j.energy.2014.10.003

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Energy 77 (2014) 932e944

Contents lists available at ScienceDirect

Energy
journal homepage: www.elsevier.com/locate/energy

CFD (computational fluid dynamics) analysis of a novel reactor design


using ion transport membranes for oxy-fuel combustion
Pervez Ahmed a, *, Mohamed A. Habib a, Rached Ben-Mansour a, Patrick Kirchen b,
Ahmed F. Ghoniem c
a
KACST TIC #32-753, KACST and Mechanical Engineering Department, KFUPM, Dhahran, 31261, Saudi Arabia
b
Department of Mechanical Engineering, University of British Colombia, Vancouver, BC, V6T 1Z4, Canada
c
Department of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge, MA, 02139, USA

a r t i c l e i n f o a b s t r a c t

Article history: Conventional two-channel ITM (ion transport membrane) reactors applied to oxy-combustion, face the
Received 2 February 2014 potential drawback of high thermal gradients and high local temperatures, which can result in mem-
Received in revised form brane damage. In such reactors, air flows on the feed side and fuel are introduced on the permeate side,
25 September 2014
where it reacts with the permeated oxygen. In this work, we propose to use a three-channel configu-
Accepted 1 October 2014
Available online 30 October 2014
ration in which a porous plate is used to separate the permeate stream from the fuel stream, allowing the
fuel to diffuse gradually into the permeate side. The gradual combustion of the fuel results in a slow
temperature rise and a more spatially uniform temperature distribution along the membrane. We model
Keywords:
Ion transport membrane
this three-channel reactor using computational fluid dynamics and compare its performance to a con-
Reactors ventional two-channel reactor. It is shown that, indeed, in case of a two-channel reactor, a high tem-
Oxy-fuel combustion perature zone is concentrated near the inlet, whereas the three-channel reactor produces a milder
temperature gradient along the reactor length. The more-uniform heat flux associated with the latter
results in a moderate temperature distribution and reduction in the wall shear stress along the channels
and the associated pressure drop. The more uniform temperature distribution should be less damaging to
the membrane. The reaction zone associated with the gradual fuel diffusion into the sweep side improves
the membrane performance by maintaining a more uniform oxygen flux.
© 2014 Elsevier Ltd. All rights reserved.

1. Introduction perovskite membranes can also be used in simultaneous oxidation


of a fuel in order to enhance the oxygen separation by reducing the
It has been demonstrated that dense perovskite membranes oxygen partial pressure on the permeate side. However, an oxida-
operating at elevated temperatures, when subjected to an oxygen tion environment (or reduction of the membrane surface) may
partial pressure gradient, exhibit high oxygen permeability [1e3]. cause chemical instability unless the elements used and the stoi-
These membranes transport oxygen from the feed side to the chiometry as well as the operating conditions are carefully chosen.
permeate side, i.e. from the high oxygen partial pressure side to low Though it is believed that membrane technology offers promising
oxygen partial pressure side without the aid of high feed pressure solution in the design of novel reactors [7], significant work re-
or electrical current, making the separation process simpler and mains in examining their operation and performance under con-
more efficient than pressure swing absorption or electrochemical ditions similar to those expected in practical applications, including
techniques, respectively. Successful development of air separation the development of detailed and rigorous models.
units incorporating ITMs can be a significant step towards novel O2 Coroneo et al. [8] performed CFD modeling of inorganic mem-
production and oxy-combustion systems [4,5]. Numerous in- branes for gas separation. The analysis showed that the transport
vestigations are being conducted to enhance the chemical stability mechanism has great impact on the permeation rate. For the case of
and performance of the material under challenging operational hydrogen permeation, molecular diffusion is the predominant
conditions, e.g., higher temperature and fuel oxidation/membrane mechanism at high pressure. Habib et al. [5] reviewed the perfor-
reduction environment [6]. Apart from oxygen separation these mance of combustors utilizing ion transport membranes for oxy-
fuel combustion, and their integration in power cycles. Wang
* Corresponding author. Tel.: þ966 13 860 7869. et al. [9] studied the permeation of oxygen in a tubular Ba0.5Sr0.5-
E-mail address: pervezahmed@kfupm.edu.sa (P. Ahmed). Co0.8Fe0.2O3-d membrane, confirming that at constant temperature,

http://dx.doi.org/10.1016/j.energy.2014.10.003
0360-5442/© 2014 Elsevier Ltd. All rights reserved.
P. Ahmed et al. / Energy 77 (2014) 932e944 933

