Sei sulla pagina 1di 11

Energy & Fuels 2007, 21, 1493-1503 1493

Sensitivity Analysis of a Fixed Bed Combustion Model


Robert Johansson,* Henrik Thunman, and Bo Leckner
Department of Energy and EnVironment, Chalmers UniVersity of Technology,
SE-412 96 Göteborg, Sweden

ReceiVed October 6, 2006. ReVised Manuscript ReceiVed December 28, 2006

The ability to model the combustion of biofuels in a fixed bed is evaluated by a sensitivity analysis. The
analysis treats the uncertainty of model parameters related to heat transport, reaction rates, and composition of
volatiles. The scatter of the parameters is estimated from the differences between several published correlations.
The results are compared with measurement data and possible model simplifications are discussed. It is shown
that the bed model is able to reproduce the ignition rate and the maximum temperature. Prediction of these
properties is relatively insensitive to the uncertainty of most of the parameters. Gas concentrations within the
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

bed are more difficult to predict. They are greatly influenced by the composition of the volatile gas released
Downloaded via KOREA INST ENERGY & RESEARCH on April 23, 2019 at 07:19:41 (UTC).

during devolatilization. Also the composition of the volatile gas has a significant influence on the ignition of
the gas in the model, affecting the ignition rate, particularly at low airflows. Moreover, the investigation shows
that treatment of radiation can be simplified, and the number of gas species included in the model can be
restricted without significant losses of model generality.

Introduction other publications1,2,10 refer to the velocity of the front, but this
Biomass is burned in fixed beds in grate furnaces and property depends on fuel density. The reaction fronts in beds
domestic boilers, common techniques for small- to medium- with different fuel densities have different velocities, while the
scale heat production. Modeling of these beds improves the ignition rates in general are comparable, as seen in Figure 3,
understanding of the combustion process and can support the which shows experiments on both wood pellets and wood chips.
design of furnaces and the investigation of problems related to In a number of papers,11-13 the influence of the devolatilization
the fuel bed. These models also provide boundary conditions rate, air flow, moisture content, particle size, fuel density, bed
at the surface of the bed for the modeling of combustion in the porosity, and heating value on temperature, conversion rate, and
freeboard. gas composition was examined. Other works2,14 looked at the
Several bed models applied to conversion of biofuels can be influence of particle size, heating value, and density on the
found in the literature.1-7 These models have been used to ignition rate. One model15 assumes local thermal equilibrium
investigate bed temperature, reaction rates, and ignition rate between gas and solid, and the combustion was approximated
during the conversion process. The ignition rate, Ignbed, is with a single-step reaction based on char kinetics. The authors
defined as the velocity of the reaction front, ufront, multiplied primarily focused on differences between stoichiometric re-
by the initial density of the bed, Fbed, according to eq 1. The gimes, but they also examined the sensitivity of the ignition
velocity of the front is calculated by dividing the vertical rate to variations of the heat of reaction, pre-exponential factor,
distance between two points in the bed (z2 - z1) by the and activation energy of the combustion reaction. In another
difference in time ∆t when these points reach a temperature of work,7 the combustion of straw was modeled, and a parametric
800 K. study was carried out on the solid conductivity, bed porosity,
heat capacity of both the solid and gas, rate of devolatiliza-
z2 - z 1 tion, straw diameter, heat- and mass-transfer coefficients,
Ignbed ) Fbed ) Fbed ufront (1) pyrolysis yields, tar combustion rate, and mixing of gaseous
∆t
species. The results showed that, within the ranges investigated,
This definition is consistent with that used in earlier works,7,8,9 these parameters do not affect the results more than 10%.
although it was called reaction front rate in one case.7 Some However, the parametric ranges were not related to the span of
* To whom correspondence should be addressed. Phone: + 46 31 772 the data found in the literature, and in general, the ranges were
5249. Fax: +46 31 722 3592. E-mail: robert.johansson@me.chalmers.se. narrow.
(1) Fatehi, M.; Kaviany, M. Int. J. Heat Mass Transfer 1997, 40, 2607- The present work is focused on the evaluation of uncertainties
2620.
(2) Shin, D.; Choi, S. Combust. Flame 2000, 121, 167-180. in the modeling of biofuel combustion in fixed beds with the
(3) Di Blasi, C. Chem. Eng. Sci. 2000, 55, 2931-2944.
(4) Yang, Y. B.; Goh, Y. R.; Zakaria, R.; Nasserzadeh, V.; Swithenbank, (10) Saastamoinen, J. J.; Taipale, R.; Horttanainen, M.; Sarkomaa, P.
J. Waste Manage. 2002, 22, 369-380. Combust. Flame 2000, 123, 214-226.
(5) Thunman, H.; Leckner, B. Fuel 2003, 82, 275-283. (11) Yang, Y. B.; Yamauchi, H.; Nasserzadeh, V.; Swithenbank, J. Fuel
(6) Bruch, C.; Peters, B.; Nussbaumer, T. Fuel 2003, 82, 729-738. 2003, 82, 2205-2221.
(7) Zhou, H.; Jensen, A. D.; Glarborg, P.; Jensen, P. A.; Kavaliauskas, (12) Yang, Y. B.; Sharifi, V. N.; Swithenbank, J. Fuel 2004, 83, 1553-
A. Fuel 2005, 84, 389-403. 1562.
(8) Gort, R. On the propagation of a reaction front in a packed bed. (13) Yang, Y. B.; Ryu, C.; Khor, A.; Yates, N. E.; Sharifi, V. N.;
Ph.D. Thesis, Universiteit Twente, Enschede, The Netherlands, 1995. Swithenbank, J. Fuel 2005, 84, 2116-2130.
(9) Rönnbäck, M.; Axell, M.; Gustavsson, L.; Thunman, H.; Leckner, (14) Thunman, H.; Leckner, B. Proc. Comb. Inst. 2005, 30, 2939.
B. Progress in Thermochemical Biomass Conversion, Bridgewater, A. V.; (15) Fatehi, M.; Kaviany, M. Combust. Flame 1994, 99, 1-17.
Ed; Blackwell Science: Oxford, 2001, 743-757.
10.1021/ef060500z CCC: $37.00 © 2007 American Chemical Society
Published on Web 03/22/2007
1494 Energy & Fuels, Vol. 21, No. 3, 2007 Johansson et al.

