Sei sulla pagina 1di 24

This article was downloaded by: [University of Tasmania]

On: 30 August 2014, At: 20:28


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Combustion Theory and Modelling


Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/tctm20

CFD simulation of wood chip


combustion on a grate using an
EulerEuler approach
a

D. Kurz , U. Schnell & G. Scheffknecht

Institute of Combustion and Power Plant Technology (IFK),


University of Stuttgart , Pfaffenwaldring 23, 70569 , Stuttgart ,
Germany
Published online: 23 Sep 2011.

To cite this article: D. Kurz , U. Schnell & G. Scheffknecht (2012) CFD simulation of wood chip
combustion on a grate using an EulerEuler approach, Combustion Theory and Modelling, 16:2,
251-273, DOI: 10.1080/13647830.2011.610903
To link to this article: http://dx.doi.org/10.1080/13647830.2011.610903

PLEASE SCROLL DOWN FOR ARTICLE


Taylor & Francis makes every effort to ensure the accuracy of all the information (the
Content) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or
arising out of the use of the Content.
This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/termsand-conditions

Combustion Theory and Modelling


Vol. 16, No. 2, 2012, 251273

CFD simulation of wood chip combustion on a grate using an


EulerEuler approach
D. Kurz , U. Schnell and G. Scheffknecht
Institute of Combustion and Power Plant Technology (IFK), University of Stuttgart, Pfaffenwaldring
23, 70569 Stuttgart, Germany

Downloaded by [University of Tasmania] at 20:28 30 August 2014

(Received 30 May 2011; final version received 27 July 2011)


Due to the increase of computational power, it is nowadays common practice to use
CFD calculations for various kinds of firing systems in order to understand the internal
physical phenomena and to optimise the overall process. Within the last years, biomass
combustion for energy purposes has gained rising popularity. On an industrial scale,
mainly grate firing systems are used for this purpose. Generally, such systems consist
of a dense-packed fuel bed on the grate and the freeboard region above, where in the
field of numerical modelling, it is common practice to use different sub-models for both
zones. To avoid this, the objective of this paper is the presentation of a numerical model
including a detailed three-dimensional description of the fuel bed and the freeboard
region within the same CFD code. Because of the implementation as an Eulerian
multiphase model, both zones are fully coupled in terms of flow and heat transfer, and
appropriate models for the treatment of turbulence, radiation, and global reactions are
presented. The model results are validated against detailed measurements of temperature
and gaseous species close to the bed surface and within the radiative section of a 240
kW grate firing test facility.
Keywords: CFD; EulerEuler; grate firing; wood combustion; biomass

Nomenclature
Latin symbols
av
AMag
cp
dp
Di
Dp
f
g
h
hevap
hb
H

Specific contact area between gas and particle phase, 1/m


Magnussen coefficient
Specific heat capacity, J/(kgK)
Particle diameter, m
Diffusion coefficient of gas species, m2 /s
Particle-mixing coefficient, m2 /s
Mechanism parameter of carbon oxidation
Gravitation constant, 9.81 m/s2
Specific enthalpy, J/kg
Evaporation enthalpy, J/kg
Bond enthalpy, J/kg
Enthalpy flow, J/s

Corresponding author. Email: kurz@ifk.uni-stuttgart.de

ISSN: 1364-7830 print / 1741-3559 online



C 2012 Taylor & Francis
http://dx.doi.org/10.1080/13647830.2011.610903
http://www.tandfonline.com

Downloaded by [University of Tasmania] at 20:28 30 August 2014

252
I
Ib
k
ka
lw
m
np
Nu
p
r
r
R
Pr
Re
Rep
s
s
T
u
Vw
Y

D. Kurz et al.
Radiation intensity, W/(m2 sr)
Black body intensity, W/(m2 sr)
Turbulent kinetic energy, m2 /s2
Absorption coefficient, 1/m
Wake length, m
Mass, kg
Number of particles per unit volume
Nusselt number
Pressure, N/m2
Rate coefficient, var.
Reaction rate, kg/s
Universal gas constant, 8314.47 J/(kmolK)
Prandtl number
Reynolds number
Reynolds number of flow around particles
Direction vector, m
Source term of the transport variable , var.
Temperature, K
Velocity, m/s
Wake volume, m3
Mass fraction, kg/kg

Greek symbols
Heat transfer coefficient between gas and particle phase, W/(m2 K)
gp

Dissipation rate of turbulent energy, m2 /s3

Volume fraction, m3 /m3


gp
Specific heat transfer coefficient between gas and particle phase, W/(m3 K)

Thermal conductivity, W/(mK)


Effective dynamic viscosity of the gas phase, kg/(ms)
g
Laminar dynamic viscosity of the gas phase, kg/(ms)
g,l
Turbulent dynamic viscosity of the gas phase, kg/(ms)
g,t

Stoichiometric coefficient, kg/kg

Physical density, kg/m3

Turbulent Schmidt/Prandtl number for


Shear stress tensor, N/m2
ij

Subscripts
a
bl
c
dw
F
g
mw
p
Prod

Ash
Boundary layer
Char
Dried wood
Fuel
Gas phase
Moist wood
Particle phase
Products