the oxygen permeation flux increases with increasing the partial include transport phenomena of momentum, heat and species with
pressure of oxygen at the shell side. In their recent work, Ben- chemical reactions. The flow is laminar in all channels of both co-
Mansour et al. [10] studied the characteristics of oxy-fuel com- feed and near isothermal OTRs. The Reynolds number for the high-
bustion in an oxygen transport reactor and found that the heat of est flow conditions was 500. Species transport and chemical re-
reaction is mostly convected to the feed side and that fuel con- actions are occurring due to combustion of methane fuel. In addition,
version is reduced with increasing the mass flow rates. Mancini and there is oxygen transport across the ITM (Ion Transport Membrane)
Mitsos [11] developed an intermediate fidelity model of the reactor, and transport of gaseous fuel across the porous membrane to control
a compromise between detailed oxygen transport models [12], the fuel distribution and eventually the oxy-combustion process in
simplified models [13], and reduced fuel oxidation kinetic models. order to obtain isothermal conditions in the ITM and hence reduce
In their investigations, Teraoka et al. demonstrated that, at the the thermal stresses in the fragile (thin) ITM. Given the complexity of
same temperatures, the oxygen flux obtained using LSCF mem- the transport and reactive processes, and since we are developing
branes are 2e4 times higher as compared to zirconia stabilized- the model for multi-dimensional (2D and 3D) geometries, we opted
membranes. Their studies paved the way to research on synthesis to start with simplified but accurate models in order to understand
and analysis of LSCF membranes [14]. Farooqui et al. [15] modeled the physical processes better and be able to properly design the OTR
cylindrical ion transport reactor for co-current and counter current for best conditions discussed above. Given the above assumptions,
flows and found that oxygen to fuel mass ratio increases as the the transport model equations for 2D-steady-state laminar flow
percentage of CO2 increases in the sweep gas. In a series of papers, with species transport can be expressed as:
Hong et al. [16e18] developed accurate and comprehensive anal-
ysis of the coupling between heat and mass transfer on both sides Conservation of mass : V$ðrUÞ ¼ 0 (1)
of the membrane. They developed an oxygen flux expression
combining the impact of adsorption/desorption and ion diffusion,  
Conservation of Momentum : V$ rUU ¼ Vp þ mV2 U (2)
as well as the contribution of gas-phase chemistry and surface fuel
reactions in the case when fuel is flown on the permeate side. Flux
measurements [19] and oxygen distributions in the permeate side, Conservation of Energy :
as well as extensive validation of these models were presented in !
  X !   (3)
Ref. [20]. rCp f U$VT ¼ V$ keff VT  hi J i þ teff $U þSh
One of the advantages of perovskite membranes is the high pu- i
rity of oxygen produced in the separation process that can be used in
!
several applications. In Refs. [21e23] the application of mixed- where keff is the effective thermal conductivity, J i is the diffusion
conducting perovskite membranes in reactors for oxidative flux of source i. The term (keffVT) represents energy transfer due to
P !
coupling of methane was presented. In these reactors the membrane conduction, ( hi J i ) represents species diffusion, (teff $U) repre-
provides the required oxygen in ionic form to prevent deep oxida- i
sents the viscous dissipation and Sh includes the heat of reaction
tion of the methane, as would occur in a conventional co-fed cata- along with the radiation source term [35]. The heat of reaction Sh,rxn
lytic reactor. The membrane not only eliminates the need for an air is calculated as:
separation units, it also provides better products yield [9,13,24e26].
Some membrane materials are chemically unstable in a fuel- X h0
i
oxidative or membrane-reducing environment [27]. Moreover, Sh;rxn ¼  Ri
i
Mi
phase decomposition or membrane degradation leads to a reduction
in the oxygen flux and failure [28]. The development of membranes
where h0i and Ri is the enthalpy of formation and the volumetric
with high stability in a reducing environment is important.
rate of creation of species i respectively.
In most previous studies, disc shaped dense perovskite mem-
branes [13,22,27,29] were used in the analysis. While convenient in  
Conservation of Species : V$ðrUYi Þ ¼ V$ rDi;m VYi þRi þSi (4)
laboratory work, these disc shaped membranes are not represen-
tative of practical systems [27]. Two-channel (2-channel) hollow where Ri is the net rate of production of species i due to the reaction
tubular membranes [30,31] have also been explored for separation and Si the source/sink term that takes into account the mass flow of
and reaction. In our previous work on 2-channel reactors [32,33], species across the membrane; a sink on the feed side and a source
we found that when feeding fuel at the inlet, the temperature ex- on the permeate side. The diffusion coefficient in the gaseous
periences a spike and most of the fuel is consumed there. This fast mixture, Di,m is calculated as [35].
and uncontrolled reaction raises the risk of membrane failure. The
need to achieve a uniform temperature along the membrane mo- 1  Xi
tives the use of multiple channels, e.g., the three channel (3- Di;m ¼ ! (5)
P Xi
channel) configuration introduced in this paper. Mancini et al.
Di;j
[34] have also proposed a multi-chamber ITM reactor to try and j;jsi
address the issue as well. The present study aims at investigating,
The source/sink term, which is non-zero only at the boundaries,
numerically, the characteristics of this concept and the methods to
accounts for the oxygen permeation across the membrane as
manage the thermal conditions in reactive membrane operation.
follows:
2. Numerical modeling 8 J $A
>
> O2 cell
<þ V
> At permeate side
Co-feed (2-channel) and the concept of near isothermal reactor cell
Si ¼ (6)
(3-channel) are investigated in the present study (see Fig. 1). The co- >
> J $A
>
:  O2 cell
feed reactor consists of two channels separated by an ITM while the At feed side
Vcell
near isothermal reactor is divided into 3 channels using an ITM and a
porous plate for oxygen separation and fuel feeding respectively. The For the heat transfer we are considering all modes of heat
physical processes in these OTRs (Oxygen Transport Reactors) transfer including surface and gas radiation. Initially we neglected
934 P. Ahmed et al. / Energy 77 (2014) 932e944

Fig. 1. (a) Co-feed reactor or 2-channel reactor. (b) Near isothermal reactor or 3-channel reactor.

radiation but found out that the temperature levels are very high. oxygen permeation through an LSCF ITM can be written in terms of
After including radiation the peak temperature dropped by the oxygen flux as [13]:
200e400 K depending on the fuel concentration. Therefore we
opted to include the radiative heat transfer by solving the full  0:5  00 0:5
radiative transfer equations which takes care of both surface and Dv kr 0
PO  PO2
2
gas radiation. The radiative transfer equation for non-scattering JO2 ¼  0:5  0:5  0:5  0:5  (8)
participating media with a unity index of refraction can be writ- 00
2Lkf PO2 0
PO þ P 0 þ P
00
Dv
2 O 2 O 2
ten as:
In Ref. [13], the global values of the partial pressure were used.
   4  
! ! sT ! ! In the work of Hong et al. [16e18], it was shown that the local value
V$I r ; s ¼ k I r ; s (7)
p at the membrane surface must be used in order to separate the
contributions of the mass transfer resistances and those associated
! !
where I is the radiation intensity and are respectively r ; s the with the permeation process. This modification was validated
position vector in a given coordinate system and the direction along experimentally in Hunt et al. [20].The flux equation represented by
which the intensity is resolved. equation (8) has been validated in our earlier work and the details
Regarding the chemical reactions modeling, two phenomena of Dv, Kr, Kf constants can be found in Ref. [36].
need to be considered: the transport of oxygen through the ITM and In terms of the combustion modeling we have opted for the
the combustion of methane and oxygen on the permeate side of the generalized finite-rate model [Fluent 6.3]. This approach is based
reactor. The oxygen permeation through the ITM we are consid- on the solution of transport equations for species mass fractions.
ering a “zero order” model which accurately the overall permeation The reaction rates appear as source terms in the species transport
process including the adsorption and dissociation of oxygen on the equations are calculated from Arrhenius rate expressions. The
air (feed) side, the solid state diffusion of oxygen ions and electrons chemical kinetic mechanisms used in these reactions use the
through the thickness of the membrane and the recombination of FLUENT database. These mechanisms have been checked against
oxygen on the permeate side of the membrane. This model initially other databases such as CHEMKIN which is widely accepted in the
published by Xu and Thomson [13]. This zero order model for chemical engineering and combustion community. Due the
P. Ahmed et al. / Energy 77 (2014) 932e944 935

