Sunteți pe pagina 1din 11

CFD simulation of a chemical-looping fuel reactor utilizing solid fuel

Kartikeya Mahalatkar
b,c,n
, John Kuhlman
a,b
, E. David Huckaby
a
, Thomas OBrien
a
a
National Energy Technology Lab., 3610 Collins Ferry Rd., Morgantown, WV 26507, USA
b
West Virginia University, Department of Mechanical and Aerospace Engineering, Morgantown, WV 26506, USA
c
ANSYS Inc., 3647 Collins Ferry Rd., Suite A, Morgantown, WV 26505, USA
a r t i c l e i n f o
Article history:
Received 5 May 2010
Received in revised form
9 April 2011
Accepted 18 April 2011
Available online 28 April 2011
Keywords:
Computational uid dynamics
Chemical reactors
Multiphase reactions
Ilmenite
Fluidization
Numerical analysis
a b s t r a c t
A computational uid dynamic (CFD) study has been carried out for the fuel reactor for a new type of
combustion technology called chemical-looping combustion (CLC). CLC involves combustion of fuels by
heterogeneous chemical reactions with an oxygen carrier, usually a granular metal oxide, exchanged
between two reactors. There have been extensive experimental studies on CLC, however CFD
simulations of this concept are quite limited. In the present paper we have developed a CFD model
for the fuel reactor of a chemical-looping combustor described in the literature, which utilized a
Fe-based carrier (ilmenite) and coal. An Eulerian multiphase continuum model was used to describe
both the gas and solid phases, with detailed sub-models to account for uidparticle and particle
particle interaction forces. Global reaction models of fuel and carrier chemistry were utilized. The
transient results obtained from the simulations were compared with detailed experimental time-
varying outlet species concentrations (Leion et al., 2008) and provided a reasonable match with the
reported experimental data.
& 2011 Elsevier Ltd. All rights reserved.
1. Introduction
Due to the threat of global warming (Solomon et al., 2007),
immense importance is being placed on developing technologies
for producing power without the release of greenhouse gases
such as CO
2
. Although many alternative energy sources have been
proposed, the continued use of fossil fuels still seems to be
essential. Therefore ways to reduce CO
2
emissions from the
combustion of fossil fuels have to be developed. Presently several
technologies, including oxy-fuel combustion, post-combustion
capture from ue gases, pre combustion capture, as well as
technologies like CLC, are being demonstrated for CO
2
capture
(Metz et al., 2005). However most technologies other than CLC
will lead to a signicant increase in the cost of electricity,
consuming a large portion of the energy they generate to separate
gases (Metz et al., 2005).
Chemical-looping combustion (CLC) is a novel combustion
technology that has recently gained great attention since it
inherently produces a concentrated CO
2
stream (Knoche and
Richter, 1968; Richter and Knoche, 1983; Ishida et al., 1987;
Lyngfelt et al., 2001; Ryu et al., 2001; Ada nez et al., 2004; Andrus
et al., 2006). The CLC system usually consists of two uidized bed
reactors: an air reactor (AR) and a fuel reactor (FR) (Fig. 1). The
fuel is oxidized in the FR by contacting hot granular metal oxides.
Most of the reported work to date utilizes gaseous fuels, which
can react directly with the carrier. However, in the present
simulation a solid fuel is used (coal). It rst devolatilizes and
the remaining char is gasied by the recycled H
2
O/CO
2
stream
used to uidize the fuel reactor. The devolatilization and gasica-
tion products (CO and H
2
) are then oxidized by the hot metal
oxide. The reduced metal particles are returned to the AR where
they are re-oxidized by air. The AR is typically a transport reactor.
At its exit the oxidized carrier is separated by a cyclone and
returned to the FR. The net chemical reaction and energy release
is identical to that of the conventional combustion of the fuel. The
energy spent on circulation of the carrier (the only energy cost of
separation) is very small (0.3%) in comparison with the total
energy released (Lyngfelt, et al., 2001). The exhaust stream of the
FR consists, mainly, of CO
2
and H
2
O. The H
2
O can be easily
condensed, resulting in a pure CO
2
gas, which can be pressurized
and sequestered. CLC holds signicant promise as a next genera-
tion combustion technology as it has the potential to allow near-
zero CO
2
emissions with very little effect on the efciency of
power plants and, hence, electricity cost (Ishida and Jin, 1996).
Several energy and exergy analyses of CLC systems have been
reported, suggesting that power plant efciencies greater than
50% can be achieved along with nearly complete CO
2
capture
(Ishida et al., 1987; Ishida and Jin, 1994; Wolf et al., 2001; Marion,
2006, Andrus, et al., 2006).
Contents lists available at ScienceDirect
journal homepage: www.elsevier.com/locate/ces
Chemical Engineering Science
0009-2509/$ - see front matter & 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2011.04.025
n
Corresponding author at: West Virginia University, Department of Mechanical
and Aerospace Engineering, Morgantown, WV 26506, USA. Tel.: 1 419 944 1391.
E-mail address: kar982@gmail.com (K. Mahalatkar).
Chemical Engineering Science 66 (2011) 36173627
Signicant research is currently underway in the design of CLC
reactors, and several small scale (300 W to 65 kW) test plants
have been constructed (Johansson et al., 2006; Abdulally et al.,
2010; Lyngfelt and Thunman, 2005; De Diego et al., 2007; Son and
Kim, 2006; Kolbitsch et al., 2009). There has been extensive study
of the properties of carriers, an important aspect of CLC technol-
ogy (Johansson, 2007). It is critical that the rates of reaction of the
carrier, both reduction by the fuel and oxidation by air, to be high
at the operating temperature of the reactors. Also, it is desirable
to have a low attrition rate to minimize carrier make-up costs, a
potentially signicant operating cost. It should also be economical
and environmentally benign. Various studies have been carried
out to characterize the reduction and oxidation behavior of metal
oxides with various gases such as CO, H
2
and CH
4
(Johansson
et al., 2004; Johansson, 2007; Ryu et al., 2001; Ishida and Jin,
1996; Mattisson et al., 2001; Ada nez et al., 2004; Siriwardane
et al., 2007; Chandel et al., 2009). Models based on shrinking core
and changing grain size have been used to accurately represent
the chemical kinetics of the metal oxides.
CFD provides a method to analyze the interaction between
uid mechanics and chemical kinetics. Fluid mechanical effects,
such as bubble formation in the fuel reactor and fuel leakage into
the air reactor, could affect their performance. Despite the large
number of publications in the past few years on the experimental
studies of CLC, very little attention has been paid to CFD studies of
such systems. Jung and Gamwo (2008), Deng et al. (2009) and
Krugel-Emden et al. (2010) have performed CFD studies of fuel
reactors using gaseous fuels. No CFD simulations involving solid
fuels have been previously reported. In the present study we
present a two-dimensional CFD simulation of a FR and demon-
strate the ability of multiphase CFD to accurately simulate the
recent experiments of Leion et al. (2008) using solid fuels. It is
envisioned that the CFD models developed and validated herein
will be utilized in design studies of experimental coal CLC systems
being developed at the National Energy Technology Lab (NETL),
Morgantown, West Virginia, USA.
2. The experimental study
The data to be simulated were provided by an experimental
study of Leion et al. (2008) in which coal was consumed in a
uidized bed reactor. The reactor has a tapered/conical section at
the bottom and tall constant radius vertical section (Fig. 2). The
operating gas velocity was above the minimum uidizing velocity
of the bed material. The uidizing gas was 50% steam and 50% N
2
.
The reactor was initially heated to the desired operating
temperature (1223 K) and then coal particles were released
into the top section of the reactor. During the injection of coal, it
was observed that some of the coal particles became stuck in the
injector and some particles were elutriated due to disintegration.
The estimated loss of coal was as high as 50% (Leion, H. and A.
Lyngfelt, Private communication, 2009), however the exact
amount of coal loss is uncertain. In the present simulations the
injected coal mass was therefore set to be 50% of that reported by
Leion et al. (2008), which was the adjustment required so that the
integrated gas production agreed with the reported values. Steam
was introduced from the distributor, as part of the uidization
ow, to gasify the char remaining after the devolatilization
process. The metal oxide used was ilmenite (which was assumed
to be Fe
2
O
3
supported on TiO
2
in the present simulations). Table 1
provides details of the Leion et al. (2008) experimental study.
3. The CFD model
The granular ows are usually modeled using approach that can
be broadly classied into two: (a) continuum or Eulerian approach
and (b) Lagrangian or discrete particle approach.
In the continuum or Eulerian approach, each particle variable
(such as velocity, temperature, mass, etc.) is averaged over a
region that is large in comparison with the particle size and/or
spacing. Thus only the bulk behavior of the solids is accounted for.
Due to this averaging process, constitutive relationships for
viscous stress, heat transfer, etc. are required to complete the
model of the particle phase and its interactions with the gas phase
as well as other solid phases. In the Lagrangian or discrete particle
approach, the trajectories of the particle motion are directly
estimated by doing a force and inertia balance on the particles.
Fig. 1. Schematic diagram of a chemical-looping combustion (CLC) system
(Lyngfelt et al., 2001).
Fig. 2. Lower portion of the coarse computational mesh.
Table 1
Simulated fuel reactor properties (Leion et al., 2008).
Diameter of bed 1030 mm
Flow rate at 0 1C and 1 atm 600 ml/min
Mass of metal oxide particles 40 g
Mass of injected coal (50% of injected amount
from Leion et al. (2008))
0.1 g
Mean diameter of metal particles 105 mm
Mean diameter of coal particles 150 mm
Mean density of metal oxide particles (including support) 4500 kg/m
3
Initial density of coal particles 2000 kg/m
3
Initial bed height 35 mm
K. Mahalatkar et al. / Chemical Engineering Science 66 (2011) 36173627 3618
In the present analysis, the continuum (Eulerian) approach has
been used to model both the gas and solid phases. The continuum
models have been used for modeling granular ows for more than
three decades and have been well documented (Gidaspow, 1992).
Numerous CFD studies using both continuum and discrete mod-
eling approach to study uidized bed with two or more different
type of solid particles have shown that CFD can capture the
experimental trends in a reasonable fashion (Van Wachem et al.,
2001; Huilin et al., 2003; Gera et al., 2004; Cooper and Cornonella,
2005; Beetstra et al., 2007; Reddy and Joshi, 2009).
The commercial CFD software FLUENT
TM
has been used in the
present work. For completeness the set of equations that have
been solved in the present analysis is documented. The details of
the rheological model in both the dilute kinetic regime and the
dense frictional ow regime are provided. The details of the
momentum interaction term (drag law) between the gas and
solid phase as well as the equations for calculation of granular
temperature are also summarized.
The continuity equation for phase q is given as (Ansys-Fluent
Inc. 2006; Syamlal et al., 1993)
@
@t
a
q
r
q
rUa
q
r
q
v
!
q

