You are on page 1of 13

Fuel 208 (2017) 522–534

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Full Length Article

Dynamic model of calcium looping carbonator using alternating


bubbling beds with gas switching
Shady Emad, Ahmed A. Hegazi, Salah H. El-Emam, Farouk M. Okasha ⇑
Mechanical Power Engineering Department, Faculty of Engineering, Mansoura University, El-Mansoura 35516, Egypt

h i g h l i g h t s

 Dynamic model of carbonator is presented to simulate CO2 capture.


 Configuration of alternating fluidized beds with gas switching is investigated.
 A novel scheme that enables smooth gas switching as well as energy saving is proposed.
 Parametric studies are conducted to assess key parameters.
 Experimental validation indicates acceptable reliability of model.

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

Article history: Carbon capture and storage (CCS) has been globally gaining popularity as a viable greenhouse gases mit-
Received 29 September 2016 igation strategy throughout the last decade. Calcium looping (CaL) is an emerging technology to capture
Received in revised form 16 June 2017 carbon dioxide from flue gases of fossil fueled power plants exploiting the reversible gas-solid reaction
Accepted 12 July 2017
between the carbon dioxide (CO2) and calcium oxide (CaO) to form calcium carbonate (CaCO3) in a flu-
idized bed. In this paper, the concept of calcium looping application using alternating fluidized beds is
explored as well as a proposed scheme to allow smooth gas switching between reactors using an inter-
Keywords:
mediate third reactor. The latter scheme enables energy saving as well. A dynamic model of a bubbling
Carbon capture
Global warming
bed carbonator, the key reactor in the capture process, is presented. The model incorporates both hydro-
Calcium looping dynamics and chemical kinetics to provide more reliable predictions. The model has been validated with
Fluidized bed experimental results obtained at combustion lab, Mansoura University, Egypt using an atmospheric flu-
idized bed reactor of 10.5 cm inner diameter. The key parameters have been investigated to check for sys-
tem sensitivity. Bed temperature has a non-monotonic effect on CO2 capture efficiency. Maximum CO2
capture efficiency was found to occur around a temperature of 675 °C. Capture efficiency increases with
either decreasing fluidization velocity or increasing bed particle size due to enhanced mass transfer and
increased residence time. These findings almost accord with published data. Also, the average CO2 cap-
ture efficiency was found to increase with increasing static bed height up to a certain limit. Further
increase in bed height doesn’t considerably affect the capture efficiency. The proposed model can be used
as a design tool that would enable the optimization and commercialization of calcium looping.
Ó 2017 Elsevier Ltd. All rights reserved.

1. Introduction Fossil fuel-based emissions of CO2 may be originated from both


stationary (e.g. power plants) and non-stationary systems (e.g.
Global warming has many consequences including sea water automobile). However, power generation sector is responsible for
level increase, agriculture and fisheries disruption, atmospheric the largest amounts of CO2 emissions [1]. Current projections for
warming, and prevalence of different diseases such as malaria. Glo- global energy demands still point to the dependence on fossil fuels
bal warming is mainly caused by anthropogenic emissions of so to meet >85% of the world’s energy needs. Therefore, the scientific
called greenhouse gases (mainly carbon dioxide (CO2)). The burn- community agrees that the solution for alleviating CO2 emissions
ing of fossil fuels constitutes the major source of CO2 emissions. for the short- to midterm lies in a portfolio of strategies, including
carbon capture and storage [2].
Carbon capture and storage, or CCS, is a chain of techniques that
⇑ Corresponding author.
enables the continuous use of conventional fuels while preventing
E-mail address: faroukok@mans.edu.eg (F.M. Okasha).

http://dx.doi.org/10.1016/j.fuel.2017.07.049
0016-2361/Ó 2017 Elsevier Ltd. All rights reserved.
S. Emad et al. / Fuel 208 (2017) 522–534 523

Nomenclature

Symbol Description, unit r0 initial grain radius, m


Ab cross sectional area of bubble phase, m2 S surface area of solid particles at any time, m2/g
Ae cross sectional area of emulsion phase, m2 S0 initial surface area of solid particles, m2/g
Ar Archimedes number, Dimensionless T operating temperature, K
Cb gas molar concentration in bubble phase, mol/m3 t time, s
Ce gas molar concentration in emulsion phase, mol/m3 t lim switching time, s
C CO2 CO2 molar concentration, mol/m3 u0 fluidization velocity, m/s
C CO2 ; eq CO2 molar concentration at equilibrium, mol/m3 ub rise velocity of bubble phase, m/s
C CO2;in CO2 molar concentration at bed inlet, mol/m3 ub The effective gas velocity through the bubble phase, m/s
C CO2;out CO2 molar concentration at bed outlet, mol/m3 ubm mean rise velocity of bubble phase, m/s
db bubble diameter, m ubr rise velocity of single bubble, m/s
dbm mean bubble diameter along bed, m ue rise velocity of emulsion gas, m/s
dp particle diameter, m umf minimum fluidization velocity, m/s
E reaction activation energy, J/mol Vg volume of gas in the control volume, m3
g gravitational acceleration, m/s2 Ws mass of solid particles in bed, kg
Heb expanded (bubbling) bed height, m X conversion ratio of solid sorbent, Dimensionless
Hmf bed height at minimum fluidization, m X av g average conversion ratio of solid sorbent, Dimensionless
Hs static bed height, m X max;N maximum sorbent capacity after N cycles, Dimension-
K be mass interchange coefficient between bubble and emul- less
sion phase, s1 xCO2 ;in carbon dioxide mole fraction at bed inlet, Dimensionless
MW molecular weight, kg/mol xCO2 ;out carbon dioxide mole fraction at bed outlet, Dimension-
MW s average molecular weight of solid particles in bed (CaO less
+ CaCO3), kg/mol z axial distance measured from air distributor, m
N number of calcination/carbonation cycles, Dimension-
less Greek symbols
n order of reaction, Dimensionless emf bed voidage fraction at minimum fluidization, Dimen-
PCO2 CO2 partial pressure, Pa sionless
PCO2;eq CO2 partial pressure at equilibrium, Pa qg gas density, kg/m3
Q gas;in volumetric flow rate of inlet gas, m3/s qs solid particles density, kg/m3
R specific reaction rate, s1 d the fraction of bed consisting of bubbles, Dimensionless
RCO2 rate of consumption of CO2, mol=m3gas  s l dynamic viscosity, N  s=m2
Rep;mf reynolds number at minimum fluidization, Dimension- / solid particles sphericity, Dimensionless
less gcapture CO2 capture efficiency, Dimensionless
r grain radius at any time, m