goal of finding simplifications and assisting in further model The surface of the bed is a moving boundary that follows the
development. Special emphasis is put on the evolution of gas shrinking bed and exchanges heat with the surroundings by
concentration throughout the fuel bed and on the ability to radiation. The bed is ignited by a radiating flame, whose temperature
reproduce measured concentrations inside the bed. The model is 1400 K with an effective area that is equal to the bed area. Once
deals with the same combustion system as the bed models just the reaction front has reached a few centimeters down into the bed,
the flame is removed, and the surroundings are now assumed to
mentioned and many of the submodels and correlations applied radiate back with a temperature equal to that of the gases flowing
are the same or similar to the ones included in these models. out from the bed but with an upper limit of 1200 K. The gas
To simulate the conversion process, these models have to treat temperature and the gas concentrations at the bed’s surface are
a complex set of equations with a large number of parameters. assumed to have zero gradients, a Neumann condition, and the
The parameters are based on data and correlations found in the pressure is specified to be 1 bar.
literature, but there is some uncertainty in their values; therefore, The grate on which the bed rests exchanges heat by conduction
there is a need to know how they influence the results. An and radiation with the solids and transports heat by radiation and
improved understanding of the impact of these parameters is convection/conduction to the surroundings. The convective part of
important for the quantitative and qualitative evaluation of model this heat is assumed to be absorbed by the entering air, and it is
results. This work treats the sensitivity of the modeling results therefore returned to the bed by the air flow. The velocity and
due to possible variation of the parameters related to effective temperature of the entering air are specified. The initial temperature
of the gas and solids in the bed and the temperature of the inlet air
reaction rates and heat transfer. The uncertainty of the param-
is 300 K.
eters is estimated by comparison of several literature sources.
Procedure and Investigated Parameters. The sensitivity of the
The influence of the parameters on the ignition rate, temperature
parameters included in the model has been investigated for two
of the gas and solids, and gas concentration is evaluated, and inlet air velocities: 0.05 m s-1 in Case I and 0.15 m s-1 in Case II.
the results are compared with measured data. The influence of Both velocities correspond to substoichiometric conditions; all
the composition of the volatile gas is also addressed and possible oxygen is consumed during the propagation of the front in the bed.
model simplifications are discussed. Stoichiometric velocities, for which enough oxygen is supplied to
convert all the ignited fuel, are in the range of 0.2-0.4 m s-1. The
Model velocities investigated are in the lower range of primary air velocity
in grate furnaces.17-18 The limits of the parameters cover the ranges
The case selected for the sensitivity analysis consists of a bed of the correlations found in literature. Only one parameter at a time
of spherical biofuel particles with an initial size of 1 cm and a has been examined. For each parameter, a reference value has been
moisture content of 10% (on a wet basis). Combustion air is fed chosen, which is applied when other parameters are investigated.
from below and ignition takes place at the surface of the bed, which The limits are summarized in Table 1 and graphically shown in
has an initial height of 0.3 m. This creates a conversion situation Appendix 2, together with correlations from literature. In addition,
that can be described in one dimension. After ignition, a reaction the influence of the composition of the volatile gas has been
front propagates down into the bed; the fuel dries and devolatilizes, investigated.
and the char, together with the released volatiles, burn as long as The parameters seen in Table 1 were chosen as follows.
there is oxygen available. The speed and profile of the reaction Mixing Rate of Gaseous Species. Mixing controls the homoge-
front is stable during most of the propagation period, and the neous reaction rates at high temperatures where the kinetics are
conversion mode is suitable for the comparison of models and fast. Three correlations for mixing have been compared. The first
measurements. This case has previously been thoroughly investi- rate19 was derived using the Ergun equation as a starting point.
gated both in experimental work8-10 and in modeling.1-7 The second14 correlation is a modification where the viscous term
The one-dimensional transient model employed in this work has been removed. This correlation is used as a reference. The third
consists of the transport equations for energy and species, together correlation7 is based on diffusion across a characteristic length,
with the continuity equation of the gaseous phase, and regards the chosen as the diameter of a particle. The difference between the
bed as a porous medium. The energy equation is expressed for the three mixing rates is very large. The rates have not been validated
solid and the gaseous phases including heat transfer between them. by experimental data, although the turbulent part in the first
The energy equation of the solid phase accounts for radiation and correlation was compared with CFD simulations. The third expres-
conduction, while convection, dispersion, and diffusion of species sion is highly dependent on the particle size: it was used to model
with different specific heats are treated in the gas phase. Eight straw combustion with the characteristic length taken as the straw
species are included in the gaseous phase: CO, CO2, H2O, H2, CH4, diameter, resulting in a significant rate, despite being small
and CxHyOz in the volatile gas and O2 and N2 in the air. The gas compared with the other expressions. The first expression also
species are transported by convection and dispersion. The density depends highly on particle size and yields the highest rate. The
of the gas is given by the ideal gas law. Drying and devolatilization rates of the third and first correlations have been chosen as the
are modeled with Arrhenius expressions, while both diffusion and lower and upper limits, respectively.
kinetics are considered in the conversion of the char. The EffectiVe Surface Area. The effective surface area is that of the
homogeneous rates of conversion are the minimums of the kinetic solids per unit of bed volume available for heat transfer and char
and mixing rates. These conversion steps are expressed by source conversion. It is important for heat transport between the phases
terms in the transport equations of energy and species. The equations and char combustion. The maximum effective area (the whole
are discretized in space according the finite volume method surface area of the particles per unit of bed volume) serves as
following the hybrid scheme. The time discretization is fully implicit reference value. This value is commonly used in the modeling of
to ensure numerical stability. Care has been taken to select step fixed beds. However, the particles in a packed bed are in close
sizes, making the results independent of the time and space contact and, to some extent, cover each other, and the entire surface
discretization. Shrinkage of the bed caused by conversion of the area may not contribute to the effective area. Therefore, the lower
fuel is accounted for by recalculation of the sizes of the compu-
tational cells at each time step. The equations are summarized in
Appendix 1, and a more detailed description of the bed model can (17) van der Lans, R. P.; Pedersen, L. T.; Jensen, A.; Glarborg, P.; Dam-
Johansen, K. Biomass Bioenergy 2000, 19, 199-208.
be found elsewhere.16 (18) Kuo, J. T.; Hsu, W.-S.; Yo, T.-C. J. Energy Resour. Technol. 1997,
119, 120-128.
(16) Johansson, R.; Thunman, H.; Leckner, B. Influence of intra-particle (19) Yang, Y. B.; Goh, Y. R.; Nasserzadeh, V.; Swithenbank, J. Presented
gradients in modelling of fixed bed combustion.Combust. Flame 2007, in at the 3rd International Symposium on Incinaration and Flue Gas Treatment
press. Technology, Brussels, Belgium, 2001.
Fixed Bed Combustion Model Energy & Fuels, Vol. 21, No. 3, 2007 1495