Combustion Theory and Modelling

253

Superscripts

Downloaded by [University of Tasmania] at 20:28 30 August 2014

Effective value, considering the local volume fraction

1. Introduction
The generation of heat and electricity by utilisation of biomass is becoming a major
alternative to fossil fuel combustion due to the possibility of CO2 neutral energy production.
Grate firing systems are the most common way for the combustion of municipal solid waste
or different kinds of biomass in large-scale applications. The fuel is transported through the
furnace by a moving grate and primary air is supplied from underneath. The solid material
is partially combusted, producing mainly H2 O, CO2 and unburned species like CO and
different hydrocarbons in the adjacent freeboard region. Burnout is assured by additional
secondary air in the over-bed region. The remaining solid products are bottom ash from the
grate and fly ash from the flue gas filters.
The numerical description of grate firing systems, either for municipal solid waste
or different kinds of biomass, has gained some interest within the last years. Basically
completely different modelling approaches are used, as indicated by the literature survey in
[1], whereas it is most common to separate the bed region from the freeboard modelling.
The model of the bed region therefore acts as a preprocessor, supplying inlet conditions
(e.g. velocity, temperature and/or gas species concentration along the surface of the bed)
for the actual CFD simulation of the freeboard region.
The simplest way to describe the bed region is to apply experience or measurement based
correlations, partially linked with energy and mass balances, to determine combustion rates
and/or gas species release as a function of the position on the grate (e.g. [29]).
Another approach for the treatment of the fuel bed is a cascade of perfectly stirred
reactors to describe the fuel conversion and gas species evolution in different zones of the
grate (e.g. [1014]). In most of these approaches, one reactor per primary air zone is used.
The approaches mentioned above neglect the influence of flow phenomena and locally
varying properties like temperature and concentrations within the packed bed, whereas
several authors perform more detailed hydrodynamic simulations for the bed region as well
to account for these effects. To describe the movement and especially the heterogeneous
reactions within the fuel bed, either a Lagrangian representation with discrete particles is
considered (e.g. [1518]), or the description of the solid fuel by means of an additional,
continuous Eulerian phase (e.g. [1924]) is applied.
Even for models with the same fundamental approach, huge differences exist in the
precision and even in the inclusion or neglect of specific physical phenomena. Despite
all these differences, the common core of most of the cited models is the disadvantage
of independent bed and freeboard region in terms of flow, turbulence and heat transfer,
due to the separation into two sub-models. As a consequence, these models have to apply
further simplifying assumptions, for instance the temperature, velocity and/or turbulence
profile on the bed surface or even the shape of the bed. Due to this partially empirical
character, the resulting overall model is not generally applicable to different grate firing
systems. Additionally, those models are rarely validated against experimental data. If at
all, the model results are in most cases only compared with measurements in the radiative
section or at the chimney. However, especially for the fuel bed model, a validation with
measurements within or close to the bed surface is preferred.
The scope of the present paper is the presentation of a numerical model which accounts for the different interactions of bed and freeboard region in detail, by considering

254

D. Kurz et al.

Downloaded by [University of Tasmania] at 20:28 30 August 2014

the whole combustion chamber in a single, three-dimensional CFD code. Therefore, a


multiphase approach has to be used, considering the physics and reactions of solid and
gaseous species simultaneously, as well as their multiple interactions. Despite the enormous increase in computational power within the last decade, Lagrangian approaches for
a three-dimensional simulation of an industrial-scale grate firing system still exceed the
computational resources [24]. Hence, for the description of wood chip combustion on a
moving grate, the implementation as an EulerEuler multiphase approach has been chosen.
The following sections introduce the overall mathematical model, and the comparison
of simulation results with detailed experimental measurements at different positions within
the combustion chamber of a 240 kW test facility will be presented.

2. Governing equations
In the following, the governing transport equations for mass, species, momentum and heat
transfer for two interpenetrating continua, namely the gas and the particle phase, in a
Cartesian coordinate system are introduced. In the bed region, only part of the volume is
available for each phase, resulting in the usage of effective values, indicated by a circumflex.
For example, the effective density of the gas phase is given by
g = g g

g + p = 1

(1)

with the physical density g , calculated by the ideal gas law, and the local volume fraction
of the gas phase g . The closure constraint implies that the volume fractions of both phases
sum up to unity.
As transient effects are not of interest for the investigation at hand, the time dependent
term is not considered in the transport equations.

2.1. Transport equations for the gas phase


Continuity

(g ug,j ) = sm,g
xj

(2)

with the effective density of the gas phase g , the gas velocity ug,j in coordinate direction
j and the source term sm,g due to conversion processes during the combustion of the solid
phase.

Momentum

g,ij

p
(g ug,j ug,i ) =
+ g gi g
+ F (ug,i )
xj
xj
xi

(3)

Combustion Theory and Modelling

255

where the shear stress tensor for a Newtonian fluid in presence of a particle phase is given
by

g,ij = g g

ug,i
ug,j
2 ug,k
+
ij
xj
xi
3
xk


with

g = g,l + g,t

(4)

Downloaded by [University of Tasmania] at 20:28 30 August 2014

with the effective viscosity g of the turbulent flow. The last term on the right-hand side of
Equation (3) represents the resistance of fluid flow in a porous medium and is calculated
by Erguns equation. Here, a modified version is used with adapted coefficients valid for
particles with rough surfaces [25]:
2



180 1 g
4 1 g
p
=
g,l ug,i +
u2g,i .
F (ug,i ) =
xi
g3 dp2
g3 dp

(5)

Species transport

(g ug,j Ygi ) =
xj
xj

g Ygi
Y xj


+ sYgi

(6)

where Ygi indicates the mass fraction of gas species and sYgi is the source/sink term accounting for concentration changes of each individual species during drying, devolatilisation and
combustion processes. In the present reaction model, seven different gas species are considered: N2 , O2 , H2 , H2 O, CO, CO2 and Cx Hy .
Enthalpy

(g ug,j hg ) =
xj
xj

g hg
h xj


+ gp (Tp Tg ) + sh,g

(7)

with the specific enthalpy of the gas phase hg and the convective heat exchange between the
gas and particle phases. The source term sh,g accounts for contributions from homogeneous
reactions, mass transfer from the particle phase, and radiative heat exchange.

2.2.

Transport equations for the particle phase

Continuity

(p up,j ) = sm,p
xj

(8)

with the effective density of the particle phase p , the particle velocity up and the sink term
sm,p due to consumption of solid mass during drying, devolatilisation and char combustion.
By means of the following relation, gas and particle continuity are linked:
sm,g = sm,p .

(9)

256

D. Kurz et al.

Momentum
Analogous to the gas phase, the conservation of momentum for the particle phase can be
described in the following manner [21, 22, 23]:

Downloaded by [University of Tasmania] at 20:28 30 August 2014

p,ij

(p up,j up,i ) =
+ p gi + AGrate .
xj
xj

(10)

The last term accounts for random movement of the particles caused by mechanical disturbances of the grate and other random sources. At present no accurate models are at hand,
neither for the shear stress of the particles p nor for AGrate . Therefore, it is common sense
to use a constant velocity of the particle phase in the grate direction, whereas the vertical
velocity is calculated by Equation (8) [2123, 26, 27]. The effect of enhanced mixing in
the bed due to grate movement is included by the particle-mixing model presented by the
group at the Sheffield University Waste Incineration Centre (SUWIC) (e.g. [2123, 27, 28])
as shown in the following sections.

