Вы находитесь на странице: 1из 12

Chemical Engineering Science 59 (2004) 931 942

www.elsevier.com/locate/ces

Modeling and simulation for the methane steam reforming enhanced by


in situ CO2 removal utilizing the CaO carbonation for H2 production
Deuk Ki Leea; , Il Hyun Baekb , Wang Lai Yoonc
a Division

of Civil and Environmental Engineering, Gwangju University, Gwangju 503-703, Republic of Korea
of Energy and Environmental Research, Korea Institute of Energy Research, Daejon 305-343, Republic of Korea
c Energy Conversion Process Research Center, Korea Institute of Energy Research, Daejon 305-343, Republic of Korea

b Division

Received 30 July 2003; received in revised form 8 December 2003; accepted 13 December 2003

Abstract
The transient behavior of catalytic methane steam reforming (MSR) coupled with simultaneous carbon dioxide removal by carbonation
of CaO pellets in a packed bed reactor for hydrogen production has been analyzed through a mathematical model with reaction experiments
for model veri8cation. A dynamic model has been developed to describe both the MSR reaction and the CaO carbonation-enhanced MSR
reaction at non-isothermal, non-adiabatic, and non-isobaric operating conditions assuming that the rate of the CaO carbonation in a local
zone of the packed bed is governed by kinetic limitation or by mass transfer limitation of the reactant CO2 . Apparent carbonation kinetics
of the CaO pellet prepared has been determined using the TGA carbonation experiments at various temperatures, and incorporated into the
model. The resulting model is shown to successfully depict the transient behavior of the in situ CaO carbonation-enhanced MSR reaction.
The e<ects of major operating parameters on the transient behavior of the CaO carbonation-enhanced MSR have been investigated using
the model. The bed temperature is the most important parameter for determining the amount of CO2 removed by carbonation of CaO,
and at temperatures of 600 C, 650 C, 700 C and 750 C, the CO2 uptake is 1.43, 2.29, 3.5 and 5:09 mol-CO2 =kg-CaO, respectively.
Simultaneously with the increase in CO2 uptake with increasing temperature, the corresponding amounts of hydrogen produced are 1.56,
2.54, 3.91 and 5:63 mol-H2 /kg-CaO, at the same temperatures as above. Operation at high pressure, high steam to methane feed ratio,
and the decreased feed rate at a given temperature are favorable for increasing the degree of the overall utilization of CaO pellets in the
reactor bed, and for lowering the CO concentration in the product.
? 2004 Elsevier Ltd. All rights reserved.
Keywords: Methane steam reforming; CaO carbonation; CO2 removal; H2 production; Modeling; Packed bed

1. Introduction

CO + H2 O CO2 + H2 ;

Methane steam reforming (MSR) is a major route for the


industrial production of H2 . The three main reactions in a
MSR reactor are represented by following equations (Xu
and Froment, 1989):

FH298 = 41:1 103 kJ=kmol:

CH4 + H2 O CO + 3H2 ;
FH298 = 206:2 103 kJ=kmol;

(1)

CH4 + 2H2 O CO2 + 4H2 ;


FH298 = 164:9 103 kJ=kmol:

(2)

Corresponding author. Tel.: +82-62-670-2394;


fax: +82-62-670-2192.
E-mail address: dklee@gwangju.ac.kr (D.K. Lee).

0009-2509/$ - see front matter ? 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2003.12.011

(3)

Reforming reactions (1) and (2) are highly endothermic, and


thermodynamically favored by high temperature and low
pressure. On the other hand, the watergas shift (WGS) reaction given by Eq. (3) is favored at low temperature, but
it has no pressure dependence. MSR is generally operated
at a temperature of 750900 C due to the overall endothermic nature of the reactions (Hufton et al., 1999). Although
high-temperature operation is indispensable for a substantial
conversion of CH4 , it facilitates the reverse WGS reaction,
giving the product gas containing 810% CO on a dry basis. For the purpose of obtaining the product gas with less
CO and more H2 , it is conventional that the MSR product
gas is fed to another reactor where the temperature is kept
as low as 300400 C for the WGS reaction to take place