Table 1. Parameter values for the simulations


parameter ref value min value max value
mixing rate of
gaseous species rmix ) 0.63
(ref 14)
(
1.75ug(1 - )
dp ) []
min
Ci
Ωi
rmix ) ln(mO2,0/mO2,τm)
(ref 7)
2Dab
dp2

(ref 19)
(
rmix ) 0.65 150
Dab(1 - )2/3
dp2
+
dp )
1.75ug(1 - )1/3
min[CiΩi]

effective surface area Ap Ap Ap


Asur ) Asur ) 0.6 Asur )
Vp Vp Vp
bed porosity 0.5 0.4 0.6
axial dispersion 0.5dpu 0.6 × ref 1.2 × ref
Deff,ax ) 0.8Dm +

0.5dpcp,gFgu
kg,eff,ax ) 0.8kg,m +

(ref 23)
solid conductivity 0.2 W/mK for virgin wood 0.6 × ref 1.5 × ref
λc ) (1 - c)(1.47 + Ts1.11 × 10-3)
(ref 28)
extinction coefficient scaled projected surface area 0.7 × ref 1.3 × ref
(ref 26)
heat and mass 0.5 × ref 1.5 × ref
transfer coefficients Nu ) 2 + 1.1Pr1/3Re0.6
Sh ) 2 + 1.1Sc1/3Re0.6
(ref 23)
shrinkage during 60 vol % 40 vol % 70 vol %
devolatilization

limit of this parameter is set to 60% of the entire area, which extinction coefficient are chosen to 70 and 130% of the reference
accounts for significant contact areas of particles. value. The conductivity of the solids is calculated from a mass-
Bed Porosity. The porosity of the bed varies with packing. A weighted average of the conductivities of virgin wood and char.
typical value for a packed bed is around 0.4,20-22 but some data The conductivity of char depends on temperature.28 From an
obtained from a bed of biofuels indicate a slightly higher porosity.9 evaluation of literature data, the parameter range is estimated to
The reference value is chosen to 0.5 with lower and upper limits be 60-150% of the reference conductivity.
of 0.4 and 0.6. Heat- and Mass-Transfer Coefficients. The heat-transfer coef-
Axial Dispersion. Gas flow through a packed bed gives rise to ficient controls the temperature difference between the phases.
dispersion, and the effective diffusion of heat and mass is much Therefore it influences ignition of both the gas and solids. The mass-
higher than the molecular counterpart. Increased heat transport transfer coefficient has a direct effect on the rate of char conversion
enables the ignition of the gas earlier in the front, since the gas in the diffusion-controlled regime. The reference values for the heat-
reaches the ignition temperature sooner. The reference values for
and mass-transfer coefficients are calculated according to correla-
effective heat and mass diffusion of the gas are taken from a
tions and fitted to an extensive amount of data from different
standard correlation23 often applied in the modeling of fixed bed
experiments.23 They have been used in earlier work.1,2,5 The
combustion.4,7,14 These values have been compared with empirical
minimum is half of the reference, and the maximum is 1.5 times
correlations for granular packed beds24 and packed bed reactors.25
The range of variation is found to be 60-120% of the reference the reference value. This range covers almost all experimental data,
value. forming the basis of the correlation.23
Extinction Coefficient and Solid ConductiVity. Both radiative and Shrinkage during DeVolatilization. Shrinkage during devolatil-
conductive heat transport are important in the solid phase. They ization affects char conversion, heat transfer between gas and solids,
affect the temperature of the gas and solids and the propagation of and mixing of gaseous species because these processes depend on
the reaction front. The radiative heat transfer in the bed is modeled the particle size. It also affects heat transfer in the bed: larger
with a two-flux model, an approach used in other bed models.2,4,5 shrinkage reduces the distance between the hot combustion zone
The extinction coefficient is equal to the projected surface area of and the unreacted wood. During devolatilization, a biofuel particle
the particles per unit bed volume, corrected by a scaling factor to shrinks, usually to a size corresponding to about 40% of the initial
account for dependent scattering.26 The influence of radiation has volume, but the shrinkage depends on the temperature and wood
been investigated not only by changes of the extinction coefficient type.29 The shrinkage is varied between 40 and 70 vol %, and the
but also by comparison of the two-flux model with a radiative reference value is 60 vol %.
conductivity model,27 in which a radiative part is added to the Composition of Volatile Gas. The influence of the composition
conductivity of the solids. The lower and upper limits of the of the volatile gas has been investigated by simulations with
compositions taken from related works.3,30 The calculation method31
(20) Zou, R. P.; Yu, A. B. Powder Technol. 1996, 88, 71-79. for the composition in the reference case involves two experimental
(21) Sederman, A. J.; Alexander, P.; Gladden, L. F. Powder Technol.
2001, 117, 255-269. ratios, which depend on the specific surface area of the fuel
(22) Montillet, A.; Le, Coq, L. Powder Technol. 2001, 121, 138- particles. Therefore, the influence of variations in these experimental
148. ratios has also been investigated. No attempt has been made to
(23) Wakao, N.; Kaguei, S. Heat and Mass Transfer in Packed Beds; estimate the upper or lower limits of these concentrations because
Gordon and Breach: New York, 1982.
(24) Guedes, de Carvalho, J. R. F.; Delgado, J. M. P. Q. Chem. Eng.
Sci. 2005, 60, 365-375. (28) Thunman, H.; Leckner, B.; Niklasson, F.; Johnsson, F. Combust.
(25) Jacobsen, H. A.; Lindborg, H.; Handeland, V. Comput. Chem. Eng. Flame 2002, 129, 30-46.
2002, 26, 333-357. (29) Davidsson, K. O.; Pettersson, J. B. C. Fuel 2002, 81, 263-270.
(26) Singh, B. P.; Kaviany, M. Int. J. Heat Mass Transfer 1992, 35, (30) Bryden, K. M.; Ragland, K. W. Energy Fuels 1996, 10, 269-
1397-1405. 275.
(27) Singh, B. P.; Kaviany, M. Int. J. Heat Mass Transfer 1994, 37, (31) Thunman, H.; Niklasson, F.; Johnsson, F.; Leckner, B. Energy Fuels
2579-2583. 2001, 15, 1488-1497.
1496 Energy & Fuels, Vol. 21, No. 3, 2007 Johansson et al.