n
p 1
_ m
pq
_ m
qp
1
where a
q
is the volume fraction of the qth phase and _ m
pq
is the
mass transfer rate from the pth to qth phase. Three phases were
considered in these calculations: one gas phase and two solid
phases. Each phase consists of a number of species. For example,
the gas phase may consist of CH
4
, CO
2
, H
2
O, N
2
, etc. and a solid
phase may consist of Fe
2
O
3
, Fe
3
O
4
, etc. A transport equation is
solved for each species:
@
@t
a
q
r
q
Y
iq
rUa
q
r
q
v
!
q
Y
iq

n
p 1

m
j 1
_ m
qp
ij
_ m
pq
ji
2
where Y
iq
is the mass fraction of species i in the qth phase and _ m
qp
ij
is the mass transfer rate from the jth species of the pth phase to
ith species contained in the qth phase.
The momentum equation for the gas phase is given as
@
@t
a
g
r
g
v
!
g
rUa
g
r
g
v
!
g
v
!
g

a
g
rprUt
g
a
g
r
g
g
!

n
s 1
R
!
sg
_ m
sg
v
!
sg
_ m
gs
v
!
gs
3
where R
,
sg
b
sg
v
!
s
v
!
g
represents the momentum transfer
between the sth solid phase and the gas phase due to drag and
the following term is due to mass transfer.
The momentum equation for the sth solid phase is given as
@
@t
a
s
r
s
v
!
s
rUa
s
r
s
v
!
s
v
!
s

a
s
rprUt
s
a
s
r
s
g
!

n
r 1
R
!
rs
_ m
rs
v
!
rs
_ m
sr
v
!
sr
4
The uid stress tensor is given as
t
g
m
g
a
g
r v
!
g
r v
!
T
g
a
g
l
g
rU v
!
g
I 5
(Here l
g
0) and the granular solid phase stress tensor, arising
from particleparticle collisions, is given as
t
s
p
s
I m
s
a
s
r v
!
s
r v
!
T
s
a
s
l
s
rU v
!
s
I 6
Here p
s
is the solid-pressure and m
s
is the granular viscosity.
Also, the uidsolid and soliduid momentum exchange terms
are opposite but equal, b
sg
b
gs
; the model proposed by Gidaspow
(1992), a combination of the pressure drop loss suggested by Wen
and Yu (1966) and the Ergun equation, is used. When a
g
40.8
(dilute regions), the uidsolid exchange coefcient b
sg
is
b
sg

3
4
C
D
a
s
a
g
r
g
9 v
!
s
v
!
g
9
d
s
a
2:65
g
, where
C
D

24
a
g
Re
10:15a
g
Re
0:687
_ _
7
When a
g
o0.8,
b
sg
150
a
s
1a
g
m
g
a
g
d
2
s
1:75
r
g
a
s
9 v
!
s
v
!
g
9
d
s
8
The solids stress is based on Lun et al. (1984) and accounts for
collision between particles:
p
s
a
s
r
s
Y
s
2r
s
1e
ss
a
2
s
g
0,ss
Y
s
9
Here g
0,ss
is the radial distribution function at contact, given as
g
0,ss
1a
s
=a
s,max

1=3
_ _
1
10
The maximum packing fraction a
s,max
was 0.63.
The solid viscosity is
m
s

4
5
a
s
r
s
d
s
g
0,ss
1e
ss

Y
p
_ _
1=2

10r
s
d
s

Y
s
p
p
96a
s
1e
ss
g
0,ss
1
4
5
g
0,ss
a
s
1e
ss

_ _
2
11
and the solids bulk viscosity is
l
s

4
3
a
s
r
s
d
s
g
0,ss
1e
ss

Y
p
_ _
1=2
12
To calculate the granular temperature, equilibrium is assumed
between production and dissipation of the random kinetic energy,
therefore,
t
s
: ru
!
s
g
Ym
f
ls
13
where the dissipation through collisions is
g
Ym

