Você está na página 1de 13

Energy Conversion and Management 86 (2014) 10101022

Contents lists available at ScienceDirect

Energy Conversion and Management


journal homepage: www.elsevier.com/locate/enconman

CFD analysis of bubble hydrodynamics in a fuel reactor


for a hydrogen-fueled chemical looping combustion system
Atal Bihari Harichandan, Tariq Shamim
Institute Center for Energy (iEnergy), Department of Mechanical and Materials Engineering, Masdar Institute of Science and Technology, P.O. Box 54224, Abu Dhabi, United
Arab Emirates

a r t i c l e

i n f o

Article history:
Received 4 September 2013
Accepted 10 June 2014
Available online 10 July 2014
Keywords:
CFD simulation
Chemical looping combustion
Fuel reactor
Kinetic theory of granular ow
Eulerian method

a b s t r a c t
This study investigates the temporal development of bubble hydrodynamics in the fuel reactor of a
hydrogen-fueled chemical looping combustion (CLC) system by using a computational model. The model
also investigates the molar fraction of products in gas and solid phases. The study assists in developing a
better understanding of the CLC process, which has many advantages such as being a potentially promising candidate for an efcient carbon dioxide capture technology.
The study employs the kinetic theory of granular ow. The reactive uid dynamic system of the fuel
reactor is customized by incorporating the kinetics of an oxygen carrier reduction into a commercial computational uid dynamics (CFD) code. An Eulerian multiphase treatment is used to describe the continuum two-uid model for both gas and solid phases. CaSO4 and H2 are used as an oxygen carrier and a fuel,
respectively. The computational results are validated with the experimental and numerical results available in the open literature. The CFD simulations are found to capture the features of the bubble formation,
rise and burst in unsteady and quasi-steady states very well. The results show a signicant increase in the
conversion rate with higher dense bed height, lower bed width, higher free board height and smaller oxygen carrier particles which upsurge an overall performance of the CLC plant.
2014 Elsevier Ltd. All rights reserved.

1. Introduction
The concern about the effect of greenhouse gases on the global
climate changes is growing worldwide. Anthropogenic carbon
dioxide (CO2) emissions are the major contributor and a signicant
part of it is due to the combustion of various carbonaceous fuels in
power generating plants. Accordingly, the development of novel
technologies to control CO2 emissions with minimum energy
losses has become a topic of strategic interest for researchers. In
addition to developing renewable energy conversion technologies,
the use of CO2 capture and storage (CCS) with the use of fossil fuels
will make a signicant contribution to the reduction of CO2 emissions. Many CO2 capture methods such as amine scrubbing, oxyfuel combustion, and decarbonization of fuels, are currently under
investigation. Chemical looping combustion (CLC) introduced by
Richter and Knoche [1] is an efcient technology in which CO2 is
inherently separated during the combustion process.
A CLC system consists of two separated reactors (air and fuel
reactors). A suitable oxygen carrier is oxidized (storing oxygen)
in the air reactor and is transported to the fuel reactor using a
Corresponding author. Tel.: +971 (2) 810 9158; fax: +971 (2) 810 9901.
E-mail address: tshamim@masdar.ac.ae (T. Shamim).
http://dx.doi.org/10.1016/j.enconman.2014.06.027
0196-8904/ 2014 Elsevier Ltd. All rights reserved.

cyclone. In the fuel reactor, the metal oxide is reduced and the
released oxygen reacts with the fuel. Carbon-dioxide and water
(H2O) vapors are the two ue gases generated in the fuel reactor.
Afterwards, H2O is condensed from the ue gases and an uncontaminated stream of CO2 is captured and stored for later use.
Fig. 1 shows a schematic of CLC process. The abbreviations
MeO/Me represent the metal oxide in oxidized/reduced forms,
respectively. In the recent past, this novel technology has been
studied extensively [213] focusing on different distinct aspects
such as technical feasibility, economic assessments, incorporation
of the CLC system in power generation cycles, numerical simulation of the system with solid and gas fuels, and the experimental
improvement of the CLC system.
In a CLC system, the distribution of solidgas particles in the
reactors (which are generally based on the uidized bed reactor
design) is highly complex and can be better understood by using
computational uid dynamics (CFD) models based on the uid
dynamics and reaction kinetic mechanisms. The CFD models can
be used to simulate the behavior of the reactants and the products
in the reactors during unsteady and quasi-steady states by using
the conservation principles of mass, momentum, energy and species. The difculties of capturing the ow physics in a relatively
complex circulating uidized bed (CFB) require signicantly large

A.B. Harichandan, T. Shamim / Energy Conversion and Management 86 (2014) 10101022

1011

Nomenclature
CD
Di,mix
Dturb
egg
F
~
Fa
~
F lift;a
~
F v m;a
g0,gg
G
I2S
~
J ai
Kgl
_ ab
m
_ ab
m
P
~
Rab
Rai
Reg
Sa
Sai
Scturb
tg
u0g

vr,g

~
va
~
v ab

drag function
diffusion coefcient of the mixture
turbulent diffusivity
coefcient of restitution for particle collision
factor that depends on CD and Reg
external body force
lift force
virtual mass force
radial distribution function
acceleration due to gravity
second invariant of the stress tensor
diffusion ux of species i
multi-phase exchange coefcient
mass transfer from phase a to phase b
mass transfer source between phases
pressure
interaction force between phases
net rate of production of homogeneous species i by
chemical reaction for phase a
relative Reynolds number
source term in continuity equation
rate of creation
turbulent Schmidt number
particulate relaxation time
uctuating velocity of the particles
terminal velocity correlation for the granular phase
velocity of phase a
inter-phase velocity