Species transport

(p up,j Ypi ) =
xj
xj



p Ypi
Dp
+ sYpi
xj

(11)

where Ypi represents the mass fraction of solid species moist wood (mw), dry wood (dw),
char (c) and ash (a) in the particle phase. The source term sYpi comprises the change in
concentration due to the heterogeneous reactions drying, devolatilisation and char combustion.
In general, solid species are not subject to diffusional transport [29]. In order to account
for riddling of the fuel bed, the particle-mixing model introduces diffusional transport of
solid species by including an experimentally determined particle-mixing coefficient Dp .

Enthalpy

(p up,j hp ) =
xj
xj



p Tp gp (Tp Tg ) + sh,p
xj

(12)

with the specific particle phase enthalpy hp , the convective heat transfer between the gas
and particle phases gp and the source term sh,p comprising radiative effects, heterogeneous
reactions and enthalpy transfer via mass transfer to the gas phase. The combined thermal
conductivity p consists of two parts, the conductivity of the solid material p0 and additional thermal transport caused by random movement of the particles pm according to the
particle-mixing approach:
p = p0 + pm .

(13)

The transport coefficients describing the particle-mixing p , Dp and pm in the fuel bed
have to be estimated. By assuming a Particle Prandtl Number P rp and a Particle Schmidt

Combustion Theory and Modelling

257

Number Scp of unity Yang et al. [21, 22, 23] deduce the following correlations:
p = p Dp

pm = p cp,p = p cp,p Dp .

and

(14)

Downloaded by [University of Tasmania] at 20:28 30 August 2014

The particle-mixing coefficient Dp can be approximated by measurements and is affected by


the physical properties of the bed material, grate type and operating conditions of the furnace
[21]. Yang et al. performed lab scale as well as full scale experiments for different grates
and bed materials [22]. For the lab scale (1:15), the deduced particle-mixing coefficients
were in the range of 3.3108 to 6.0106 m2 /s, whereas in the full scale measurements,
coefficients in the range of 6.9106 to 1.8104 m2 /s have been reported. In their model
validation, Yang et al. utilised values for the particle-mixing coefficient in the range of
1.0106 to 5.5105 m2 /s [22, 27].
2.3. Convective heat transfer in the bed
The specific convective heat transfer coefficient gp is calculated from the heat transfer
coefficient gp and the specific contact area av between the gas and particle phases:
gp = av gp .

(15)

Under the assumption of spherical particles, the surface area of the particle phase with
respect to the local bed volume is given by
av =

6(1 g )
.
dp

(16)

The heat transfer coefficient gp is calculated from an empirical correlation for the Nusselt
number. Wakao and Kaguei evaluated numerous experimental results for the convective
heat exchange of gas and particle phases within a fixed bed and deduced the following
formula [30]:
Nu =

gp dp
= 2.0 + 1.1 P r 1/3 Rep0.6 .
g

(17)

P r represents the Prandtl number of the gas and Rep the Reynolds number of the flow
around the particles of the bed, given as:
Rep =

ug up |g
dp |
.
g,l

(18)

2.4. Radiative heat transfer


Numerous publications dealing with fixed bed combustion adapt the thermal conductivity
to account for particle radiation within the bed (e.g. [10, 19, 28, 3138]). Here, no use of this
simplification is made and radiation effects of both phases are explicitly modelled, which
is assumed to give more realistic results [39]. Furthermore, this is necessary to account for
the influences between the freeboard region and the fuel bed in terms of radiation effects,
which is supposed to have a strong impact on the overall combustion process.

258

D. Kurz et al.

Assuming that scattering effects can be neglected, the radiative transport equation can
be written as
dI (x, s)
= ka,g (x)Ib,g (x) + ka,p (x)Ib,p (x) (ka,g (x) + ka,p (x))I (x, s)
ds

(19)

Downloaded by [University of Tasmania] at 20:28 30 August 2014

where ka,g (x) and ka,p (x) are the absorption coefficients and Ib,g (x) and Ib,p (x) represent
the black body radiation of gas and particle phase, respectively.
The calculation of the gas phase absorption coefficient ka,g (x) is realised via a polynomial approach, depending on the local CO2 and H2 O concentration [40]. The particle
phase absorption coefficient ka,p (x) is determined by the use of the following correlation
proposed in [39]:
ka,p =

2.5.

1
ln (g ).
dp

(20)

Turbulence modulation

For the description of turbulence a modification of the widely used standard k,  model of
Launder and Spalding [41] is introduced. The adapted transport equations for the turbulent
kinetic energy k and the dissipation rate  for multiphase flow are [42, 43]:

(g uj k) =
xj
xj

(g uj ) =
xj
xj




g,t
+ g,l
k
g,t
+ g,l





k
xj

xj




+ P 
+ sk

(21)


2
+ C1 P C2 + s
k
k

(22)

where the production term P is given by


 


 
ug,j 2 2 ug,k 2
1 ug,i
P = g,t
+

2 xj
xi
3 xk

(23)

and the effective turbulent viscosity is calculated from:


g,t = C g

k2
.


(24)

The source terms sk and s account for the interaction of turbulent gas flow and the
presence of solid particles, known as turbulence modulation. It is generally accepted that
large particles augment the turbulence intensity of the gas phase, whereas small particles
attenuate it [4446]. On the other hand, dense packing of solids is supposed to dampen the
turbulence intensity [43].
Here, the approach of Bolio and Sinclair is used [47]. The model was invented to
account for turbulence production due to wakes and vortex shedding downstream of large
particles:
sk =

np k
.
ep

(25)

Combustion Theory and Modelling

259

It depends on the number of particles per unit volume np ,


np =

6 p
dp3

(26)

and the eddy/particle-interaction time ep ,



ep = MIN

l 2 g
le
, e
|
ug up | g,t


(27)

Downloaded by [University of Tasmania] at 20:28 30 August 2014

where the characteristic eddy length-scale is given by:


le = C0.75

k 1.5
.