the CO2 emissions from polluting the atmosphere. It begins with (900–950 °C respectively, allowing for efficient heat recovery in
the capture of CO2 from fossil fuel combustion or industrial pro- process heating or steam cycle of a power generation system.
cesses. Next, the captured CO2 is compressed and transported via Moreover, CaL can be integrated with cement manufacturing so
ships or pipelines. Finally it is stored in suitable geological forma- that a near zero-waste process can be achieved [5,6].
tions. An overview of different CCS technologies can be found in Hirama et al. [7] patented separation of carbon dioxide from
[3,4]. However, the most critical step in the CCS chain that deter- gases containing it by contacting the gas mixture with metal based
mine the feasibility of a certain technique is the capture step [2]. oxides (e.g. calcium oxide) to form metal carbonate. The metal
One of the promising technologies that has shown some poten- oxide is then regenerated at higher temperatures in a second con-
tial advantages in terms of net efficiency and cost of CO2 avoided tacting zone where heat is supplied. The application of CaL as a
on both lab and pilot scale is carbon dioxide capture by absorp- post-combustion CO2 capture process with dual fluidized bed
tion/regeneration process with calcium oxide, known as calcium was then proposed and applied by Shimizu et al. [8]. Since then,
looping as shown in Fig. 1. Both the carbonation and calcination a lot of research has been done to further analyze and develop
reactions are carried out at high temperatures (600–700 °C and the process [5,9,10]. Moreover, several projects have been estab-
lished to assess its feasibility on both lab and pilot scales [11–14].
Fig. 2 shows two different dual fluidized bed configurations that
can be used in the calcium looping process where heat is supplied
in the calciner by oxyfuel combustion. Most research works have
focused on interconnected fluidized beds with solids circulation
between carbonator and calciner, Fig. 2(a). However, this configu-
ration has three main drawbacks.
First, it requires using circulating fluidized bed with relatively
high inlet gas velocities which increases attrition and diminishes
the possibility of reusing sorbent particles in subsequent cycles.
Second, solids exiting the carbonator may include unreacted cal-
cium oxide particles and solids exiting the calciner may include
calcium carbonate particles which affects the capture efficiency.
Third, plugging can occur in solid circulation lines between the
Fig. 1. Calcium looping process. two reactors. In order to solve these problems, the present work
524 S. Emad et al. / Fuel 208 (2017) 522–534

dynamic nature posing restrictions on fuel handling and switching.


This makes the proposed system most feasible for gaseous fuels.
In the present study, a proposed scheme for CaL application
using alternating beds with smooth gas switching is presented.
Then, a mathematical model for a bubbling bed reactor operating
in carbonation mode, where capturing of CO2 takes place, is devel-
oped. The fluidization phenomenon is studied and its effects on
mass transfer between different phases and residence time of gas
molecules in the reactor are investigated. The model pays due
attention to the kinetic parameters controlling the reaction rate
and its dependence on operating conditions (e.g. operating temper-
ature and CO2 partial pressure). Model calculations yield axial con-
centration profiles of carbon dioxide in the bed phases. Model
predictions are compared with experimental measurements
obtained in combustion lab. Parametric study has been carried
out to explore the key parameters that affect the capture process.

2. A smooth switching scheme

One of the main challenges in the application of gas switching


configuration is to have control of reactors temperatures while
Fig. 2. Dual fluidized bed configurations.
switching takes place. For example, after calcination ends (i.e.
almost all CaCO3 are decomposed to CaO), switching should occur
and carbonation is supposed to start. However, as calcination tem-
has been dedicated to study the mode of alternating beds with gas perature is significantly higher than carbonation temperature, the
switching, Fig. 2(b). The classical advantages of bubbling beds over bed has to be first cooled before carbonation can take place. In
circulating beds include less attrition rate, better temperature dis- order to overcome this problem with minimum energy penalties,
tribution and flexibility to wide range of particles size. BFBs are the following novel configuration is proposed, see Fig. 3.
also much simpler as solids circulation system is avoided, more This configuration consists of two similar main reactors plus a
details can be found elsewhere [15]. smaller reactor operating as intermediate, as follows.
The adopted configuration consists of two separate fluidized
bed (bubbling-bubbling) where carbonation and calcination reac- (1) Reactor A is operating as carbonator at a temperature of
tions take place periodically in each reactor. In the carbonator, flue 650 °C and reactor B is operating as calciner at 950 °C while
gases are admitted to the bed where calcium oxide particles cap- reactor C is kept at 650 °C (no reaction).
ture carbon dioxide from the gases mixture. When almost CaO par- (2) When switching is required, oxygen is fed to reactor B so
ticles get converted into CaCO3 (i.e. the bed is no longer capable of that O2 is preheated. Preheated O2 is redirected to reactor
capturing CO2), supplying flue gases is stopped and fuel with oxy- A (with fuel) in order to heat it up (from 650 to 950 °C)
gen is admitted (oxyfuel combustion) to provide heat and operate and to start calcination. Meanwhile flue gases are directed
the reactor as a calciner. During calcination, a pure stream of CO2 is to reactor C in order that CO2 capture continues till reactor
released and can be further compressed and transported for geo- B is cooled down from 950 °C to 650 °C and get ready for
logical sequestration. After CaO is regenerated in the calciner, car- next carbonation period.
bonation process starts again and the cycle is repeated periodically. (3) Reactor B is operating as carbonator at a temperature of 650
To allow continuous capture, dual alternating fluidized beds are and reactor A is operating as calciner at 950 °C. Reactor C is
used with gas switching between them. This concept has a heated using (O2 + Fuel) and decomposition occurs.

Fig. 3. Proposed scheme for smooth gas switching.


S. Emad et al. / Fuel 208 (2017) 522–534 525

1000 depending on the operating conditions. Baker [16] proposed the


900 following equation to estimate equilibrium conditions where oper-
ating temperature is the only independent variable;
800
Bed Temperature (oC)

