Está en la página 1de 14

Chemical Engineering Science 55 (2000) 2931}2944

Dynamic behaviour of strati"ed downdraft gasi"ers


Colomba Di Blasi*
Dipartimento di Ingegneria Chimica, Universita& degli Studi di Napoli **Federico II++, P.le V. Tecchio, 80125 Napoli, Italy
Received 20 November 1998; received in revised form 13 September 1999; accepted 8 November 1999
Abstract
A one-dimensional unsteady model is formulated for biomass gasi"cation in a strati"ed concurrent (downdraft) reactor. Heat and
mass transfer across the bed are coupled with moisture evaporation, biomass pyrolysis, char combustion and gasi"cation, gas-phase
combustion and thermal cracking of tars. Numerical simulation has allowed to predict the in#uence of model parameters, kinetic
constants and operational variables on process dynamics, structure of the reaction front and quality of the producer gas. In particular,
two di!erent stabilization modes of the reaction front have been determined. For high values of the air-to-fuel ratio and of the primary
pyrolysis rate, the process is top-stabilized, resulting in a high conversion e$ciency and good gas quality. As the air #ow is decreased
below a critical limit value, the reaction front becomes grate-stabilized. The two di!erent con"gurations are largely determined by the
gas-phase combustion of volatile pyrolysis products. Finally, the predictions of the gas composition and the axial temperature pro"les
are in agreement with experimental data. 2000 Elsevier Science Ltd. All rights reserved.
Keywords: Mathematical modelling; Simulation; Packed bed; Gasi"cation; Biomass
1. Introduction
Biomass and waste are widely recognized to be a major
potential for energy production. Gasi"cation enables
conversion of this material into combustible gas, mech-
anical and electrical power and synthetic fuels and chem-
icals. In principle, the gasi"cation units employed for coal
can also be applied for biomass and waste, but signi"cant
di!erences exist between the two fuel categories. Coal
pyrolysis yields 60}80% char the balance coming from
gases and tars. Updraft, countercurrent gasi"ers are well
suited for the conversion of low-reactive char into gas
(around 90% of the coal gasi"ed in the world makes use
of this con"guration (Hobbs, Radulovic & Smoot, 1993)).
When biomass is pyrolyzed, gases and tars represent
70}90% of the total mass fed, whereas only 30}10% is
a highly reactive char. In updraft air gasi"cation, the
oxygen is consumed at the grate, essentially through
partial combustion of char. The resulting hot gases cause
char gasi"cation and biomass pyrolysis. The relatively
low temperatures and the absence of oxygen result in
large amounts of tars in the producer gas.
* Tel.: 39-081-7682232; fax: 39-081-2391800.
E-mail address: diblasi@unina.it (C. D. Blasi)
Downdraft gasi"ers, characterized by the concurrent
#ows of solid fuel and gas, show little #exibility with
respect to the fuel moisture content and size. However,
these units are usually preferred for small-scale processes
(below 500 kWe) because they produce a cleaner gas,
resulting in a less complicated cleaning process. In the
classical design, the packed bed is supported across
a constriction (throat), where most of the gasi"cation
takes place. Air is fed above this zone and, due to the
restriction, the #ow is highly turbulent, thus favouring
mixing. The more recent version is strati"ed (or open-
core), where there is no restriction and the bed is sup-
ported on a grate. This design has been shown to cope
well with problems deriving from the gasi"cation of loose
materials, that is, poor oxygen distribution with conse-
quent instabilities and ash melting (Buekens, Bridgwater,
Ferrero & Maniatis, 1990).
Di!erences between the countercurrent and concur-
rent gasi"cation are also due to the processes which
determine the composition of the producer gas. In the
"rst case, the quality of the gas is largely dependent on
the presence of devolatilization products. On the con-
trary, the heating value of the producer gas of concurrent
gasi"ers is determined by the amount of carbon monox-
ide and hydrogen present. Both species are formed by the
cracking of tars and the slow heterogeneous gasi"cation
0009-2509/00/$- see front matter 2000 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 0 9 - 2 5 0 9 ( 9 9 ) 0 0 5 6 2 - X
Nomenclature
A
G
pre-exponential factor (Eqs. (18)}(22), (24), (25),
(27))
B
E
permeability to gas #ow, m`
c speci"c heat, J/kg K
C molar concentration, kmol/m`
d particle diameter, m
D reactor diameter, m
D
G
di!usion coe$cient, m`
E
G
activation energy, kJ/mol
h heat transfer coe$cient, W/m` K
H
G
speci"c species enthalpy, kJ/kg
k
K
mass transfer coe$cient, m/s
kH
K
maximum value of the mass transfer coe$cient,
m/s
m
+
moisture evaporation rate, kg/m` s
M molecular weight
p gas pressure, kPa
p
TQ
vapor pressure expressed by the
Clausius}Clapeyron equation
Pr particle Prandtl number
r
A
current particle radius, m
R initial particle radius
R universal gas constant
Re particle Reynolds number
R
H
reaction rate, kg/m` s
Sc particle Schmidt number
t time, s
temperature, K
; total moisture content, d.b., kg/kg
;
E
gas velocity, m/s
;
Q
solid velocity, m/s
< particle volume, m`
=
?
air feed rate, kg/h
=
@
biomass feed rate, kg/h
z space, m
Greek letters
: stoichiometric coe$cient
[ stoichiometric coe$cient
stoichiometric coe$cient
Ah reaction enthalpy, kJ/kg
c porosity
zH thermal conductivity, W/mK
moisture (evaporation) enthalpy, kJ/kg
j viscosity, kg/ms
v
N
particle density number, 1/m
j
A"
constant bed density in the combustion/gasi"-
cation zone, kg/m`
j
G
gas phase (i"O
`
, CO,
2
) mass concentration
(mass/gas volume), kg/m`
j
I
apparent solid (k"B, M) density (mass/total
volume), kg/m`
o Stephan}Boltzmann constant
, ash content of the biomass, % of initial dry
mass
c
G
rate of species production (devolatilization),
kg/m` s
correction factor for the solid/gas heat transfer
coe$cient
Subscripts
B biomass (wood)
c1 tar combustion
c2 methane combustion
c3 carbon monoxide combustion
c4 hydrogen combustion
c5 char combustion
C CHAR
CH
"
methane
CO carbon monoxide
CO
`
carbon dioxide
E equilibrium
g total volatiles (vapour#gas)
gw gas/wall
g1 carbon dioxide gasi"cation
g2 hydrogen gasi"cation
g3 steam gasi"cation
H
`
hydrogen
H
`
O steam
i gas-phase chemical species
j chemical reaction
k solid-phase chemical species
M moisture
O
`
oxygen
p1 primary pyrolysis
p2 secondary pyrolysis
s solid
sg solid/gas
sw solid/wall
TAR
v vapour
w wall
wg water gas shift
0 ambient value
of char, so that the size of the reduction zone and the
residence time of the gasifying agents play a fundamental
role on the performances of the gasi"er. These are, on the
other hand, dependent not only on the size of the gasi"er
but also on the position where the reaction zone stabil-
izes (Reed & Markson, 1985).
Kinetics-free, equilibrium models can predict the exit
gas composition, given the solid composition and the
2932 C. Di Blasi / Chemical Engineering Science 55 (2000) 2931}2944
Table 1
Model equations
* biomass
c

ct
#
c;
Q

cz
"!R
N
, (1)
* moisture
c
+
ct
#
c;
Q

+
cz
"!m
+
, (2)
* gas-phase species
c
c
G
ct
#
c(
G
;
E
)
cz
"
c
cz
cD
G

E
c>
G
cz
#M
G

H
v
GH
R
H
#c
NG
,
i"O
`
, H
`
, CO, CO
`
, CH
"
, j"c1}c4, g1}g3, wg, (3}7)
* steam
c
c
'
`
'
ct
#
c(
'
`
'
;
E
)
cz
"M
'
`
'