complexity of the problem and the number of equations being 4. Boundary conditions
solved over a fine grid, we used a single step reaction mechanism in
the present calculation. The one-step kinetic reaction mechanism is Steady state laminar flow is assumed in all channels. The mass
solved on the permeate side of the ITM for combustion of CH4 with flow rate of the air and the fuel streams is kept constant at a value of
permeated O2. The one-step kinetic reaction rate is given by: 3E-2 kg/s and 1.5E-3 kg/s, respectively, throughout the study. Air is
assumed to consist of N2 and O2 only by mass percentage of 77 and
RCH4 ¼ k½CH4 nCH4 ½O2 nO2 (9) 23 respectively. In the bottom channel, a diluted fuel stream that
consists of CH4 and CO2 is flown. Oxidizing fuel with O2 rather than
where k is the Arrhenius reaction rate given by, air results in very high temperatures. Therefore, CO2 is added to the
  fuel stream to reduce the combustion temperature. On the other
Ea
k ¼ AT b exp (10) hand, removing N2 (79% of air by volume) reduces the volume flow
RT rate to the power equipment (e.g., turbines) which is compensated
The pre-exponential factor A ¼ 2.119  1011, the activation en- for by the added CO2. In a practical system, CO2 would be obtained
ergy Ea ¼ 2.027  108, the temperature exponent b ¼ 0 and the rate in the form of recirculated flue gases. In case of the 3-channel
exponents for CH4 and O2 are taken as nCH4 ¼ 0.2 and nO2 ¼ 1.3 reactor, the sweep gas (CO2) is passed through the middle chan-
respectively. nel to avoid any reverse flow. A value of 1E-6 kg/sis assigned for the
In terms of the flowecombustion interaction, we opted for the mass flow rate of the sweep gas (CO2) in the middle channel (i.e.
laminar finite-rate model since the flow is laminar in all our sweep channel). The temperature, oxygen partial pressures and
calculations. therefore oxygen flux along the length of the membrane are ex-
The above model equations along with the boundary conditions pected to vary. In order to achieve nearly uniform combustion
described in the next section in detail; constitute the mathematical throughout the length of the reactor, the same stoichiometric ratio
model. This model is numerically integrated using the Finite Vol- of CH4/O2 has to be maintained. For this purpose, the porous plate
ume approach widely used by many CFD (computational fluid dy- separating the sweep stream and the fuel stream is divided into 10
namics) codes. The main advantage of the Finite Volume method is parts with different porosities. The porosities of the 10 porous
the conservative approach, which assures conservation of mass, membranes range from 2.47E-20 to 2.47E-12 m2 and are assigned
momentum and species even for an approximate solution. Since ascending order of porosities from the inlet to the exit of the
this approach is already implemented in FLUENT code, we used reactor. The porosities are chosen in order to match the expected
FLUENT for the solution of the model equations. oxygen flux across the membrane and hence to obtain uniform
However, FLUENT does not have an implementation of an ITM stoichiometric ratio along the length of the reactor. The porous
membrane. This was achieved by developing the required model plates are designated as P_1, P_2, P_3…….to P_10 and are given the
code in a separate user-defined-function. This UDF (user- values of face permeabilities ranging from 2.47E-20 m2, 3.47E-13
defined-function) is compiled and linked to Fluent to be an in- m2, 1.47E-12 m2….to 2.47E-12 m2 respectively. These values are
tegral part of the computer model. This membrane model is selected to obtain the desired uniform stoichiometric ratio
coupled with Fluent as source/sink term (Si) that take into ac- throughout the length in the middle channel. Both the ITM and the
count the mass flow of oxygen across the membrane as given in porous plates are 1 mm thick. While BSCF membranes [9,29,37,38]
equation (6). yield higher fluxes, they are less stable at higher temperatures, and
The one-dimensional (1D) Porous Jump model based on Darcy's an LSCF membrane is used in this study. Perovskite membranes
law [35] is used to model the flow through the porous plate sepa- such as LSCF, BaCeO3, BaZrO3 have thermal conductivities ranging
rating the fuel stream from the sweep/permeate stream: from 2 to 12 W/m/K in the temperature range of 500e1000  C
 
m 1
Dp ¼  n þ C2 rn2 Dm (11)
a 2 Table 1
Boundary conditions used in the present study.

3. Geometry

The geometry of a 2-channel reactor and a 3-channel reactor are


shown in Fig. 1(a) and (b) respectively. In the 2-channel reactor, the
air and fuel streams are separated by an ITM. The total height of is
taken as 60 mm with each channel corresponding to 30 mm. Air
flows in the top channel over the membrane surface. A dilute fuel
stream (CH4 þ CO2) flows through the bottom channel. Fig. 1(b)
presents the 3-channel reactor concept for oxy-fuel combustion.
The reactor is divided into three channels using an ITM and a
porous plate. The top, middle and bottom channels represent air,
sweep and fuel zones, respectively, with a height of 30 mm each.
The length of the channels is 1000 mm and the total height of the
reactor is 90 mm. Air is fed into the top channel. Oxygen passes
through the ITM to the sweep side. A sweep gas (CO2) is introduced
through the middle channel. A stream containing the fuel is
introduced through the bottom channel whose outlet is blocked to
maintain the pressure of the stream. This pressurized mixture
passes through the porous membrane to reach the middle sweep
channel where it reacts with the permeating oxygen. The top and
the bottom walls are assumed adiabatic.
936 P. Ahmed et al. / Energy 77 (2014) 932e944

Fig. 2. Temperature contours of cases with radiation and without radiation for a mass flow rate of fuel mixture Mf ¼ 0.0015 kg/s and CH4 ¼ 1%.