computational efforts for complete modeling. With more advanced


computing facilities and better numerical methods, CFD is now
becoming a popular means for the CFB modeling. Jung and Gamwo
[14], Deng et al. [15], Kruggel-Emden et al. [16], Shuai et al. [17]
and Mahalatkar et al. [8] have performed the CFD studies of fuel
reactors using gaseous fuels. Numerical studies for fuel reactors
using solid fuels were reported by Wang et al. [18], Mahalatkar
et al. [19] and Garcia-Labiano et al. [20].
There have been a limited number of investigations on the
unsteady aspects of bubble formation, rise and burst in the fuel
reactor of a hydrogen-fueled CLC system. One relevant work was
performed by Deng et al. [15], who numerically simulated the bubbling uidized bed reactor of a hydrogen-fueled CLC system to
study the quasi-steady features of bubble dynamics. However,
the temporal development of bubbles inside the reactor, which is

Y ai

the local mass fraction and net rate of production of


homogeneous species i

Greek Symbols
a
volume fraction
bgd
inter-phase momentum transfer for calculating drag between solid phase and gas phase
cg
dissipation of uctuating energy due to inelastic collision
Hg
granular temperature
ka
bulk viscosity of phase a
la
shear viscosity of phase a
lg
granular shear viscosity
lg,col
granular collisional shear viscosity
lg,fr
granular frictional viscosity
lg,kinetic granular kinetic viscosity
lturb
turbulent viscosity
q
density
sa
ath phase stressstrain tensor
/f
internal friction angle
Subscripts
a,b
phases (solid and gas)
dg
diameter of particles in solid phase
g
granular phase
gas
gas phase
i, j
species
l
liquid phase

very important particularly at the beginning of reaction and the


subsequent stages before attaining the quasi-steady state, has
not been investigated. Hence, the primary objective of the present
study is to investigate the temporal development of bubble hydrodynamics of reactive ow in a fuel reactor of a hydrogen-fueled
CLC system by employing a two-dimensional multiphase CFD
model.
The present study is carried out by using calcium sulde (CaS)
as an oxygen carrier and hydrogen (H2) as a fuel. To keep the focus
on the investigation of the bubble hydrodynamics, H2was selected
as the fuel gas due to its simpler reaction kinetics. The hydrogen
fueled CLC process produces pure steam in the fuel reactor. This
steam can be expanded to a very low pressure before being used
in the gas turbine cycle of a power plant for power production. It
has an excellent performance without the formation of NOx [21].
The understanding of the bubble hydrodynamics in a hydrogenfueled reactor will also be useful for the analysis of the higher
hydrocarbon-fueled fuel reactors.
Fig. 2 represents the schematic of a CLC system with two interconnected reactors in which the reduction and the oxidation processes take place. Calcium sulfate (CaSO4) is reduced to CaS and
it is then oxidized back to CaSO4. The present study does not consider the formation of SO2 and H2S. The CLC process described in
the present study has two chemical reactions as follows:

CaSO4 4H2 ! CaS 4H2 O


reduction of oxygen carrier in the fuel reactor

CaS 2O2 ! CaSO4


oxidation of oxygen carrier in the air reactor
Fig. 1. Schematic view of chemical looping combustion process.

The reactions described in Eqs. (1) and (2) are both exothermic in
nature with low level energy releases at low-temperature region

1012

A.B. Harichandan, T. Shamim / Energy Conversion and Management 86 (2014) 10101022

2.1.2. Momentum equations


The momentum equation for phase a is:

@
aa qa~
ma r  aa qa~
ma~
ma aa rp r  sa aa qa~
g
@t
n


X
~
_ ba~
Rba m
mba  m_ ab~
mab

b1



~
Fa ~
F lift;a ~
F mm;a

a is the ath phase stressstrain tensor.


where s

2
3

sa aa la r~ma r~mTa aa ka  la r  ~maI

Here, la and ka are the shear and bulk viscosities of phase a, ~


F a is
an external body force, ~
F lift;a is a lift force, ~
F mm;a is a virtual mass
force, ~
Rba is an interactive force between phases, p is the pressure
shared by all the phases, and ~
mba is the interphase velocity. If mass
_ ba > 0), ~
mba = ~
ma ;
transfer takes place from phase b to phase a (i.e., m
_ ba < 0),
whereas for mass transfer from phase a to phase b (i.e., m
~
mba = ~
mb . Similarly, if m_ ab > 0 then ~
mab = ~
ma , if m_ ab < 0 then ~
mab = ~
mb .
Rba is inuenced by friction, pressure, cohesion, and
The value of ~
other effects. Based on the physical conditions, ~
Rba ~
Rab and
~
Raa = 0. The velocity gradients in the gaseous (primary) phase ow
F lift;a ) that act on
eld in a multiphase ow generate lift forces (~

Fig. 2. CLC with two interconnected uidized bed reactors.

(6001200 K) for the rst reaction and high-temperature region


(8001700 K) for the second reaction, respectively. The present
study considers the simulations of the fuel reactor only (it is shown
with the dotted line in Fig. 2). In this study, the reaction kinetic
for CaSO4 in Eq. (1), which is a rst-order reaction with respect
to the hydrogen partial pressure, has been calculated on the basis
of a shrinking-core model [22] with the activation energy of
151,000 J/mol and the reaction kinetic constant of 4300. In addition to the investigation of bubble hydrodynamics, the paper also
presents the effects of some important physical parameters on the
performance of a CLC system.