Figure 1. Layout of the gas measurements in the reaction front.

Figure 3. Comparison of ignition rate in simulations and experi-


ments: ×, Gort,8 wood blocks 10 mm, Xm,0 ) 0.1; O, Rönnbäck et
al.,9 wood cylinders 8 mm; 3, Saastamoinen et al.,10 wood 5-20 mm,
Xm,0 ) 0.11. The lines are as follows: reference case (dashed-dotted),
high mixing rate (solid), and low heat- and mass-transfer coefficients
(dashed). The shaded area shows a change of less than 10% of the
simulated ignition rate in the reference case. The solid straight line
represents the stoichiometric air flow based on ignited fuel.

Figure 2. Simulated mole fractions of gas species at the position of


the probe as a function of time. Parameters are those of reference case
I. The shaded area is shown as a close up in Figure 4.
they depend on experimental conditions, and therefore, the results
are hard to compare.

Results and Discussion


The reaction front descends in the bed promoted by conduc-
Figure 4. Close up of measured32 and simulated gas concentrations at
tive and radiative heat transport in the solid phase, while the the position of the probe during passage of the reaction front. Simulated
fuel is dried and devolatilized. The concentration of gases from data are from reference case I.
the drying and devolatilization increases rapidly when the front
passes at the same time that the concentration of oxygen Increased heat transport in the solid phase enhances the
decreases because of dilution. Figure 1. Figure 2 shows the gas ignition rate and is followed by a small change of the
composition at a position corresponding to the probe tip. As temperature of the solids. When the rate increases, more fuel is
the reaction front progresses downward, the surface of the bed heated per time unit, and since the heat release is more or less
moves to a position below the gas probe; then, the simulated constant for a given air flow, the temperature tends to fall. A
concentrations correspond to the gas leaving the surface of the lower ignition rate, on the other hand, means that less fuel is
bed. heated per unit time, and it is, therefore, followed by a rise in
The char formed during devolatilization reacts with oxygen, temperature. This assumes that the heat release is independent
which results in further heating of the solids and propagation of the ignition rate, which is only a reasonable approximation
of the front. In the downstream part of the front, the gas is colder when all oxygen is consumed in the reaction front. For over-
than the solid phase, and heat transfer between the phases stoichiometric conditions, there is more oxygen available. A
reduces the heating rate of the solids. Once the gas temperature higher ignition rate then means that more fuel is burned, and
is high enough for the released volatiles to ignite, the gas attains the heat release increases. Even for sub-stoichiometric condi-
a higher temperature than the solids. Oxygen is consumed by tions, a change of the ignition rate can affect the heat release if
the combustion of char and volatiles, and the concentration of the fraction of oxygen needed for the gas-phase reactions
the combustion products rises. If the air flow is low, there is changes compared to that consumed by the combustion of the
not enough oxygen to complete the char conversion during solids. This is caused by the difference in heat release per mole
propagation of the front, and a char layer builds up. This char of oxygen between heterogeneous and homogeneous combustion
is to some extent gasified by H2O and CO2. At 2200 s when reactions. The effect is, however, of minor importance.
the front reaches the grate, devolatilization is finished, and the The model predictions of ignition rate and maximum tem-
concentration of volatile gases approaches zero, Figure 2. During perature agree well with experimental data,8-10 Figure 3. Figure
the first period of the final char burnout, around 2600 s, there 4 shows the corresponding agreement between simulated and
is a peak in the concentration of CO. In the second half of this measured32 gas concentrations in the reaction front. The
conversion step, combustion is slower, and the concentration comparison of the simulated and measured gas species is not
of oxygen increases again, while the fractions of CO and CO2 straightforward because the local experimental air velocity at
decrease. the centerline of the bed, where the probe is positioned, differs
Fixed Bed Combustion Model Energy & Fuels, Vol. 21, No. 3, 2007 1497