[39e47]. For the present study, the thermal conductivity and simulation of gas separation in few previous works [8,52e54].
density of the ITM are 4 W/m/K and 6000 kg/m3 [48,49] respec- Since the 3-channel reactor is a novel concept, no experimental
tively. The temperature of all the gases at the inlet is 1173 K. The data is available for validation and this work aims to explore the
temperature of 1173 K [50] yields relatively high oxygen perme- concept through comparison with a conventional 2 channel reactor
ation fluxes without degrading LSCF stability [51].The porous configuration. Validation of a 2-channel reactor model was per-
membrane is modeled using the Porous Jump boundary condition formed previously [36]. The boundary conditions used in the pre-
[35] as described in the earlier section. CFD has been applied in the sent study are summarized in Table 1.

Fig. 3. Reaction rate contours of cases with radiation and without radiation for a mass flow rate of fuel mixture Mf ¼ 0.0015 kg/s and CH4 ¼ 1%.
P. Ahmed et al. / Energy 77 (2014) 932e944 937

Fig. 4. Comparison of normal profiles of reaction rates (a, c, e) and temperatures (b, d, f) for with radiation and without radiation cases for a mass flow rate of fuel mixture
Mf ¼ 0.0015 kg/s and CH4 ¼ 1%.

5. Results and discussions

The aim of the present study is to compare the characteristics of


a conventional 2-channel reactor and the proposed 3-channel
nearly isothermal reactor for oxy-fuel combustion. Laminar
steady flow is considered for the present study. We focus on the
flow and combustion in the middle channel (i.e. the permeate
channel) where the Reynolds number corresponding to a mass flow
rate of 1.5E-03 kg/s is < 500. In the top channel, (i.e. the air feed
channel) the Reynolds number corresponding to the air mass flow
rate of 3.0E-02 kg/s is 1676. Therefore, the flow is laminar. Many
combustion studies [32,33,55] omit the use of radiation models in
their investigations either because its role is less important or to
save computational time. The effect of radiation in the case
considered here may be important because of the use of wider
channels as shown next.

5.1. Effect of radiation

The effect of radiation is investigated by comparing the cases of


the 3-channel reactor with and without radiation. The internal
emissivities of wall and membrane are taken as 0.8. The WSGGM Fig. 5. Comparison of Wall temperatures along the ITM length for with radiation and
(Weighted Sum of Gray Gases Model) has been used to model the without radiation cases for a mass flow rate of fuel mixture Mf ¼ 0.0015 kg/s and
absorption coefficient (1/m) of the gases. CH4 ¼ 1%.
938 P. Ahmed et al. / Energy 77 (2014) 932e944

Fig. 6. Comparison of (a) radiative heat fluxes and (b) convective heat fluxes along the ITM wall for with radiation and without radiation cases for a mass flow rate of fuel mixture
Mf ¼ 0.0015 kg/s and CH4 ¼ 1%.

The temperature contours with radiation and without radiation considered, the reaction zone rates shifts further away from the
are presented in Fig. 2. The contours are presented on the same ITM at positions further downstream in the reactor.
scale and the color bar (in web version) at the bottom highlights the The higher reaction rates and temperatures without radiation
temperature variation. Without radiation, the temperature is lead to higher wall temperatures, as shown in Fig. 5. The heat flux
higher. The maximum temperature without radiation is 1335 K
whereas with radiation it is 1219 K. With radiation model included,
surface radiation and participating gaseous radiation increases the
heat transfer and thereby reduces the temperatures. Moreover, it is
observed that inclusion of radiation model results in lower reaction
rates. Fig. 3 shows the reaction rate contours with and without
radiation. The contours are presented using the same scale and
color bar (in web version). Because of the higher temperature
without radiation, the reaction rates are almost doubled, the re-
action zone is narrower and the reaction zone shifts away from the
ITM towards the porous plates. With lower temperatures in radi-
ation case, lower reaction rates are achieved as reaction rates are a
function of temperature (see equation (10)).
The temperature and reaction rate distributions across the
channel are shown at a distance of 300 mm, 600 mm and 900 mm
respectively from the entrance of the reactor are presented in Fig. 4.
Lower gas temperatures and reaction rates are observed when
radiative heat transfer is considered. The position of the reaction
zone is also noted to be dependent on the radiation, as it moves
closer to the ITM when radiation is considered. If radiation is not Fig. 8. Comparison of wall shear stresses along the ITM wall for with radiation and
without radiation cases for a mass flow rate of fuel mixture Mf ¼ 0.0015 kg/s and
CH4 ¼ 1%.

Fig. 7. Comparison of total heat fluxes along the ITM wall for with radiation and Fig. 9. Comparison of partial pressures of O2 along the ITM wall on both feed and
without radiation cases for a mass flow rate of fuel mixture Mf ¼ 0.0015 kg/s and permeate side for 2-channel and 3-channel reactors for a mass flow rate of fuel
CH4 ¼ 1%. mixture Mf ¼ 0.0015 kg/s and CH4 ¼ 1%.
P. Ahmed et al. / Energy 77 (2014) 932e944 939