121e
2
ss
g
0,ss
d
s

p
p r
s
a
2
s
Y
3=2
s
14a
and the loss to the uid is
f
ls
3b
s
Y
s
14b
When the solid fraction becomes close to the dense packing
limit, enduring frictional interactions between particles take
place, which are usually simulated as an additional frictional
stress term. However the frictional term associated with dense
granular ows was not added because of the following reasons:
i. The addition of frictional pressure and viscosity resulted in
requirement of smaller time steps and greater number of
iterations at each time step to achieve convergence of the
solution. The present simulations require months of computa-
tional time. Given that addition of the frictional model would
increase the needed computational time, it was deemed beyond
our current computational capability. Since this reactor consists
of a bubbling bed, with no packed regions, it is believed that the
rheology of the granular phase can be described adequately by
the Kinetic Theory of Granular Flow. The granular phase was
still restricted by a maximum packing constraint imposed by a
normal force, which resists extreme packing.
ii. The frictional-rheological behavior of the metal oxide material
used by Leion et al. (2008) in the experimental study is
unknown. Kronberger et al. (2004) have carried out experi-
ments with circulating uidized bed. They have shown that
particles of very similar size and density, but of different
material (FCC and glass), show dramatically different frictional
ow behavior. Langroudi et al. (2010) suggest that the
K. Mahalatkar et al. / Chemical Engineering Science 66 (2011) 36173627 3619
frictional behavior of a material is very complex and use a
Couette ow device to characterize the rheological behavior of
the material in the dense frictional ow. However such a
study for characterization of frictional-rheological model is
currently available for only a limited number of materials.
iii. The frictional models currently available in literature are
based on plasticity theory and are only a very approximate
representation of the complex frictional ow behavior. Also
the equations of popular frictional pressure models have
many coefcients that have to be assumed in an arbitrary
fashion.
The energy equation is solved for each phase and is given as
@
@t
a
q
r
q
h
q
rUa
q
r
q
v
!
q
h
q
a
q
@p
@t
t
s
: rU v
!
q
rU q
!
q
S
q

n
p 1
Q
pq
15
where h
q
is the specic enthalpy of the qth phase and q
!
q
is the
heat ux, S
q
is the source term for enthalpies due to chemical
reactions, Q
pq
is the heat transfer between gas and solid phases
and is provided by the Gunn (1978) correlation. Heat transfer
between solid phases is considered to be small and neglected.
4. The reaction scheme and rates
The following reaction mechanisms were used.
Coal devolatilization
Coal-aChar bC
Soot
cCH
4
dC
2
H
6
eCO
f CO
2
gH
2
hH
2
OiAsh 16
The coefcients (ai) in Eq. (16) are determined from the
ultimate and proximate analysis of the coal (Section 4.1 provides
details).
Watergas-shift reaction
COH
2
O-CO
2
H
2
(17)
Char gasication by CO
2
CharCH
0:0245
O
0:001531
CO
2
-2CO0:01072H
2
0:001531H
2
O
18
Char gasication by H
2
O
CharCH
0:0245
O
0:001531
H
2
O-CO1:01072H
2
0:001531H
2
O
19
Metal oxide reduction
12Fe
2
O
3
CH
4
-8Fe
3
O
4
2H
2
OCO
2
21Fe
2
O
3
C
2
H
6
-14Fe
3
O
4
3H
2
O2CO
2
3Fe
2
O
3
CO-2Fe
3
O
4
CO
2
3Fe
2
O
3
H
2
-2Fe
3
O
4
H
2
O 20
4.1. Details of the devolatilization scheme
The models by Bradley et al. (2006) and Merrick (1983) have
been utilized in this work to determine the components of
primary devolatilization. They provide models for both global
devolatilization and for individual species as well. The global
devolatilization model has been used here because of its simpli-
city and also because the devolatilization timescales are fast in
comparison with the total time scale for the combustion of coal.
Merrick (1983) proposes that coal initially breaks up into the
major components: char, tar and volatile species. The volatile
species are assumed to consist of CH
4
, C
2
H
6
, CO, CO
2
, H
2
and H
2
O.
Based on experimental studies of ve different coals, Merrick
proposes the mass fraction composition of char and tar as given in
Table 2. Bradley et al. (2006) suggest that the decomposition of
tar results in soot, CH
4
, CO, H
2,
H
2
S and HCN. In the Leion et al.
(2008) experiments, the rate controlling step is the gasication of
char. At temperatures typical of CLC, the timescales for gasica-
tion of coal are usually several minutes while the devolatilization
occurs in a matter of seconds. Because of the fast nature of
devolatilization the primary and secondary devolatilization reac-
tions are combined to obtain Eq. (16).
The following is assumed for primary and secondary devola-
tilization (Bradley et al., 2006; Merrick, 1983):
1) Methane consists of 32.7% of the hydrogen in coal.
2) C
2
H
6
consists of 4.4% of the hydrogen in coal.
3) CO consists of 18.5% of the oxygen in coal.
4) CO
2
consists of 11% of the oxygen in coal.
Based on the above assumptions, the coefcients for the
reaction given in Eq. (16) can be calculated.
According to Bradley et al. (2006) the devolatilization rate can
be expressed in the following form:
dm
k
dt
k
v,k
m
ok
m
k
21
The rate used is k
v
110
5
exp(12000/T
p
) from Bradley et al.
(2006). This results in a devolatilization time scale of around a
second at 1000 K (k
v
0.614 1/s). The calculated stoichiometric
coefcients using the devolatilization model for South African
coal used in the Leion et al. (2008) experiments are given in
Table 3.
4.2. Gasication rates
The expressions and constants for the gasication rates are
obtained from Everson et al. (2006). For a single char particle,
according to the shrinking core model:
dX=dt k 1X
2=3
; where k r
1
S
o
=1e
0
22
This implies the reaction rate (kg/(m
3
s)):
_ m
char
r
char
e
char
S
o
1e
0
r
1
1X
2=3
23
Here X is the conversion, S
o
is the initial calculated
particle surface area and e
0
is the initial porosity. The ratio was
Table 2
Mass fraction of different elements in char and tar (Merrick, 1983).
C H O
Char 0.98 0.002 0.002
Tar 0.85 0.082 0.049
Table 3
Stoichiometric coefcients for devolatilization.
Species Label Value
C a 0.4639
Soot b 0.012
CH
4
c 0.0309
C
2
H
6
d 0.0026
CO e 0.0096
CO
2
f 0.00266
H
2
g 0.0582
H
2
O i 0.0805
Ash l 0.0165
K. Mahalatkar et al. / Chemical Engineering Science 66 (2011) 36173627 3620
specied as
S
o
=1e
0
8:83 10
4
m
2
=m
3
24
For char gasication by H
2
O, r
1
is given as (also see Eq. (22))
r
1
r
H
2
O

k
H
2
O
K
H
2
O
P
H
2
O
1K
H
2
O
P
H
2
O
K
H
2
P
H
2
25a
For char gasication by CO
2
, r
1
is given as
r
1
r
CO
2