from the cross-sectional average air velocity because of the area. The effect of mixing and heat- and mass-transfer on
higher velocities close to the reactor walls. The model assumes ignition rate can also be seen in Figure 3.
that the air flow is evenly distributed, that is, that the velocity The composition of the volatile gas has an important effect
is uniform over the cross-section. The simulated data are from on the ignition rate, as will be discussed below, and it
reference case I, and the air velocity of this case corresponds significantly influences the concentration of gas species. Simu-
to the experimental air velocity at the position of the probe. lation with one of the alternative compositions30 results in small
Since the average measured air flow is higher than the local differences in the concentrations of H2O, CO, and CO2 and large
velocity at the probe, the simulated ignition rate is 28% lower differences, 20-70%, in the concentration of H2, CH4, and
than the measured one. Therefore, the stoichiometries differ. CxHyOz, whereas the second alternative composition3 gives large
The measured concentration profiles have been shifted along differences, >20%, in all gas species. Variation of the empirical
the time axis to make the initial reduction in the oxygen profile ratios in the method31 for prediction of the composition in the
coincide with the simulated initial decrease to remove differ- reference cases changes the concentration of H2O, CO, and CO2
ences caused by the variation in ignition rates and initial bed up to 15%, and the fractions of the other gas species are affected
heights. This is justified because the front propagates with a even more.
constant speed throughout most of the bed, and the temperature The mixing rate influences the homogeneous reactions and
and concentration profiles in the front remain stable during this the heat release in the gaseous phase: lower mixing rates give
period. The purpose is to compare differences in the concentra- lower heat release, and the gas attains a lower temperature and
tion profiles inside the front and not differences in the ignition less-steep temperature profile than it does at higher mixing rates.
rate. The increase in the concentration of water vapor in the The temperature profiles for the two extreme values of mixing
initial stage, followed by a release of volatile gases, is well rates in Table 1 are illustrated in Figure 6. The slower heating
reproduced by the model. The sharp increase of CO2 compared at the lower mixing rate increases the cooling or, for some
to CO, seen in the measured profiles, is also evident in the conditions, reduces the heating of the solids and thereby
simulated profiles. The trends of slightly increasing fractions decreases the ignition rate. The effect is more pronounced in
of H2 and CO, together with a decreasing fraction of H2O during the low-velocity case, Case I, where the upper and lower limits
the propagation of the reaction front and a peak in the CO of the mixing rate give significantly different ignition rates, as
concentration during the final char burnout, are seen in both shown in Figure 5a-b. The low mixing rate reduces the rates
the experiments and simulations. Furthermore, the measured of the gas reaction. Then, they consume less oxygen, which
decrease of the water vapor concentration upstream of the becomes available for char combustion, and the char is more
reaction front, predominantly resulting from gasification reac- completely converted. In Case II, this effect means that only a
tions and the water-gas shift reaction, is also captured by the thin char layer builds up, resulting in a short residence time of
model. It should be pointed out that, when the entire conversion the gas inside the bed. The gas, therefore, does not have enough
process is examined, the measured fraction of water vapor time to reach its peak temperature before it leaves the surface
continues to decrease significantly during the period of front of the bed.
propagation while the simulated fraction remains more
A higher mixing rate leads to a higher ignition rate and more
constant. The conversion of the lower hydrocarbons, mainly
fuel-rich gas. When the upper limit of the mixing rate is used,
CH4, is not well captured by a model that predicts a much faster
the stoichiometry approaches the experimental one. This results
increase of the concentration than what is seen from the
in the same trends in the simulated gas concentrations as in the
measurements.
reference case, but the fraction of water vapor is significantly
Figure 5 shows the relative difference in the ignition rates
higher and the fraction of CO is somewhat lower, as can be
and peak temperatures of the gases and solids between the
seen from a comparison of Figures 7 and 4.
reference cases and the cases with the upper or lower limits of
the parameters. The positions on the x axis where the lines reach Heat and mass transfer between the gas and solid phases
a relative horizontal value of unity, corresponding to the affects the ignition rate in two ways: lower heat transfer reduces
reference case, are calculated from the relation between refer- the cooling of the gas, and lower mass transfer reduces the rate
ence values and limits of the parameters (p) of char combustion. For the conditions investigated, the influ-
ence of the mass transfer dominates, and the lower heat- and
pref - plow mass-transfer coefficients markedly reduce the ignition rate; in
xref ) xlow + (xhigh - xlow) (2) case II, the ignition rate differs from that of the reference case
phigh - plow by more than 10%, Figure 5b. The highest values tested do not
change the ignition rate as much, and in all cases, the effect on
The rate of mixing of gaseous species is estimated because the the maximum temperatures is limited. The temperature differ-
rate changes during conversion. The correlation used for the ence between the phases is also influenced, as shown in Figure 8.
lower limit of mixing does not depend on gas velocity, while The sensitivity of the heat transport mechanisms in the solid
the other two correlations do. Therefore the x value correspond- phase, conduction and radiation, is limited, but some comments
ing to the reference case differs between cases I and II. As should be made about an alternative radiation model. In the
shown in Figure 5, the influence of most parameters is less than reference case, radiation was modeled by a two-flux model
10% (shaded area), which is a small range compared to the (Appendix I) but a radiative conductivity model was also tested
scatter in measurement data in Figure 3. This is also the case for these conditions. This model requires less computational
for the influence on the gas concentration. Only the rate of time and, for this reason, is a more efficient option. The radiative
mixing of the gaseous species and the heat- and mass-transfer conductivity model yields a somewhat lower ignition rate
coefficients affect the ignition rate and the temperatures more, (<10%) and a somewhat higher maximum temperature of the
while some of the gas species are also affected by the effective solids because of lower heat transport. The reason is that the
two-flux model allows radiative heat transport across larger
(32) Rönnbäck, M.; Samuelsson, J.; Tullin, C.; Thunman, H.; Leckner,
B. Presented at the Science in Thermal and Chemical Biomass Conversion distances than the conductivity model, where the transport only
Convention, Victoria, BC, Canada, 2004. takes place between neighboring computational cells. The steep
1498 Energy & Fuels, Vol. 21, No. 3, 2007 Johansson et al.

Figure 5. Sensitivity of parameters compared to the reference cases. Thick lines: mixing (solid), heat/mass transfer (dashed), effective area (dotted),
and conductivity (dashed-dotted), Thin lines: radiation (solid), porosity (dashed), axial dispersion (dotted), andshrinkage (dashed-dotted). Low and
high corresponds to the investigated parameter range. The shaded area represents a change of less than 10%.

gradients in the bed require several computational cells along a The choice of the kinetic rate of hydrogen combustion is
distance equal to one particle diameter. In comparison, the motivated by the low ignition temperature of volatiles from
typical gas distance between the particles is the same size as biomass, an ignition, which, according to an experimental
the particle diameter. Radiative heat transport reaches over these work,33 is believed to be initiated by heterogeneous reactions
distances, and this is better captured by the two-flux model; on the solid surface. This means that the hydrogen starts to react
however, the effects of using the less-demanding conductivity and contributes to the heating of the gaseous phase already when
model are, as already pointed out, small. leaving the fuel particles, while other gas reactions require higher
The effective area of the solids has some effects on the gas temperature. When the gas temperature increases further up in
concentration: a smaller area reduces the rate of char combus- the front, other homogeneous reactions become fast, heating
tion and gasification. This leads to lower concentrations of the the gas further. If there is a lack of hydrogen in the volatile
gasification products CO and H2 at the same time that more gas, the initial homogeneous heating of the gas is less significant,
oxygen is available for the combustion of volatiles, and the and the heating process is delayed. A higher fraction of hydrogen
concentrations of CxHyOz and CH4 also become lower. in the volatile gas gives a higher heat release in a forward
The ignition properties of the volatile gas leaving the fuel position of the front and results in a higher ignition rate. This
particles have a significant effect on the ignition rate. In the picture was supported by the small impact on the ignition rate
model, the gas is ignited by hydrogen in the volatiles. The kinetic
rate used for hydrogen combustion is high even at low (33) Grotkjær, T.; Dam-Johansen, K.; Jensen, A. D.; Glarborg, P. Fuel
temperatures, and the effective reaction rate is limited by mixing. 2003, 82, 825-833.
Fixed Bed Combustion Model Energy & Fuels, Vol. 21, No. 3, 2007 1499