38000
700 log10 Peq ðatmÞ ¼ 7:079  ð2Þ
4:574  T
600
This heterogeneous carbonation reaction takes place in two
500
sequential stages. First, an initial fast stage controlled by chemical
400 kinetics at reaction surface. The formation of the product layer
300 starts to prevent the unreacted core from contacting the reacting
gas. Then comes a much slower stage controlled by diffusion of
200
gas through the product layer of calcium carbonate. In most indus-
100 trial applications, the diffusion stage is commonly neglected and
0 calcium oxide is considered to reach its maximum conversion at
0 50 100 150 200 the end of the kinetically controlled stage.
Time(s)
The thickness of the product layer formed on the free surface of
Fig. 4. Cooling of reactor B. CaO is an important parameter to indicate the end of the fast reac-
tion stage. Also, the formation and growth of solid product affects
(4) Oxygen is fed to reactor C so that O2 is preheated and fed to the porous structure decreasing the available surface area for reac-
reactor A (with fuel) to continue calcination while reactor C tion. Many researchers investigated the average value for the crit-
cools down to 650 °C. Then, steps 1–4 are repeated with ical CaCO3 product layer thickness [17]. The theory of a critical
reactors A and B are interchanged. product layer thickness has been used extensively to study the
maximum obtainable conversion during carbonation reaction
A simplified mathematical analysis of step 2 is given in [18,19]. However, these studies lack the important effects of oper-
Appendix A. For the base case (discussed later), it is found that step ating temperature on chemical kinetics. They consider that operat-
2 continues about 2 min (Fig. 4) corresponding to 13% average con- ing temperature only determines the equilibrium partial pressure
version of calcium oxide particles in reactor C. Oxygen preheating of carbon dioxide. However, experiments indicate greater role for
during this time has been found to save fuel consumption required temperature in the reaction scenario [20].
for calcination by 2.24%. As stated above, only the fast reaction period is considered for
most practical applications. It is controlled by chemical kinetics.
The grain model for porous solids [21] is adopted here to model
3. Carbonator model description
the gas-solid reaction between CO2 and CaO.
The reaction rate for a gas-solid reaction is usually defined as a
The following assumptions are adopted throughout the follow-
specific reaction rate R, where;
ing model derivation;
dX
1- There are two main phases in the bed: emulsion and bub- R¼ ð3Þ
dtð1  XÞ
bles. Bubbles are free of particles. Clouds and wakes are con-
sidered part of the emulsion. When the reaction is under kinetic control, the specific rate can
2- Gas streams passing through the emulsion, bubble phase, be further expressed in power law form [22];
and freeboard follow a plug-flow regime. R ¼ 56ks  ðPCO2  PCO2 ;eq Þn  S ð4Þ
3- One-dimensional model, i.e. variations of gas concentrations
occur only in the vertical or axial z direction. Assuming that reaction takes place uniformly on spherical
4- Homogeneous composition of solid particles throughout the grains, the following equation can be used,
bed is assumed. This is justifiable because of the high circu-  3
lation rate in the bed. r
1X ¼ ð5Þ
5- Uniform temperature distribution is assumed throughout r0
the bed due to the higher heat capacities of proposed config- Combining Eqs. (3)-(5) would result in
uration, due to higher bulk solid densities.
6- Thermal equilibrium between flue gases and bed particles dX 5
¼ 56ks  S0  ðP CO2  PCO2 ;eq Þn  ð1  XÞ3 ð6Þ
takes place in less than 1 cm. So, reaction almost takes place dt
at constant temperature. Sun et al. [22] reported a first-order reaction changing to zero-
7- Heat generation is found to increase gases temperature by order dependence when the CO2 partial pressure exceeded 10 kPa.
175 °C. Hence, carbonator temperature is kept approxi- The kinetic constant can be evaluated using Arrhenius equation;
mately constant via introducing flue gases at a temperature  
lower than the predetermined carbonation temperature by E
ks ¼ k0  exp  ð7Þ
175 °C. Heat generation compensate for this difference. R:T
Kinetic parameters (k0 and E) are obtained from experimental
3.1. Kinetic model measurements and values reported by Sun et al. [22] are adopted
here.
Carbon capture in carbonator takes place through the exother- Barker [23] reported that the carbonation reaction, presented
mic reaction of calcium oxide with carbon dioxide as follows: by Eq. (1), is far from reversible in practice. The sorption capacity
in the fast reaction stage decreases rapidly with increasing the
CaO þ CO2 ! CaCO3 DH0r ¼ 179:2 kJ=mol ð1Þ
number of calcination–carbonation cycles. Grasa and Abanades
The carbonation reaction is a non-catalytic gas-solid reaction [24] proposed a semi-empirical expression, Eq. (8) to express the
and there exists an equilibrium partial pressure of carbon dioxide sorbent capacity after a large number of complete cycles (up to
526 S. Emad et al. / Fuel 208 (2017) 522–534

500). It is valid for different sorbents and for a wide range of oper- 3.2.3.3. Bed expansion. The fraction of bubble phase in the fluidized
ating conditions. bed (d) is proportional to the fluidization velocity of inlet gas. For
intermediate bubbles, the following expression has been proposed
1
X max;N ¼ þ Xr ð8Þ by Abanades et al. [29]:
1
ð1X r Þ
þkN
u0  umf
d¼ 5umf ub emf
ð16Þ
where k and Xr represent the deactivation constant and the ub þ 4
residual conversion, respectively. It is observed that values of
k ¼ 0:52 and X r ¼ 0:075 fit well with a wide range of sorbents Also, expanded bed height is related to Bubble fraction by Eq.
and conditions [24]. (17)
Hmf
3.2. Hydrodynamic model Heb ¼ ð17Þ
1d

Calcium oxide particles used in practice fall well in the Geldart It can be concluded from Eqs. (12)–(14), (16), and (17), that an
B category. Kunii-Levenspiel model for bubbling bed is widely iterative solution (Fig. 5) is required to evaluate the expanded bed
accepted for its simplicity and reliable results [15]. A bubbling flu- height, bubble size and velocity, and bubble fraction of the flu-
idized bed consists mainly of two phases, bubbles and emulsion. idized bed [30].
Bubbles are lean phase free of solid particles, while emulsion is a
dense phase where solid particles are assumed to be uniformly 3.3. Reactor model
distributed.
Species conservation is applied on both phases of fluidized bed
3.2.1. Minimum fluidization reactor. Assuming no accumulation of gas in control volume, the
The minimum fluidization velocity, umf , may be calculated using rate of outflow should equal the summation of rate of inflow, rate
the following expression [25]; of mass transfer to the control volume and rate of generation by
chemical reaction, Fig. 6.
qg  dp  umf qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

Rep;mf ¼ ¼ C 21 þ C 2  Ar  C 1 ð9Þ
l o For Emulsion phase, the conservation equation for carbon diox-
ide can be written as follows,
d3p qg ðqs qg Þg
where: Ar ¼ l2  
dC e
C 1 and C 2 are constants with values of 27.2 and 0.0408, respec- Ce þ Dz  Ae  umf ¼ C e  Ae  umf þ K be ½C b  C e   Ab  Dz
tively as suggested by Grace [26]. dz
Broadhurst and Becker [27] proposed a correlation to estimate  RCO2  Ae  Dz  emf ð18Þ
bed porosity at minimum fluidization conditions, emf , where;
 0:021
qg
emf ¼ 0:586/0:72 Ar0:029 ð10Þ
qs

3.2.2. Emulsion phase


The emulsion phase is assumed to be at minimum fluidization
condition. Hence, the superficial rise velocity of emulsion gas is
considered to be the same as ðumf Þ

3.2.3. Bubble phase


3.2.3.1. Bubble size. Bubbles size can be estimated using the tradi-
tional Darton’s correlation [28] where,

db ¼ 0:54½u0  umf 0:4  ½z0:8  g 0:2 ð11Þ


The mean bubble diameter along the bed may be estimated by
integrating Eq. (11) from z ¼ 0 to z ¼ Heb as given below;

dbm ¼ 0:3½u0  umf 0:4  ½Heb 0:8  g 0:2 ð12Þ

3.2.3.2. Bubble rise velocity. For Bubbles in bubbling bed, the rise
velocity is given by the following equation:
ub ¼ ½u0  umf  þ ubr ð13Þ
where ubr is the rise velocity of a single bubble estimated using
the expression reported by Kunii-Levenspiel [15] as follows,
0:5
ubr ¼ 0:711½gdb  ð14Þ
The effective superficial gas velocity through the bubble phase
can be defined from the gas balance in a cross section of the bed
where;
u0  ð1  dÞumf
ub ¼ ð15Þ
d Fig. 5. Iterative calculation of hydrodynamic parameters.
S. Emad et al. / Fuel 208 (2017) 522–534 527

Switching time, tlim , is an important parameter in the investi-


gated configuration of alternating beds with gas switching. For
convenience, it is taken as the period after which the capture effi-
ciency falls below 85% of its maximum value (an arbitrary limit).