H
v
'
`
'H
R
H
#
c
cz
cD
T

E
c>
'
`
'
cz
#m
+
#c
N'
`
'
, j"c1}c4, g1}g3, wg, (8)
* vapor-phase tar
c
c
2
ct
#
c(
2
;
E
)
cz
"
c
cz
cD
2

E
c>
2
cz
#v
2
R
N
!R
N`
, (9)
* nitrogen

`
`
"
E
!
G$`
`

G
, (10)
* total gas continuity
c
c
E
ct
#
c(
E
;
E
)
cz
"
G

H
v
GH
M
G
R
H
#m
+
#(1!v
!
)R
N
, (11)
i"N
`
, O
`
, H
`
, CO, CO
`
, CH
"
, H
`
O, j"c1}c4, g1}g3, wg,
* solid-phase energy
c(
G

G
H
G
)
ct
"
c
cz
zH
Q
c
Q
cz
#
c(;
Q

G
H
G
)
cz
!
H
R
H
AH
H
!Q
QE
#Q
QU
!m
+
, (12)
H
G
"c
QG
(
Q
!
"
), i"B, C, M, j"c5, g1}g3, p1,
* gas-phase energy
c
c(
G

G
H
G
)
ct
"
c
cz
zH
E
c
E
cz
!
c(;
E

G
H
G
)
cz
#Q
QE
#Q
EU
!
H
R
H
AH
H
, (13)
h
G
"c
EG
(
E
!
"
), i"N
`
, O
`
, H
`
, CO, CO
`
, CH
"
, H
`
O, T,
j"c1}c4, wg, p2,
Q
QE
"h
QE
v
N
(
Q
!
E
), Q
QU
"
4h
QU
D
(
U
!
Q
),
Q
EU
"
4h
EU
D
(
U
!
E
),
* ideal gas law, modi"ed Darcy law
P"

E
R
E
M
E
, (14)
B
E
j
cP
cz
";
Q
!;
E
. (15)
equilibrium temperature, but they cannot be used for
reactor design. Only in a very few cases chemical reaction
kinetics and transport phenomena have been properly
coupled to model conventional (Groeneveld & van
Swaaij, 1980) and strati"ed (Manurung & Beenackers,
1994) downdraft gasi"ers. However, the description of
the `#aming pyrolysisa (Reed & Markson, 1985), that is
biomass pyrolysis and combustion of volatile pyrolysis
products, is still based on equilibrium models or on
highly simpli"ed treatments, which assume the existence
of a stable combustion zone with in"nitely oxygen con-
version rate near the air inlet. Furthermore, these models
describe steady-state conditions. Thus, they do not allow
the prediction of the dynamic behaviour of strati"ed
gasi"ers and of the di!erent modes of stabilization of the
reaction front and thus the size of the reduction zone.
This study proposes a more advanced, dynamic model,
which includes "nite rate kinetics for biomass pyrolysis
and combustion of char, gaseous species and tars. The
model was used to simulate the e!ects of changes in
transport coe$cients, chemical kinetics and operating
conditions on the performances of the gasi"cation pro-
cess, in view of reactor design and optimization, espe-
cially in relation to the dynamic behaviour of downdraft
reactors, i.e. to top- or grate-stabilized operation.
2. Mathematical model
The strati"ed gasi"er model is based on mass and
energy balances for the solid phase and mass and energy
balances for the gas phase, written for a one-dimensional,
unsteady state system. Species considered are: oxygen,
nitrogen, hydrogen, steam, carbon dioxide, carbon mon-
oxide, methane and hydrocarbons (which also include
tars). The pressure drop in the reactor is modelled using
the generalized Darcy law but, given the large bed per-
meabilities, simulations have been carried out with the
assumption of constant pressure (the gas velocity is deter-
mined from the continuity equation and the density of
the mixture from the ideal gas law). Model equations are
listed in Table 1. As the concurrent #ows of solid and gas
descend across the reactor, several processes take place,
namely, moisture evaporation, biomass pyrolysis, char
combustion and gasi"cation, combustion of the gases
and thermal cracking of the tars, in accordance with the
schematic representation reported in Fig. 1. The main
features of submodels applied for these processes can be
derived from Table 2 and the relevant approximations
made are listed below.
Recently, moisture evaporation has been described in
some detail for a single biomass particle (Di Blasi, 1998).
For conditions similar to those of "xed-bed gasi"cation,
it has been shown that both gas-phase and liquid-phase
transport phenomena can play a controlling role. How-
ever, the single-particle e!ects are usually neglected in the
C. Di Blasi / Chemical Engineering Science 55 (2000) 2931}2944 2933
Fig. 1. Schematic of the strati"ed concurrent (downdraft) gasi"er.
Table 2
Moisture evaporation, chemical reactions and transport coe$cients
Evaporation/condensation
m
+
"v
N
k
K
(
T
!
'
`
'
), (16)
p
T
/p
TQ
"exp[(17.884!0.1423
Q
#23.63;10``
Q
)1.0327!67.41
;10`
Q
)"`3]. (17)
Pyrolysis reactions
B
I
N
Pv
A
CHAR#v
%
GAS