2. Mathematical model
The study employs the kinetic theory of granular ow with an
Eulerian approach. The continuity, momentum and energy equations are solved for each phase whereas a single pressure is shared
by all the phases.

solid particles. These forces are more substantial for bigger particles. In the present model, the lift forces on solid particles are insignicant due to their smaller resisting surface area and in
comparison to larger drag forces. So, these forces are not included
in the model. The virtual mass force (~
F v m;a ) is considered when a
solid phase (secondary, b) moves faster relative to the gas phase
(primary, a). The solid particles are likely to experience a virtual
mass force due to the inertia of gas phase caused by the fast moving particles (bubbles or droplets). As the gas phase is denser than
the solid phase, this force is not very important and is not incorporated in the present model. Furthermore, the adhesive forces
between particles were not considered for the present simulations.
2.1.3. Species transport equations
The local mass fraction (Y ai ) of ith species in a multiphase ow is
predicted by the species conservation equation, which is described
as:



@  a a a
q a Y i r  qa aa~
ma Y ai r  aa~Jai aa Rai aa Sai
@t
n
X


_ bi aj  m
_ aj bi R
m

b1

2.1. Hydrodynamic model


The general governing equations considered for the unsteady
and multiphase ow are given as follows:

2.1.1. Continuity equations


The continuity equation for phase a is:
n
X
@
aa qa r  aa qa~
ma m_ ba  m_ ab Sa
@t
b1

where Rai is the rate of formation of identical species i by chemical


_ aj bi is the mass transfer source between species
reaction for phase a, m
i and j from phase a to b, and R is the heterogeneous reaction rate. In
addition, aa is the volume fraction for phase a and Sai is the rate of forJ a is the diffusion ux of species i which is present due to difmation.~
ferential concentration and temperature of species. The mass
diffusion due to differential concentration is modeled by using the
dilute approximation (also called Ficks Law) as dened in Eq. (7):

where aa, qa and ~


ma are the density and velocity of phase a, respec_ ba represents the mass transfer from the bth to ath phase,
tively. m
_ ab symbolizes the mass transfer from phase a to phase b.
and m
The last term Sa is the source term. In a heterogeneous reaction,
the exchange of mass, momentum and heat between the gassolid
phases are considered in the continuity equation.


~
J ai  qDi;mix

lturb
Scturb

rY ai

Scturb is selected to be 0.7 for calculations.


2.1.4. Energy equations
The energy equation for phase a is:

@
aa qa ha r  aa qa ua ha rka rT a Q ab Sab hab
@t

1013

A.B. Harichandan, T. Shamim / Energy Conversion and Management 86 (2014) 10101022

where h, k and Q are enthalpy, thermal conductivity of mixture and


heat exchange between gas phase and solid phase, respectively. Qab
is the heat transfer when phase b changes to phase a. The heat
exchange between two phases is a function of the temperature difference and is expressed as:

Q ab hab T a  T b

where

A a4:14
l
(
and B

where hab is the heat transfer coefcient. Also, Qab = Qba.

tg
2.1.5. Kinetic theory of granular ow
A random granular motion of particles is likely to occur due to
the collisions of solid particles inside the reactor. The transport
equation for the solid phase [23] is described by the following
equation:


@
2
g : r~
ug r:kg rHg
ag qg Hg r:ag qg Hg ug  pgI ag s
@t
3
10
 c  3bggas Hg
Here, Hg is the granular temperature and is given by: Hg 13 u0g u0g ,
where u0g is the uctuating velocity of the particles and can be
derived from u0g V g  ug , and pg is the solid pressure. bggas is the
interphase momentum transfer for calculating the drag between
two phases. The well-known Ergun equation [24] is considered for
agas < 0.8 to describe the dense region in the reactor:

bggas 150

1  agas r g lgas

agas d2g

1:75

qgas ag jugas  ug j
dg

11

If agas > 0.8, the drag coefcient is calculated by the equation proposed by Wen and Yu [25]:

bggas

3 jugas  ug j 2:65
Cd
agas
4
dg

12

where

Cd

24
1
Re

0:15Re0:687 ; Re 6 1000

0:44;

and Re

Re > 1000

jugas  ug jagas qgas dg

lgas

13

14

2.2. Closure models


2.2.1. Interphase exchange coefcient
The constitutive closure models suitable for the above mentioned governing equations are described below:
The soliduid interchange exchange coefcient in its general
form is expressed as:

K gl
f

ag qg f

15

tg

C D Reg al
24m2r;g

16

where f is a factor that depends on the drag function (CD) and relative Reynolds number (Reg) [26].

and C D

Reg

mr;g

4:8
0:63 p
Reg =mr;g

ql dg j~
mg  ~
ml j
ll

19

!2

0:8a1:28
; al 6 0:85
l
a2:65
;
al > 0:85
l

qg d2g
18ll

21

where tg is the particulate relaxation time.