3.3.1. Numerical solution


The carbonator model has been implemented in MATLAB 2014b
[31]. Fig. 7 shows a flow chart of the algorithm used. After the
introduction of the input values, the hydrodynamic and kinetic
model are run to calculate the parameters required for Eqs. (21)
and (22). Then, this system of equations is solved at each time step
with the condition that (at z ¼ 0; C b ¼ C e ¼ C inlet ) to give the aver-
age CO2 exit concentration (C exit ). Since the beginning, results
clearly indicated that the rate of change of CO2 concentration in
emulsion is so rapid at the bottom of bed close to the air distributor
and then gets very slow with height. Solution of the system using
fixed step methods would either lead to inaccuracy of results in
Fig. 6. Schematic of bubbling bed control volume.
case of using relatively large step size or high consumption of time
and calculation power in case of using too small step size all over
the bed. The use of adaptive step size method would solve this
Eq. (18) can be rearranged to give;

dC e K be d emf
¼  ½C b  C e    RCO2 ð19Þ
dz umf 1  d umf
The last term in Eq. (19) (RCO2 ) is the rate of consumption of CO2
due to chemical reaction and is given by Eq. (20)
" #  
mol 1 dNCO2 1  emf qs dX
RCO2 ¼ ¼  ð20Þ
m3gas :s V g dt emf MW s dt

This leads to:


 
dC e K be d 1  emf qs dX
¼ : ½C b  C e    ð21Þ
dz umf 1  d umf MW s dt

It should be noted that RCO2 is time-dependent as dX


dt
is calculated
using Eq. (6) which depends on the initial conversion of solid sor-
bent before each time step.

o For Bubble Phase

As bubbles are free of solid particles, therefore no chemical


reaction takes place in the bubble phase. Conservation of carbon
dioxide in bubble element can be written as:

dC b K be
¼   ½C b  C e  ð22Þ
dz ub
Although time is not a variable in Eq. (22), this equation is also
time-dependent as the CO2 concentration in emulsion phase
depends on reactivity of sorbent, Eq. (21), which varies with time
as stated above.
The average conversion (or conversion ratio), X av g , of sorbent at
any time can be calculated using this following expression:
Z
MW s t 1  xCO2;in
X av g ¼  Q gas;in C CO2;in  C CO2;out  dt ð23Þ
Ws 0 1  xCO2;out

This should be calculated at the end of each time step and used
to calculate RCO2 ns the following time step.
A common performance parameter used in applications of car-
bon dioxide capture is called CO2 capture efficiency, gcapture . It is
defined as the ratio between the number of moles of captured
CO2 to the number of moles of CO2 entering the carbonator.

N CO2;Captured C CO2;out 1  xCO2;in


gcapture ¼ ¼1  ð24Þ
NCO2;Entering C CO2;in 1  xCO2;out
Fig. 7. Flowchart of the carbonator model.
528 S. Emad et al. / Fuel 208 (2017) 522–534

problem. The algorithm used would reduce the integration step 16


size (less than 0.1 mm) where the rate of change is high to spot this
Bubble Phase
change accurately and relatively increase step size where the rate 14
is low to save computational time and power. Emulsion Phase

CO 2 mole fraction (%)


Detailed description of solution procedure is given below; 12
Average
10
1- To deal with the time-dependent term in Eq. (21), dX dt
is eval-
uated using Eq. (6) by substituting X (in the right-hand side) 8
with X av g at the beginning of the time step and assuming it
to be relatively constant during this time step. 6
2- The system of differential equations becomes a system of
ODEs (z is the only independent variable). Then, they are 4
solved starting from the distributor (z = 0) to the end of
the bed (z ¼ Heb ). As stated above, using an algorithm with 2
fixed step size is not effective since some points may be
missed where signal’s frequency is greater than the solver’s 0
frequency and calculation time may be wasted where sig- 0 0.1 0.2 0.3 0.4 0.5
nal’s frequency is too low. The used procedure is based on Height (m)
the Dormand–Prince Runge Kutta 45 integration method (a) t = 15 minutes
[32]. It uses six function evaluations to calculate fourth-
and fifth-order accurate solutions. The difference between 16
these solutions is then taken to be the error of the (fourth- Bubble Phase
14
order) solution and step size is altered accordingly. This Emulsion Phase

CO 2 mole fraction (%)


method is already implemented in the MATLAB built-in
12 Average
function, ODE45.
3- At the end of the time step, X av g;ðtþDtÞ is evaluated using Eq. 10
(23) and the time derivative of X is calculated using the
 
X X 8
expression dX dt
¼ av g;ðtÞ Dtav g;ðtþDtÞ . This value is then compared
with the value obtained by Eq. (6). If the difference between 6

obtained values of dX dt
is within acceptable tolerance,
calculations are approved and performance parameters are 4
evaluated. However, if this difference is larger than toler-
ance, time step is reduced and calculations have to be 2
repeated. Note that, if the difference is considerably smaller
than tolerance, solution proceeds with an increased time 0
step to save calculation time.
0 0.1 0.2 0.3 0.4 0.5
4- Steps 1-to-3 are repeated till the end of simulation time Height (m)
span. (b) t = 30 minutes

4. Results and discussion 16


15
The model is capable of predicting carbon dioxide mole fraction
CO 2 mole fraction (%)

and concentration at any height and at any instant, Fig. 8. It can be 14


observed that axial distribution is very similar in the first two fig-
13
ures (at 15 and 30 min), however considerable difference can be
easily detected in Fig. 8(c) indicating the transient nature of pre- 12
sent configuration. For illustration purposes, a fluidization column 11
with inner diameter of 10.5 cm loaded with lime particles of static
height of 15 cm operating at fluidization velocity of 0.8 m/s and 10
temperature of 650 °C is considered as the base case. Inlet gas is 9 Bubble Phase
composed of carbon dioxide and nitrogen with mole fractions of
15% and 85%, respectively. Results indicate that carbonation reac- 8 Emulsion Phase
tion taking place in emulsion phase is so rapid that CO2 mole frac- 7 Average
tion in emulsion gas decreased from 15% to about 2% in less than
two centimeters. After that the calcium oxide is primarily reacting 6
with the CO2 transferred from bubbles to emulsion. 0 0.1 0.2 0.3 0.4 0.5
Height (m)
4.1. Model validation with experimental results (c) t = 45 minutes