#v
2
AR, (p1)
AR
I
N`
PGAS
`
(p2)
R
N
"A
N
exp

!
E
N
R
Q

, (18)
R
N`
"cA
N`
exp

!
E
N`
R
E

2
. (19)
Gas-phase combustion
(tar) CH
```
O
""```
#0.867O
`
I
A
PCO#0.761H
`
O, (c1)
CH
"
#1.5O
`
I
A`
PCO#2H
`
O (c2)
2CO#O
`
I
A`
P2CO
`
, (c3)
H
`
#O
`
I
A"
P2H
`
O, (c4)
R
H
"A
H
exp

!
E
H
R
E

E
C"`
G
C
'
`
, j"c1, c2, i", CH
"
, (20)
R
A`
"A
A`
exp

!
E
A`
R
E

C
''
C"``
'
`
C
'
`
'
, (21)
R
A"
"A
A"
exp

!
E
A"
R
E

C
'
`
C
'
`
. (22)
Gas-phase water gas shift
CO#H
`
O
I
UE
& CO
`
#H
`
, (wg)
R
UE
"ck
UE

C
''
C
'
`
'
!
C
''
`
C
'
`
K
#

, (23)
k
UE
"A
UE
exp

!
E
UE
R
E

, K
#
"A
#
exp

E
#
R
E

. (24)
K
#
"A
#
exp

E
#
R
E

. (25)
Heterogeneous reactions of char
CH
?
O
@
#O
`
I
A`
P

2!2![#
:
2
CO#

2#[!
:
2
!1

CO
`
#:2H
`
O, (c5)
CH
?
O
@
#CO
`
I
E
P2CO#[H
`
O#

:
2
![

H
`
, (g1)
CH
?
O
@
#

2!
:
2
#[

H
`
I
E`
PCH
"
#[H
`
O, (g2)
CH
?
O
@
#(1![)H
`
O
I
E`
PCO#

1![#
:
2
H
`
, (g3)
formulation of reactor models, because the characteristic
times of moisture evaporation are orders of magnitude
shorter than those of char gasi"cation. Therefore, in most
cases the process is assumed to take place instanta-
neously (for instance Yoon, Wei & Denn, 1978) or to be
a di!usion limited process (Hobbs et al., 1993). The latter
treatment is retained in this study with the use of an
empirical expression for the vapour pressure, to take into
account the di!erences in evaporation rates for capillary
and bound water (Di Blasi, 1998) (p
TQ
(
Q
) is the
Clausius}Clapeyron expression).
It is well known that pyrolysis of large biomass par-
ticles is a process controlled by heat and mass transfer
(for instance, Di Blasi, 1996a). More speci"cally, the
conversion times are largely determined by the rate of
inward heat conduction, whereas product distribution is
highly dependent on intra-particle residence times of
volatiles and thus on the extent of secondary cracking.
Given the very steep temperature pro"les along the gasi-
"er axis in the concurrent con"guration, it is often (Bliek,
1984) assumed that intra-particle heat transfer resist-
ances are much lower than extra-particle axial resist-
ances. Thus, the particle is assumed to be thermally thin.
In this study, this treatment is still retained but apparent
2934 C. Di Blasi / Chemical Engineering Science 55 (2000) 2931}2944
Table 2 (Continued)
R
H
"v
N
C
G
(1/k
K
)#(1/k
H
)
, k
H
"A
H
exp

!
E
H
R
Q

, j"c5, g1!g3,
i"O
`
, CO
`
, H
`
, H
`
O (26}27)

r
A
R
`
"(1!,)X#,, (28)
X"
;
Q
;
Q"
, (29)
v
N
(r
A
)"
3(1!c)
r
A
, (30)

A"
c;
Q
cz
"!M
!
(R
A`
#R
E
#R
E`
#R
E`
). (31)
Solid/gas heat transfer coe$cient, mass transfer coe$cient
h
QE
"
2.06c
NE

E
;
E
c
Re"```Pr``, (32)
k
K
"
2.06c
NE
;
E
c
Re"```Sc``. (33)
Properties
zH
E
"cz
E
, zH
Q
"cz
PE
#c
z
Q
(z
Q
/(dz
PQ
)#1.43(1!1.2c))
, (34)
z
PE
"4o0.05`
E
, (35)
z
PQ
"4o0.85`
Q
, (36)
z
Q
"0.0013#0.05(
Q
/1000)#0.63(
Q
/1000)`
J
ms K
(37)
z
E
"4.8;10"'``
E
J
ms K
(38)
j"1.98;10`(
E
/300)``
kg
sm
(39)
Table 3
Devolatilization data: gas composition is expressed as percent of the
initial dry biomass
I II III
v
'
0.33 0.41 0.33
v
'
0.48 0.4 0.5
v
'
0.19 0.19 0.2
Gas 1 CO 7.5% 5.5% 11%
CO
`
13% 10.5% 11%
H
`
1% 0.02% 1%
CH
"
1.5% 1% 2%
H
`
O 25% 23% 25%
Gas 2 CO 9.5% 9.5% 13%
CO
`
5.7% 5.7% 3%
CH
"
3.8% 3.8% 4%
kinetics are used for the pyrolysis process. A one-step
global reaction is considered, where the fractions of
gases, tars and chars (Di Blasi, 1993) produced and the
gas composition should be speci"ed. Tars undergo sec-
ondary cracking in the void spaces of the bed (one-step
global reaction), to produce secondary gases, whose com-
position should again be speci"ed. The size (volume) and
the solid velocity of the spherical particles remain con-
stant as the bed density decreases during moisture evap-
oration and biomass devolatilization. Consequently, the
porosity of this portion of the bed varies. However, this
e!ect is not taken into account as single-particle simula-
tions (Di Blasi, 1997) show that porosity variation exerts
a negligible in#uence on the devolatilization character-
istics.
For primary wood pyrolysis, apparent activation ener-
gies are reported to be in the range 63}125 kJ/mol de-
pending on the wood species, sample size and heating
conditions (Roberts, 1970). Reference values of the ac-
tivation energy and pre-exponential factor are chosen as:
A
N
"1.516;10` s, E
N
"105 kJ/mol, determined
for beech wood cylinders 1}2 cm thick. Kinetic constants
for tar cracking from Liden, Berruti and Scott (1988) are
used. While the composition of the secondary gas has
been estimated on the basis of literature data obtained
for wood (Boroson, Howard, Longwell & Peters, 1989),
an experimental study (Di Blasi, Signorelli, Di Russo
& Rea, 1999) has been carried out to determine
primary pyrolysis product yields and gas composition for
conditions corresponding to downdraft gasi"cation.
Packed-beds (diameter 4 cm) of biomass particles have
been exposed to di!erent radiation intensities, which
correspond to maximum (surface) temperatures in the
bed in the range 650}1000 K. A continuous nitrogen #ow
across the bed reduces the residence times of volatile
products and makes the activity of extra-particle second-
ary reactions negligible. As expected, devolatilization
characteristics, which are representative of intra-particle
primary and secondary reactions, have been found to
depend both on the heating conditions and the biomass
type. Assuming a surface temperature of 850 K, three sets
of devolatilization data ha ve been considered for the
simulation of downdraft gasi"cation (Table 3). The "rst
data set (I) refers to 0.5}1 cm thick wood chips, the
second (II) to rice husks and the third (III) has been
considered for a parametric investigation of the e!ects of
variations in pyrolysis products on the gasi"cation char-
acteristics.
Combustion of volatile products is an important pro-
cess in downdraft gasi"cation. Here, the treatment pro-
posed by Bryden and Ragland (1996) for the "xed-bed
combustion of biomass is adopted. Hydrocarbons, which
include tars and methane, react with oxygen to form
water vapour and carbon monoxide, according to
a "nite-rate, global reaction. The formation of the reac-
tion intermediate, carbon monoxide, is important for the
correct prediction of the ignition delay time. Global,