ignition rate by 20% in Case I and 14% in Case II. A sensitivity


test was made with lower reaction rate of hydrogen.34 With this
rate, hydrogen combustion becomes important only at a tem-
perature where the other volatile gases also contribute to
combustion, and the impact of the hydrogen fraction on the
ignition rate almost disappears.
The significant difference between the simulated and mea-
sured profiles of CH4 can be related to two factors in the
model: the low combustion rate and the fraction of CH4 in the
volatile gas. The kinetic rate is the lowest of the homogeneous
combustion rates employed, and therefore, the conversion of
the other combustible species in the volatile gas, CxHyOz, CO,
and H2, consumes the oxygen in the kinetically controlled
region. If the kinetic rate of CxHyOz is applied for CH4 as well,
the concentration of CH4 decreases, but it remains higher than
the experimental concentration. The fraction of CH4 in the
volatile gas is higher than the fraction of CxHyOz, although the
experiments indicate the opposite if both species have similar
combustion rates. The composition used in another work3 has
Figure 6. Temperature profiles in the reaction front. (a) Cases I and a low fraction of CH4 and a high fraction of CxHyOz compared
(b) II for different mixing rates (see Table 1) of gaseous species in the
homogeneous reactions. Solid lines represent the temperature of the
to the reference composition. The use of this composition gives
solids, and dashed lines respresent the temperature of the gas. profiles of CH4 and CxHyOz that are closer to the experimental
ones. It is, however, hard to exactly point to the cause for the
discrepancies between the model and measurements. The ability
of the model in the prediction of these profiles is limited.
Because of the difficulties in the quantitative prediction of
gas concentrations and the significant influence of the uncer-
tainty of parameters on these concentrations, the number of
species in the present model formulation could be questioned.
The major combustion species, CO, CO2, and H2O, are
important for the gasification reactions and the water-gas shift
equilibrium, but the purpose of including the combustible species
CH4 and CxHyOz is mainly to simulate the heat release in the
gas phase, which is important for both the ignition rate and
temperatures. It should be possible to treat them as one species
and still maintain their role as a heat source. The concentration
of hydrogen is crucial because it is a source of ignition in this
Figure 7. Comparison of measured32 and simulated gas concentrations model. The ignition of the volatile gas needs to be studied
at the position of the probe during passage of the reaction front. further. Hydrogen is also a part of the water-gas shift
Simulated data are from Case I with homogeneous mixing according equilibrium.
to Yang.
There is, as apparent from literature data, significant uncer-
tainty in the composition of the volatile gas. The model used
to predict the composition31 is based on conservation of species
and energy, together with two experimental parameters. This

(37) Kanury, A. M. Introduction to Combustion Phenomena; Gordon


and Breach Science Publishers: New York, 1977.
(38) Howard, J. B.; Williams, G. C.; Fine, D. H. Proceedings of the
14th Symposium (International) on Combustion; The Pennsylvania State
University, State College, PA; The Combustion Institute: Pittsburgh, PA,
1973; pp 975-986.
(39) Dryer, F. L.; Glassman, I. Proceedings of the 14th Symposium
Figure 8. Temperature profiles in the reaction front in Case I for (International) on Combustion; The Pennsylvania State University, State
different heat and mass transfer coefficients. Solid lines represent the College, PA; The Combustion Institute: Pittsburgh, PA, 1973; 987-1003.
temperature of the solids, and dashed lines respresent the temperature (40) Smoot, L. D.; Smith, P. J. Coal Combustion and Gasification;
of the gas. Plenum Press: New York, 1985.
(41) Biba, M.; Macák, J.; Klose, E.; Malecha, J. Ind. Eng. Chem. Process
Des. DeV. 1978, 17, 92-98.
by changes in the fraction of other gas species, while the (42) Dullien, F. A. L. Porous Media Fluid Transport and Pore Structure;
hydrogen fraction was kept constant. A change of the hydrogen Academic Press Inc.: San Diego, CA, 1992.
fraction in the volatile gas from the reference value of around (43) TenWolde, A.; McNatt, J. D.; Krahn, L. Thermal Properties of Wood
and Panel Products for Use in Buildings; Report No. DOE/USDA-21697/
0.003-0.015, as in one of the other models,3 increases the 1; Oak Ridge National Library: Oak Ridge, TN, 1988.
(44) Koch, P. Wood Sci. 1969, 1, 203-214.
(34) Jones, W. P.; Lindstedt, R. P. Combust. Flame 1988, 73, 233-249. (45) MacLean, J. D. Trans. Am. Soc. Heat. Ventilating Eng. 1941, 47,
(35) Modest, M. F. RadiatiVe Heat Transfer; Academic Press: San Diego, 323-354.
CA, 2003; pp 456-458. (46) Thunman, H.; Leckner, B. Biomass Bioenergy 2002, 23, 47-54.
(36) Chan, W. C. R.; Kelbon, M.; Krieger, B. B. Fuel 1985, 64, 1505- (47) Bird, R. B.; Stewart, W. E.; Lightfoot, E. N. Transport Phenomena;
1513. Wiley: New York, 2002; p 441
1500 Energy & Fuels, Vol. 21, No. 3, 2007 Johansson et al.