The apparatus used in this work is a bubbling fluidized-bed Fig. 8. Axial profiles of CO2 mole fraction at different times (T = 650 °C, u0 = 0.8 m/s).

reactor with a fluidization column of 10.5 cm ID and 4 m height,


as shown in Fig. 9. The fluidization gases are distributed using a
nozzle-type plate. Three heaters with 5 kW are used to preheat plenum to the freeboard. Temperatures have been measured in
the fed gases. Four rotameters are used to measure the gases flow the bed using type K thermocouples with limiting error of ±0.4%
rates. Two taps are used to measure the pressure drop from the of the temperature. The manufacturer specified thermocouples
S. Emad et al. / Fuel 208 (2017) 522–534 529

10 pleted. The temperature controller is regulated to have the pre-


1. Fluidization column designed temperature of bed materials. When the bed
2. Gas distributor temperature stabilizes at the test temperature (e.g. 650 °C), the
flow rates of nitrogen-carbon dioxide mixture (15% CO2) are regu-
3. Pressure tap lated to the pre-design values to allow for the required fluidization
4. CO2 velocity. At this point the carbonation process starts and the con-
centration of CO2 in outlet gases is measured and recorded. The
5. N2
measurements continue until the concentration of CO2 restores
6. Control valve its value in inlet gases, i.e. 15%.
7. Rotameter Fig. 10 shows typical experimental measurements and model
3 predictions of CO2 exit mole fraction with time, for the base case
8. Heater
1 described is ⁄⁄Sec. 4. The results indicate that the bed of calcium
9. Preheater oxide initially capture carbon dioxide with an acceptable effi-
10. Hopper ciency  72% as shown in Fig. 11. The efficiency of the process
8 maintains the same level until a certain point after which the
CO2 concentration quickly builds up to attain its value in the fed
2 gases whereas capture efficiency falls down close to zero.
Figs. 10–12 imply that carbonation process is very fast during the
initial period where the available surface area is still very large
7 3 and reaction is chemical kinetic controlled. The carbonation, then,
passes by a transition period to a very slow reaction rate. Actually,
9 the final slow stage is diffusion controlled. This behavior can be

6
16

CO 2 exit mole fraction (%) 14 Model


12 Experiment
5 4
10

Fig. 9. Schematic diagram of the experimental set up. 6

4
limiting error as ±0.4% of the temperature. A temperature con-
2
troller is used to maintain the bed temperature within the preset 0 1000 2000 3000 4000
temperature range by switching on or switching off the electric
Time (s)
heaters based on the bed-thermocouple signal. Measurement of
gases concentrations has been carried out using KANE455 gas ana- Fig. 10. CO2 exit mole fraction vs. time (T = 650 °C, u0 = 0.8 m/s).
lyzer. The measurement accuracy is ±0.3% for CO2.
The sorbent used is a limestone coming from Jabal al-Tair in
Minya, Egypt. The chemical composition of the used limestone is 80
reported in Table 1. The limestone has a density of 2560 kg/m3
CO 2 capture efficiency (%)

70
and particle size of 0.425–0.600 mm. The corresponding minimum
fluidization velocities are 0.19 m/s and 0.1 m/s at 27 °C and 650 °C 60
bed temperature, respectively.
50
The technique pursued to perform an experimental test may be
described in the following. The fluidization air is preheated before 40
it is delivered through the distributor plate to the bed. Feeding of
air continues until the bed temperature reaches the calcination 30
Model
temperature, 950 °C using electric heaters. During the calcination 20
process CaCO3 converts to CaO by losing CO2 according to the Experiment
reaction, 10

CaCO3 ! CaO þ CO2 ð25Þ 0


0 1000 2000 3000 4000
The bed is maintained at the calcination temperature until the Time (s)
CO2 concentration indicates nearly zero by the gas analyzer. At this
Fig. 11. CO2 capture efficiency vs. time (T = 650 °C, u0 = 0.8 m/s).
point the calcination of limestone particles has been fully com-

Table 1
Chemical composition of the sorbent (wt%).

SiO2 Al2O3 Fe2O3 MgO CaO Na2O K2O LOI*


2.68 1.32 0.82 1.34 45.3 1.21 0.47 46.87

LOI: loss of ignition.


530 S. Emad et al. / Fuel 208 (2017) 522–534

80 90

Average CO 2 capture efficiency (%)


70
80
Sorbent conversion ratio (%)

60
70
50

40 60

30 Model at
50
Model
20 Model at
Experiment
40 Model at
10
Experiment at
0 30
0 1000 2000 3000 4000 550 600 650 700 750
Time (s) Temperature (°C)

Fig. 12. Sorbent conversion ratio vs. time (T = 650 °C, u0 = 0.8 m/s). Fig. 14. Average CO2 capture efficiency vs. bed temperature.

attributed to the greater molar volume of the product CaCO3 the reaction (Eq. (6)). From 550 °C to 675 °C, the chemical kinetic
(36.9 cm3/g) compared to that of the reactant CaO (16.9 cm3/g). term is dominant leading to an increase in average efficiency and
Layers of CaCO3 are building up that impede the diffusion of CO2 after that the increasing equilibrium concentration becomes more
inside the particles. For this reason the complete carbonation of dominant resulting in a decrease in average efficiency. An opti-
the particle couldn’t be possible. mum temperature of 675 °C has also been reported by Mostafavi
Comparisons between experimental results and model predic- et al. [33]
tions given in Figs. 10–12 indicate satisfying. Fig. 14 also indicates that decreasing fluidization velocity moves
Parametric studies have been conducted to assess influences of the peak slightly to the left. As fluidization velocity decreases, the
operating parameters on carbonator behavior by evaluating gas molecules have increased residence time in the bed (i.e. have
capture efficiency at different bed temperatures, fluidization veloc- more time to react with solid sorbent) which reduces the effect
ities, particle sizes, and static bed height. of decreased reaction rate at lower temperatures.

4.2. Effect of operating temperature 4.3. Effect of fluidization velocity

Fig. 13 shows the model results at different temperatures from Different fluidization velocities have been tested (from 0.4 to
550 °C to 700 °C. The average capture efficiency versus bed tem- 1.2 m/s) to investigate the effect of hydrodynamics on the reactor,
perature is shown in Fig. 14 along with experimental results. The Fig. 15. It can be observed from Fig. 16 that the average efficiency
maximum average capture efficiency is found to occur around a increases with decreasing the fluidizing velocity due to increased
bed temperature of 675 °C. To understand the existence of such residence time and enhanced mass transfer from bubbles to emul-
an optimum bed temperature, we need to keep in mind that oper- sion. Meanwhile, Kunii and Levenspiel [15] stated that bubbling
ating temperature has two opposing effects. The positive effect is bed starts to switch to slugging when dbm =dbed > 0:6. At a fluidiza-
that increasing temperature enhances the chemical kinetics of tion velocity of 0.4, dbm =dbed  0:36, while at a velocity of 1.2 the
the carbonation reaction (Eq. (7)) and the negative effect is that ratio  0.93 indicating that slugging takes place resulting in poor
increasing temperature also increases the equilibrium concentra- mass transfer.
tion and partial pressure of carbon dioxide which slows down However it should be noted that for the same amount of CO2,
decreasing fluidization velocity would require larger system (e.g.