k
CO
2
K
CO
2
P
CO
2
1K
CO
2
P
CO
2
K
CO
P
CO
25b
The coefcients for the two gasication rates are listed in
Table 4. Liu and Niksa (2004) provide a similar gasication model
based on LangmuirHinshelwood rate expressions and suggest
that the gasication rates for different types of coal can vary
signicantly, and therefore there is a need for tuning of the
gasication rates at one temperature. In the present simulations
the rates for the South African coal were tuned at one specic
operating temperature of 1223 K (i.e., coefcients k
CO
2
and k
H
2
O
were increased by a factor of 2 in comparison with values
provided by Everson et al. (2006) and then these rates were used
to simulate the reactor at other operating temperatures).
4.3. Metal oxide reduction rates
The iron oxide (hematite) reduction rates for CO and H
2
were
obtained from Mattisson et al. (2005). The reduction rate for CH
4
was obtained from Mattisson et al. (2001). Since the experiments
utilized a large excess of hematite, it was assumed that only
reduction to magnetite need be included in the chemistry
scheme. These reaction rates were determined from TGA experi-
ments. A simple one-step chemical reaction was assumed, as
given by Eq. (20). Experimental studies by Ryu et al. (2001) and
Mattisson et al. (2005) suggest that the porous grain model
(Szekely et al., 1976) can be used to explain the progress of these
heterogeneous reactions. They also suggest that the outer bound-
ary layer diffusion and pore diffusion of gases are fast and can
therefore be neglected. The controlling step in the heterogeneous
reaction is the slow chemical reactions on the surface of the
pores. The progress of reactions can be described by Eq. (22),
where
X
mm
red
m
ox
m
red
26
Here m
ox
is the mass when the particle is fully oxidized to
hematite and m
red
is the mass when particle is fully reduced to
magnetite. Also R
o
(m
ox
m
red
)/m
ox
. Differentiating Eq. (26) leads
to
dX
dt

1
m
ox
m
red
dm
dt

1
R
o
m
ox
dm
dt
27
The loss in mass of the metal oxide is due to the loss of oxygen,
so that
dmdm
O
2
MW
O
2
dn
O
2
Using these results, the reaction rates (kg/(m
3
s)) of the fuel
gases with iron oxide can be written as (Mahalatkar et al., 2010;
Mattisson et al., 2005)
_ m
H
2

k
H
2
R
o
2MW
O2
r
avg
e
s
Y
Fe
2
O
3
Y
Fe
3
O
4

n
Fe
2
O
3
MW
Fe
2
O
3
n
Fe3O4
MW
Fe3O4
_ _
1X
2=3
MW
H
2
where
k
H
2

3bkC
H
2
C
H
2
,eq

r
m
r
o
; k k
o
e
E=RT
b 3, k
o
2:3 10
3
1=s and E 24kJ=mol 28
_ m
CO

k
CO
R
o
2MW
O2
r
avg
e
s
Y
Fe
2
O
3
Y
Fe
3
O
4

n
Fe2O3
MW
Fe2O3
n
Fe3O4
MW
Fe3O4
_ _
1X
2=3
MW
CO
where
k
CO

3bkC
CO
C
CO,eq

r
m
r
o
; k k
o
e
E=RT
, b 3,
k
o
6:2 10
4
1=s and E 20kJ=mol 29
_ m
CH4

k
CH4
R
o
2MW
O2
r
avg
e
s
Y
Fe2O3
Y
Fe3O4

n
Fe2O3
MW
Fe2O3
n
Fe3O4
MW
Fe3O4
_ _
1X
Y
CH4
Y
CH4,TGA
MW
CH4
30
where, k
CH
4
5:33 10
4
1=s and Y
CH
4
,TGA
0:1.
The reaction rate for C
2
H
6
was assumed to be same as CH
4
because of unavailability of experimental data. Although tar
cracking and heterogeneous reduction is probably an important
practical aspect of CLC, no attempt has been made to model this
process in these simulations.
4.4. Watergas shift reaction rates (Bustamante et al., 2004)
The reaction is given as
COH
2
O2CO
2
H
2
(31)
Bustamante et al. (2004) studied the reverse watergas-shift
reaction and reported the reverse reaction rate and equilibrium
constant. This reverse reaction rate was given in the form
rkmol=m
3
s dCO=dt k
o
e
E=RT
H
2

a
CO
2

b
32
K
eq
exp4:334577:8=T 33
The forward rate is calculated from the backward rate using
the equilibrium constant. The net reaction rate can then be
written as
r dCO=dt k
o
e
E=RT
H
2

a
CO
2

1
K
eq
e
E=RT
H
2
O CO
_ _
kmol
m
3
s
_ _
34
The coefcients given by Bustamante et al. (2004) are provided
in Table 5.
4.5. Simulating a particle of changing size and density
In coal combustion the particle diameter and density generally
both vary with the extent of conversion. At high temperatures
(T41800 K) the combustion reactions are so fast that reaction
Table 4
Numerical values for coefcients in gasication reactions (Everson et al., 2006).
Pre-exponential factor Exponential factor
k
CO2
3.9610
4
1/s 109 kJ/mol
K
CO2
8.3710
5
1/Pa 16 kJ/mol
K
CO
1.910
5
1/Pa
k
H2O
22. 1/s 212 kJ/mol
K
H2O
9.5410
2
1/Pa 69 kJ/mol
K
H2
9.3610
5
1/Pa
Table 5
Coefcients for watergas-shift reaction rates.
k
o
2.1710
7
E (kJ/mol) 192.9
a 0.5
b 1
K. Mahalatkar et al. / Chemical Engineering Science 66 (2011) 36173627 3621
with surrounding gas takes place mainly at the outer surface and
the particle density tends to remain constant while the diameter
changes with time (Smoot and Smith, 1985). At the lower
temperatures characteristic of uidized beds (To1300 K), the
gasication reactions being slow, it is thought that the reacting
gas has enough time to penetrate through the pores and the
reactions take place throughout the volume of the particle
(Syamlal and Bissett, 1992).
Liu and Niksa (2004) have used the following model in their
Carbon Burnout Kinetics Model for gasication (CBK/G) (see Eqs.
(35) and (36) in Liu and Niksa, 2004):
r
r
0

m
m
0
_ _
a
and
d
p
d
p0

m
m
0
_ _
1=3
r
r
0
_ _
1=3
_
35
where m is the mass of a particle and m
o
is its initial mass. They
provide a single example case of a 60 mm particle where they
estimate the parameter a to be around 1. This implies d
p
d
p0
and
(r/r
0
)(m/m
0
), which were used in this study.
Based on this reasoning, it has been assumed that the particle
diameter is constant, but the density is a linear function of
conversion, X:
r
p