2.2.2. Fluidsolids drag force
The motion of solid particles through a viscous uid inside the
fuel reactor often experiences a resistance due to interphase drag
forces between the solidliquid phases. The solidgas interactions
are dened by the interphase momentum exchange and the drag
correlation established on the settling of beds and the terminal
velocity of uids as:

F gasg F ggas

q
2
3lgas ag agas 
0:63 Reg =mr;g 4:8
2
4mr;g dp

pg ag qg Hg 2qg 1 egg a2g g 0;gg Hg

18

23

where egg is the coefcient of restitution for particle collisions, g 0;gg


is the radial distribution function, and Hg is the granular temperature that is proportionate to the kinetic energy of the uctuating
particle motion.
2.2.4. Radial distribution function
The prospect of collisions between granules in a dense solid
granular phase medium is described by the radial distribution
function. This function describes the non-dimensional distance
between granules (assumed to be spherical).

g0

"

 #1
3
ag 1=3
1
5
ag;max

24

2.2.5. Solid shear stresses


The momentum exchange due to particle translation and collision inside the reactor gives rise to the solid shear stress tensor
that has components of the shear and the bulk viscosities. For granular particles with the maximum solid-volume-fraction, plastic
transition of particles occurs and an additional component of frictional-viscosity is considered to account for it. Thus, the solid shear
viscosity is expressed as:

lg lg;col lg;kinetic lg;fr


4
5


q
0:5 A  0:06Reg 0:06Reg 2 0:12Reg 2B  A A2

22

2.2.3. Solids pressure


For the cases of granular ows having smaller solid volume
fractions than the maximum possible value (compressible regime),
a solid pressure is considered for the pressure gradient term in the
momentum equation, which is composed of a kinetic term and a
particle collision term as:

lg;col ag qg dg g 0;gg 1 egg


17

20

lg;kinetic
lg;fr

ag dg qg Hg p
63  egg

pg sin /f
p
2 I2S

25
 1=2

ag

2
1 1 egg 3egg  1ag g 0;gg
5

26

27

28

1014

A.B. Harichandan, T. Shamim / Energy Conversion and Management 86 (2014) 10101022

/ and I2D are the angle of internal friction and second invariant of
the stress tensor respectively.
2.2.6. Granular temperature
The diffusion coefcient for granular energy kg is given by:

kg

2
150qg dg Hg p
6
1 ag g 0 1 e 2qg c2g dg g 0 1 e
5
3841 eg 0
r


Hg

29

p
"

cg 31  e2 a2g qg dg g 0 Hg

4
dg

r!

Hg

#
 rug

3. Results and discussions

30

where cg is the dissipation of uctuating energy due to inelastic


collision.
2.3. Numerical considerations
A nite volume method based on phase-coupled SIMPLE (PCSIMPLE) algorithm [27] has been used to solve the unsteady multiphase ow problems. The commercial CFD software code FLUENT
was used for solving the coupled equations by adopting a second
order upwind differencing scheme. The numerical solutions were
obtained by using a convergence criterion of 105 for each scaled
residual component and a time step of 5  104 s.
Fig. 3 shows the schematic and grids used for the fuel reactor.
The width of the reactor was 0.25 m. The computational domain
of the reactor was discretized by 2500 quadrilateral cells. This
numerical grid was selected based on the ndings reported by
many researchers that a numerical cell size of 10 times larger than
the particle-size can capture the bubble hydrodynamics in uidized bed reactors accurately [28]. The grid independence test performed in the present work also has the similar observations like
that of Gelderbloom et al. [28]. The fuel reactor was lled with
metal oxide particles up to a height of 0.4 m. Only H2 fuel gas
was fed to the reactor at the bottom section. The dispersed k  e
turbulence model was used. The heat transfer coefcient between
the gas phase and the solid phase used for the present simulation
were proposed by Gunn [24]. The restitution coefcient between
the solid particles was 0.9. The boundary conditions were xed
by specifying the inlet velocity and the outlet pressure values at

Fig. 3. Schematic and grid of the fuel reactor.

the reactor inlet and outlet respectively. In the CFB reactor, the
soliduid mixture behaves as a uid medium due to introduction
of pressurized uid through the particulate medium. For simplicity, no-slip condition at wall was used for both the phases. However, the case of solid particulate having some nite slip along
the wall can also be simulated by employing different values of
specularity coefcient in the numerical model. The convective
terms are evaluated by using a second order QUICK scheme. The
model parameters used for the base case in the present study are
similar to those used by Deng et al. [15].

A multiphase CFD model based on the kinetic theory of granular


ow has been numerically simulated to describe the hydrodynamics and chemical reactions of dense gassolid ows. The solid particles are considered to be smooth, spherical, inelastic and
undergoing binary collisions. The shrinking-core model [22] for
the grain geometry is considered for calculating the reaction kinetics in the fuel reactor. 100 wt.% of H2 is fed as the feed fuel gas into
the reactor through a distributor. The oxygen carrier present in the
reactors static bed and the supplied fuel gas at the reactor inlet are
highly mixed by the momentum of upward moving gas bubbles.
The reaction takes place between the gaseous reactant (H2) and
the desorbed oxygen (O2) of the oxygen carrier. In the process,
H2O in the gaseous form is produced and oxygen carrier is reduced.
The simulations are carried out for different dimensions of the
reactor and for different particle sizes of the granules.
Fig. 4 shows the temporal proles of gas phase mole fractions of
H2 (reactant) and H2O (product) along the centerline of the reactor
(x = 0 cm) and (a) at a height of 30 cm from the inlet (y = 30 cm)
(dense bed region) and at (b) the outlet. For validation of the simulation results, the results reported by Deng et al. [15] with similar
simulation parameters are also shown in the gure. One of the differences in this work and Deng et al. [15] is the use of more accurate discretization scheme (second order) in the present
simulations. Another difference is the use of pressure outlet
boundary condition at the reactor outlet in the present simulations
as opposed to the use of outow boundary condition by Deng et al.
[15]. The temporal proles of reactant and product molar fractions
are found to be in good agreement with the results of Deng et al.
[15]. The oscillations of reactant and product observed in
Fig. 4(a) are due to the movement of bubbles and high reaction
rates in the dense bed region. The reactant (H2) and product
(H2O) molar fractions oscillate around 0.68 and 0.32, respectively.
In the initial phase (up to 1.0 s), the reactant mole fraction (H2)
decreases rapidly from unity. Beyond 1.0 s, the reactant mole fraction does not decrease and it oscillates around 0.7. Similar variation of the product mole fraction is observed, which increases
rapidly from zero to 0.32 at around 1.0 s. The result shows that
the reaction reached quasi-steady state after 1.0 s. At the reactor
outlet, the reactant and product mole fraction proles achieve
steady values of 0.65 and 0.35 respectively after 3.8 s. The constant
proles of H2 and H2O at the outlet are expected because of
unavailability of metal oxides in the upper portion of the reactor.
Thus, the conversion rate of the fuel gas is 35% under the simulation condition.
To develop a better understanding of the hydrodynamic and
reacting environment in the fuel reactor, Fig. 5 shows the development of solid volume fraction proles in the fuel reactor from 0 to
1.0 s. As the reaction starts, smaller bubbles grow larger by merging with the adjacent smaller bubbles. Merging of bubbles is also
observed at the upper portion of the dense uidized bed, just
below the interface of static bed and the free board regions. In
the dense bed region, the bubbles formed near the distributer tend
to move upward and, as a result, two columns of vertically offset