16
T= 550 Exp. at 550 20
CO2 exit mole fraction (%)

14 u = 0.4 m/s u = 0.6 m/s u = 0.8 m/s


T= 600 Exp. at 675 18 u = 1.0 m/s u = 1.2 m/s Exp. at 0.4
T= 650 Exp. at 700
12 Exp. at 0.8 Exp. at 1.2
CO2 exit mole fraction (%)

16
T= 675
14
10 T= 700

12
8
10
6 8

6
4
4
2
2
0 0
0 500 1000 1500 2000 2500 3000 3500 4000 0 1000 2000 3000 4000 5000 6000
Time (s) Time (s)

Fig. 13. CO2 exit mole fraction vs. time at different bed temperatures (u0 = 0.8 m/s). Fig. 15. CO2 exit mole fraction at different fluidization velocities (T = 650 °C).
S. Emad et al. / Fuel 208 (2017) 522–534 531

85 100
Model

Average CO2 capture efficiency (%)


Average CO2 capture efficiency (%)
90
80 Experiment 80
70
75
60
70 50
40
65
30 Model
20 Experiment
60
10
55 0
0.4 0.6 0.8 1 1.2 300 400 500 600 700 800 900 1000
Fluidization velocity (m/s) Parcle diameter ( m)
Fig. 16. Average CO2 capture efficiency vs. fluidization velocity (at T = 650 °C). Fig. 18. Average CO2 capture efficiency vs. particles diameter (T = 650 °C,
u0 = 0.8 m/s).

wider reactor or multiple reactors in parallel) to handle the


required flow rate of flue gases. So a compromise between perfor- ble effect of particle size on the kinetic parameters, more investiga-
mance and capital cost would be required. tions on the porous structure of lime particles is required to fully
describe the dependency of capture efficiency on particles size.
4.4. Effect of bed particle size
4.5. Effect of static bed height
Increasing particle size was found to result in a decrease in exit
CO2 mole fraction and an increase in the capture efficiency of the In order to understand the effects of static bed height on carbon
carbonator as shown in Figs. 17and 18. It should be noted that dioxide capture, the model has been run at different heights (from
increasing particle size results in an increase in minimum fluidiza- 0.05 m to 0.90 m). Fig. 19 shows model prediction for the mole
tion velocity. Consequently larger particles lead to less bubbling fraction of carbon dioxide at the exit of reactor. Increasing bed
bed as long as the fluidization velocity is kept constant. This is height increases carbonation time due to increasing active solid
analogous to decreasing fluidization velocity with the same parti- inventory. Also, reaction is allowed to proceed longer and flue
cle size as discussed above. For example, the corresponding mini- gases get in contact with active particles for more time as they pass
mum fluidization velocity for 300 lm and 500 lm particles are through the bed which improves the capture efficiency. However,
0.04 and 0.1 m/s, respectively. So, for the same fluidization velocity Fig. 20 shows that further increase of bed height has little effect
of 0.8 m/s, the mean bubble diameter in case of smaller particles is on capture efficiency improvement. This is due to the fact that bub-
larger (8 cm for 300 lm particles compared to 5 cm for 500 lm bles continue to expand and coalesce with height, so the higher
particles), resulting in smaller mass transfer coefficient. However, zones of the bed suffer from poor mass transfer coefficient and
this conclusion is limited to the particles size range under consid- probably the bed turns into slugging. This can be checked for using
eration of the present study, as further increases in particle diam- the condition that dbm =dbed > 0:6 stated above. At a static bed
eter would reduce the total reacting surface area significantly. This height of 15 cm, dbm =dbed  0:5, however at a height of 60 cm the
can be observed from gradual decrease of the slope of the curve in ratio is higher than unity indicating that slugs are found in consid-
Fig. 18. erably large portion of the bed. Kunii and Levenspiel [15] gave a
Moreover, changing particle size is expected to change the por- similar condition of slugging describing a critical slugging height
ous structure of solid reactant leading to a change in kinetic after which slugging can take place as long as superficial gas veloc-
parameters. Although Bhatia and Perlmutter [34] reported negligi- ity is higher than minimum slugging velocity as follows;
16

14 22
H = 0.05 m Hs= 0.15 m
s
CO 2 exit mole fraction (%)

Hs= 0.30 m Hs= 0.45 m


CO 2 exit mole fraction (%)

12 Hs= 0.60 m Hs= 0.75 m


18
Hs= 0.90 m
10
14
8
dp = 300 m
6 dp = 500 m 10
4 dp = 700 m
dp = 900 m 6
2

0 2
0 1000 2000 3000 4000 5000 6000 0 3000 6000 9000 12000 15000 18000
Time (s) Time (s)

Fig. 17. CO2 exit mole fraction at different particles sizes (T = 650 °C, u0 = 0.8 m/s). Fig. 19. CO2 Exit mole fraction at different bed heights (T = 650 °C, u0 = 0.8 m/s).
532 S. Emad et al. / Fuel 208 (2017) 522–534

80 The dependency of performance on different input parameters


has been discussed. Operating temperature has major effects as it
Average CO2 capture efficiency (%)

determines kinetic constants and equilibrium conditions. Opti-


75 mum operating temperature of around 675 °C is obtained. Flu-
idization velocity determines the residence time and mass
transfer coefficients. An increase in particles size affects minimum
70
fluidization conditions and causes the bed to be less bubbling with
better mass transfer. However further investigations on the effect
65 of particle size on kinetic constants are required. Bed height has lit-
Model tle effects on the capture efficiency except at relatively low heights.
Experiment
60 Appendix A.