(28)

The change in turbulent kinetic energy per eddy/particle-interaction k is given by


k =

1
g Vw (ug 2 up 2 )
2

(29)

with the wake volume Vw taken as half of the volume of an ellipsoid, characterised by the
wake length lw , which is taken from the correlation from Rimon and Cheng [48]:
Vw =

dp2
6

with

lw

lw
= 1.41 log(Re) 1.93.
dp

(30)

The additional term in the transport equation for the dissipation rate  is calculated from:
s = C3 sk


.
k

(31)

The constants of the model are given in Table 1. Within the dense packed bed region, the
distance between adjacent particles is assumed to be too small to allow for the development
of wakes inside the bed. Therefore, the above mentioned model of Bolio and Sinclair is only
utilised in the uppermost region of the fuel bed, where wakes downstream of the particles
are likely to extend into the freeboard region.

2.6. Mathematical modeling of the combustion process


A global seven-step reaction model has been developed for the combustion of wood chips,
as indicated in Figure 1.
Table 1. Constants of the k,  model.
C

C1

C2

C3

0.09

1.44

1.92

1.2

1.0

1.3

Downloaded by [University of Tasmania] at 20:28 30 August 2014

260

D. Kurz et al.

Figure 1. Schematic representation of the reaction model.

Drying
It is assumed that moist wood is heated up, mainly by radiation from the freeboard region,
and the resulting evaporation rate (reaction (s1)) is determined from an enthalpy balance
of the particle phase:

revap =

H p (Tp ) H p (Tsat )
hev + hb

(32)

where Tsat is the saturation temperature, taken as 373.15 K, and hev and hb are the evaporation and bond enthalpy, respectively.

Devolatilisation
The mathematical description of the devolatilisation process of a solid fuel (reaction (s2))
comprises two aspects. On the one hand, the net devolatilisation rate in terms of a mass
flux from the particle phase to the gas phase has to be modelled, and on the other hand, the
composition of the released gas has to be determined as well.
The devolatilisation rate is described with a simple one-step model, using an Arrhenius
expression:


rdevol = rs2 mdw

with

Es2
rs2 = k0,s2 exp
R Tp


.

(33)

Combustion Theory and Modelling

261

The kinetic parameters are taken from [49]:

Downloaded by [University of Tasmania] at 20:28 30 August 2014

k0,s2 = 5300 s1

Es2
= 8567 K.
R

and

(34)

The composition of the volatile gas is identified using proximate and elementary analysis
of the fuel. Assuming that the fuel mainly consists of the chemical elements C, H and
O, three mass balances can be formulated and the amount of three volatile species can
be calculated accordingly. Utilising the experimentally observed ratios of CO/CO2 and
Cx Hy /CO2 published by [50], two additional volatile species can be determined. The
following five values on a mass basis have been determined: Cx Hy 20.1%; CO 43.7%; CO2
18.2%; H2 O 17.8% and H2 0.2%.
In terms of the reaction enthalpy of the devolatilisation, large uncertainty prevails in
the literature, not only of the absolute value but even for the treatment as an exothermic or
endothermic overall process. Comparison of published data is given for example in [51
53]. According to [53] and [54], the competing endothermic and exothermic subprocesses
more or less compensate each other for which reason several authors dealing with thermal
biomass conversion simply neglect the reaction enthalpy of the devolatilisation process [49,
55, 56]. This assumption is made here as well.

Char reactions
Under the concept of char reactions, several processes are summarised. On the one hand,
the oxidation of fixed carbon with oxygen is accounted for, whereas on the other hand,
gasification reactions with carbon dioxide or water vapour need to be considered as well.
(1 + f ) C + O2

2f CO + (1 f ) CO2

(35)

(s4a) :

C + CO2

2 CO

(36)

(s4b) :

C + H2 O

CO + H2 .

(37)

(s3) :

The temperature dependent ratio of the resulting products from the char oxidation reaction
is given by [57] as:


CO
6240
2f
= 2500 exp
=
CO2
1f
Tbl

(38)

where the temperature of the boundary layer Tbl is estimated by the arithmetic mean of the
gas and particle temperatures.
The effective reaction rate for each char reaction is calculated as the harmonic mean of
the limiting resistances, namely physical diffusion and chemical kinetics:

r=

1
rph

1
+
rch

1
.

(39)

262

D. Kurz et al.

According to [30] the physical diffusion rate of a gaseous species i within a packed bed
can be described as:
rph =


Di 
2 + 1.1 Rep0.6 P r 1/3 .
dp

(40)

Instead of Di , for simplification the binary diffusion coefficient of oxygen in nitrogen is


used [58]:

Downloaded by [University of Tasmania] at 20:28 30 August 2014

DO2 = 3.49 104

Tbl
T0

1.75

p0
p

(41)

which is based on reference conditions T0 (1600 K) and p0 (1bar).


The chemical reaction rate rch is calculated via an Arrhenius expression:

rch = k0,ch Tp exp


Ech
.
R Tp

(42)

The kinetic constants for the char reactions are given in Table 2. The rates are calculated
using the specific surface area of the particle phase (Equation 16):
rchar = av r mc .

(43)

Homogeneous reactions
A global model for the homogeneous reactions is used, as indicated in Figure 1. For
simplicity, a two-step model for the oxidation of higher hydrocarbons, as presented in [61],
is chosen. Hydrogen from the heterogeneous pyrolysis reaction is oxidised to water vapour.
(g1) :

Cx Hy +

(g2) :
(g3) :

y
O2 x CO +
H2 O
2
4
2
1
CO + O2 CO2
2
1
H2 + O2 H2 O.
2

(44)
(45)
(46)

Turbulence/chemistry interaction is described by the eddy dissipation model of Magnussen


and Hjertager [62]:
rgj = AMag



YP rod

YF YO2
mF MIN
,
, 0.5
.
k
|F | |O2 |
|P rod |

(47)

Table 2. Kinetic parameters of char oxidation and gasification.


Reaction
(s3)
(s4a)
(s4b)

Reactand

k0,i [m/(sK)]

Ei /R [K]

Source

O2
CO2
H2 O

1.715
3.42
3.42

9000
15600
15600

[59]
[60]
[60]

Combustion Theory and Modelling

263

When applying this model to the relatively slow carbon monoxide oxidation, additionally a
kinetic rate proposed by Dryer and Glassmann [61] is calculated, and the minimum of both
rates is taken as the effective one. For further details please refer to [63].