5.2. Comparison between the 2-channel and 3-channel reactors

The partial pressure of oxygen on the feed and permeate sides of


the ITM for both the 2-channel (2C) and 3-channel (3C) reactors are
presented in Fig. 9. The corresponding oxygen permeation fluxes
are presented in Fig. 10.The partial pressure of oxygen on the feed
side of the membrane starts at a high value of 0.21 bar (i.e. at the
atmospheric pressure). It decreases slightly and remains almost
constant along the membrane length for both cases. On the other
hand, the partial pressure of oxygen on the permeate side starts at
zero at the inlet and increases quickly to a value of approximately
0.01 bar. However, a closer look at the plot reveals that it is slightly
higher in case of the 3-channel reactor. For a significant length of
the ITM in the 2-channel reactor the partial pressure of oxygen on
the permeate side is lower compared to the 3-channel reactor case.
The difference in partial pressures of oxygen drives the oxygen flux
across the membrane. The oxygen permeation flux across the
membrane is presented in Fig. 10. The flux in the 2-channel case is
Fig. 10. Comparison of oxygen permeation flux along the ITM wall on permeates side high in the first part and gradually decreases. As the oxygen partial
for 2-channel and 3-channel reactors for a mass flow rate of fuel mixture pressure builds up on the permeate side, the flux begins to reduce.
Mf ¼ 0.0015 kg/s and CH4 ¼ 1%.
On the other hand, the partial pressure of oxygen on the permeate
side in the 3-channel reactor is relatively higher. As the partial
on the ITM is presented in Fig. 6. Fig. 6(a) and (b) shows the radi- pressure starts increasing, the oxygen flux reduces accordingly. The
ative heat flux and the convective heat flux for the cases without average permeation flux in the 2-channel reactor is only 20%
and with radiation. The negative heat flux implies energy leaving (approximately) more than the 3-channel reactor. Nevertheless,
the gas, therefore, with radiation the magnitude of the flux is less unlike the 2-channel reactor, the oxygen permeation flux in case of
than without radiation. Hence, lower ITM wall temperatures (Fig. 5) 3-channel case is uniform all along the length. This is useful in
are observed while using radiation model. Radiation contributes achieving uniform stoichiometric ratio (using porous membrane to
significantly to the overall heat fluxes, and raises the convective feed the fuel) and uniform combustion all along the reactor length.
heat flux due to additional heat transfer by surface and partici- Fig. 11 shows the temperature contours for the 2-channel and 3-
pating gaseous radiation. As a result, the total heat flux, presented channel reactors. The corresponding reaction contours are pre-
in Fig. 7, is lower, explaining the lower temperatures observed in sented in Fig. 12. The contour levels for both cases are compared on
previous sections. the same scale and the color bar (in web version) highlights the
The shear stresses along the ITM wall are shown in Fig. 8. These temperature variations. These temperature contours shown in
shear stresses increase exponentially in the first 50 mm of the Fig. 11 indicate that in the case of a 2-channel reactor the maximum
reactor and remain almost constant beyond this point. Higher shear temperature is 1228 K which is about 12 K (only 1%) higher than
stresses are predicted when radiation model is included in that in the 3-channel case. However, the 2-channel reactor yields a
modeling the reactors, thus providing more realistic results of the high temperature zone that is concentrated near the inlet, whereas
reactor operations. the 3-channel reactor produces a longer milder high temperature

Fig. 11. Comparison of contours of temperature for 2-channel and 3-channel reactors for a mass flow rate of fuel mixture Mf ¼ 0.0015 kg/s and CH4 ¼ 1%.
940 P. Ahmed et al. / Energy 77 (2014) 932e944

Fig. 12. Comparison of contours of reaction rates for 2-channel and 3-channel reactors for a mass flow rate of fuel mixture Mf ¼ 0.0015 kg/s and CH4 ¼ 1%.

zone distributed over much of the reactor length. The reaction length. Moreover, it is apparent from these reaction contours that in
contours, presented in Fig. 12 indicate that the reaction takes place case of 2-channel reactor the reaction takes place close to the
only in the first half of the reactor in the 2-channel case while in the membrane whereas in the 3-channel reactor it is further away. As
3-channel reactor the reaction occurs throughout the reactor we move down the reactor length, the reaction rate decreases in 2-

Fig. 13. Comparison of normal profiles of reaction rates (a, c, e) and temperatures (b, d, f) for 2-channel and 3-channel reactors for a mass flow rate of fuel mixture Mf ¼ 0.00015 kg/s
and CH4 ¼ 1%.
P. Ahmed et al. / Energy 77 (2014) 932e944 941

Fig. 14. Comparison of ITM wall temperatures for 2-channel and 3-channel reactors for Fig. 16. Comparison of total heat flux for 2-channel and 3-channel reactors on both the
a mass flow rate of fuel mixture Mf ¼ 0.0015 kg/s and CH4 ¼ 1%. feed side and the permeate side for a mass flow rate of fuel mixture Mf ¼ 0.0015 kg/s
and CH4 ¼ 1%.