C. Di Blasi / Chemical Engineering Science 55 (2000) 2931}2944 2935
"nite rate reactions are also considered for the combus-
tion of carbon monoxide and hydrogen. However, for
computational simplicity, instead of an in"nitely fast rate
of hydrogen consumption, a high pre-exponential factor
and a low activation energy have been introduced for
a second-order reaction rate (E"83 kJ/mol, A"
10` m`/s/mol). Tars are modelled as hydrocarbons
CH
```
O
""```
(Bryden & Ragland, 1996), with molecu-
lar weight equal to 95 (Corella, Aznar, Delgado & Aldea,
1991).
As for homogeneous reactions, the water gas shift
reaction is also considered and is described by "nite rate
kinetics, with the equilibrium constant reported by Yoon
et al. (1978) and the kinetic constants derived by Biba,
Macak, Klose and Malecha (1978).
Combustion and gasi"cation reactions of char are het-
erogeneous and are described by the unreacted core,
shrinking particle model, where a steep reaction zone
propagates through the isothermal char. Two mecha-
nisms responsible for the global reaction rate are con-
sidered: di!usion through the gas "lm, surrounding the
particle, and intrinsic chemical kinetics. To account for
the simultaneous e!ects of the two resistances, an e!ec-
tive volumetric reaction rate is introduced, based on the
assumption of a linear dependence of the reaction rate on
the oxidizing/gasifying species concentration. As a conse-
quence of the heterogeneous reactions, the particle dia-
meter shrinks and the density of the bed (and porosity)
remains constant, causing a gradual decrease in the solid
velocity. The minimum size of the particle and conse-
quently the maximum particle density number (v
N
(Eq.
(23))) depend on the ash content of the solid (Hobbs et al.,
1993).
For simplicity, chars are assumed to consist of pure
carbon (:"["0) and the products of heterogeneous
combustion to be carbon dioxide only ("1), though
this assumption is not valid in general (Hobbs et al.,
1993). The intrinsic kinetics of wood char combustion are
described according to the data determined for cellulosic
fuels (Kashiwagi & Nambu, 1992). Kinetics of char gasi"-
cation by carbon dioxide and steam are those reported
by Groeneveld and van Swaaij (1980). Hydrogasi"cation
is usually negligible under atmospheric conditions and its
rate is assumed to be slower by three orders of magnitude
compared with carbon dioxide and steam gasi"cation
(Hobbs et al., 1993).
Literature correlations are used for the e!ective ther-
mal conductivity of the bed (Goldman, Xieu, Oko, Milne
& Essenhigh, 1984), the e!ective bed-to-wall heat transfer
coe$cients (Hobbs et al., 1993) (not reported in Table 2),
the solid/gas heat transfer and the mass transfer coe$-
cients (Gupta & Thodos, 1963). However, the solid/gas
heat transfer coe$cient estimated from non-reacting
system data can exceed the experimental values in
reacting gasi"ers (Hobbs et al., 1993), as a consequence
of unsteady heat transfer. Therefore, the experimental
correlation is multiplied by empirical factors () with
values in the range 0.02}1 (Cho & Joseph, 1981;
Radulovic, Ghani & Smoot, 1995). In this study the
reference value for is taken equal to 1. Some di$culties
also exist in the application of the Gupta and Todos
correlation for the mass transfer coe$cient, mainly be-
cause of the di!erent scales of the conversion units (in
particular, there is large uncertainty for low Reynolds
numbers) and again the changes introduced by chemical
reactions. Apart from the correlation suggested by Bhat-
tacharya, Salam, Dudukovic and Joseph (1986), in all
cases the mass transfer coe$cient decreases as the par-
ticle size is increased, but di!erences become very high as
the particle diameter reduces to very low values. This is
critical, because in the unreacted-core, shrinking particle
model, the maximum temperature predicted in the oxida-
tion/reduction zone becomes highly dependent on the
correlation used. On the other hand, at high temper-
atures, "lm transfer becomes the controlling resistance.
To avoid unrealistic temperature values, corrective fac-
tors, which limit the maximum value of k
K
(Bhattacharya
et al., 1986) or reduce its value for all conditions (for
instance, 0.5 in Goldmann et al., 1984) have been sugges-
ted. Both procedures have been applied to the Gupta and
Todos correlation, used in this model. Given the proper
values of the correction factor or its maximum allowed in
the course of the reaction, the two procedures do not
result in any change from the qualitative point of view in
the predictions of process characteristics. Therefore, only
the e!ects associated with the changes in the maximum
value of k
K
(indicated as kH
K
) on the prediction of the
gasi"cation process will be discussed. The reference value
chosen for kH
K
is 0.045 m/s.
The variation of the gas thermal conductivity and
viscosity with temperature is described as in Purnomo,
Aerts and Ragland (1990). A constant value is assumed
for di!usivities (0.2;10" m`/s) and speci"c heats of
gaseous species (evaluated for a temperature of 1000 K)
and speci"c heats of biomass and char (1.34 kJ/kg) (Di
Blasi, 1996b). All the thermochemical data referred to or
listed in this section, with the inclusion of the devolatiliz-
ation data set I, will be indicated in the following as
reference data (a summary of kinetic constants is re-
ported in Table 4).
The numerical solution of the model equations is
based on operator splitting procedures and "nite-di!er-
ences approximations. The reactor is divided into a set of
elementary cylindrical cells, whose cross sections co-
incide with the reactor cross section, while the height can
be variable. The solution procedure is divided into three
stages, corresponding to chemical reaction processes,
heat exchange (between phases and with the reactor wall)
and transport phenomena. For each time step, in the "rst
two stages, the solution is calculated, for each control
volume, of linearized ordinary di!erential equations, by
means of a "rst-order implicit Euler method. In the third
2936 C. Di Blasi / Chemical Engineering Science 55 (2000) 2931}2944
Table 4
Reference value for kinetic constants
Pyrolysis
A
NJ
(s) 1.516;10` Roberts (1970)
E
NJ
(kJ/mol) 105 Roberts (1970)
A
N
`
(s) 4.28;10' Liden et al. (1988)
E
N
`
(kJ/mol) 107 Liden et al. (1988)
Ah
N