Downloaded by [University of Tasmania] at 20:28 30 August 2014

3.

Numerical solution

The above mentioned transport equations are solved in the finite volume code AIOLOS
which has been developed at the Institute of Combustion and Power Plant Technology
(IFK), University of Stuttgart. The SIMPLE method, described by [64], is utilised for
velocitypressure coupling, whereas the pressure interpolation scheme of [65] prevents
decoupling of velocities and pressure on the non-staggered grid. The calculation of the
pressure correction equation was done using the SIP method. An upwind scheme is applied
for the calculation of convective fluxes and the solution of all other transport equations was
carried out by a SOR solver. The radiation equation is solved using a discrete ordinates
method. Concerning implementation issues the reader is referred to [63] and [66]. The
model parameters are compiled in Table 3.

4. Experimental set-up
For validation purposes of the presented numerical model, detailed measurements were
conducted at a 240 kW grate firing test facility. A schematic plot of the test facility is given
in Figure 2. The fuel is forwarded into the combustion chamber with a conveying screw.
Primary air from two different wind boxes is supplied from underneath the grate. Transport
and riddling of the fuel is accomplished by five rows of periodically pushing grate bars.
Burnout is assured by means of six arrays of secondary air nozzles in the flue gas duct.
Validation is carried out by means of temperature measurements in the central cross-section
of the combustion chamber and the flue gas duct, as well as measurements of the gaseous
species oxygen, carbon monoxide and carbon dioxide at four different positions above the
fuel bed in different cross-sections as indicated in Figure 2. The fuel input consisted of
wood chips (spruce wood with bark) with a high moisture content of almost 50%. The net
input of fuel energy was 281 kW.

5. Results and discussion


The computational grid, consisting of about 150 000 cells, as well as the hydrodynamic
boundary conditions are shown in Figure 3. The preset flows of primary air (PA) and
secondary air (SA) are assigned to their specific positions, whereas the calculated leakage
air is assumed to enter the furnace mainly through the fuel inlet. Due to the refractory
Table 3. Model parameters.
Particle volume fraction within the bed
Particle mixing coefficient
Initial particle diameter
Physical density of the particle phase
Thermal conductivity of the particle phase
Evaporation enthalpy of water
Bond enthalpy of water in wood
Magnussen coefficient of homogeneous reactions

p
Dp
dp
p
p,0
hev
hb
AMag

0.65 m3 /m3
3.0105 m2 /s
0.02 m
770 kg/m3
0.2 W/(m K)
2257 kJ/kg
400 kJ/kg
1.0

Downloaded by [University of Tasmania] at 20:28 30 August 2014

264

D. Kurz et al.

Figure 2. Schematic representation of the test facility with measurement positions.

lining, the walls of the combustion chamber are assumed adiabatic, with the exception of
the uppermost duct which is water cooled.
Due to the strong influence of the particle phase on the overall system and especially the
severe impact on the local behaviour in the case of existence/non-existence of the particle
phase, especially the particle phase enthalpy as well as the particle phase species needed
to be under-relaxed rather strongly. After 50 000 iterations, the numerical solution of all
transport equations converged to normalised residuals of less than 105 . In addition, the
overall balances of enthalpy and elements C, H and O was closed with an error of less than
0.1% related to the inflowing values. As the implementation is highly optimised for parallel
vector computers, a computational time of less than two hours on a NEC-SX8 platform
using eight CPUs could be achieved.
The computed temperature field in the centre section of the furnace is shown in Figure 4.
One can easily distinguish the relatively cold zone close to the fuel inlet where drying of the
fuel mainly by radiation from the surrounding hot freeboard region occurs. During pyrolysis,
combustible gases and char are released which leads to an increasing temperature due to
oxidation reactions. The maximum temperature is observed in the flue gas duct where the
secondary air supply results in the burnout of the gas, leading to an additional temperature
increase. After two thirds of the grate length, the char content of the residual solids is
depleted and the resulting ash is cooled by radiative and convective processes, leading to a
temperature decrease towards the ash outlet.

Downloaded by [University of Tasmania] at 20:28 30 August 2014

Combustion Theory and Modelling

265

Figure 3. Computational grid.

The results are validated against the four measurement points within the flue gas duct,
as well as the six positions in the combustion chamber, referred as measurement line M200
(see Figure 4). Comparison of simulated and measured temperatures on measurement line
M200 is given in Figure 5, where the temperatures are plotted against the horizontal distance
from the left side of the furnace. For the experimental results, the range of the measured
values is given in terms of error bars. It is observed that simulated temperatures are slightly
higher than the measured ones. This might be due to the assumption of adiabatic walls,

Figure 4. Simulated temperature profile in the centre cross-section.

266

D. Kurz et al.
1100

Messung
Simulation

900
800
700
600
500
400

0.3

0.5

0.7
0.9
1.1
Position auf M200 [m]

1.3

1.5

Figure 5. Simulated and measured temperatures on M200.

whereas in reality some heat removal over the walls by radiation and convection is likely
to occur. Nevertheless, the overall temperature profile within the combustion chamber is
depicted quite well by the model. Simulated and measured temperatures within the flue gas
duct are compared in Figure 6. Especially for the measurement position FG3, the simulated
temperature is considerably higher than the measured counterpart. As indicated by the
horizontal cross-section through the measurement points in the flue gas duct (Figure 7),
the measurement point FG3 is located in an area of large temperature gradients. Therefore,
even a slightly different position of the thermocouple or unsteady phenomena will result
in a considerably lower temperature, which might explain part of the deviation from the
1300

experiment
simulation

1200
Temperature [C]

Downloaded by [University of Tasmania] at 20:28 30 August 2014

Temperature [C]

1000

1100
1000
900
800
700
600

FG1

FG2
FG3
Points of measurement

Figure 6. Simulated and measured temperatures in the flue gas duct.