seems to be a reasonable approach, but the experimental data therefore strongly depends on this fraction. For all parameters,
should be taken from a relevant application. except the mixing rate of gaseous species and the heat- and
mass-transfer coefficients, the maximum temperatures of gas
Conclusions and solids are affected by less than 10%. The two-flux radiation
model applied can be substituted with a less-demanding
The influence of a number of parameters on a one- conductivity radiation model without causing more than a 10%
dimensional model describing a bed of biofuel that is ignited change in the ignition rate or temperatures.
at the surface, while air is fed from below, has been investigated. The trends in most of the gas concentration profiles agree
The parameters examined are related to heat transfer and reaction well with the experimental data, but quantitative differences are
rates, as well as to the composition of the volatile gas. The hard to evaluate because of the unevenly distributed gas flow
uncertainty of the parameter values has been estimated by found in the experiments. In most cases, the sensitivity of the
comparison of several literature sources. The results have been gas concentration to the investigated parameters is less than
compared to scatter in measured data, and model simplifications 10%. The mixing rate, heat- and mass-transfer coefficients, and
have been discussed. effective area are more important parameters, with up to a 25%
Comparison of the model with measurements shows that both change for some species. The gas concentrations are more
the ignition rate and temperature are well predicted. Within the sensitive to the composition of the volatile gas, which varies
parameter ranges investigated, it is only the mixing rate of widely in published experiments. The significant uncertainty
gaseous species and the heat- and mass-transfer coefficients that in the prediction of the gas concentrations does not motivate
give more than a 10% change of the ignition rate. The onset of too much detail in the submodels describing the gas species.
ignition of the volatile gas has proven to be important for the
ignition rate, especially at low air velocity. In this process, the Acknowledgment. The work was financed by the Nordic
fraction of hydrogen in the volatiles plays a key role in the model Energy Research Program, CECOST, and the Swedish Energy
because of its high kinetic combustion rate. The ignition rate Agency.

Appendix 1

Equations A1-A3 are the transport equations for the energy in the gaseous and solid phases, together with the equation of the
gaseous species. The radiation exchange between solid particles is modeled with a two-flux model according to eqs A10-A12, and
relations A13 and A14 are the rates of drying and devolatilization. Relations A15-A18 describe char conversion. Equations A19
and A20 are homogenous reactions. The mixing reaction, A16, is only used for combustion reactions rg,1-4 because rg,5-6 describes
the water-gas shift equilibrium, where the reactants are assumed to be well mixed. The kinetic constants for the backward reaction
rg,6 have been derived from the equilibrium constant.37 Equation A21 is Ergun’s equation for rough particles, expressing the pressure
drop over the bed, and the internal surface area and particle volume are calculated with eqs A22 and A23. Equations A24 and A25
are relations for the heat- and mass-transfer coefficients, and eqs A26-A29 describe physical properties.

Table A1. Computational Details


equation ref

(A1)
∂Fgcp,gTg
∂t
+
∂ugFgcp,gTg
∂z
)
∂z

(
λg,eff
∂Tg
∂z ) + AsurhT (Ts - Tg) +
(∑ (
i

∂z
DeffMicp,iTgCtot
∂Xi
∂z
)) + S Tg

(A2)
∂Fscp,s(1 - )Ts ∂usFscp,sTs ∂
∂t
+
∂z
)
∂z
(λs(1 - ))
∂Ts
∂z (
- AsurhT (Ts - Tg) + STs + Srad )
(A3)
∂Fi ∂ugFi ∂
∂t
+
∂z
)
∂z
DeffFg
∂Yi
∂z (
+ SFi )
(A4) S Tg ) ( ∑ ∆h r
j
j g,j) +( ∑∑ΩMr
l k
k k h,lAsurcp,kTk) + Sdryingcp,H2OTs + ( ∑S
n
devYdev,ncp,nTs)

(A5) STs ) ( ∑ ∆h r
l
l h,lAsur) + Sdrying∆Hvap + Sdev∆Hdev

Ydev,iSdev
(A6) S Fi ) ( ∑ Ω r
j
i,j g,j) +
Mi
+( ∑Ω r l
i,l h,lAsur)

Sdrying
If i ) H2O, the term must be added.
MH2O

Vp,0
(A7) Sdrying ) (1 - ) r
Vp drying
Vp,0
(A8) Sdev ) (1 - )
Vp
∑r dev,i
Fixed Bed Combustion Model Energy & Fuels, Vol. 21, No. 3, 2007 1501

Table A1 (Continued)
equation ref

(A9) Sh ) ∑ (r
l
h,lAsurMC)

∂ + 35
(A10) Srad ) π (J - J- )
∂z
1 ∂J+ ω
) (1 - ω)Jb - J+ + (J+ + J-)
2 ∂κ 2
1 ∂J- ω
) -(1 - ω)Jb + J- - (J+ + J-)
2 ∂κ 2
Jb ) σTs4/π

∫ Kdz
z
(A11) κ)
0

K ) φ(1 - )3/(2dp)
26
(A12) φ ) 1 + 1.84(1 - ) - 3.15(1 - )2 + 7.2(1 - )3

(A13) rdrying ) YmFs,01.6 × 1027 exp(-25 000/Ts)


36
(A14) rdev,1 ) YvFs,dry1.3 × 108 exp(-16 875/Ts)

rdev,2 ) YvFs,dry2.2 × 108 exp(-16 009/Ts)

rdev,3 ) YvFs,dry1.1 × 107 exp(-14 602/Ts)

CO2
(A15) ΩcombC + O2 f 2(Ωcomb - 1)CO + (2 - Ωcomb)CO2 rh,1 ) (1 - R)Ωcomb
1/kc1 + 1/hm
CCO2
C + CO2 f 2CO rh,2 )
1/kc2 + 1/hm
CH2O
C + H2O f CO + H2 rh,3 )
1/kc3 + 1/hm
C H2
C + 2H2 f CH4 rh,4 )
1/kc4 + 1/hm

2(1 + 4.3 exp[-3390/Ts])


(A16) Ωcomb ) 28
2 + 4.3 exp[-3390/Ts]

SdevYdev,H2
(A17) R1 )
MO2AsurhmCO2

if R1 > 1 f R ) 1
if R1 < 1 f R ) R1
kc1 ) 1.715 Tgexp(-9000/Tg) 28

(A18) kc2 ) kc3 ) 3.42Tg exp(-15 000/Tg) 28

kc4 ) 3.42 × 10-3Tg exp(-15 600/Tg) 28


(A19) rg,i ) min[rkin,i rmix,i] for combustion reactions 14

rmix,i ) 0.63 ( dp,h ) []


1.75ug(1 - )
min
Ci
Ωi

(A20) CO + 1/2O2 f CO2 rkin,1 ) 1.3 × 108CO20.5CH2O0.5CCO exp(-15 100/Tg) 38

C1.16H4 + 1.58O2 f 1.16CO + 2H2O rkin,2 ) 1.585 × 1010CO20.8CCH40.7 exp(-24 392/Tg) 39

C6H6.2O0.2 + 2.8O2 f 3CO + 3.1H2 rkin,3 ) 20 700Tg CO2CCHO 0.3 0.5


exp(-9650/Tg) 40

H2 + /2O2 f H2O rkin,4 ) 10 CH2CO2 exp(-5050/Tg)


1 11
3
CO + H2O f H2 + CO2 rkin,5 ) 2.78CH2OCCO exp(-1510/Tg) 41
H2 + CO2 f CO + H2O rkin,6 ) 93.69CCO2CH2 exp(-5604/Tg) 37

∆P (1 - )2µgug (1 - )Fgug2
(A21) ) 180 +4 42
∆z 3
d 2
3d
p p
1502 Energy & Fuels, Vol. 21, No. 3, 2007 Johansson et al.