(kJ/kg) !420 Di Blasi (1996b)


Ah
N
`
(kJ/kg) 42 Di Blasi (1996b)
Char combustion
A
A
`
(s) 5.67;10" Kashiwagi and Nambu (1992)
E
A
`
(kJ/mol) 160 Kashiwagi and Nambu (1992)
Ah
A`
(kJ/kg) 2.5;10` Kashiwagi and Nambu (1992)
Char gasi"cation
A
E
"A
E`
(m`/mol s) 7.92;10" Groeneveld and van Swaaij (1980)
A
E`
(m`/mol s) 79.2 Groeneveld and van Swaaij (1980)
E
E
"E
E`
"E
E`
(kJ/mol) 218 Groeneveld and van Swaaij (1980)
Ah
E
(kJ/kg) !9.3;10' Groeneveld and van Swaaij (1980)
Ah
E`
(kJ/kg) 7.2;10' Groeneveld and van Swaaij (1980)
Ah
E`
(kJ/kg) !6.4;10' Groeneveld and van Swaaij (1980)
Water-gas shift
A
#
(!) 0.0265 Yoon et al. (1978)
E
#
(kJ/mol) 65.8 Yoon et al. (1978)
A
UE
(m`/mol s) 2.78 Biba et al. (1978)
E
UE
(kJ/mol) 12.6 Biba et al. (1978)
Ah
UE
(kJ/mol) 41 Biba et al. (1978)
Gas-phase combustion
A
A
"A
A`
((m`/mol)"`/sK) 9.2;10' Bryden and Ragland (1996)
E
A
"E
A`
(kJ/mol) 80 Bryden and Ragland (1996)
Ah
A
(kJ/kg) 17 473 Bryden and Ragland (1996)
Ah
A`
(kJ/kg) 50 190 Bryden and Ragland (1996)
A
A`
((m`/mol)"``/sK) 10`' Bryden and Ragland (1996)
E
A`
(kJ/mol) 166 Bryden and Ragland (1996)
Ah
A`
(kJ/kg) 10 107 Bryden and Ragland (1996)
A
A"
((m`/s mol) 1;10 Bryden and Ragland (1996)
E
A"
(kJ/mol) 42 Bryden and Ragland (1996)
Ah
A"
(kJ/kg) 142919 Bryden and Ragland (1996)
Estimated.
step the transport equations, after discretization with the
hybrid scheme, are solved through a semi-implicit pro-
cedure, that is, each conservation equation is implicit in
the corresponding variable being conserved, whereas the
other variables are taken as the last available values. For
all stages, the equations for the condensed-phase vari-
ables are solved "rst, followed by those of gaseous com-
ponents and enthalpies. All the simulations presented
here were carried out with a time step of 1;10" s and
space steps of 0.001 m, giving grid-independent solutions
for the reference values of the process parameters.
3. Results
Simulations have been carried out for biomass par-
ticles 1 cm thick, with an ash content of 10.5% and an
initial moisture content of 10% d.b. (dry basis). A pilot
scale unit (ID 0.45 and 0.50 m high (Manurung &
Beenackers, 1994) is considered, where the bed density is
200 kg/m` and the void fraction is 0.5. Ignition of the bed
is caused by hot air. A steady-state scenario (time equal
zero is referred to this initial condition), corresponding to
a reaction zone located near the top of the reactor, has
been chosen as initial condition for the simulations by
varying model parameters (, kH
K
, devolatilization data),
kinetic constants (pyrolysis, gasi"cation, combustion)
and the two operational variables: the biomass feed rate
and the air-to-biomass ratio. A"rst set of simulations has
been carried out by varying the biomass feed rate (=
@
) in
the range 7}18 kg/h for an air-to-fuel ratio of 1.5. The
e!ects of the air-to-fuel ratio (=
?
/=
@
) have also been
investigated for variations in the range 0.6}2.6 for a bi-
omass feed rate of 18 kg/h. In this section, the e!ect of
C. Di Blasi / Chemical Engineering Science 55 (2000) 2931}2944 2937
Fig. 2. A. Axial pro"les of solid and gas temperature and gas velocity
(reference data, =
@
"18 kg/h, =
?
/=
@
"1.5). B. Axial pro"les of spe-
cies molar fractions (reference data, =
@
"18 kg/h, =
?
/=
@
"1.5).
Fig. 3. A. Structure of the reaction front: oxygen molar fraction, char
mass fraction and temperature pro"les (reference data, =
@
"18 kg/h,
=
?
/=
@
"1.5). B. Structure of the reaction front: axial pro"les of
reaction rates (reference data, =
@
"18 kg/h, =
?
/=
@
"1.5).
these parameters on the process characteristics are dis-
cussed. Finally, a comparison between model predictions
and experimental measurements is given.
The main characteristics of the gasi"cation process can
be observed from Figs. 2 and 3, which report the axial
pro"les of the main variables of interest at steady condi-
tions for =
@
"18 kg/h and =
?
/=
@
"1.5. The leading
edge of the reaction zone is stabilized slightly below the
feeding section (about 2 cm), so that the gasi"er is top
stabilized. The air and biomass co-fed at the top of the
reactor constitute a non-reactive fuel/oxidant mixture,
which makes the #ows homogeneous, providing a uni-
form air/fuel distribution to the zones below. This region
is followed by the evaporation of moisture and the partial
charring of the biomass particles, essentially because of
radiative/conductive heat transfer through the solid-
phase. Indeed, oxygen soon meets the steep char front,
causing its exothermic combustion and a large increase
in the gas velocity. As a consequence of the increase in the
solid-phase temperature, the gasi"cation rates also attain
a local maximum. However, the endothermicity of this
process and the convective cooling of the inlet gas
quench the heterogeneous consumption of char. Then air
mixes with volatile products, downstream of the charring
front, resulting in a pre-mixed #ame. Here, when the
gaseous mixture attains temperatures su$ciently high,
combustion of volatile pyrolysis products becomes an
important mechanism to provide heat for further char
gasi"cation, thus increasing the hydrogen and carbon
monoxide content of the gas. Consequently, the particle
size and the solid velocity continuously decrease,
whereas the particle density number increases. Homo-
geneous, gas-phase combustion partially destroys the
tars and, through high temperatures, also favours their
cracking.
It should be noted that the moisture evaporation front
is very steep, probably as a consequence of the highly
simpli"ed description (di!usional resistance only), which
does not include intra-particle temperature and moisture
gradients. The endothermicity of the process has been
found to contribute signi"cantly to limit the maximum
temperature. Biomass devolatilization is also character-
ized by spatial temperature and solid-phase species
gradients much larger than those observed for the up-
draft con"guration. Indeed, the concurrent solid and gas
#ow towards the combustion zone makes the heat trans-
fer di$cult towards the virgin solid region. Di!erences
between the solid and gas temperatures are signi"cant at
2938 C. Di Blasi / Chemical Engineering Science 55 (2000) 2931}2944
Fig. 4. Conversion characteristics (particle volume, solid velocity and
particle density number pro"les) for several biomass feed rates (refer-
ence data, =
?
/=
@
"1.5).
Fig. 5. Composition of the (dry) gas and yield of unreacted char (as
% of the amount formed from pyrolysis) for several biomass feed rates
(reference data, =
?
/=
@
"1.5).
Fig. 6. Axial temperature pro"les for several times (reference data,
=
@
"11 kg/h, =
?
/=
@
"1).
the leading edge of the reaction zone, but they tend to
disappear in the reduction zone. Among the gasi"cation
products, the most abundant is CO, followed by H
`
and
CH
"
. The tar content of the gas is very small (0.2% of
the mass of the gaseous e%uents), while H
`
O and CO
`
are still present in signi"cant amounts. On the whole, the
processes of primary and secondary pyrolysis and the
homogeneous and heterogeneous combustion reactions
take place simultaneously, with high rates mainly along
a distance of about 0.2 m.
The main characteristics of the process remain un-
altered as the biomass feed rate is varied in the range
7}18 kg/h. However, as =
@
is decreased, the charring
front becomes successively steeper and the amount of
unreacted carbon discharged at the grate lower. The
di!erences in the temperature pro"les are small, but in
the "rst part of the reacting bed the gas temperature is
successively higher (as =
@
is decreased). The particle and
conversion characteristics (axial, steady pro"les of par-
ticle volume, solid velocity and particle density number)
and the gas composition are shown in Figs. 4 and
5 (=
?
/=
@
"1.5) for di!erent biomass feed rates. It can
be seen that complete particle burn-out is attained for
=
@
)7 kg/h. Given the better conversion, the quality of
the gas slightly improves as =
@
is decreased, with a tar
content that goes to zero for =
@
)11 kg/h.
Signi"cant changes are introduced in process dynam-
ics, structure of the reaction front and gas quality if the
air-to-fuel ratio is varied. For the reference values of the
thermochemical properties, kinetic constants and trans-
port coe$cients, a top-stabilized process is possible only