FG4

Combustion Theory and Modelling

267

Downloaded by [University of Tasmania] at 20:28 30 August 2014

Figure 7. Temperature profile in the cross-section of measurement points FG1 to FG3.

simulated value. Also the penetration depth of the secondary air nozzle flows might be
underestimated in the model, resulting in an overestimation of the gas temperature at
measurement points FG1 to FG3. At position FG4 where strong mixing due the deflection
of the flue gas stream has occurred the simulated temperature corresponds well to the
measured one.
Validation of the reaction model is carried out by comparison of measured gas concentrations at the points of measurement M1 to M4 with the simulation results (see Figure 2).
The concentration plots of oxygen and carbon monoxide on a dry basis in the centre crosssection of the furnace, and the positions of the measurement points M1 to M4, are given
in Figures 8a and 8b. According to the oxygen profile, one can easily identify the section

Figure 8a. Simulated O2 profile in the centre cross-section.

D. Kurz et al.

Figure 8b. Simulated CO profile in the centre cross-section.

of high fuel conversion in the second third of the grate. Due to the already mentioned
low penetration depth of the secondary air nozzles in the flue gas duct, the low oxygen
concentrations extend into the water cooled region of the flue gas duct. The high carbon
monoxide concentrations above the grate accrue from devolatilisation and char reaction

18
Oxygen concentration [vol.-%,dry]

Downloaded by [University of Tasmania] at 20:28 30 August 2014

268

experiment
simulation

16
14
12
10
8
6
4
2
0

M1

M2
M3
Points of measurement

Figure 9a. Comparison of simulated and measured O2 profiles.

M4

Carbon monoxide concentration [vol.-%,dry]

24

269

experiment
simulation

22
20
18
16
14
12
10
8
6
4
2
0

M1

M2
M3
Points of measurement

M4

Figure 9b. Comparison of simulated and measured CO profiles.

processes, whereas the carbon monoxide is depleted by the oxidation to carbon dioxide in
the course of the flue gas flow.
The simulated oxygen, carbon monoxide and carbon dioxide concentrations and their
measured counterparts are given in Figures 9a9c, where the gas concentration for each
measurement point is plotted over the width of the combustion chamber.
As indicated by the error bars, experimental results are subject to considerable fluctuations, especially in the vicinity of the bed surface. Reasons for that are unavoidable
unsteady phenomena, caused by riddling of the fuel due to the grate movement, and sudden
changes in the bed status like the appearance or destruction of channels.

Carbon dioxide concentration [vol.-%,dry]

Downloaded by [University of Tasmania] at 20:28 30 August 2014

Combustion Theory and Modelling

20

experiment
simulation

18
16
14
12
10
8
6
4
2
0

M1

M2
M3
Points of measurement

Figure 9c. Comparison of simulated and measured CO2 profiles.

M4

270

D. Kurz et al.

Downloaded by [University of Tasmania] at 20:28 30 August 2014

Regarding the oxygen concentration (Figure 9a), the model slightly underpredicts the
value in the main reaction zone, indicating an overestimation of the reaction process. Close
to the fuel inlet, at measurement point M4, the simulated oxygen concentration is higher
than the measured one. The carbon monoxide (Figure 9b) and carbon dioxide (Figure 9c)
profiles correspond reasonably well to the measured values, as the simulation results are
located in the range of variations of the experimental data. Only at measurement point M4,
the carbon dioxide concentration is underestimated by the model. One possible reason might
be the above mentioned assumption for the leakage air. Assuming the origin of leakage air
through the fuel inlet might lead to the overestimation of oxygen and enhanced dilution of
the flue gas flow above the fuel inlet.

6.

Conclusions

The governing equations for mass, species and momentum conservation and heat transfer
for a multiphase system applicable to a grate firing arrangement have been presented.
Mathematical models for drying, devolatilisation and char burnout of the solid fuel, as
well as combustion reactions in the gas phase are illustrated in detail, as well as models
describing radiation and turbulence. Therefore, the whole combustion chamber, including
both the fuel bed and the adjacent freeboard region, can be simulated within one CFD
code, enabling the coupled description of gassolid interactions in terms of flow, species
transport, turbulence, reactions and heat transfer by convection and radiation.
To validate the model results, detailed measurements within a 240 kW grate firing test
facility have been performed. Emphasis was put especially on the regions close to the bed
surface where strong interactions between solid and gas reactions, as well as flow and
heat transfer phenomena are likely to occur. Experimentally observed temperatures and
gas concentrations were compared with the corresponding numerical results. Reasonably
good agreement between experimental and numerical results has been observed. For most
of the measurement points, the numerical results lie within the range of deviations of
the experimentally observed values. Also, the general trend of all measurements, both in
parallel and in transversal directions of the grate, is depicted quite well by the model.
Nevertheless, there is still room for improvement of various sub-models. Especially for the
heterogeneous reaction model the consideration of an uneven particle size distribution needs
to be implemented in order to accurately predict the burnout behaviour of the solid fuel.
To prove the general applicability, the sensitivity of the model regarding fuel properties
like moisture content as well as operational parameters will be presented and compared to
experimental findings in a subsequent paper. Furthermore, the scale-up of the presented
model towards industrial-scale arrangements needs to be validated. Therefore, detailed
measurements within a 60 MW wood chip combustion grate firing application close to
the bed surface and within the radiative section have been conducted, and comparison to
numerical results will be presented in future publications.

Acknowledgements
Funding of this work by Energie Baden-Wurttemberg (EnBW) is gratefully acknowledged. Special
thanks go to Sven Unterberger and Oliver Greil from EnBW for their cooperation, as well as our
colleagues Daniel Kilgus and Kevin Brechtel for their support and for conducting the experimental
measurements.