A.B. Harichandan, T. Shamim / Energy Conversion and Management 86 (2014) 10101022

1015

Fig. 4. Mole fractions in gas phase along the centerline of the reactor (x = 0 cm) and (a) at a height of 30 cm from the inlet (y = 30 cm) (dense bed region) and at (b) the outlet.

bubbles are formed. The smaller bubbles are noticed to trail behind
the larger ones. Due to a low pressure zone created at the wake of
the larger bubbles, the trailing smaller bubbles accelerate upward
and coalesce with the larger bubbles. Also, the upward moving larger bubbles, having higher velocity than the smaller ones, coalesce
with the neighboring smaller bubbles. In the process, the bubble
sizes grow quickly in the narrow ow passage area of the reactor
and slugs are formed. Thus, the solid particles are pushed upward
by the rising slugs. Later, the solid particles are pushed down into
the bed along the central core of the reactor as well as along the
walls of the reactor due to gravity and the differential density of
gassolid particles. These phenomena are noticeable from 0.5 s to
1.0 s. At 1.0 s, a core-annulus region is formed similar to that
observed experimentally by Clift and Grace [29] for slug ow reactors. This study used the reactors of different heights and widths
for different supercial velocities. The cross-section of the reactor
used by Clift and Grace [29] was also rectangular (similar to the
present simulation) where solid particles were pushed upward
by the rising slugs which subsequently moved back down along
the walls of the reactor forming a core-annulus section.
Fig. 6 shows the development of solid volume fraction proles
beyond 1.0 s. The bubbles and slugs have low solid volume fractions (nearly equal to zero). Due to the difference in gas velocities
inside the bubbles and the slugs, the reaction rates are different in
the respective regions. They are signicantly higher in the emulsion region due to larger concentration of solid particles. The reaction rates are lower in the fast moving bubbles. This transient
behavior of bubble dynamics is observed within the rst 1.0 s after
which the quasi-steady state condition is attained. The continuous

supply of the feed fuel gas through the distributer ensures the global mixing of gassolid particles in the reactor with the continuous
bubble formation, rise and burst. However, the core-annulus feature as observed during the unsteady period does not prevail for
longer time. This salient feature of bubble dynamics is clearly
noticeable up to 1.6 s after which the core-annulus structure bursts
into intimate mixing of gassolid particles and the quasi-steady
state condition is achieved.
Fig. 7 shows the quasi-steady state distribution of (a) solid volume fraction, and the gas phase mole fractions of (b) reactants and
(c) products at 10 s. A global mixing of the gas and particles is
observed. The qualitative representation of the bubble hydrodynamics is shown in Fig. 7(a) where the blue color1 represents the
pure gas and the red color denotes the dense gassolid mixture.
The reactant (H2) prole, shown in Fig. 7(b), implies that the H2 mole
fraction decreases linearly from about 1.0 around the distributor
(inlet) to 0.65 at the interface between the dense uidized and the
free-board regions. A sudden increase in its magnitude to a constant
value of 0.7 has been observed in the free-board region. As expected,
the reverse characteristics are observed for the product (H2O) prole
as shown in Fig. 7(c). Similar observations were also observed by
Deng et al. [15] and the proles presented in Fig. 7 are in qualitative
agreement with their results.
The study also investigated the effects of reactor geometric
dimensions and particle size on the bubble hydrodynamics and

1
For interpretation of color in Fig. 7, the reader is referred to the web version of
this article.

1016

A.B. Harichandan, T. Shamim / Energy Conversion and Management 86 (2014) 10101022

Fig. 5. Unsteady variation of solid volume fraction from time 0 to 1.0 s.

the reactor fuel conversion performance. The following sections


describe the results of these investigations.
3.1. Effect of dense bed heights
Fig. 8 shows the variations of solid volume fraction proles for
different dense bed heights at the quasi-steady state (t = 10.0 s).
The static bed heights in the uidized bed region are selected as
10, 20, 30, 40, 50 and 60 cm with a constant total bed height
(100 cm). The gure shows a qualitative representation of the bubble dynamics in the form of formation, rise and burst. In the dense
bed region, characterized by a roughly constant porosity, the
phenomena of bubble hydrodynamics are prominent in the lower
one-third part whereas in the upper two-third part, the reaction
efciency decreases [30]. Thus, the fuel gas conversion rate in the
upper zone of the dense bed is limited by the gas transfer between
bubbles and emulsion. The growth of bubble size has been
observed to decrease considerably with the increase of static bed
heights thereby increasing the reaction rate in the reactor. However, the bubble size shows an increasing trend for the reactors
with lower (<30 cm) and higher (>50 cm) static bed heights.