for =
?
/=
@
'1.2 (for =
@
)19 kg/h). For lower values,
both homogeneous and heterogeneous combustion rates
do not attain su$ciently high values for the stabilization
of the reaction front. The two processes appear to be
closely connected and give rise to a front propagating
with a constant speed towards the bottom of the gasi"er,
where eventually extinction takes place. An example of
this situation is shown in Fig. 6, through the axial
temperature pro"les, simulated for several times,
(=
?
/=
@
"1 and =
@
"11 kg/h). The lower the air-to-
fuel ratio, the faster the propagation rate of the reaction
front and the shorter the extinction time. As an increase
in the biomass feed rate is associated with an increase in
the rate of solid (char and ash) discharge at the bottom of
the gasi"er, high values of =
@
also result in an extinction
process with dynamics qualitatively similar to those pre-
sented in Fig. 6 (for instance, for =
@
"20 kg/h and
=
?
/=
@
"1.4).
For a top-stabilized reactor, as the air-to-fuel ratio is
increased, no signi"cant changes are observed at the
leading edge of the reaction zone. This is understandable
as the characteristics of this zone are determined essen-
tially by the rate of moisture evaporation and primary
pyrolysis kinetics, which are not varied. However, the
size of the reaction zone becomes successively narrower
and the homogeneous combustion more favoured, giving
rise to signi"cant temperature increases. It should be
C. Di Blasi / Chemical Engineering Science 55 (2000) 2931}2944 2939
Fig. 7. Conversion characteristics (particle volume, solid velocity and
particle density number pro"les) for several air-to-fuel ratios (reference
data, =
@
"18 kg/h).
Fig. 8. Composition of the (dry) gas and yield of unreacted char (as
% of the amount formed from pyrolysis) as functions of the air-to-
biomass ratio (reference data, =
@
"18 kg/h: solid lines for devolatiliz-
ation data I and dashed lines for devolatilization data III).
Fig. 9. Structure of the reaction front: axial pro"les of reaction rates and
temperature for the reference data, =
@
"18 kg/h, =
?
/=
@
"2.4 (dashed
lines: devolatilization data I, solid lines: devolatilization data III).
Fig. 10. Structure of the reaction front: axial pro"les of combustion
rates and temperature for the reference data, =
@
"18 kg/h,
=
?
/=
@
"2 (dashed lines: E
N
"105 kJ/mol, solid lines: E
N
"
84 kJ/mol).
noted that the maximum temperature remains almost
constant as long as there is carbon available for en-
dothermic gasi"cation. As shown by Fig. 7, complete
burn-out of the solid particles is achieved for values of
=
?
/=
@
slightly above 1.75, corresponding to a max-
imum in the rate of char consumption and a slow com-
bustion rate of volatile species. The tar content of the gas
decreases as the air-to-fuel ratio is increased, from about
0.4% (=
?
/=
@
"1.3) to 0.1% (=
?
/=
@
"1.75) of the
total mass of the gaseous e%uents and goes to zero for
=
?
/=
@
*2.
The gas composition and the char yield discharged at
the grate are reported in Fig. 8 as functions of the air-to-
fuel ratio for the reference data and the pyrolysis charac-
teristics I (solid lines) and III (dashed lines) (=
@
"
18 kg/h). As expected, the contents of hydrogen and
carbon monoxide initially increase and then decrease, as
a consequence of the improved gasi"cation e$ciency and
the enhanced combustion rate of the combustible gas,
respectively. It can be seen that, though from the quali-
tative point of view the same trend in the producer gas
quality is predicted for both the devolatilization data
I and III, the quantitative di!erences are signi"cant. Of
course the quality of the producer gas is better when
higher amounts of carbon monoxide are assumed to be
formed from both primary and secondary reactions. The
higher concentration of combustible gas also result
in larger combustion rates and in higher temperatures
(Fig. 9).
The e!ects of the apparent pyrolysis rate on the predic-
tions of the gasi"cation process are particularly impor-
tant, because the rate of charring largely determines the
structure of the leading edge of the reaction zone. A com-
parison between the structures of the reaction zone for
two di!erent activation energies can be made through
Fig. 10 (kH
K
"0.06 m/s, =
@
"11 kg/h, =
?
/=
@
"2). The
increase in the pyrolysis rate is associated with an in-
crease in the rate of heterogeneous combustion and the
2940 C. Di Blasi / Chemical Engineering Science 55 (2000) 2931}2944
Fig. 11. Axial temperature pro"les for several times (reference data,
devolatilization data III, E
N
"63 kJ/mol, kH
K
"0.08 m/s, no methane
combustion, heterogeneous reaction rates increased by a factor of 2,
=
@
"9.5 kg/h, =
?
/=
@
"1.1).
Fig. 12. Conversion characteristics (particle volume, solid velocity and
particle density number pro"les) for several times (reference data, de-
volatilization data III, E
N
"63 kJ/mol, kH
K
"0.08 m/s, no methane
combustion, heterogeneous reaction rates increased by a factor of 2,
=
@
"9.5 kg/h, =
?
/=
@
"1.1).
presence of a successively steeper charring front. The
maximumtemperature at the leading edge of the reaction
zone slightly decreases, as a consequence of variation in
the relative position of the exothermic homogeneous
combustion zone and the exothermic/endothermic
char consumption zone. Indeed, two distinct maxima
appear in the solid-phase temperature pro"le, corre-
sponding to heterogeneous and homogeneous combus-
tion. Between the two, there is a minimum caused by
the convective cooling of the air fed at the top of
the reactor and the endothermicity of the gasi"cation
reactions. The large di!erences between the heterogen-
eous and the homogeneous combustion rates, established
in the presence of low activation energies of the primary
pyrolysis, appear to be the key feature for a grate-stabil-
ized gasi"er.
Besides extinction (low air-to-fuel ratio) and a top-
stabilized reactor (high air-to-fuel ratio), another interest-
ing behaviour is simulated for intermediate values of the
air-to-fuel ratio, as shown through Figs. 11 and 12. These
simulations have been obtained for E
N
"63 kJ/mol,
kH
K
"0.08 m/s, rates of heterogeneous reaction increased
by a factor of 2, no methane combustion and devolatiliz-
ation data III. For values of =
?
/=
@
in the range 1}1.2,
associated with the prompt formation of char, heterogen-
eous combustion also occurs at some extent near the top
of the reactor. However, because of the reduced temper-
atures, the subsequent gas-phase combustion and char
gasi"cation zones move downward. After unsteady
propagation, this front stabilizes at a certain distance
(0.22}0.20 m) from the grate, where both homogeneous
and heterogeneous reactions take place to some extent.
This mode of stabilization, observed also in the experi-
ments by Reed and Markson (1985) and indicated as
grate-stabilization, is not very e!ective. Indeed, because
of the reduced size of the homogeneous combus-
tion/gasi"cation zone, the tar content of the gas is high
and char conversion is low. It is important to point out
that a grate-stabilized process is simulated only if a sig-
ni"cant axial temperature gradient is established, so that
a temperature su$ciently high for the stabilization of the
reaction front is attained at a certain distance from the
feed section. In the absence of such a gradient and for
low air-to-fuel ratios extinction takes place (for instance,
Fig. 6).
The limit value of the air-to-fuel ratio for a top stabil-
ized reactor is a!ected by the intrinsic kinetics of the
heterogeneous reactions. Thus, it moves to lower values
as the activation energy of the combustion reaction is
decreased and/or the activation energy of the gasi"cation
reactions is increased. Activation energies of the combus-
tion reaction in the range 117}159 kJ/mol do not give rise
to variations in the characteristics of the gasi"cation
process, given a su$ciently high value of the air-to-fuel
ratio (*1.2). However, this parameter in#uences the
dynamic behaviour of the reaction front (stabilization)
for low air-to-fuel ratios. Activation energies of the gasi"-
cation reactions below the reference value again do not
result in signi"cant changes in the gasi"cation process.
However, higher values are associated with higher tem-
peratures, given the reduced activity of endothermic char