932

D.K. Lee et al. / Chemical Engineering Science 59 (2004) 931 942

prior (Hufton et al., 1999). To obtain the H2 product stream,


the eKuent is then cooled and fed to a multicolumn pressure
swing adsorption (PSA) process.
During industrial operation of the MSR process, it is customary to release the product CO2 , which is a greenhouse
gas with potential to contribute to global warming, to the environment. To be ready for possible obligation to decrease
the amount of CO2 released, it is important for industries to
develop means to capture/sequestrate CO2 from their processes. As far as the MSR for the production of H2 is concerned, in situ capture of CO2 from the MSR reaction system
has advantages of not only providing a chance to sequestrate the greenhouse gas instead of its release to the atmosphere, but also bringing the following separation-enhanced
reaction merits: higher CH4 conversion and H2 yield than
those equilibrium-limited at a given condition. Provided that
a CO2 -acceptor is available in the reaction zone, the conversion of CH4 to CO2 through Eq. (2) is enhanced, as is
that produced via CO intermediate.
Several studies concerning MSR or WGS enhanced by in
situ separation of CO2 employing adsorbents or chemical
acceptors are available in the literature. Using dolomite as
CO2 acceptor in a Muidized bed reactor containing Ni-based
catalyst, Brun-Tsekhovoi et al. (1986) reported a high enhancement of CH4 conversion to H2 . Han and Harrison
(1994) investigated H2 production via the WGS reaction
using dolomite as a CO2 acceptor in the temperature range
of 500600 C. At the most favorable conditions, the total
concentration of carbon oxides in the product gas was low
as 50 ppm. A concept called sorption-enhanced reaction
process (SERP) was suggested by Circar and co-workers
(Carvill et al., 1996; Hufton et al., 1999; Waldron et al.,
2001) for H2 production by MSR using a packed column
containing an admixture of MSR catalyst and adsorbent
to remove CO2 from the reaction zone. In the SERP, the
adsorbent, a potassium carbonate promoted hydrotalcite
over which CO2 could reversibly adsorb in the temperature range of 300500 C in the presence of excess steam,
was periodically regenerated by PSA. They reported that
the SERP concept allowed direct production of high-purity
H2 ( 95 mol%) at high methane to hydrogen conversion
( 80%) with dilute CH4 and trace carbon oxide impurities
at 450 C and 4:8 bar. Once adsorbents have been saturated
with CO2 , the reaction-enhancement e<ect does not exist any
more. Therefore, such methods as pressure-, concentration-,
and temperature-swing operations, or reactive regeneration are developed to regenerate the spent adsorbents (Xiu
et al., 2002). MSR employing hydrotalcite-based CO2 adsorbents was studied experimentally and theoretically in
detail recently (Ding and Alpay, 2000; Xiu et al., 2002).
Ding and Alpay (2000) analyzed the transient behavior of
a tubular reactor when a Ni-based catalyst was admixed
with the adsorbent. Considerable enhancement of the CH4
conversion was experimentally demonstrated, and predicted
by a mathematical model. Xiu et al. (2002) reported that
hydrogen-enriched stream with traces of CO2 and CO could

be produced by a sorption-enhanced MSR process with


reactive regeneration of adsorbent where the purge step
was performed at a low temperature of 400 C (compared
with the reaction temperature of 450 C) with 10% H2 in
nitrogen at atmospheric pressure. It is noticeable above
that the CO2 removal-enhanced MSR processes mostly employ hydrotalcite-based sorbents. Reversible CO2 uptake
capacity of the hydrotalcite-based sorbent was reported to
be about 0:45 mol=kg at 400 C after 10 cycles of adorption/desorption (Hufton et al., 1999). Such a low uptake
capacity of the adsorbent means a very short period of service, requiring frequent regeneration of the spent sorbents
even in a large reactor.
The application of CaO as a CO2 -acceptor to MSR in a
8xed bed reactor was reported by Balasubramanian et al.
(1999), demonstrating that hydrogen with a purity of more
than 95% could be produced in a single step MSR process
using dolomite particles as a CaO-based CO2 -acceptor in
the reactor at 650 C and above. Later, Ortiz and Harrison
(2001) reported the e<ect of regeneration conditions of the
spent dolomite as a function of temperature and regeneration gas composition on the performance in the subsequent
H2 production cycle, showing only moderate activity loss
under most of the regeneration conditions. The carbonation
reaction of CaO with CO2 is represented by
CaO + CO2 CaCO3 ;
FH298 = 178:8 103 kJ=kmol:

(4)

CaO carbonation as a kind of non-catalytic gassolid reaction was reported to not proceed to the complete conversion
of CaO, with ultimate conversions in the range of 7080%
for fresh powder samples (Bhatia and Perlmutter, 1983).
However, the ultimate carbonation conversion of CaO after
14 cycles of CaO-carbonation/CaCO3 -calcination was reported to decrease to 20% (Abanades, 2002). Even with this
lowered carbonation conversion, the CO2 -uptake capacity
of CaO is calculated to 3:57 mol=kg, much higher than that
of the hydrotalcite-based sorbent. This large amount of CO2
uptake is of great advantage to the MSR process utilizing
the CaO carbonation for in situ CO2 removal as compared
with the process utilizing the hydrotalcite-based sorbent.
Whereas the CO2 removal-enhanced MSR reaction utilizing
hydrotalcite-based sorbents was well analyzed using mathematical models (Ding and Alpay, 2000; Xiu et al., 2002),
no studies concerning transient modeling of the MSR reaction combined with the CaO carbonation as a non-catalytic
gassolid reaction have been reported.
In this work, the transient behavior of the catalytic MSR
over a Ni-catalyst coupled with the simultaneous carbonation of CaO pellets in a packed bed reactor has been analyzed for H2 production through a mathematical model and
reaction experiments for the model veri8cation. Modeling
such a process requires information on the CaO carbonation kinetics as well as reaction kinetic models for reactions
(1)(3). CaO carbonation kinetics needed for modeling has

D.K. Lee et al. / Chemical Engineering Science 59 (2004) 931 942

been obtained using TGA experiments at various temperatures for CaO pellets prepared with inorganic binders. The
reaction kinetic model proposed by Xu and Froment (1989)
has been employed for reactions (1)(3). Throughout modeling, e<ects of major operating parameters on the transient
behavior of the CaO carbonation-enhanced MSR have been
investigated and discussed. Because the CaO carbonation
reaction generates a signi8cant amount of heat, as represented by reaction (4), the thermal e<ect in the reactor on
the reaction performance has been also determined.

933

experiment was carried out as follows: (i) heat the reactor at


a rate of 5 C=min at atmospheric pressure under the Mow of
50% H2 /He up to 800 C, at which the reactor was kept for
additional 12 h for catalyst reduction; (ii) adjust the reactor
temperature to a desired reaction temperature; (iii) supply
water and H2 to the reactor so that the molar ratio of H2 O to
H2 is the same as the desired H2 O=CH4 ratio in the reaction,
(iv) pressurize the reactor using a backpressure regulator,
(v) replace the H2 Mow with CH4 to start the MSR reaction.
3. Mathematical modeling