Fig. 9 depicts the temporal variation of mole fractions of H2 at


the outlet and the conversion rate of the feed fuel gas at 10.0 s
for different dense bed heights. It is observed from Fig. 9(a) that
the H2 mole fractions at the outlet decrease with an increase in
the static bed height. For small dense bed heights, the H2 mole
fraction at the outlet remains almost constant till the bed height
of 40 cm. However, for higher bed heights such as 50 and 60 cm,
the H2 mole fraction has been observed to vary with time due to
relatively higher concentration of oxygen carrier. Also, the time
span over which the H2 molar fraction decreases at the outlet
(before attaining a certain constant value) is bigger for reactors
with smaller dense bed heights. Fig. 9(b) implies the increase of
the conversion rate with the static bed height which is also
reected from solid volume fraction contours.
3.2. Effect of bed widths
Fig. 10 shows the variation of solid volume fraction proles for
different reactor bed widths at the quasi-steady state (t = 10.0 s).
For all the cases, the static bed height and the total height of the
reactor are kept xed at 40 cm and 100 cm, respectively. The bed

A.B. Harichandan, T. Shamim / Energy Conversion and Management 86 (2014) 10101022

1017

Fig. 6. Unsteady development of solid volume fraction contour with quasi-steady reaction rate.

Fig. 7. (a) Solid volume fraction, (b) mole fraction of H2 and (c) mole fraction of H2O at the quasi-steady state conditions (t = 10 s).

widths (w) for different cases are selected as 10, 15, 20, 25 and
30 cm. The ow Reynolds numbers in all the cases are kept similar
to the base case. The global mixing of gassolid particles is
observed to be more prominent for the reactors with smaller

widths. The velocities of gas and emulsion in these reactors are relatively higher than the reactors with the higher widths. Thus, the
gas and the emulsion tend to grow over the entire span of the reactor. Alternate layers of gas and emulsion are noticed in the smaller

1018

A.B. Harichandan, T. Shamim / Energy Conversion and Management 86 (2014) 10101022

Fig. 8. Solid volume fraction contour with different dense bed heights (H) at the quasi-steady state conditions (t = 10 s).

Fig. 9. (a) Mole fraction of H2 and (b) conversion rate for different dense bed heights of the reactor at the quasi-steady state conditions (t = 10 s).

width reactors. However, for a bed width of 10 cm, the characteristics of volume fraction contours are somewhat different as no
core-annulus region was formed and many bubbles were observed
with most of the oxygen carrier particles accumulated near the
wall. Moreover, the upper portions of the reactors with high widths
are predominantly covered by the gas particles. This causes the net
reaction rate to be lower due to unavailability of oxygen carrier
particles.
Fig. 11 shows the temporal variations of H2 mole fractions at the
outlet and the conversion rate of the feed fuel gas at the quasisteady state (t = 10.0 s) for different reactor bed widths. The results
depict that the H2 mole fraction at the outlet increases and consequently the fuel conversion rate decreases with an increase in the
reactor bed width. The H2 molar fraction at the outlet and the fuel
conversion rate become nearly constant for reactors with bed
widths P20 cm.
3.3. Effect of free board heights
Fig. 12 shows the variations of solid volume fraction contours
for different heights of free board region at the quasi-steady state

(t = 10.0 s). For these simulations, the static bed height and the
bed width of the reactor are kept xed at 40 cm and 25 cm, respectively. The heights of the free board region for different cases are
selected as 40, 60, 80, 100 and 120 cm. The results show that an
increase in the height of the free board region increases the mixing
of gassolid mixture and reduces the bubble size. This increases
the H2 conversion rate in the reactor. The reactor free board height
also affects the gas concentration prole in the fuel reactor mainly
due to differences in the uid dynamics of the dense bed and the
free board regions. The metal oxide particle concentration in the
free board region decreases with an increase in the reactor height.
However, this increases the reaction rate in this region due to a
better gassolid contact than in the dense bed.
Fig. 13 shows the temporal variations of H2 mole fractions at the
outlet and the conversion rate of the feed fuel gas at the quasisteady state (t = 10.0 s) for different reactor free board heights.
The results show that the H2 mole fraction at the outlet decreases
with an increase in the reactor free board region height. As
expected, the H2 molar fraction decreases rapidly from unity to
certain constant values. These observations are also similar to the
experimental results reported by Kolbitsch et al. [31] and Abad

A.B. Harichandan, T. Shamim / Energy Conversion and Management 86 (2014) 10101022

1019

Fig. 10. Solid volume fraction contour with different bed widths: (a) w = 10 cm, (b) w = 15 cm, (c) w = 20 cm, (d) w = 25 cm, (e) w = 30 cm at the quasi-steady state conditions
(t = 10 s).