channel reactor due to drop in the fuel concentration. On the


contrary, fuel is continuously being fed in the 3-channel reactor channel case may lead to failure or performance penalty. Li et al.
using porous plates that can overcome the drawback of using a 2- reported that phase decomposition or membrane degradation
channel reactor. Moreover, the reaction zone near the ITM in 2- leads to reduction in oxygen permeation fluxes causing membrane
channel reactor may be a limiting factor for the stability of the failure [28].
membrane. The reaction rate contours and reaction rate profiles show that
Fig. 13 shows the reaction rate profiles and the corresponding in case of the 2-channel reactor the reaction takes place close to the
temperature profiles at a distance of 300 mm, 600 mm, and ion transport membrane reducing the oxygen partial pressure and
900 mm respectively from the entrance of both the 2-channel and increasing permeation flux in the first part of the reactor. In the 3-
3-channel reactors. Fig. 13(a) clearly indicates that in case of 2- channel case, by adjusting the amount of CH4 in the fuel mixture
channel reactor the reaction takes place closer to the membrane (fed through the porous membrane), the reaction zone is shifted
surface, which results in higher temperatures than in the 3-channel away from the membrane thereby reducing its potential harmful
reactor. The higher the reaction rate is associated with higher ox- effects on the material.
ygen consumption, causing its partial pressure to drop (as shown in Fig. 14 compares the ITM wall temperature on the permeate side
Fig. 9) leading to higher oxygen flux (as shown in Fig. 10). As we for the 2-channel and 3-channel reactor cases. It is clear that the
move down the reactor length i.e. at 600 mm (as shown in temperature is more uniform along the length of the reactor in the
Fig. 13(c)) the reaction rate is lower in case of the 2-channel reactor 3-channel case. This may be advantageous, as a uniform tempera-
than in the 3-channel reactor and so is the temperature (shown in ture will reduce differential thermal expansion, which can lead to
Fig. 13(d)). At 900 mm, the reaction rate (shown in Fig. 13(e)) is zero stressing of the reactor structure or practical challenges such as
in the 2-channel reactor whereas in the 3-channel reactor the re- reactor sealing. The heat flux is plotted in Fig. 15, with the radiative
action rate is still finite. The fact that this action ceases halfway heat flux and convective heat shown in Fig. 15(a) and (b) respec-
down the reactor in the 2-channel case contributes to the reduction tively. Fig. 15(a) indicates that the radiative heat flux varies more in
in the temperature at the outlet. On the other hand, continuous the 2-channel reactor. On the other hand, the convective heat flux,
reaction in the 3-c case yields a higher temperature at the outlet. as shown in Fig. 15(b), exhibits higher values in the 3-channel
The closer proximity of the reaction zone to the ITM in the 2- reactor. The higher convective heat flux in the 3-channel reactor

Fig. 15. Comparison of (a) radiative heat flux and (b) convective heat flux for 2-channel and 3-channel reactors on both the feed side and the permeate side for a mass flow rate of
fuel mixture Mf ¼ 0.0015 kg/s and CH4 ¼ 1%.
942 P. Ahmed et al. / Energy 77 (2014) 932e944

Fig. 17. Comparison of contours of mass fraction of CH4 for 2-channel and 3-channel reactors for a mass flow rate of fuel mixture Mf ¼ 0.0015 kg/s and CH4 ¼ 1%.

Fig. 18. Comparison of contours of mass fraction O2/CH4 ratio for 2-channel and 3-channel reactors for a mass flow rate of fuel mixture Mf ¼ 0.0015 kg/s and CH4 ¼ 1%.

results in more heat loss and therefore accounts for lower mem- distribute the fuel over the length of the reactor. This extra degree
brane temperature. The total surface heat flux, as shown in Fig. 16, of freedom while adding complexity to the design is also beneficial.
is more uniform in the 3-channel reactor.
Fig. 17 presents the mass fraction contours of the fuel (CH4) in 6. Conclusions
the two cases considered. The contours are compared on the same
scale and the color bar (in web version) at the bottom highlights the In this study, a two-dimensional, CFD (computational fluid dy-
fuel variation. Almost all the fuel is consumed in the first half of the namics) model was used to predict the characteristics of oxy-fuel
reactor in the 2-channel case whereas the fuel supply and con- combustion in two different oxygen transport reactors. Details of
centration is distributed over most of the length in the 3-case. The the fluid flow, distribution of species and temperature variations
fuel supply determines the local O2/CH4 ratio as shown in Fig. 18, were predicted for a conventional 2-channel reactor and a novel 3-
and hence the location of the reaction zone. The gradual supply of channel reactor employing a porous plate to introduce the fuel
the fuel through the porous wall in the 3-channel reactor results in gradually into the sweep channel where oxygen permeates through
an extended reaction zone throughout and temperature distribu- an opposing membrane. We find that the 3-channel reactor pro-
tion as observed above. The introduction of a porous plate (with duces a more uniform temperature distribution along the length of
multiple sections and different porosity), provides the flexibility to the ITM compared to the conventional 2-channel reactor. Moreover,
P. Ahmed et al. / Energy 77 (2014) 932e944 943