Combustion Theory and Modelling

271

References
[1]
[2]
[3]
[4]
[5]

Downloaded by [University of Tasmania] at 20:28 30 August 2014

[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]

C. Yin, L.A. Rosendahl, and S.K. Kr, Grate-firing of biomass for heat and power production,
Progr. Energy Comb. Sci. 34 (2008), pp. 725754.
R. Scharler, Entwicklung und Optimierung von Biomasse-Rostfeuerungen durch CFD-Analyse,
University of Graz, 2001.
C. Yin, L. Rosendahl, S.K. Kr, S. Clausen, S.L. Hvid, and T. Hille, Mathematical modeling
and experimental study of biomass combustion in a thermal 108 MW grate-fired boiler, Energy
Fuels 22 (2008), pp. 13801390.
N. Griselin and X.S. Bai, Particle dynamics in a biomass-fired furnace Predictions of solid
residence changes with operation, IFRF Comb. J. article number 200009 (2000).
W. Dong and W. Blasiak, CFD modeling of Ecotube system in coal and waste grate combustion,
Energy Conv. Managem. 42 (2001), pp. 18871896.
T. Klasen and K. Gorner, Numerical calculation and optimisation of a large municipal solid
waste incinerator plant, in Proceedings of the 2nd International Symposium on Incineration
and Flue Gas Treatment Technologies, Sheffield, UK, 1999.
K. Gorner and T. Klasen, Modelling, simulation and validation of the solid biomass combustion
in different plants, Progr. Comp. Fluid Dyn. 6 (2006), pp. 225234.
S. Kim, D. Shin, and S. Choi, Comparative evaluation of municipal solid waste incinerator
designs by flow simulation, Combust. Flame 106 (1996), pp. 241251.
M. Huttunen, L. Kjaldman, and J. Saastamoinen, Analysis of grate firing of wood with numerical
flow simulation, IFRF Comb. J. article number 200401 (2004).
C. Wolf, Erstellung eines Modells der Verbrennung von Abfall auf Rostsystemen unter besonderer Berucksichtigung der Vermischung, University Duisburg-Essen, 2005.
M. Beckmann, Mathematische Modellierung und Versuche zur Prozessfuhrung bei der Verbrennung und Vergasung in Rostsystemen zur thermischen Ruckstandsbehandlung, University
of Clausthal, 1995.
T. Gruber, Vorgange bei der Verbrennung von Hausmull auf dem Rost, University of Berlin,
1993.
G. Brem, R. Gort, and L.B.M. van Kessel, Theoretical and experimental modelling of municipal
solid waste incineration, in Ruckstnde aus der Mullverbrennung, M. Faulstich, ed., EF-Verlag,
Berlin, 1990.
M. Rovaglio, D. Manca, G. Biardi, and J. Falcon, Dynamic modelling of waste incineration
systems: A startup procedure, Comp. Chem. Eng. 18 (1994), pp. 361368.
B. Peters, U. Muller, and L. Krebs, A principal approach to model furnace processes for waste
incineration, Env. Comb. Tech. 2 (2001), pp. 383402.
C. Bruch, Modellierung der Festbettverbrennung in automatischen Holzfeuerungen, University
of Zurich, 2001.
B. Peters, A. Dziugys, H. Hunsinger, and L. Krebs, An approach to qualify the intensity of
mixing on a forward acting grate, Chem. Eng. Sci. 60 (2005), pp. 16491659.
E. Simsek, B. Brosch, S. Wirtz, V. Scherer, and F. Krull, Numerical simulation of grate firing
systems sing a coupled CFD/Discrete Element Method (DEM), Powder Tech. 193 (2009), pp.
266273.
J. Strohle, A. Austegard, M. Grnli, and T. Pettersen, Two-dimensional numerical simulation of a moving bed of wood, in Science in Thermal and Chemical Biomass Conversion,
A. Bridgewater and D. Boocock, eds., CPL Press, 2006.
Y.R. Goh, C.N. Lim, R. Zakaria, K.H. Chan, G. Reynolds, Y.B. Yang, R.G. Siddall, V.
Nasserzadeh, and J. Swithenbank, Mixing, modelling and measurements of incinerator bed
combustion, J. Inst. Chem. Eng., Trans. I. Chem. E, Part B 78 (2000), pp. 2132.
Y.B. Yang, Y.R. Goh, R. Zakaria, V. Nasserzadeh, and J. Swithenbank, Mathematical
modelling of MSW incineration on a travelling bed, Waste Managem. 22 (2002), pp.
369380.
Y.B. Yang, C.N. Lim, J. Goodfellow, V.N. Sharifi, and J. Swithenbank, A diffusion model for
particle mixing in a packed bed of burning solids, Fuel 84 (2005), pp. 213225.
Y.B. Yang, V.N. Sharifi, and J. Swithenbank, Numerical simulation of municipal solid waste
Incineration in a moving-grate furnace and the effect of waste moisture content, Progr. Comp.
Fluid Dyn. 7 (2007), pp. 129142.
S. Li, Y. Ding, D. Wen, and Y. He, Modelling of the behaviour of gassolid two-phase mixtures
flowing through packed beds, Chem. Eng. Sci. 61 (2006), pp. 19221931.

272
[25]
[26]
[27]
[28]
[29]

Downloaded by [University of Tasmania] at 20:28 30 August 2014

[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]