Fig. 11. (a) Mole fraction of H2 and (b) conversion rate at for different bed widths of the reactor at the quasi-steady state conditions (t = 10 s).

et al. [30] for the chemical-looping combustion of methane. The


results show that the reactors with lower free board heights experience faster decrease of molar fraction of feed fuel gas at the start
of the simulation. The conversion rate of H2 increases with an
increase in the reactor free board heights.
3.4. Effect of particle sizes
Fig. 14 shows the variation of solid volume fraction contour for
different particle sizes at the quasi-steady state (t = 8.0 s). For all
the cases, the static bed height, free board height and the width
of the reactor are kept xed at 40 cm, 60 cm and 25 cm, respectively. The particle sizes for different cases are selected as 0.1,
0.2, 0.3, 0.4 and 0.5 lm. The results show that the global mixing
of gassolid particles is signicantly higher for the reactors with
smaller particle sizes. Furthermore, the gas and emulsion velocities

in the reactors with smaller particles are also relatively higher.


Thus, the gas and emulsion tend to spread over a large part of
the reactor. However, the reactors with bigger particles, in spite
of having more oxygen carriers in the dense bed region, experience
very low reaction rates due to improper mixing of gassolid
particles.
Fig. 15 shows the temporal variations of H2 mole fraction at the
outlet and the conversion rate of the feed fuel gas at the quasisteady state (t = 8.0 s) for different particle sizes. The results show
that the H2 mole fraction at the outlet is higher for cases with larger particle sizes. Hence, the conversion rates for these cases are
lower as shown in Fig. 15(b). This is due to a relatively lower mixing of the gassolid particles which gives rise to lower reaction
rates. The results also indicate that the feed fuel gas conversion
rate can further be increased by using nano-size oxygen carrier
particles.

1020

A.B. Harichandan, T. Shamim / Energy Conversion and Management 86 (2014) 10101022

Fig. 12. Solid volume fraction contour with constant bed width (w = 25 cm) and different free board heights: (a) h = 40 cm, (b) h = 60 cm, (c) h = 80 cm, (d) h = 100 cm, (e)
h = 120 cm at the quasi-steady state conditions (t = 10 s).

Fig. 13. (a) Mole fraction of H2 and (b) conversion rate for different free board heights of the reactor at the quasi-steady state conditions (t = 10 s).

Fig. 14. Solid volume fraction contour with different particle sizes: (a) Dp = 0.1 lm, (b) Dp = 0.2 lm, (c) Dp = 0.3 lm, (d) Dp = 0.4 lm, (e) Dp = 0.5 lm at the quasi-steady state
conditions (t = 8 s).

A.B. Harichandan, T. Shamim / Energy Conversion and Management 86 (2014) 10101022

1021

Fig. 15. (a) Mole fraction of H2 and (b) conversion rate for different particle sizes in the reactor at the quasi-steady state conditions (t = 8 s).

4. Conclusions

References

A CFD model has been developed and employed for the simulations of the fuel reactor in a chemical looping combustion (CLC)
system. The model included the reaction kinetics of fuel and metal
oxide. The governing equations were solved by using a commercial
CFD code. The numerical model was employed to investigate the
temporal development of the bubble hydrodynamics and the
unsteady aspects of bubble formation, rise and burst in the fuel
reactor of a hydrogen-fueled CLC system. The molar fractions of
the product in gas and solid phases were also investigated by the
numerical model.
The computational results were validated qualitatively and
quantitatively with the numerical results of Deng et al. [15] and
the experimental results of Clift and Grace [29]. The bubble
dynamics and the ow patterns were in good agreement with
the literature results. A low conversion rate of the feed fuel gas
was obtained which is due to combined effect of low bed temperature, bigger particle size, smaller dense bed height and larger bed
width of the reactor. All these parameters led to fast and large bubbles in the reactor which reduced the reaction rate and hence the
conversion rate of the feed fuel gas. The low fuel conversion can
be improved by (i) the use of an oxygen carrier with small particle
sizes and high reactor temperature; (ii) the use of a different oxygen carrier material, for example Ni; and (iii) the recirculation of
the unutilized hydrogen to the fuel reactor.
The paper also presents the results of a parametric study based
on the reactor dimensions and particle sizes. The results show that
a signicant increase in the fuel conversion rate can be obtained
with a higher dense bed height, a lower bed width, a higher free
board height and smaller oxygen carrier particles. However, additional cost must be considered for larger dense bed and higher free
board heights. The smaller oxygen carrier particles (preferably of
nano-size) are a preferred option for enhancing the overall performance of CLC plant as they do not generate large bubbles and a
proper mixing of gassolid particles can very well be established.
The direct numerical simulation of the present multiphase reactive model will be considered for the future work in order to get
better bubble hydrodynamics.