Time required to bring down the temperature of fluidized bed


55
0 0.2 0.4 0.6 0.8 1 reactor from 950 to 650 °C:
Static bed height (m) Considering a batch of solids originally at temperature T pi flu-
idized by gas entering at temperature T gi , the following assump-
Fig. 20. Average CO2 capture efficiency vs. bed height (T = 650 °C, u0 = 0.8 m/s). tions are made;

1- The temperature of solids is independent of location in the


 0:5 0:175 bed (due to high internal circulation rates and large heat
ums ¼ umf þ 0:07 ðg  dbed Þ &zs ¼ 0:6ðdbed Þ
capacity).
In present case, ums  0:2 m=s and zs  40 cm. 2- The temperature of gas at bed exit (T ge ) at any time is equal
to bed solids temperature at that time, i.e. thermal equilib-
rium is achieved before bed exit.
4.6. Switching time
Applying heat balance on whole bed during a short time inter-
As stated before, switching time is an important parameter in val dt;
the operation of the present configuration and has been taken as
the period of time after which the capture efficiency drops below ðheat gained by gasÞ ¼ ðheat lost by solidsÞ
85% of the maximum value obtained at time zero. For instance,
dT p
in the base case instantaneous CO2 capture efficiency at time zero _ g  C pg  ðT ge  T gi Þ ¼ ms  C ps 
m
74%, Fig. 10. When it falls below 0.85 of this value (i.e. dt
0.85*74%  63%), switching takes place. In this case, switching time Using assumption #2 and rearranging;
was found to be 2100s 35 min. Switching time is shown in Table 2
dT p
for some suggested design choices along with average CO2 capture qg :u0 :C pg :ðT p  T gi Þ ¼ qs :Hs :ð1  es Þ:C ps :
efficiency and average sorbent conversion. This table would give
dt
valuable insights to the designer in order to build an efficient Integrating the previous equation with (T p ¼ T pi att ¼ 0) gives;
and cost-effective carbon dioxide capture facility. Average sorbent  
T p  T gi qg  C pg u0
conversion at switching time can be considered as a utilization fac- ¼ exp   t
T pi  T gi qs  C ps ð1  es Þ  Hs
tor. Lower values indicate that capture efficiency decreases rapidly
before most of the sorbent gets converted. Maximum average cap- Rough estimation of fuel consumption in calciner:
ture efficiency and maximum sorbent conversion occurs at lower    
velocities and higher beds. However, this would require multiple heal liberated by heat to
¼
reactors to handle the required flow of exhaust gases and more sor- fuel combustion solids
bent inventories. A compromise between performance and eco-   
heal to decompose heat to
nomics would be required in most cases. þ þ
CaCO3 fuel
 
heat to
þ
5. Conclusion gases

A dynamic model has been presented to evaluate carbon diox- mf  LHV  gcomb ¼ ms  C ps  ðT calc  T carb Þ þ ms  DHr
ide capture using the carbonation reaction between calcium oxide þ mf  C pf  ðT calc  T amb Þ
(from lime) and carbon dioxide. It describes carbonator perfor- þ mf  ð1 þ AFÞ  C pg  ðT calc  T gi Þ
mance at different conditions. The model makes a coupling
between hydrodynamics and kinetics to give more realistic ½ms  C ps  ðT calc  T carb Þ þ ms  DHr 
insights. It can be used to make design choices such as bed sizing mf ¼
ðLHV  gcomb  C pf  ðT calc  T amb Þ  ð1 þ AFÞ  C pg  ðT calc  T gi ÞÞ
and determine optimum operating conditions that maximize the
capture efficiency. Carbonator model can be integrated with the Without Oxygen preheating;
whole plant simulations to predict thermal efficiency penalties of
the carbon capture process. T gi ¼ T amb ffi 300 K
A novel scheme is proposed to allow smooth gas switching With Oxygen preheating;
between reactors using an intermediate third small reactor. The
"Z Z #
later scheme serves in energy saving as well. The scheme can save 1 t¼t preheat t¼t final

about 2.24% of fuel consumption necessary for calcination process T gi ¼  T preheat  dt þ T amb  dt
time t¼0 t¼t preheat
by preheating the oxygen required for combustion.
S. Emad et al. / Fuel 208 (2017) 522–534 533

Table 2
Switching time, average CO2 capture efficiency and average sorbent conversion at selected cases (For all; dbed ¼ 10:5 cm;dp ¼ 500 lm and T ¼ 675
C).

u0 (m/s) Hs (cm) tlim (min) gavg (%) Xavg (%)


#1 0.4 15 55 83.80 58.37
#2 0.4 30 118 84.43 62.94
#3 0.4 45 182 84.69 64.84
#4 0.4 60 246 84.83 65.91
#5 0.8 15 26 74.75 49.36
#6 0.8 30 58 76.36 56.30
#7 0.8 45 91 77.01 59.51
#8 0.8 60 124 77.36 61.39
#9 1.2 15 18 64.25 44.48
#10 1.2 30 41 66.45 52.25
#11 1.2 45 64 67.38 56.04
#12 1.2 60 89 67.89 58.41

Appendix B. hydrogen production. Chem Eng Res Des 2011;89:836–55. http://dx.doi.org/


10.1016/j.cherd.2010.10.013.
[6] Romeo LM, Catalina D, Lisbona P, Lara Y, Martínez A. Reduction of greenhouse
Derivation of Eq. (22) is given below; gas emissions by integration of cement plants, power plants, and CO2 capture
  systems. Greenh Gases Sci Technol 2011;1:72–82. http://dx.doi.org/10.1002/
dX d NCaCO3 1 d ghg3.5.
¼ ¼  ðNCaCO3 Þ ðB1Þ [7] Hirama T, Hosoda H, Kunihiro K, Shimizu T. Separating carbon dioxide from
dt dt NCaOinitial NCaOinitial dt
gases containing it. GB 2291051 (A); 1996.
[8] Shimizu T, Hirama T, Hosoda H, Kitano K, Inagaki M, Tejima K. A twin fluid-bed
d d d reactor for removal of CO2 from combustion processes. Chem Eng Res Des
ðNCaCO3 Þ ¼ ðNCO2 Þ ¼ ðNCO2  NCO2out Þ 1999;77:62–8. http://dx.doi.org/10.1205/026387699525882.
dt dt captured dt in
[9] Abanades JC, Anthony EJ, Wang J, Oakey JE. Fluidized bed combustion systems
¼ C CO2  Q gasin  C CO2out  Q gasout ðB2Þ integrating CO2 capture with CaO. Environ Sci Technol 2005;39:2861–6. http://
in
dx.doi.org/10.1021/es0496221.
[10] Stanmore BR, Gilot P. Review-calcination and carbonation of limestone during
Other Species ðexcluding CO2 ÞMolar Balance ði:e: N2 þ H2 O þ . . .Þ thermal cycling for CO2 sequestration. Fuel Process Technol 2005;86:1707–43.
http://dx.doi.org/10.1016/j.fuproc.2005.01.023.
NðN2 þH2 Oþ...Þin ¼ NðN2 þH2 Oþ...Þout [11] Lu DY, Hughes RW, Anthony EJ. Ca-based sorbent looping combustion for CO2
C ðN2 þH2 Oþ...Þin  Q gasin ¼ C ðN2 þH2 Oþ...Þout  Q gasout capture in pilot-scale dual fluidized beds. Fuel Process Technol
P P ðB3Þ 2008;89:1386–95. http://dx.doi.org/10.1016/j.fuproc.2008.06.011.
*C i ¼ xi  RuPT ; xiin ¼ xiout ¼ 1 [12] Wang W, Ramkumar S, Li S, Wong D, Iyer M, Sakadjian BB, et al. Subpilot
)ð1  xCO2 Þ  Q gasin ¼ ð1  xCO2out Þ  Q gasout demonstration of the Carbonation-Calcination Reaction (CCR) process: high-
in temperature CO2 and sulfur capture from coal-fired power plants. Ind Eng
Chem Res 2010;49:5094–101. http://dx.doi.org/10.1021/ie901509k.
Substitution of Eq. (B2) in Eq. (B3) gives: [13] Arias B, Diego ME, Abanades JC, Lorenzo M, Diaz L, Martínez D, et al.
Demonstration of steady state CO2 capture in a 1.7 MWth calcium looping
d 1  xCO2 pilot. Int J Greenh Gas Control 2013;18:237–45. http://dx.doi.org/10.1016/j.
ðNCaCO3 Þ ¼ C CO2  Q gasin  C CO2out  Q gasin  in