r
p,max
1X for XoX
c
r
p,min
for X4X
c
_
36
The density of the unburned coal, r
p,max
, has been assumed to
be 2000 kg/m
3
. The density has been limited to a minimum value,
r
p,min
100 kg/m
3
, resulting in a maximum value of X
c
of 0.95
(or 95%). The conversion of the particle X was computed from
the ratio of mass of ash to char in a given computational cell. The
density of the particle is assumed to remain constant within a
given computational cell and is calculated based on the average
conversion of all particles within the cell. The density can
however vary from one computational cell to another.
5. The initial and boundary conditions
The initial bed height of 35 mm (Table 1) was determined from
the geometry of the reactor, the density of iron oxide and the total
mass of iron oxide initially in the bed. The coal was assumed to be
initially fully mixed with the iron oxide. In the experiments the
entire reactor was placed in an oven at 950 1C (Leion et al., 2007,
Leion et al., 2008). To simulate these conditions the walls of the
reactor were assumed to be at 950 1C and the inlet gas was
assumed to be at 950 1C temperature. The bed material was
initially assumed to be at 950 1C. The metal oxide used in the
experiments was ilmenite (Leion et al., 2008). In the simulations it
was assumed that ilmenite is Fe
2
O
3
supported on TiO
2
. The initial
mass fraction of Fe
2
O
3
was 60% and rest was TiO
2
. The gas velocity
at the inlet (or distributor plate of the reactor) was assumed to be
constant and corresponds to a ow of 600 ml/min at 0 1C and
1 atm (Leion et al., 2008). The gas outlet was assumed to be at
constant atmospheric pressure.
6. Numerical parameters and grid convergence study
The CFD domain for the geometry of the FR is shown in Fig. 2.
The grid used in the simulation was an axi-symmetric quadrilat-
eral grid generated in GAMBIT
TM
software. In Fig. 2, for display
purposes, the mesh has been mirrored about its axis. Two
different grid sizes were used. The cell count in the ne mesh
exceeds that in coarse mesh by a factor of 2.25. The typical
vertical cell dimension at different regions of the mesh is
provided in Fig. 2. The time step size for the ne mesh was
reduced by a factor of two for numerical stability. The discretiza-
tion scheme and other numerical parameters are provided in
Table 6. Fig. 3 compares the time varying outlet gas concentra-
tions obtained from coarse and ne mesh. A temperature of
950 1C and 50% inlet steam concentration was used for these
simulations. The maximum difference in the gas concentration
obtained from coarse and ne mesh occurs at the peak concen-
tration during devolatilization (Fig. 3). The maximum difference
in the gas concentration is around 5%, but at most times the error
is signicantly less. The coarse mesh was used in all subsequent
simulations as the computational time required for the ne mesh
is signicantly larger.
7. Results
Fig. 4 shows instantaneous velocity vector and volume fraction
contour plots for both iron oxide and unreacted char, respectively,
at approximately 175 s into the simulation. The time instant of
175 s is chosen because the devolatilization products have left the
system by this time and the rates constraining gasication
reactions are dominant. (Since the experiments were batch,
Table 6
Summary of numerical parameters.
Coarse mesh Fine mesh
No. of cells 3300 7425
Spatial discretization scheme for the momentum equations Second order upwind Second order upwind
Spatial discretization scheme for each phase and species equation First order upwind First order upwind
Time integration scheme First order implicit First order implicit
Time step size 0.001 s 0.0005 s
Iterations per time step 20 20
Fig. 3. Comparison of the calculated time development of the CO
2
and CO
concentrations using a coarse and a ne mesh (temperature of 950 1C and H
2
O
concentration of 50%).
K. Mahalatkar et al. / Chemical Engineering Science 66 (2011) 36173627 3622
time-average results, as are usually shown, have no real mean-
ing.) The regions of low solids volume fraction (high gas volume
fraction) display the bubble. Larger bubble diameter will allow
reacting gases to bypass the emulsion phase and can allow
combustible gases to escape without reacting with the metal
oxides. Velocity vector plots in Fig. 4a and b clearly show a high
velocity central jet, which causes the solids to spurt. Based on
numerical simulations, the experimental system behaves some-
where between a spouted bed and a bubbling bed. In spouted
beds a strong central jet passes through the core of the reactor. In
the present system the uidizing velocity is not strong enough to
form a penetrating central jet but it does form a series of bubbles
rising rapidly along the central core of the reactor. As in spouted
beds, the particles rise through the central core, however, dragged
in the wakes of the rising bubbles. The solid particles subse-
quently return back down through an outer annular low velocity
region, which encircles the high velocity central core consisting of
bubbles. The particles used in the experiments are Geldart B type
particles.
Comparing phase contour plots of metal oxide and coal (Fig. 4a
and b) it is clearly observed that the region of large volume
fraction of metal oxide is also the region of large volume fraction
for coal indicating that both the coal and iron oxide particles
remain well mixed during the simulation. Also the solids velo-
cities are the same order of magnitude for both coal and metal
oxide, which further establishes the well mixed nature of the
particles. The mixing of metal oxide and coal is essential for high
efciency of the reactor. If coal particles segregate in the upper
portion of the bed due to their progressively lower density the CO
and H
2
released by gasication will not contact the carrier
particles, resulting in partial combustion and, hence, a reduction
in combustion efciency.
The bubbles are formed at the inlet and rise mainly along the
central axis of the reactor. This causes a higher concentration of
solids, mainly iron oxides, along the side walls of the reactor
(Fig. 4a). Fig. 5a shows the mass fraction of steam, the primary
gasifying agent, constituting 50% of the inlet gas. As expected the
peak mass fraction of H
2
O occurs in the bubble region. The drop in
its mass fraction in the emulsion phase is due, in part, to
gasication of coal. Fig. 5b shows the mass fraction of CO
2
. It is
high in the emulsion phase, where it is produced, and low in the
bubble regions. Only a small amount of CO
2
moves through the
bubble phase. Fig. 5c shows the H
2
mass fraction, a product of coal
gasication. It is high in the emulsion region, where it is
produced, and can further react with the carrier. Only a small
amount of H
2
is present in the bubble regions and hence it is
likely that only a small percentage of H
2
bypasses the bed through
the bubble phase.
In the experiments the concentration of outlet ue gases was
measured after 40 s of time delay (Leion et al., 2007). This was
primarily the time required for gases to ow through the pipe
connecting the reactor and gas analyzer. Experiments by Taylor
(1954) demonstrate that there can be signicant apparent diffu-
sion of gases when they travel through pipes. This can lead to
signicant changes in peak concentration of gases. To account for
apparent diffusion of gases during the travel between reactor and
gas analyzer, the length of the simulated reactor was extended to
10 m. In the simulations the outlet gas concentration has been
Fig. 4. Velocity vectors (m/s) and volume fractions of iron oxide and coal particles at 175 s, temperature of 1273 K and 50% steam concentration. (a) Iron oxide and (b) Coal.
Fig. 5. Instantaneous gas species mass fractions at 175 s, temperature of 1273 K and 50% steam concentration. (a) H
2
O, (b) CO
2
and (c) H
2
.
K. Mahalatkar et al. / Chemical Engineering Science 66 (2011) 36173627 3623
measured at a height of 3 m (Fig. 6) to approximate 40 s of time
delay in measurement of experimental concentrations. A consid-
erable amount of apparent diffusion of gases occurs during this
period and this affects the peak concentrations of devolatilization
gases such as CH
4
. Also, the diffusive effects damp out oscillations
in concentration of ue gases that are observed close to the top of
the bed because of bubble eruption and other uctuations.
The peak CO
2
concentration predicted by the simulations
agrees well with the experiments (Fig. 6 and Table 7). The
computed char burnout time is close to the experimentally
observed 15 min. The peak CH
4
and CO concentrations predicted
by the simulations differ by 30% in comparison with experimental
values. These differences may be attributed to the assumptions
made in the modeling of the reactor as well as the uncertainty in
the injection of coal mass as mentioned in Section 2.
In the operation of the current CLC reactor, the gasication
reaction is the slowest and it requires several minutes to com-
pletely gasify the coal. The gasication can occur through two
reaction pathways as indicated by Eqs. (18) and (19). Fig. 7 shows
the char consumption rates (kmol/m
3
/s) due to gasication by
both steam and CO
2
. The char consumption rate due to steam is at
least 30 times higher than CO
2
(Fig. 7). This is because of two
reasons: (i) the concentration of steam is higher and (ii) the
reaction rates for steam gasication are also higher.
Fig. 8 shows the variation in the hematite and magnetite mass
during the coal combustion phase. There is a rapid change in the
mass at time close to zero seconds. This is due to the devolatiliza-
tion of coal and the reaction of the volatiles with the metal oxide.
Both hematite and magnetite reach a stable asymptotic value
with further consumption of coal. Also hematite is present in
excess amounts in the system and therefore the maximum drop
in mass of hematite is only around 20%. Table 8 compares the
mass of carbon input into the system and the integrated carbon
mass leaving the system over the entire simulated time. The mass
balance is within 0.5% error. This error is mainly attributed to the
time integration procedure used for evaluating the total carbon
mass owing through the exit.
7.1. Effect of higher operating temperature
Fig. 9 shows the effect of operating temperature on the
average coal reaction rates, dened as (Leion et al., 2008)
R
avg

