airflow for larger diameter fuel, and the transition from Regime 2 to Regime 3 is extended to higher airflow. The reason for this has to be further investigated. Parameters of interest are devolatilization rate compared to the ignition rate and heat transfer between particles and primary airflow. (5) In modern combustion equipment, the air supply is divided between primary air to the bed and secondary air to the produced gas. The primary air is kept substoichiometric to enhance reduction of NO. The secondary air is important as a measure for mixing during gas phase combustion. This means that the most important regions for the primary air are the low velocity regimes. Knowledge of the influence of fuel characteristics on the regimes is important for the division between primary and secondary air under various load conditions. However, depending on the design (e.g. batch-wise combustion or steady-state combustion on a transporting grate) the build-up of char or not is also an important issue to consider. ACKNOWLEDGEMENT This work has been supported by the Swedish National Energy Administration, which is gratefully acknowledged. REFERENCES 1. La Fontaine, H. & Reed, T. B. (1993) An Inverted Downdraft Wood-Gas Stove and Charcoal Producer, Energy from Biomass and Wastes XV, Washington D.C. 2. Gort, R. (1995) On The Propagation of a Reaction Front in a Packed Bed, Ph D Thesis, Universiteit Twente, Enschede. 3. Shin, D. & Choi, S. (2000) The Combustion of Simulated Waste Particles in a Bed, Combustion and Flame Vol. 121, pp. 167-180. 4. Fatehi, M. & Kaviany, M. (1994) Adiabatic Reverse Combustion in a Packed Bed, Combustion and Flame Vol. 99, pp. 1-17. 5. Kuo, J. T., Hsu, W.-S. & Yo, T.-C. (1997) Effect of Air Distribution on Solid Fuel Bed Combustion, Journal of Energy Resources Technology Vol. 119, pp. 120-128. 6. Horttanainen, M. V. A., Saastamoinen, J. J. & Sarkomaa, P. J. (1999) Ignition Front Propagation in Packed Beds of Wood Particles, The Swedish-Finnish Flame Day, International Flame Research Foundation, Vxj. 7. Thunman, H. & Leckner, B. (2000) Ignition and propagation of a reaction front in cross-current bed combustion of wet biofuels, Fuel (in press). 8. MacLean, J. D. (1941) Thermal Conductivity of Wood, Transaction American Society of Heating and Ventilation Engineers, Vol. 47, pp. 323-354. 9. Howard, J. B., Williams, G. C. & Fine, D. H. (1972) Kinetics of Carbon Monoxide Oxidation in Postflame Gases. Fourteenth Symposium (International) on Combustion, The Combustion Institute, pp. 975-986. 10. Saastamoinen, J. J., Huttunen, M. & Kjldman, L. (1998) Modelling of Pyrolysis and Combustion of Biomass Particles, ECCOMAS 98, John Wiley & Sons, Ltd. pp. 814-819.

Você também pode gostar