D. Kurz et al.
F.A. Dullien, Porous Media: Fluid Transport and Pore Structure, Academic Press, San Diego,
1992.
Y.B. Yang, V. Nasserzadeh, J. Goodfellow, Y.R. Goh, and J. Swithenbank, Parameter study on
the incineration of municipal solid waste fuels in packed beds, J. Inst. Energy 75 (2002), pp.
6680.
Y.B. Yang, J. Goodfellow, V.N. Sharifi, and J. Swithenbank, Investigation of biomass combustion systems using CFD techniques: A parametric study of packed-bed burning characteristics,
Progr. Comp. Fluid Dyn. 6 (2006), pp. 262271.
Y.B. Yang, R. Newman, V. Sharifi, J. Swithenbank, and J. Ariss, Mathmatical modelling of
straw combustion in a 38 MWe power plant furnace and effect of operating conditions, Fuel
86 (2007), pp. 261273.
B. Peters, A detailed model for devolatilization and combustion of waste material in packed
beds, in 3rd European Conference on Industrial Furnaces and Boilers (INFUB), 1521 April
1995, Lisbon, Portugal.
N. Wakao and S. Kaguei, Heat and Mass Transfer in Packed Beds, Gordon and Breach, New
York, 1982.
S. Dasappa and P.J. Paul, Gasification of char particles in packed beds: Analysis and results,
Int. J. Energy Res. 25 (2001), pp. 10531072.
H. Thunman and B. Leckner, Thermal conductivity of wood models for different stages of
combustion, Biomass Bioenergy 23 (2002), pp. 4754.
K.M. Bryden and M.J. Hagge, Modeling the combined impact of moisture and char shrinkage
on the pyrolysis of a biomass particle, Fuel 82 (2003), pp. 16331644.
B. Peters and C. Bruch, Drying and pyrolysis of wood particles: Experiments and simulation,
J. Analyt. Appl. Pyrolysis 70 (2003), pp. 233250.
Y.B. Yang, C. Ryu, A. Khor, V.N. Sharifi, and J. Swithenbank, Fuel size effect on pinewood
combustion in a packed bed, Fuel 84 (2005), pp. 20262038.
Y.B. Yang, V.N. Sharifi, and J. Swithenbank, Numerical simulation of the burning characteristics of thermally-thick biomass fuels in packed-beds, J. Inst. Chem. Eng., Trans. I. Chem. E,
Part B 83 (2005), pp. 549558.
R. Johansson, Modelling elements in conversion of solid fuels Fixed bed bombustion and
gaseous radiation, Chalmers University of Technology, Goteborg, Sweden, 2008.
C. Ghabi, H. Bentchia, and M. Sassi, Two-dimensional computational modeling and simulation
of wood particles pyrolysis in a fixed bed reactor, Comb. Sci. Technol. 180 (2008), pp. 833853.
D. Shin and S. Choi, The combustion of simulated waste particles in a fixed bed, Combust.
Flame 121 (2000), pp. 167180.
K. Schack, Berechnung der Strahlung von Wasserdampf und Kohlendioxid, Chem. Ing. Technik
42 (1970), pp. 53104.
B.E. Launder and D.B. Spalding, The Numerical computation of turbulent flows, Comp.
Methods Appl. Mech. Eng. 3 (1974), pp. 269289.
H. Qi, Euler/Euler Simulation der Fluiddynamik zirkulierender Wirbelschichten, RWTH
Aachen, 1997.
Y. Zhang and J.M. Reese, Gas turbulence modulation in a two-Fluid model for gassolid flows,
Am. Inst. Chem. Eng. 49 (2003), pp. 30483065.
R.A. Gore and C.T. Crowe, Effect of particle size on modulating turbulent intensity, Int. J.
Multiphase Flow 15 (1989), pp. 279285.
Z. Yuan and E.E. Michaelidis, Turbulence modulation in particulate flows A theoretical
approach, Int. J. Multiphase Flow 18 (1992), pp. 779785.
C.T. Crowe, T.R. Troutt, and J.N. Chung, Numerical models for two-phase turbulent flows,
Ann. Rev. Fluid Mech. 28 (1996), pp. 1143.
E.J. Bolio and J.L. Sinclair, Gas turbulence modulation in the pneumatic conveying of massive
particles in vertical tubes, Int. J. Multiphase Flow 21 (1995), pp. 9851001.
Y. Rimon and S. Cheng, Numerical solution of a uniform flow over a sphere at intermediate
Reynolds numbers, Phys. Fluids 12 (1969), pp. 949959.
T. Willner and G. Brunner, Pyrolysis kinetics of wood and wood components, Chem. Eng.
Technol. 28 (2005), pp. 12121225.
H. Thunman, F. Niklasson, F. Johnsson, and B. Leckner, Composition of volatile gases and
thermochemical properties of wood for modeling of fixed or fluidized beds, Energy Fuels 15
(2001), pp. 14881497.

Combustion Theory and Modelling


[51]
[52]
[53]
[54]
[55]
[56]

Downloaded by [University of Tasmania] at 20:28 30 August 2014

[57]
[58]
[59]
[60]
[61]
[62]
[63]
[64]
[65]
[66]

273

H.C. Kung and A.S. Kalelkar, On the heat of reaction in wood pyrolysis, Combust. Flame 20
(1973), pp. 91103.
C. Gomez Daz, Understanding biomass pyrolysis kinetics: Improved modeling based on
comprehensive thermokinetic analysis, University of Catalunya, 2006.
C. Di Blasi, Modeling chemical and physical processes of wood and biomass pyrolysis, Progr.
Energy Combust. Sci. 34 (2008), pp. 4790.
A. Demirbas and G. Arin, An overview of biomass pyrolysis, Energy Sources 24 (2002), pp.
471482.
K.M. Bryden and K.W. Ragland, Numerical modeling of a deep, fixed bed combustor, Energy
Fuels 10 (1996), pp. 269275.
C. Di Blasi, Modeling wood gasification in a countercurrent fixed-bed reactor, Am. Inst.
Chem. Eng. 50 (2004), pp. 23062319.
J. Arthur, Reactions between carbon and oxygen, J. Trans. Faraday Soc. 47 (1951), pp. 164178.
M.N. Anany, Numerical modelling of combustion processes at elevated pressures, University
of Stuttgart, 2010.
D.D. Evans and H.W. Emmons, Combustion of wood charcoal, Fire Safety J. 1 (1977), pp.
5766.
B.S. Brewster, S.C. Hill, P.T. Radulovic, and L.D. Smoot, in Fundamentals of Coal Combustion
for Clean and Efficient Use, L.D. Smoot, ed., Elsevier, Amsterdam, 1993, pp. 567706.
F.L. Dryer and I. Glassmann, High-temperature oxidation of CO and CH4 , in Proceedings of
the 14th Symposium (Int.) on Combustion, 1972.
B.F. Magnussen and B.H. Hjertager, On mathematical modeling of turbulent combustion with
special emphasis on soot formation and combustion, in Proceedings of the 16th Symposium
(Int.) on Combustion, 1976, pp. 719729.
R. Schneider, Beitrag zur numerischen Berechnung dreidimensional reagierender Stromungen
in industriellen Brennkammern, University of Stuttgart, 1997.
S.V. Patankar, Numerical Heat Transfer and Fluid Flow, Hemisphere Publishing Corporation,
1980.
A.W. Date, Complete pressure correction algorithm for the solution of incompressible Navier
Stokes equations on a non-staggered grid, Numer. Heat Transfer 29 (1996), pp. 441458.
J. Strohle, Spectral modelling of radiative heat transfer in industrial furnaces, University of
Stuttgart, 2002.

Potrebbero piacerti anche