gasi"cation.
The maximum value of the e!ective mass transfer
coe$cient, kH
K
, and the solid/gas heat transfer coe$cient
(factor ) exert a signi"cant in#uence on the structure of
the reaction zone. Fig. 13 shows the conversion charac-
teristics for three values of kH
K
(=
?
/=
@
"2, =
@
"
11 kg/h, reference data). For very high kH
K
values, the
heterogeneous reactions predominate: particle burnout
and oxygen consumption take place along a very thin
region, the solid temperature presents a very sharp peak
at the leading edge of the reaction zone (char combus-
tion) followed by a rapid decay (char gasi"cation). As
C. Di Blasi / Chemical Engineering Science 55 (2000) 2931}2944 2941
Fig. 13. Conversion characteristics (particle volume, solid velocity and
particle density number pro"les) for three values of kH
K
(reference data,
=
@
"11 kg/h, =
?
/=
@
"2).
Table 5
Comparison between the predicted and measured composition of the producer gas
% vol Groeneveld and van Swaaij (1980) Walawender, Chern and Fan (1985) Wang and Kinoshita (1994) This study
CO 17 18}20 18}20 20.3}18.5
H
`
14 12}16 10}14 16.8}9.8
CO
`
13.6 14}16 9.6}11 15.3}9.4
CH
"
0.9 2.5 2.5}4.8 4.5}2.4
N
`
46.5 53}45 52}54 43}60
=
?
/=
@
* 1}2 * 1.4}2.2
carbon dioxide is the only product of char combustion
(the predominating reaction) and the rather thin zone of
high temperature does not allow the tars to be destroyed,
the quality of the gas is poor (low contents of carbon
monoxide and hydrogen and high amounts of tars).
As kH
K
is reduced to very low values, heterogeneous com-
bustion/gasi"cation processes become controlled by
chemical kinetics, so that the rates of heterogeneous
reactions become comparable or slower than those of
the gas-phase combustion. Indeed, for low values of the
maximum k
K
, oxygen is burned mainly by volatile prod-
ucts, the reaction zone enlarges, tars are completely
cracked (or burned) and the quality of the gas is good.
The partial combustion of the volatile products also
results in a wide region of high temperature, though the
peak at the leading edge of the reaction is slightly re-
duced. Variations in the solid/gas heat transfer coe$-
cient, through the factor , result in changes in the
reaction front and gasi"cation characteristics qualitat-
ively similar to those discussed in relation to kH
K
.
Another important point is the comparison between
the model predictions and the experimental measure-
ments. It is well known (Britten, 1988) that the predic-
tions of maximum temperature and other parameters,
such as extension of the combustion and gasi"cation
zone, are signi"cantly a!ected by parameter values,
which are not generally well known, in particular heat
and mass transfer coe$cients. However, it is also well
known (Amundson & Arri, 1978, Yoon et al., 1978;
Hobbs et al., 1993) that overall conversion rates and
e%uent gas composition are remarkably insensitive to
model details. Indeed, in agreement with previous "nd-
ings, it has been found that the gas composition is mainly
determined by the composition of the pyrolysis gas and
the equilibrium conditions of the water gas shift reaction.
As can be seen from Table 5, the use of reference data and
the set of devolatilization data I (wood chips) give rise to
predictions of producer gas composition very close to the
values reported for woody biomass by several references.
It is believed that the agreement can be further improved
by taking into account the variation in the product
distribution and gas composition with the reaction tem-
perature.
Contrary to the case of gas composition, the literature
does not report extensive measurements of the temper-
ature pro"le in downdraft gasi"ers. Few exceptions are
the two-dimensional temperature measurements, made
available by Wang and Kinoshita (1994) for a con-
ventional (throated) gasi"er and the measurements by
Manurung and Beenackers (1994) for a strati"ed, rice
husk downdraft gasi"er. This second case is considered
for comparison. With respect to the reference data, the
initial particle diameter has been chosen as 0.5 cm, the
initial ash content as 16% and the devolatilization char-
acteristics, indicated as data set II, have been used. Fur-
thermore, the activation energy of the apparent primary
pyrolysis degradation rate has been taken as 63 kJ/mol
because a reduction in the particle size should be asso-
ciated with a faster degradation rate (no indication is
currently available on the apparent degradation kinetics
of rice husks at high temperatures). Also, the value of
kH
K
has been taken equal to 0.02 m/s. In agreement with
experimental observation (Manurung & Beenackers,
1994), optimal gasi"cation conditions, in terms of ther-
mal e$ciency, are simulated for intermediate values
(about 1.7) of the air-to-fuel ratio, when the char conver-
sion is high and the calori"c value of the producer gas is
at the maximum (molar composition consisting of about
20% carbon monoxide, 15% hydrogen, 13% carbon
dioxide and minor fractions of methane). Measured and
2942 C. Di Blasi / Chemical Engineering Science 55 (2000) 2931}2944
Fig. 14. Comparison between axial temperature pro"les as predicted
by the model (solid and dashed lines) and measured by Manurung and
Beenackers (1994) (symbols) for rice husk gasi"cation (=
@
"10.5 kg/h,
=
?
/=
@
"1.3).
predicted temperature pro"les (Fig. 14) show some dis-
agreement mainly at the bottom of the gasi"er, probably
caused by the absence of grate heat losses in the model.
Also, it should be noted that the predicted temperature
pro"le is signi"cantly dependent on the fraction of wall
heat losses, f, retained from the correlation proposed for
industrial gasi"er (Hobbs et al., 1993), a parameter which
cannot be assigned a priori. However, on the whole, the
agreement is acceptable.
4. Conclusions and further developments
A mathematical model has been formulated, which
includes the description of all the main chemical and
physical processes taking place during the "xed-bed
downdraft gasi"cation of lignocellulosic fuels. Probably,
the most innovative aspects of the model, with respect to
the state of the art, are represented by the description of
the #aming pyrolysis process through "nite rate kinetics
of primary (apparent) pyrolysis, secondary cracking of
tars and combustion of carbon monoxide, hydrogen, tars
and methane. These features, coupled with the formula-
tion of transient equations for heat and mass transport
across the bed, have allowed the key characteristics of
downdraft gasi"cation to be analysed. In particular, nu-
merical simulation has been applied to investigate the
e!ects of the two most important operational variables,
i.e. biomass feed rate and air-to-fuel ratio, on process
dynamics, composition and quality of the producer gas
and overall e$ciency of the process.
The model predictions reproduce well, from the quali-
tative point of view, the dynamic behaviour and the
steady-state con"gurations, on dependence on the air/
fuel feed rate, of downdraft wood gasi"ers. Gas-phase
combustion and primary pyrolysis appear to play a con-
trolling role for the mode of stabilization of the reaction
front, that is, for a top- or grate-stabilized reactor. From
the quantitative point of view, the description of the
devolatilization stage (product yields and gas composi-
tion) and the correlations/values of the transport coe$-
cients are crucial points. A `reasonablea selection for
these variables leads to good quantitative agreements in
terms of gas compositions, but marginal agreement is still
shown for the temperature pro"les. Model validation has
always been carried out only to a limited extent in "xed-
bed gasi"er modelling (Hobbs et al., 1993), because of the
very few experimental results available in the literature
(often, not enough information is reported for compari-
son with model predictions). This is particularly true for
the downdraft gasi"cation of wood. Therefore, model
validation should be further addressed, possibly in con-
junction with an adequate experimental program.
There are numerous other aspects which need further
investigation in the concurrent wood gasi"cation. The
comprehensive model presented in this study has been
applied only for a pilot-scale gasi"er fed with a "xed-
property lignocellulosic biomass, by varying the operat-