Table A1 (Continued)
equation ref
Ap(1 - ) 6(1 - )
(A22) Asur ) )
Vp dp

(A23) Vp ) Vp,0 - Vp,0θdrying


Ym,0
Ym,θ
1-(Ym
Ym,0 )
- Vp,0θdev 1-
Yv
Yv,0 ( )
- Vp,0θcomb 1-
Yc
Yc,0 ( )
Vp,ash Ym,0
θcomb ) 1 - - θdrying - θdev
Vp,0 Ym,θ
Vp,0Fs,0Yash
( Ym,0
)Yash,0

[ ( )]
Vp,ash ) ) Vp,0 1 - θdrying - θdev
Ym Ym,θ Yc,0
Vp,0Fs,0Yc,θ/ Vp,0 1 - θdrying - θdev
Ym,θ

λg
(A24) hT,0 ) (2 + 1.1Pr1/3Re0.6) 23
dp
(Sdev + Sdrying)cp,g/Asur
hT )
exp[(Sdev + Sdrying)cp,g/Asur/hT,0] - 1

Dab
(A25) hm,0 ) (2 + 1.1Sc1/3Re0.6) 23
dp
(Sdev + Sdrying)/Asur/Fg
hm )
exp[(Sdev + Sdrying)/Asur/Fg/hm,0] - 1

(A26) λg,eff ) 0.8λg +0.5dpcp,gFgug/ 23

Deff ) 0.8DAB + ugdp/ 23

(A27) DAB ) DAB,ref ( )


Tg
298
1.75
37

′′′ + cp,cmc′′′)
(cp,moistmdry
(A28) cp,s )
mdry′′′ + mc′′′
cp,sdry + 4190Ym/(1 - Ym)
cp,moist ) +A
1 + Ym/(1 - Ym)

(23.55Ts - 1320Ym/(1 - Ym) - 6191)Ym 43


A)
(1 - Ym)
where cp,sdry ) 4.206T - 37.7 44

′′′ + λcmc′′′)
(λdrmdry
(A29) λs )
′′′ + mc′′′
mdry
λc ) (1 - )(1.47 + Ts1.11 × 10-3) 28

Appendix 2
This section provides a comparison of correlations from different literature sources. Parameter ranges are indicated by the shaded
areas in Figures A1-A5. Figure A1 shows the axial dispersion. The solid conductivities for char and virgin wood are shown by
Figures A2 and A3, respectively. The extinction and heat-transfer coefficients are shown in Figures A4 and A5, respectively.

Figure A2. Solid conductivity for char: Thunman et al., ref 28, solid
Figure A1. Axial dispersion: Wakao, ref 23, solid line; Guedes et lines; Chan + radiative contribution, ref 36, thick lines; MacLean, ref
al., ref 24, dashed line; Skaare et al., ref 25, dashed dotted line. 45, thin lines.
Fixed Bed Combustion Model Energy & Fuels, Vol. 21, No. 3, 2007 1503

J+, J- ) radiation intensity in positive and negative direction, W


m-2
K ) extinction coefficient, [-]
M ) molar mass, kg mol-1
S ) source term, unit dependent on equation
R ) the gas constant, J mol-1 K-1
T ) temperature, K
V ) volume, m3
X ) molar fraction, [-]
Y ) mass fraction, [-]
cp ) specific heat, J kg-1 K-1
dp ) diameter of particle, m
hm ) mass transfer coefficient, m s-1
hT ) heat transfer coefficient, W m-2 K-1
kc ) reaction rate constant, char conversion, m s-1
Figure A3. Solid conductivity for virgin wood: reference 0.2 W/mK, m ) mass, kg
solid line; Thunman and Leckner, ref 46, dashed line; Maclean, ref p ) parameter value, unit depends on parameter
45, dotted line. r ) reaction rate, unit depends on equation
t ) time, s
u ) superficial velocity, m s-1
ufront ) reaction front velocity, m s-1
x ) position on vertical axis
z ) bed height, m
Ω ) stoichiometric coefficient, [-]
R ) fraction oxygen consumed by hydrogen combustion, [-]
 ) porosity, [-]
θ ) shrinkage factor, [-]
κ ) optical thickness, [-]
λ ) conductivity, W m-1 K-1
µg ) dynamic viscosity gas, Pa s
F ) density, kg m-3
σ ) Stefan-Boltzmann constant, W m-2 K-4
φ ) dependent scattering scaling factor, [-]
Figure A4. Extinction coefficent: Singh and Kaviany, ref 26, solid ω ) scattering albedo, [-]
line; Chin and Choi, ref 2, dashed line. ∆h ) heat of reaction, J mol-1
∆H ) heat of conversion, J kg-1
Subscripts
0 ) initial or inlet
ash ) ash, dry wood basis
bed ) property of the bed
c ) char
comb ) char combustion
dev ) devolatilization
dry ) dry wood
eff ) effective
g ) gas
h ) heterogeneous reaction
j ) homogeneous reaction j
k ) gas species k in heterogeneous reaction
l ) heterogeneous reaction l
Figure A5. Heat transfer coefficient: Wakao, ref 23, solid line; Bird, n ) gas species n in devolatilisation
ref 47, dashed line. m ) moisture, wet wood basis
p ) particle
Nomenclature ref ) reference, 298 K
s ) solid
Ap ) surface area of a particle, m2 sur ) surface
Asur ) internal surface area, m2 m-3 tot ) total
C ) molar concentration of gas species, mol m-3 v ) volatiles, dry wood basis
DAB ) binary diffusion coefficient, m2 s-1 vap ) evaporation
Deff ) effective diffusion coefficient, m2 s-1
Ignbed ) ignition rate, kg m-2 s-1 EF060500Z

Potrebbero piacerti anche