1
m
tot
m
t
t
38
where m
tot
is the total mass of carbon to be converted at all times
and m
t
is the total mass of carbon converted up until time t, the
time from start of experiments/simulation. The simulations are in
agreement within experimental variations observed at most
operating conditions showing rapid increase with increasing
temperature. Recall that the Everson gasication rates had to be
increased by a factor of two to obtain agreement at 950 1C. This
was not further adjusted for the simulations at the higher or
lower temperatures; the rates increased due to their Arrhenius
temperature dependence.
Fig. 6. Calculated concentrations of CO
2
, CO and CH
4
at a height of 3 m compared
to the experimental data of Leion et al. (2008). Temperature is 1223 K and 50%
steam concentration is used.
Table 7
Peak concentration from experiment and simulation.
Peak concentration
experimental (%)
Peak concentration
simulation (%)
CH
4
0.036 0.049
CO
2
0.078 0.081
CO 0.047 0.032
Fig. 7. Gasication rate of char (kmol/(m
3
s) of char). (a) Steam gasication rate
and (b) CO
2
gasication rate.
Fig. 8. Time variation of hematite (Fe
2
O
3
) and magnetite (Fe
3
O
4
) mass. Tempera-
ture of 1223 K and 50% steam concentration.
K. Mahalatkar et al. / Chemical Engineering Science 66 (2011) 36173627 3624
7.2. Effect of steam concentration
The effect of the variation of steam concentration in the
uidization gas on the average reaction rates is shown in
Fig. 10. Once again the simulation results are generally within
the observed experimental variations and predict the slight
increase in reaction rate with increasing H
2
O concentration due
to increased rate of char gasication.
8. Summary and conclusions
A CFD model for simulation of coal combustion in the fuel
reactor of chemical-looping combustion systems has been devel-
oped. The solid particles are modeled as a continuum uid.
Chemical kinetic models have been assembled from the literature
for the reactions between the iron oxide and fuel gases as well as
the devolatilization and char gasication of coal. The CFD model
has been used to simulate experiments of Leion et al. (2008). The
model was able to predict the outlet concentrations of CO
2
, CO
and CH
4
. The reactor performance at different operating tempera-
tures was captured in a reasonable manner. The changes in
average coal conversion rates due to change in steam concentra-
tion were also captured in a reasonable manner. This demon-
strates that CFD modeling can be an effective approach in the
design of such a reactor.
In it is hoped that simulation accuracy could be improved
through the following: (a) improvement to drag laws particularly
inclusion of polydisperse drag laws, (b) experimental character-
ization of the rheological behavior of the solids in the dense
frictional regime, (c) inclusion of improved and more accurate
devolatilization and gasication models, (d) characterization of
change in coal particle density and diameter as well as particle
breakup during combustion, or (e) inclusion of more complex
chemical mechanisms (for both homogeneous and heterogeneous
reactions).
Nomenclature
AR air reactor
b coefcient in the metal oxide reduction reaction
C
D
drag coefcient
C
CO
concentration of carbon monoxide
C
CO,eq
equilibrium concentration of carbon monoxide
C
H
2
concentration of H
2
C
H
2
,eq
equilibrium concentration of H
2
d
s
particle diameter (m)
e
ss
coefcient of restitution
E activation energy for reduction of Fe
2
O
3
to Fe
3
O
4
(kJ/mol)
FR fuel reactor
g
!
acceleration due to gravity (m/s
2
)
g
0,ss
radial distribution function
I identity matrix or tensor
k
o
pre-exponential factor for gasication reaction
K
eq
equilibrium constant
m mass of particle (kg)
m
i
mass of the ith species (kg)
m
ox
mass of particle in fully oxidized (ox) form (kg)
m
red
mass of particle in fully reduced (red) form (kg)
_ m
char
rate of consumption of char (kg/(m
3
s))
_ m
H
2
rate of consumption of hydrogen (kg/(m
3
s))
Table 8
Mass balance of carbon in reactor after complete char burnout.
Total carbon mass injected through coal Total integrated carbon mass exiting at 3 m as
gases over the entire simulated time
% Error
Carbon 0.125 g (or 62.5% of coal) 0.12435 g (CO
2
COsootCH
4
C
2
H
6
) 0.52%
Fig. 9. Average coal reaction rate (see text) at different operating temperatures
(50% steam concentration).
Fig. 10. Average coal reaction rate (see text) at different steam concentrations,
temperature of 1223 K.
K. Mahalatkar et al. / Chemical Engineering Science 66 (2011) 36173627 3625
_ m
CO
rate of consumption of carbon monoxide (kg/(m
3
s))
_ m
qp
ij
mass transfer rate from the jth species of the pth phase
to ith species contained in the qth phase (kg/(m
3
s))
_ m
pq
mass transfer from the pth phase to qth phase
(kg/(m
3
s))
_ m
rs
mass transfer from the rth phase to sth phase (kg/(m
3
s))
m
0k
ultimate yield of the volatiles (kg)
m
k
amount of volatiles already evolved (kg)
M mass of Fe
3
O
4
per unit volume (kg/m
3
)
M
ox
mass per unit volume of metal oxide in fully oxidized
(ox) form (kg/m
3
)
M
red
mass per unit volume of metal oxide in fully reduced
(red) form (kg/m
3
)
MW
i
molecular weight of the ith species (kg/kmol)
MW
CH
4
molecular weight of methane (kg/kmol)
MW
CO
molecular weight of carbon monoxide (kg/kmol)
MW
H
2
molecular weight of hydrogen (kg/kmol)
MW
Fe
2
O
3
molecular weight of hematite (kg/kmol)
MW
Fe
3
O
4
molecular weight of magnetite (kg/kmol)
MW
O
2
molecular weight of oxygen (kg/kmol)
n number of particles per unit volume or particle density
n
i
number of moles of the ith species
p gas pressure (N/m
2
)
p
s
solids pressure (N/m
2
)
P
H
2
partial pressure of H
2
(N/m
2
)
P
H
2
O
partial pressure of H
2
O (N/m
2
)
r
o
grain radius of a particle (m)
r
1
reaction rate for the gasication reactions (m/s)
R universal gas constant
R
o
oxygen carrying capacity
Re Reynolds number
R
,