dt in 1  xCO2out ijggc.2013.07.014.
! [14] Charitos A, Rodríguez N, Hawthorne C, Alonso M, Zieba M, Arias B, et al.
1  xCO2 Experimental validation of the calcium looping CO2 capture process with two
¼ Q gasin  C CO2  C CO2out  in
ðB4Þ circulating fluidized bed carbonator reactors. Ind Eng Chem Res
in 1  xCO2out 2011;50:9685–95. http://dx.doi.org/10.1021/ie200579f.
[15] Kunii D, Levenspiel O. Fluidization engineering. Butterworth-Heinemann;
1991.
Ws [16] Baker EH. The calcium oxide-carbon dioxide system in the pressure range 1–
NCaOinitial ¼ ðNCaO þ NCaCO3 Þat any time ¼ ðB5Þ 300 atmospheres. J Chem Inf Model 1962:464–70. http://dx.doi.org/10.1017/
MW s
CBO9781107415324.004.
Substituting Eq. (B4) and Eq. (B5) in (B1) gives: [17] Alvarez D, Abanades JC. Determination of the critical product layer thickness in
the reaction of CaO with CO2. Ind Eng Chem Res 2005;44:5608–15. http://dx.
!
dX MW s 1  xCO2 doi.org/10.1021/ie050305s.
¼  Q gasin  C CO2  C CO2out  in
ðB6Þ [18] Alonso M, Rodríguez N, Grasa G, Abanades JC. Modelling of a fluidized bed
dt Ws in 1  xCO2out carbonator reactor to capture CO2 from a combustion flue gas. Chem Eng Sci
2009;64:883–91. http://dx.doi.org/10.1016/j.ces.2008.10.044.
Integrating Eq. (B6) over time leads to the following; [19] Romano MC. Modeling the carbonator of a Ca-looping process for CO2 capture
from power plant flue gas. Chem Eng Sci 2012;69:257–69. http://dx.doi.org/
Z ! 10.1016/j.ces.2011.10.041.
MW s t 1  xCO2
X av g ¼  Q gasin  C CO2  C CO2out  in
dt [20] Li Z, Fang F, Tang X, Cai N. Effect of temperature on the carbonation reaction of
Ws 0
in 1  xCO2out CaO with CO2. Energy Fuels 2012;26:2473–82. http://dx.doi.org/10.1021/
ef201543n.
[21] Szekely J, Evans JW. Gas-solid reactions. Academic Press; 1976.
[22] Sun P, Grace JR, Lim CJ, Anthony EJ. Determination of intrinsic rate constants of
References the CaO-CO2 reaction. Chem Eng Sci 2008;63:47–56. http://dx.doi.org/
10.1016/j.ces.2007.08.055.
[1] Statistics I. CO2 emissions from fuel combustion-highlights 2016. [23] Barker R. The reversibility of the reaction CaCO3 ¢ CaO + CO2. J Appl Chem
[2] Metz B, Davidson O, De Coninck HC, Loos M, Meyer LA. IPCC, 2005: IPCC special Biotechnol 1973;23:733–42.
report on carbon dioxide capture and storage. Prepared by Working Group III [24] Grasa GS, Abanades JC. CO2 capture capacity of CaO in long series of
of the Intergovernmental Panel on Climate Change. Cambridge, United carbonation/calcination cycles. Ind Eng Chem Res 2006;45:8846–51. http://
Kingdom New York, NY, USA 2005;442. dx.doi.org/10.1021/ie0606946.
[3] Boot-Handford ME, Abanades JC, Anthony EJ, Blunt MJ, Brandani S, Mac Dowell [25] Wen CY, Yu YH. A generalized method for predicting the minimum fluidization
N, et al. Carbon capture and storage update. Energy Environ Sci velocity. AIChE J 1966;12:610–2.
2014;7:130–89. http://dx.doi.org/10.1039/C3EE42350F. [26] Grace JR. Contacting modes and behaviour classification of gas–solid and other
[4] Abanades JC, Arias B, Lyngfelt A, Mattisson T, Wiley DE, Li H, et al. Emerging two-phase suspensions. Can J Chem Eng 1986;64:353–63.
CO2 capture systems. Int J Greenh Gas Control 2015;40:126–66. http://dx.doi. [27] Broadhurst TE, Becker HA. Onset of fluidization and slugging in beds of
org/10.1016/j.ijggc.2015.04.018. uniform particles. AIChE J 1975;21:238–47.
[5] Dean CC, Blamey J, Florin NH, Al-Jeboori MJ, Fennell PS. The calcium looping [28] Darton RC, LaNauze RD, Davidson JF, Harrison D. Bubble growth due to
cycle for CO2 capture from power generation, cement manufacture and coalescence in fluidised beds. Trans Inst Chem Eng 1977;55:274–80.
534 S. Emad et al. / Fuel 208 (2017) 522–534

[29] Abanades JC, Anthony EJ, Lu DY, Salvador C, Alvarez D. Capture of CO2 from [33] Mostafavi E, Sedghkerdar MH, Mahinpey N. Thermodynamic and kinetic study
combustion gases in a fluidized bed of CaO. AIChE J 2004;50:1614–22. http:// of CO2 capture with calcium based sorbents: experiments and modeling. Ind
dx.doi.org/10.1002/aic.10132. Eng Chem Res 2013;52:4725–33. http://dx.doi.org/10.1021/ie400297s.
[30] Okasha F. Modeling of liquid fuel combustion in fluidized bed. Fuel [34] Bhatia SK, Perlmutter DD. Effect of the product layer on the kinetics of the CO2-
2007;86:2241–53. http://dx.doi.org/10.1016/j.fuel.2007.01.014. lime reaction. AIChE J 1983;29:79–86. http://dx.doi.org/10.1002/
[31] The MathWorks Inc. MATLAB 2014b 2014. aic.690290111.
[32] Dormand JR, Prince PJ. A family of embedded Runge-Kutta formulae. J Comput
Appl Math 1980;6:19–26. http://dx.doi.org/10.1016/0771-050X(80)90013-3.

You might also like