ing conditions (air and/or biomass feed rate). Also, the
in#uences of few key parameters (primary gas composi-
tion, adjustable factors in the heat and mass transfer
coe$cients and pyrolysis kinetic constants) have been
examined only for a narrow range of values, in order to
assess their role in the process dynamics. A successive
study should provide extensive sensitivity analyses for
the model parameters, the physico-chemical properties of
feedstocks and the plant size. Further development can
also be important in relation to the accuracy of the
di!erent process submodels. For instance, though the
submodel for the primary pyrolysis process, proposed in
this study, is more accurate than those currently applied
in biomass gasi"er modelling, it is believed that single-
particle e!ects in the description of the devolatilization
stage during gasi"cation deserve a more accurate treat-
ment. Indeed, while the assumption of a thermally thin
particle was found to work well for coal gasi"cation,
where the amount of volatiles released is relatively low,
for the present case it may result in a too rapid biomass
conversion and wrong predictions of the position of the
reaction front in strati"ed downdraft reactors or in incor-
rect evaluation of the size of the pyrolysis zone in updraft
reactors.
Single-particle e!ects in the description of primary
pyrolysis and moisture evaporation are also important
for a correct prediction of the tar evolution and thus for
the quality of the producer gas. Indeed, the simulations
have shown that the maximum temperature at the lead-
ing edge of the reaction zone is highly a!ected by the rate
of moisture evaporation. Too low temperatures may
cause that incompletely devolatilizated char enters the
reduction zone, tars are produced there and remain un-
converted in the gas. Also, low temperatures may cause
incomplete conversion of tars in the oxidation zone.
C. Di Blasi / Chemical Engineering Science 55 (2000) 2931}2944 2943
It should be noted that a correct prediction of the
tar content of the gaseous e%uents is important for an
e!ective design of the gasi"er, the selection of the optimal
operation condition and the choice of the most adequate
gas cleaning procedure.
Another important factor for the conversion process is
the char reactivity (gasi"cation and combustion), which
in the model is described by the unreacted core model, on
the basis of the e!ective mass transfer coe$cient and the
intrinsic kinetics. Critical points in this approach, which
needs further investigation, are represented by the in"-
nitely thin reaction zone and by the absence of the char
concentration in the rates of the gasi"cation reactions.
Finally, more reliable input data are needed for the
simulation of the gasi"cation process, in relation to both
transport coe$cients and intrinsic reaction kinetics. In-
deed, these have been, for large part, investigated under
thermogravimetric conditions, which do not reproduce
the true gasi"cation conditions.
Acknowledgements
The research was funded in part by the European
Commission in the framework of the Non Nuclear En-
ergy Programme (JOULE III), Contract JOR3-CT95-
0021.
References
Amundson, N. R., & Arri, L. E. (1978). Char gasi"cation in a counter-
current reactor. A.I.Ch.E. Journal, 24, 87.
Bhattacharya, A., Salam, L., Dudukovic, M. P., & Joseph, B. (1986).
Experimental and modelling studies in "xed-bed char gasi"cation.
Industrial & Engineering Chemical Process Design and Development,
25, 988.
Biba, V., Macak, J., Klose, E., & Malecha, J. (1978). Mathematical
model for the gasi"cation of coal under pressure. Industrial & Engin-
eering Chemical Process Design and Development, 17, 92.
Bliek, A. (1984). Ph.D. thesis, Mathematical modelling of a cocurrent xxed
bed coal gasixer. The Twente University of Technology, The Neth-
erlands.
Boroson, M. L., Howard, J. B., Longwell, J. P., & Peters, A. W. (1989).
Products yields and kinetics from the vapor phase cracking of wood
pyrolysis tars. A.I.Ch.E. Journal, 35, 120.
Britten, J. A. (1988). Extinction phenomena in countercurrent packed-
bed coal gasi"ers: A simple model for gas production and char
conversion rates. Industrial & Engineering Chemistry Research, 27,
197.
Bryden, K. M., & Ragland, K. (1996). Numerical modeling of a deep,
"xed-bed combustor. Energy & Fuels, 10, 269.
Buekens, A. G., Bridgwater, A. V., Ferrero, G. L., & Maniatis, K. (1990).
Commercial and marketing aspects of gasi"ers. Commission of the
European Communities, EUR 12736.
Cho, Y. S., & Joseph, B. (1981). Heterogeneous model for moving-bed
coal gasi"cation reactors. Industrial & Engineering Chemical Process
Design and Development, 20, 314.
Corella, J., Aznar, M. P., Delgado, J., & Aldea, E. (1991). Steam
gasi"cation of cellulosic wastes in a #uidized bed with downstream
vessels. Industrial & Engineering Chemistry Research, 30, 2252.
Di Blasi, C. (1993). Modeling and simulation of combustion processes
of charring and non-charring solid fuels. Progress in Energy and
Combustion Science, 19, 71.
Di Blasi, C. (1996a). Kinetic and heat transfer control in the slow and
#ash pyrolysis of solids. Industrial & Engineering Chemistry Re-
search, 35, 37.
Di Blasi, C. (1996b). Heat, momentum and mass transfer through
a shrinking biomass particle exposed to thermal radiation. Chemical
Engineering Science, 51, 1121.
Di Blasi, C. (1997). In#uences of physical properties on biomass de-
volatilization characteristics. Fuel, 76, 957.
Di Blasi, C. (1998). Multi-phase moisture transfer in the high-temper-
ature drying of wood particles. Chemical Engineering Science, 53,
353.
Di Blasi, C., Signorelli, G., Di Russo, C., & Rea, G. (1999). Product
distribution from pyrolysis of wood and agricultural residues. Indus-
trial & Engineering Chemistry Research, 38, 2216.
Goldman, J., Xieu, D., Oko, A., Milne, R., & Essenhigh, R. H. (1984).
A comparison of predictions and experiment in the gasi"cation of
anthracite in air and oxygen-enriched/steam mixtures. 20th Interna-
tional Symposium on Combustion. The Combustion Institute, Pit-
tsburgh, p. 1365.
Groeneveld, M. J., & van Swaaij, W. P. M. (1980). Gasi"cation of
char particles with CO
`
and H
`
O. Chemical Engineering Science,
35, 307.
Gupta, A. S., & Thodos, G. (1963). Direct analogy between mass and
heat transfer to beds of spheres. A.I.Ch.E. Journal, 9, 751.
Hobbs, M. L., Radulovic, P. T., & Smoot, L. D. (1993). Combustion and
gasi"cation of coal in "xed-beds. Progress in Energy and Combustion
Science, 19, 505.
Kashiwagi, T., & Nambu, H. (1992). Global kinetic constants for
thermal oxidative degradation of a cellulosic paper. Combustion and
Flame, 88, 345.
Liden, A. G., Berruti, F., & Scott, D. S. (1988). A kinetic model for the
production of liquids from the #ash pyrolysis of biomass. Chemical
Engineering Communications, 65, 207.
Manurung, R. K., & Beenackers, A. A. C. M. (1994). Modeling and
simulation of an open-core downdraft moving bed rice husk gasi"er.
Advances in Thermochemical Biomass Conversion. London: Blackie
A & P, p. 288.
Purnomo, Aerts, D. J., & Ragland, K. W. (1990). Pressurized downdraft
combustion of wood chips, Twenty-third International Symposium on
Combustion. The Combustion Institute (pp. 1025}1032).
Reed, T. B., & Markson, M. (1985). Biomass gasi"cation reaction
velocities. In R. P. Overend, T. A. Milne, L. K. Mudge, Funda-
mentals of Thermochemical Biomass Conversion (p. 951). Amsterdam:
Elsevier.
Radulovic, P. T., Ghani, M. U., & Smoot, L. D. (1995). An improved
model for "xed-bed coal combustion and gasi"cation. Fuel, 74, 582.
Roberts, A. F. (1970). A review of the kinetic data of the pyrolysis of
wood and related substances. Combustion and Flame, 14, 261.
Yoon, H., Wei, J., & Denn, M. M. (1978). A model for moving bed coal
gasi"cation reactors. A.I.Ch.E. Journal, 24, 885.
Walawender, W. P., Chern, S. M., & Fan, L. T. (1985). Wood chip
gasi"cation in a commercial downdraft gasi"er. In R. P. Overend, T.
A. Milne & L. K. Mudge, Fundamentals of Thermochemical Biomass
Conversion (p. 911). Amsterdam: Elsevier.
Wang, Y., & Kinoshita, C. M. (1994). Temperature "elds in downdraft
biomass gasi"cation, Proceedings of the International Conference on
Advances in Thermochemical Biomass Conversion (p. 280). London:
Blackie A & P.
2944 C. Di Blasi / Chemical Engineering Science 55 (2000) 2931}2944

También podría gustarte