sg
momentum transfer between the sth solid phase and the
gas phase (N/m
3
)
R
,
rs
momentum transfer between the rth phase and the sth
phase (N/m
3
)
S
o
initial calculated particle surface area (m
2
/m
3
)
t time (s)
T temperature (K)
v
!
g
velocity of gas phase (m/s)
v
!
s
velocity of solid phase (m/s)
v
!
q
velocity of qth phase (m/s)
v
!
sg
relative velocity between gas phase and sth solid
phase (m/s)
v
!
rs
relative velocity between sth and rth solid phase (m/s)
X conversion based on fully reduced state
Y
CH
4
mass fraction of CH
4
in the fuel reactor
Y
CH
4
_TGA
mass fraction of CH
4
in thermo-gravimetric analyzer
experiments
Y
iq
mass fraction of species i in the qth phase
Y
Fe
2
O
3
mass fraction of hematite
Y
Fe
3
O
4
mass fraction of magnetite
r gradient operator
a coefcient in the watergas reaction
a
g
volume fraction of gas phase
a
q
volume fraction of the qth phase
a
p
volume fraction of the pth phase
a
s
volume fraction of the solid phase
b coefcient in the watergas reaction
b
sg
the uidsolid exchange coefcient
e
char
volume fraction of char
e
0
initial porosity of the particle
g
Ym
dissipation through collision (kJ/(m
3
s))
l
g
bulk viscosity of the gas phase (N s/m
2
)
l
s
bulk viscosity of the solid phase (N s/m
2
)
f
ls
dissipation in uid (kJ/m
3
/s)
f
s
sphericity of the particle
m
g
shear viscosity of gas phase (N s/m
2
)
m
s
shear viscosity of solid phase (N s/m
2
)
r
m
molar density of the particle (mol/m
3
)
r
g
density of gas phase (kg/m
3
)
r
s
density of solid phase (kg/m
3
)
r
q
density of the qth phase (kg/m
3
)
t
g
gas phase stress tensor (N/m
2
)
t
s
solid phase stress tensor (N/m
2
)
Y granular temperature
n
i
stoichiometric coefcients of ith species in a reaction
n
Fe
2
O
3
stoichometric coefcient for Fe
2
O
3
(n
Fe
2
O
3
12)
n
Fe
3
O
4
stoichometric coefcient for Fe
3
O
4
(n
Fe
3
O
4
8)
Acknowledgment
The authors gratefully acknowledge the nancial support of
the U.S. Department of Energy, Carbon Sequestration and Gasi-
cation Programs administered at the National Energy Technology
Laboratory. The rst author (KM) acknowledges support provided
through RDS Contract DE-AC26-04NT41817.
References
Abdulally, I., Andrus, H., Thibeault, P., Chiu, J., 2010. Alstoms chemical looping
combustion coal power technology development prototype. In: Proceedings of
the First International Chemical-looping Combustion Conference.
Ada nez, J., de Diego, L.F., Garca-Labiano, F., Gaya n, P., Abad, A., 2004. Selection of
oxygen carriers for chemical-looping combustion. Energy Fuels 18, 371377.
Andrus, H.E., Burns, G., Chiu, J.H., Liljedahl, G.N., Stromberg, P.T., Thibeault, P.R.,
2006. Hybrid Combustion-Gasication Chemical Looping Coal Power Technol-
ogy Development Phase III Final Report, Alstom Power Inc. PPL-08-CT-25,
Contract DE-FC26-03NT41866, U.S. Department of Energy, National Energy
Technology Laboratory.
Ansys-Fluent Inc., 2006. Fluent Users Manual Version 6.3.
Beetstra, R., van der Hoef, M.A., Kuipers, J.A.M., 2007. Drag force of intermediate
Reynolds number ow past mono and bi-disperse array of spheres. AIChE J. 53
(2), 489501.
Bradley, D., Lawes, M., Park, H., Usta, N., 2006. Modeling of laminar pulverized coal
ames with speciated devolatilization and comparisons with experiments.
Combust. Flame 144, 190204.
Bustamante, F., Enick, R.M., Cugini, A., Killmeyer, R., Howard, B.H., Rothenberger,
K.S., Ciocco, M., Morreale, B., Chattopadhyay, S., Shi, S., 2004. Kinetic of the
homogeneous reverse watergas shift reaction at high temperature. AIChE J.
50 (5), 10281041.
Chandel, M.K., Hoteit, A., Delebarre, A., 2009. Experimental investigation of some
metal oxides for chemical looping combustion in a uidized bed reactor. Fuel
88, 898908.
Cooper, S., Cornonella, C., 2005. CFD simulations of particle mixing in binary
uidized bed. Powder Tech. 151 (13), 2736.
De Diego, L.F., Garc ia-Labiano, L., Gaya n, P., Celaya, J., Palacios, J.M., Ada nez, J.,
2007. Operation of a 10 kW
th
chemical-looping combustor during 200 h with a
CuOAl
2
O
3
oxygen carrier. Fuel 86, 10361045.
Deng, Z., Xiao, R., Jin, B., Song, Q., 2009. Numerical simulation of chemical looping
combustion process with CaSO
4
oxygen carrier. Int. J. Greenhouse Gas Control
3, 368375.
Everson, R., Neomagus, H., Kasaini, H., Njapha, D., 2006. Reaction kinetics of
pulverized coal-chars derived from inertinite-rich coal discards: gasication
with carbon dioxide and steam. Fuel 85, 10761082.
Gera, D., Syamlal, M., OBrien, T., 2004. Hydrodynamics of particle segregation in
uidized beds. Int. J. Multiphase Flow 30, 419428.
Gidaspow, D., 1992. Multiphase Flow and Fluidization. Academic Press.
Gunn, D.J., 1978. Transfer of heat or mass to particles in xed and uidized beds.
Int. J. Heat Mass Transfer 21, 467476.
Huilin, L., Yurong, H., Gidaspow, D., 2003. Hydrodynamic modelling of binary
mixture in a gas bubbling uidized bed using the kinetic theory of granular
ow. Chem. Eng. Sci. 58 (7), 11971205.
Ishida, M., Zheng, D., Akehata, T., 1987. Evaluation of a chemical-looping combus-
tion power-generation system by graphic exergy analysis. Energy 12,
147154.
Ishida, M., Jin, H., 1994. A fundamental study of a new kind of medium material for
chemical-looping combustion. Chem. Eng. Jpn. 27, 296301.
Ishida, M., Jin, H., 1996. Chemical looping combustor without NOx formation. Ind.
Eng. Chem. 35, 24692472.
Johansson, J., Mattisson, T., Ryde n, M., Lyngfelt, A., 2006. Carbon capture via
chemical-looping combustion and reforming. In: Proceedings of the Interna-
tional Seminar on Carbon Sequestration and Climate Change, Rio de Janeiro,
October 2427, 2006.
K. Mahalatkar et al. / Chemical Engineering Science 66 (2011) 36173627 3626
Johansson, M., Mattisson, T., Lyngfelt, A., 2004. Investigation of Fe
2
O
3
with
MgAl
2
O
4
for chemical looping combustion. Ind. Eng. Chem. 43, 69786987.
Johansson, M., 2007. Screening of Oxygen-Carrier Particles based on Iron-,