[1] Richter HJ, Knoche KF. Reversibility of combustion processes. ACS Symposium
Series 1983;235:7185.
[2] Abad A, Mattison T, Lyngfelt A, Ryden M. Chemical looping combustion in a
300 W continuously operating reactor system using a manganese based
oxygen carrier. Fuel 2006;85:117485.
[3] Johansson E, Mattisson T, Lyngfelt A, Thunman H. A 300 W laboratory reactor
system for chemical-looping combustion with particle circulation. Fuel
2006;85:142838.
[4] Mattisson T, Garcia-Labaiano F, Kronberger B, Lyngfelt A, Adanez J, Hofbauer H.
Chemical-looping combustion using syngas as fuel. Int J Greenhouse Gas
Control 2007;1(2):15869.
[5] Song Q, Xiao R, Deng Z, Zhang H, Shen L, Xiao J, et al. Chemical-looping
combustion of methane with CaSO4 oxygen carrier in a xed bed reactor.
Energy Convers Manage 2008;49(11):317887.
[6] Forero CR, Gayan P, de Diego LF, Abad A, Garcia-Labiano F, Adanez J. Syngas
combustion in a 500 Wth chemical-looping combustion system using an
impregnated Cu-based oxygen carrier. Fuel Process Technol 2009;90(12):
14719.
[7] Dennis JS, Scott SA. In situ gasication of a lignite coal and CO2 separation
using chemical looping with a Cu-based oxygen carrier. Fuel 2010;7(89):
162340.
[8] Mahalatkar K, Kuhlman J, Huckaby ED, OBrien T. Computational uid dynamic
simulation of chemical looping fuel reactors utilizing gaseous fuels. Chem Eng
Sci 2011;66:46979.
[9] Hassan B, Shamim T, Ghoniem AF. A parametric study of multi-stage chemical
looping combustion for CO2 capture power plant. In: Proceedings of the ASME
2012 Summer Heat Transfer Conference, Puerto Rico; 2012. Paper ASME201258597.
[10] Moldenhauer P, Ryden M, Mattisson T, Lyngfelt A. Chemical-looping
combustion and chemical-looping with oxygen uncoupling of kerosene with
Mn- and Cu- based oxygen carriers in a circulating uidized bed 300 W
laboratory reactor. Fuel Processing Technol. 2012;104:37889.
[11] Wang B, Zhao H, Zheng Y, Liu Z, Yan R, Zheng C. Chemical looping combustion
of a Chinese anthracite with Fe2O3-based and CuO-based oxygen carriers. Fuel
Processing Technol 2012;96:10415.
[12] Abad A, Gayan P, de Diego LF, Garcia-Labiano F, Adanez J. Fuel reactor
modelling in chemical-looping combustion of coal: 1. model formulation.
Chem Eng Sci 2013;87:27793.
[13] Peltola P, Ritvanen J, Tynjl T, Hyppnen T. Model-based evaluation of a
chemical looping combustion plant for energy generation at a pre-commercial
scale of 100MWth. Energy Convers Manage 2013;76:32331.
[14] Jung J, Gamwo IK. Multiphase CFD based models for chemical looping
combustion process: fuel reactor modeling. Powder Technol 2008;183:4019.
[15] Deng Z, Xiao R, Jin B, Song Q. Numerical simulation of chemical looping
combustion process with CaSO4 oxygen carrier. Int J Greenhouse Gas Control
2009;3:36875.
[16] Kruggel-Emden H, Rickelt S, Stepanek F, Munjiza A. Development and testing
of an interconnected multiphase CFD-model for chemical looping combustion.
Chem Eng Sci 2010;65:473245.
[17] Shuai W, Guodong L, Huilin L, Juhui C, Yurong H, Jiaxing W. Fluid dynamic
simulation in a chemical looping process with two interconnected uidized
beds. Fuel Processing Technol 2011;92:38593.
[18] Wang X, Jin B, Zhong W, Zhang Y, So Min. Three-dimensional simulation of a
coal gas fueled chemical looping combustion process. Int J Greenhouse Gas
Control 2011;5:1498506.

Acknowledgements
The work is nancially supported by the MIT/Masdar Institute
Collaborative Research funds (grant # 10MAMA1).

1022

A.B. Harichandan, T. Shamim / Energy Conversion and Management 86 (2014) 10101022

[19] Mahalatkar K, Kuhlman J, Huckaby ED, OBrien T. CFD simulation of a


chemical-looping fuel reactor utilizing solid fuel. Chem Eng Sci 2011;66:
361727.
[20] Garcia-Labiano F, de Diego LF, Gayan P, Abad A, Adanez J. Fuel reactor
modelling in chemical-looping combustion of coal: 2. Simulation Optim Chem
Eng Sci 2013;87:17382.
[21] Jin H, Ishida M. A novel gas turbine cycle with hydrogen-fueled chemical
looping combustion. Int J Hydrogen Energy 2000;25:120915.
[22] Zafar Q, Abad A, Mattisson T, Gevert B. Reaction kinetics of freeze granulated
NiO/MgAl2O4 oxygen carrier particles for chemical-looping combustion.
Energy Fuels 2007;21:6108.
[23] Patil DJ, Annaland MS, Kuipers JAM. Critical comparison of
hydrodynamic models for gas-solid uidized beds-Part I: bubbling
gas-solid uidized beds operated with a jet. Chem Eng Sci 2005;60:
5772.
[24] Gunn DJ. Transfer of heat or mass to particles in xed and uidized beds. Int J
Heat Mass Transfer 1978;21:46776.

[25] Wen CY, Yu HY. Mechanics of uidization. Chem Eng Prog Symp Ser
1966;62:100.
[26] Syamlal M, OBrien TJ. Computer simulation of bubbles in a uidized bed.
AlChE Symp Ser 1989;85:2231.
[27] Vasquez SA, Ivanov VA. A phase coupled method for solving multiphase
problems on unstructured meshes. In: Proceedings of ASME FEDSM00: ASME
2000 Fluids Engineering Division Summer Meeting, Boston; 2000
[28] Gelderbloom SJ, Gidaspow D, Lyczkowski RW. CFD simulations of bubbling/
collapsing uidized beds for three geldart groups. AlChE J. 2003;49:84458.
[29] Clift R, Grace JR. Continuous Bubbling and Slugging. London: Academic Press;
1985.
[30] Abad A, Adanez J, Garcia-Labiano F, de Diego LF, Plata A, Gayan P. Modeling of
the chemical-looping combustion of methane using a Cu-based oxygencarrier. Combust Flame 2010;157:60215.
[31] Kolbitsch P, Proll T, Hofbauer H. Modeling of a 120 kW chemical looping
combustion reactor system using a Ni-based oxygen carrier. Chem Eng Sci
2009;64:99108.

Você também pode gostar