while in the 2-channel reactor the high temperature zone is References


concentrated in the first part and closer to the membrane surface,
the 3-channel reactor produces a longer reaction zone with a [1] Islam MS, Cherry M, Catlow CRA. Oxygen diffusion in LaMnO3 and LaCoO3
perovskite-type oxides: a molecular dynamics study. J Solid State Chem
milder temperature gradient. The location of the reaction zone and 1996;124:230e7.
its temperature are controlled by the porosity of the plate. The exit [2] Bouwmeester HJM, Kruidhof H, Burggraaf AJ. Importance of the surface ex-
temperature of the 3-channel reactor is higher which could change kinetics as rate limiting step in oxygen permeation through mixed-
conducting oxides. Solid State Ionics 1994;72:185e94. Part 2.
improve the thermodynamics efficiency an ITM reactor based cycle. [3] Steele BCH. Oxygen ion conductors and their technological applications. Mater
Future studies will consider the effects of fuel pyrolysis, the impact Sci Eng B 1992;13:79e87.
of the channel height, flow rates and porous plate designs, with a [4] Bernardo P, Drioli E, Golemme G. Membrane gas separation: a review/state of
the art. Ind Eng Chem Res 2009;48:4638e63.
particular focus on optimizing the 3-channel reactor operation. [5] Habib MA, Badr HM, Ahmed SF, Ben-Mansour R, Mezghani K, Imashuku S,
et al. A review of recent developments in carbon capture utilizing oxy-fuel
combustion in conventional and ion transport membrane systems. Int J En-
Acknowledgments ergy Res 2011;35:741e64.
[6] Sunarso J, Baumann S, Serra JM, Meulenberg WA, Liu S, Lin YS, et al. Mixed
ioniceelectronic conducting (MIEC) ceramic-based membranes for oxygen
The authors would like to acknowledge the support provided by separation. J Membr Sci 2008;320:13e41.
King Fahd University of Petroleum & Minerals and KACST-TIC on [7] Smith A, Klosek J. A review of air separation technologies and their integration
with energy conversion processes. Fuel Process Technol 2001;70:115e34.
CCS during the course of this work under Project No.CCS_5. [8] Coroneo M, Montante G, Giacinti Baschetti M, Paglianti A. CFD modelling of
inorganic membrane modules for gas mixture separation. Chem Eng Sci
2009;64:1085e94.
Nomenclature [9] Wang H, Cong Y, Yang W. Oxygen permeation study in a tubular Ba0.5Sr0.5-
Co0.8Fe0.2O3-^I0 oxygen permeable membrane. J Membr Sci 2002;210:259e71.
[10] Ben-Mansour R, Habib M, Badr H, Nemitallah M. Characteristics of oxy-fuel
Acell area of the cell [m2] combustion in an oxygen transport reactor. Energy Fuels 2012;26:4599e606.
a absorption coefficient [11] Mancini ND, Mitsos A. Ion transport membrane reactors for oxy-combustion
e part I: intermediate-fidelity modeling. Energy 2011;36:4701e20.
Cp heat capacity [J/kg-K] [12] van Hassel Bart A. Oxygen transfer across composite oxygen transport
Di,m diffusion coefficient of mixture species i [m2/s] membranes. Solid State Ionics 2004;174:253e60.
Dv diffusion coefficient of oxygen vacancies [cm2/s] [13] Xu SJ, Thomson WJ. Oxygen permeation rates through ion-conducting
perovskite membranes. Chem Eng Sci 1999;54:3839e50.
Di,j binary mass diffusion coefficient of species i [m2/s] [14] Hsieh H. Inorganic membranes for separation and reaction. Elsevier; 1996.
ED, Er, Ef activation energies [J/kg-mol] [15] Farooqui AE, Habib MA, Badr HM, Ben-Mansour R. Modeling of ion transport
I radiation intensity, which depends on position and reactor for oxy-fuel combustion. Int J Energy Res 2013;37:1265e79.
[16] Hong J, Kirchen P, Ghoniem AF. Interactions between oxygen permeation and
direction
homogeneous-phase fuel conversion on the sweep side of an ion transport
JO2 oxygen permeation flux [mol/m2-s] membrane. J Membr Sci 2013;428:309e22.
kio surface exchange coefficient [m/s] [17] Hong J, Kirchen P, Ghoniem AF. Laminar oxy-fuel diffusion flame supported by
kr surface-exchange reaction reverse-rate constant an oxygen-permeable-ion-transport membrane. Combust Flame 2013;160:
704e17.
[mol cm2 s1] [18] Hong J, Kirchen P, Ghoniem AF. Analysis of heterogeneous oxygen exchange
kf surface-exchange reaction forward-rate constant and fuel oxidation on the catalytic surface of perovskite membranes. J Membr
[cm atm0.5 s1] Sci 2013;445:96e106.
[19] Kirchen P, Apo DJ, Hunt A, Ghoniem AF. A novel ion transport membrane
keff effective thermal conductivity reactor for fundamental investigations of oxygen permeation and oxy-
L membrane thickness [m] combustion under reactive flow conditions. Proc Combust Inst 2013;34:
Mi molecular weight of species i [kg/Kmol] 3463e70.
[20] Hunt A, Dimitrakopoulos G, Kirchen P, Ghoniem AF. Measuring the oxygen
m,_ Mf mass flow rate [kg/s] profile and permeation flux across an ion transport membrane and the
n refractive index development and validation of a multistep surface exchange model. J Membr
p pressure [Pa] Sci 2014;468:62e72.
0 , P [21] Kao Y, Lei L, Lin Y. A comparative simulation study on oxidative coupling of
PO 1 partial pressure of oxygen at the feed side [Pa] methane in fixed-bed and membrane reactors. Ind Eng Chem Res 1997;36:
00 2
PO2 , P2 partial pressure of oxygen at the permeate side [Pa] 3583e93.
! [22] ten Elshof JE, Bouwmeester HJM, Verweij H. Oxidative coupling of methane in
r position vector
a mixed-conducting perovskite membrane reactor. Appl Catal A Gen
Si, Sm source/sink term [kg/m3-s], mass source term [kg/m3-s] 1995;130:195e212.
! !0 [23] Balachandran U, Dusek J, Mieville RL, Poeppel R, Kleefisch M, Pei S, et al. Dense
S, S direction vector, scattering direction vector
ceramic membranes for partial oxidation of methane to syngas. Appl Catal A
T temperature [K] Gen 1995;133:19e29.
U, V superficial velocity [m s1] [24] Liu S, Gavalas GR. Oxygen selective ceramic hollow fiber membranes. J Membr
Sci 2005;246:103e8.
Vcell volume of cell [m3]
[25] Schiestel T, Kilgus M, Peter S, Caspary KJ, Wang H, Caro J. Hollow fibre
Xi mole fraction of species perovskite membranes for oxygen separation. J Membr Sci 2005;258:1e4.
Yi mass fraction of species i [e] [26] Tan X, Pang Z, Li K. Oxygen production using La0. 6Sr0. 4Co0. 2Fe0. 8O3a
4i,j mixture rule constant for species i in species j [e] (LSCF) perovskite hollow fibre membrane modules. J Membr Sci 2008;310:
550e6.
f phase function [27] Zeng Y, Lin YS, Swartz SL. Perovskite-type ceramic membrane: synthesis,
U0 solid angle oxygen permeation and membrane reactor performance for oxidative
r density [kg m3] coupling of methane. J Membr Sci 1998;150:87e98.
[28] Li S, Jin W, Huang P, Xu N, Shi J, Lin YS. Tubular lanthanum cobaltite perov-
m dynamic viscosity [N s m2] skite type membrane for oxygen permeation. J Membr Sci 2000;166:51e61.
mO2 oxygen vacancy potential [J/mol] [29] Shao Z, Yang W, Cong Y, Dong H, Tong J, Xiong G. Investigation of the
ss scattering coefficient permeation behavior and stability of a Ba0.5Sr0.5Co0.8Fe0.2O(3-d) oxygen
membrane. J Membr Sci 2000;172:177e88.
s StefaneBoltzmann constant (5.669  1008 W/m2 e K4) [30] Santos L, Moraes C, Hughes R. Characterization of hollow fibre membranes for
a permeability (m2) oxygen permeation and partial oxidation reactions. Braz J Pet Gas 2011;5.
a1...a10 face permeabilities of porous membranes 1,2….10 (equal [31] Tan X, Liu Y, Li K. Preparation of LSCF ceramic hollow-fiber membranes for
oxygen production by a phase-inversion/sintering technique. Ind Eng Chem
lengths) [m2]
Res Ind Eng Chem Res 2004/12/07;44:61e6.
li thermal conductivity of species i [W/m/K]
944 P. Ahmed et al. / Energy 77 (2014) 932e944