Manganese-, Copper- and Nickel Oxides for use in Chemical-Looping Tech-
nologies. Chalmers University of Technology, G oteborg, Sweden.
Jung, J., Gamwo, I., 2008. Multiphase CFD-based models for chemical looping
combustion process: fuel reactor modeling. Powder Technol. 183, 401409.
Kolbitsch, P., Pr oll, T., Bolhar-Nordenkampf, J., Hofbauer, H., 2009. Characterization
of chemical looping pilot plant performance via experimental determination
of solids conversion. Energy Fuels 23, 14501455.
Knoche, K.F., Richter, H., 1968. Verbesserung der Reversibilit at von Verbrennung-
sprozessen. Brennst.-Warme-Kraft 20, 205210.
Kronberger, B., Johansson, E., Lofer, G., Mattisson, T., Lyngfelt, A., Hofbauer, H.,
2004. Two-compartment uidized bed reactor for CO
2
capture by chemical
looping combustion. Chem. Eng. Technol. 27 (12), 13181326.
Krugel-Emden, H., Rickelt, S., Stepanek, F., Munjiza, A., 2010. Development and
testing of an interconnected multiphase CFD-model for chemical looping
combustion. Chem. Eng. Sci. 65 (16), 47324745.
Langroudi, M., Turek, S., Ouazzi, A., Tardos, G., 2010. An investigation of frictional
and collisional powder ows using a unied constitutive equation. Powder
Technol. 197 (1), 91101.
Leion, H., Mattisson, T., Lyngfelt, A., 2007. The use of petroleum coke as fuel in
chemical-looping combustion. Fuel 86, 19471958.
Leion, H., Mattisson, T., Lyngfelt, A., 2008. Solid fuels in chemical looping
combustion. Int. J. Green House Gas Control 2, 180193.
Liu, G., Niksa, S., 2004. Coal conversion submodels for design applications at
elevated pressures: Part II Char gasication. Prog. Energy Combust. Sci. 30,
679717.
Lun, C.K.K., Savage, S.B., Jeffrey, D.J., Chepurniy, N., 1984. Kinetic theories for
granular ow: inelastic particles in Couette ow and slightly inelastic particles
in general ow eld. J. Fluid Mech. 140, 223256.
Lyngfelt, A., Lecknor, B., Mattisson, T., 2001. A uidized bed combustion process
with inherent CO
2
separation: an application of chemical looping combustion.
Chem. Eng. Sci. 56, 31013113.
Lyngfelt, A., Thunman, H., 2005. Construction and 100 h of operational experience
of a 10 kW chemical-looping combustor. In: Thomas, D.C., Benson, S.M. (Eds.),
Carbon Dioxide Capture for Storage in Deep Geologic Formations. Elsevier,
Amsterdam, pp. 625645.
Mahalatkar, K., Kuhlman, J., Huckaby, E.D., OBrien, T., 2010. Computational uid
dynamic simulations of chemical looping fuel reactors utilizing gaseous fuels.
Chem. Eng. Sci. 66 (3), 469479.
Marion, J.L., 2006. Technology options for controlling CO
2
emissions from fossil
fueled power plants. In: Proceedings of the Fifth Annual Conference on Carbon
Capture and Sequestration, Alexandria, VA, May 810, 2006.
Mattisson, T., Lyngfelt, A., Cho, P., 2001. The use of iron oxide as an oxygen carrier
in chemical-looping combustion of methane with inherent separation of CO
2
.
Fuel 80, 19531962.
Mattisson, T., Abanades, J., Lyngfelt, A., Abad, A., Johansson, M., Ada nez, J., Garcia-
Labiano, F., de Diego, L.F, Gayan, P., Kronberger, B., Hofbauer, H., Luisser, M.,
Palacios, J.M., Alvares, D., Orjala, M., Heiskanen, V.P., 2005. Capture of CO
2
in
Coal Combustion ECSC Coal RTD Programme Final Report, ECSC-7220-PR125.
Merrick, D., 1983. Mathematical models for thermal decomposition of coal: 1. The
evolution of volatile matter. Fuel 62 (5), 534539.
Metz, B., Davidson, O., de Coninck, H., Loos, M., Meyer, L., IPCC Special Report on
Carbon Capture and Storage. Intergovernmental Panel on Climate Change,
Working Group III, 2005.
Reddy, R., Joshi, J., 2009. CFD modeling of solid-liquid uidized beds of mono and
binary particle mixtures. Chem. Eng. Sci. 64 (16), 36413658.
Richter, H.J., Knoche, K.F., 1983. Reversibility of combustion processes. In: Gaggioli,
R.A. (Ed.), ACS Symposium Series: Vol. 235. Efciency and Costing. Second Law
Analysis of Processes. American Chemical Society, Washington, DC, pp. 7185.
Ryu, H., Bae, D., Han, K., Lee, S., Jin, G., Choi, J., 2001. Oxidation and Reduction
characteristics of oxygen carrier particles and reaction kinetics by unreacted
core model. Korean J. Chem. Eng. 18, 831837.
Siriwardane, R., Poston, J., Chaudhari, K., Zinn, A., Simonyi, T., Robinson, C., 2007.
Chemical-looping combustion of simulated synthesis gas using nickel oxide
oxygen carrier supported on bentonite. Energy Fuels 21, 15821591.
Smoot, L.D., Smith, P.J., 1985. Coal Combustion and Gasication. Plenum Press,
New York.
Solomon, S., Qin, D., Manning, M., Kristen, M., Marquis, M., Averyt, K., Tignor, M.,
Miller, H., 2007. Climate Change 2007: The Physical Science Basis, Fourth
asessment report of the intergovernmental panel on climate change.
Son, S., Kim, S., 2006. Chemical-Looping combustion with NiO and Fe
2
O
3
in a
thermo balance and circulating uidized bed reactor. Ind. Eng. Chem. 45,
26892696.
Syamlal, M., Rogers, W., OBrien, T., 1993. MFIX documentation and theory guide,
US-Department of Energy Report No. DOE/METC-94/1004.
Syamlal, M., Bissett, L., 1992. METC gasier advanced simulation model,
US-Department of Energy Report No. DOE/METC-92/4108.
Szekely, J., Evans, W., Sohn, H., 1976. GasSolid Reactions. Academic Press.
Taylor, G.I., 1954. The dispersion of matter in turbulent ow through pipes. Proc. R.
Soc. London. Ser. A, Math. Phys. Sci. 223, 1155.
Van Wachem, B.G.M., Schouten, J., Van den Bleek, C., Krishna, R., Sinclair, J., 2001.
Comparitive analysis of CFD models of dense gassolid systems. AIChE J. 47
(5), 10351051.
Wen, C.Y., Yu, H.Y., 1966. Mechanics of uidization. Chem. Eng. Prog. Symp. Ser.
62, 100.
Wolf, J., Anheden, M., Yan, J., 2001. Performance analysis of combined cycles with
chemical looping combustion for CO
2
capture. In: Proceedings of the 18th
International Pittsburgh Coal Conference, pp. 11221139.
K. Mahalatkar et al. / Chemical Engineering Science 66 (2011) 36173627 3627

S-ar putea să vă placă și