[32] Farooqui AE, Badr HM, Habib MA, Ben-Mansour R. Numerical investigation of [45] Petruzzi L, Cocchi S, Fineschi F. A global thermo-electrochemical model for
combustion characteristics in an oxygen transport reactor. Int J Energy Res SOFC systems design and engineering. J Power Sources 2003;118:96e107.
2013;38:638e51. [46] Buykx W. Specific heat, thermal diffusivity and thermal conductivity of Syn-
[33] Ben-Mansour R, Habib MA, Badr, Azharuddin HM, Nemitallah M. Character- roc, Perovskite, Zirconolite and Barium Hollandite. J Nucl Mater 1982;107:
istics of oxy-fuel combustion in an oxygen transport reactor. Energy Fuels 78e82.
2012/07/19;26:4599e606. [47] Mahato N, Sharma S, Keshri AK, Simpson A, Agarwal A, Balani K. Nano-
[34] Mancini N, Gunasekaran S, Mitsos A. A multiple-compartment ion-transport- mechanical properties and thermal conductivity estimation of plasma-
membrane reactive oxygen separator. Ind Eng Chem Res 2012;51:7988e97. sprayed, solid-oxide fuel cell components: ceria-doped, yttria-stabilized zir-
[35] F. U. Guide. Lebanon, NH: “Fluent,” Inc.; 2007. conia electrolyte. JOM;1e14.
[36] Habib MA, Ahmed P, Ben-Mansour R, Badr HM, Kirchen P, Ghoniem AF. [48] Petric A, Huang P, Tietz F. Evaluation of LaeSreCoeFeeO perovskites for solid
Modeling of a combined ion transport and porous membrane reactor for oxy- oxide fuel cells and gas separation membranes. Solid State Ionics 2000;135:
combustion. J Membr Sci 2013;446:230e43. 719e25.
[37] Zhu X, Sun S, Cong Y, Yang W. Operation of perovskite membrane under [49] Li S, Jin W, Xu N, Shi J. Synthesis and oxygen permeation properties of
vacuum and elevated pressures for high-purity oxygen production. J Membr La0.2Sr0.8Co0.2Fe0.8O3ed membranes. Solid State Ionics 1 Sep, 1999;124:
Sci 2009;345:47e52. 161e70.
[38] Nigara Y, Mizusaki J, Ishigame M. Measurement of oxygen permeability in [50] Kusaba H, Shibata Y, Sasaki K, Teraoka Y. Surface effect on oxygen permeation
CeO2 doped CSZ. Solid State Ionics 1995;79:208e11. through dense membrane of mixed-conductive LSCF perovskite-type oxide.
[39] Jun A, Yoo S, Gwon O-h, Shin J, Kim G. Thermodynamic and electrical prop- In: Solid State Ionics 15: Proceedings of the 15th International Conference on
erties of Ba0.5Sr0.5Co0.8Fe0.2O3d and La0.6Sr0.4Co0.2Fe0.8O3d for Solid State Ionics, Part II, vol. 177; 2006/10/31. p. 2249e53.
intermediate-temperature solid oxide fuel cells. Electrochim Acta 2013;89: [51] Xu SJ, Thomson WJ. Stability of La0.6Sr0.4Co0.2Fe0.8O3ed perovskite mem-
372e6. branes in reducing and nonreducing environments. Ind Eng Chem Res Ind Eng
[40] Mota L, da Silva MG, de Souza VP, Vargas H, Guimara ~es VF, Paes Júnior HR. On Chem Res. 19 Mar, 1998;37:1290e9.
the use of photoacoustic technique for monitoring the thermal properties of [52] Coroneo M, Montante G, Paglianti A. Numerical and experimental fluid-
lanthanum strontium cobalt ferriteeyttria stabilized zirconia two-layer sys- dynamic analysis to improve the mass transfer performances of PdAg
tems. Thin Solid Films 2010;519:938e42. membrane modules for hydrogen purification. Ind Eng Chem Res 2010;49:
[41] Guogang Y, Danting Y, Haijun L, Xinrong L. The effects of thermal conductivity 9300e9.
of porous layer on chemical reaction and transport processes in a SOFC anode [53] Staudacher M, Harasek M, Brinkmann T, Hilgendorff W, Friedl A. CFD-simu-
duct. In: Advances in Energy Engineering (ICAEE), 2010 International Con- lation of mass transfer effects in gas and vapour permeation modules. Desa-
ference on; 2010. p. 325e8. lination 2002;146:237e41.
[42] Kaisare N, Vlachos D. Optimal reactor dimensions for homogeneous com- [54] Goto S, Assabumrungrat S, Tagawa T, Praserthdam P. The effect of direction of
bustion in small channels. Catal Today 2007;120:96e106. hydrogen permeation on the rate through a composite palladium membrane.
[43] Mei H, Li C, Liu H, Ji S. Simulation of catalytic combustion of methane in a J Membr Sci 2000;175:19e24.
monolith honeycomb reactor. Chin J Chem Eng 2006;14:56e64. [55] Farooqui AE, Habib MA, Badr HM, Ben-Mansour R. Modeling of ion transport
[44] Yamanaka S, Hamaguchi T, Oyama T, Matsuda T, Kobayashi S-i, Kurosaki K. reactor for oxy-fuel combustion. Int J Energy Res 2013;37:1265e79.
Heat capacities and thermal conductivities of perovskite type BaZrO3 and
BaCeO3. J Alloys Compd 2003;359:1e4.

You might also like