2.1. Carbonation of CaO pellets in TGA


For the in situ CO2 removal, spherical pellets (3 mm in
diameter, surface area = 5 m2 /g, porosity = 0:045) of CaO
containing inorganic binders, feldspar (5%) and bentonite
(5%), were prepared. Carbonation conversion of the CaO
pellet was obtained using a TGA (TA Instrument). For a
CaO pellet (about 22:8 mg) placed on the TGA sample pan,
the weight change by the carbonation was measured as function of time by Mowing pure CO2 (100 ml=min) at a controlled temperature. For accurate measurement and control
of the temperature of the sample pellet during the TGA experiment, the sensing tip of a thermocouple was placed at a
position as close to the sample as possible. The durability
of the CaO pellet was tested in a separate experiment using TGA by repeating 10 cycles of carbonation (at 750 C)
followed by complete calcination (at 900 C) under Mow of
100 ml=min 10% CO2 =N2 .
2.2. MSR
A commercial Ni-based reforming catalyst (ICI 57-7, 1/8
cylindrical pellets) was used in combination with the CaO
pellets prepared. Catalyst (16:4 g) and CaO (83:6 g) pellets
were admixed, and packed into a stainless steel tube reactor of internal diameter 24 mm, and the length of the reactive packing bed was 290 mm. Ahead of the reactive bed,
spherical pellets of -alumina were packed for preheating
and even distribution of the feed. Reactor bed temperatures
were monitored with two thermocouples, one located at the
preheating zone and the other at the middle section of the reactive packing bed, and controlled with two electric heaters.
The reactor entrance was connected to the lines for feed
gases (CH4 , H2 O), reducing gas (50% H2 /He) and purging
or process gas (N2 , H2 ). The feed Mow of CH4 was regulated using a massMow controller, and steam was supplied
to the reactor via water-vaporizer using a HPLC pump. The
reactor eKuent gas for analysis was introduced to water condenser followed by a moisture trap to remove water. A GC
(Donam, DS6200) equipped with a TCD detector and a Porapak Q column was used for analysis of the dried gas sample employing a 6-port auto-sampling valve. The reaction

3.1. CaO carbonation kinetics


TGA-determined carbonation conversions of the CaO
pellet with time at temperatures 650750 C are shown in
Fig. 1. The CaO pellet exhibits a relatively low degree of
conversion at each temperature. Even at 750 C, carbonation was limited to a low 8nal conversion of about 30%.
This may be attributed to the large pellet size with the
low surface area and porosity, by which pore plugging in
the outer shell of the pellet is strongly e<ected, preventing
a large amount of intra-pellet CaO from being carbonated. A smaller pellet size would have been preferable for
higher ultimate conversion. However, this sample did not
show any sign of deterioration in the ultimate carbonation
conversion at the durability test repeating 10 cycles of carbonation/calcination in the TGA. Lee (2004) suggested an
apparent kinetic expression for the carbonation of CaO as
follows:

2
dX
X
;
(5)
= kc 1
dt
Xu
where X denotes the fractional conversion of CaO to CaCO3 ,
kc is an apparent kinetic rate constant which is dependent
50

Carbonation conversion, %

2. Experimental

Temperature, C
650
700
750
predicted

40

30

20

10

0
0

20

40

60

80

Time, min
Fig. 1. Carbonation conversions of the prepared CaO pellets under pure
CO2 .

934

D.K. Lee et al. / Chemical Engineering Science 59 (2004) 931 942

on temperature, and Xu is the ultimate conversion of CaO.


Integration of Eq. (5) leads to
X=

kc bt
;
b+t

(6)

where b is de8ned as the time taken to reach half the ultimate


conversion in the carbonation reaction at a given temperature. By substituting the relation X =Xu =2 at t =b to Eq. (6),
the ultimate conversion is given by Xu = kc b. Least-square
regression for the conversion data in Fig. 1 using Eq. (5)
gives values of kc and b at each temperature. The temperature dependencies of the two parameters have been determined as follows:
kc = 96:34 exp(12171=T );

(7)

b = 4:49 exp(4790:6=T ):

(8)

The prediction using Eq. (6) by substituting kc and b in


Eqs. (7) and (8) is in a good agreement with the conversion
data shown in Fig. 1, supporting that apparent carbonation
reaction of the CaO pellets in the TGA takes place on a rate
given by Eq. (5). Therefore, the molar rate of CO2 removal
per unit mass of CaO can be represented as


dX
1
;
(9)
rcbn =
MCaO dt
where MCaO is molecular weight of CaO. Eq. (9) is given
independent of the concentration of CO2 . Bhatia and Perlmutter (1983) reported that the rate of CaO carbonation was
independent of CO2 partial pressures, except for a slight effect at a very early stage of the carbonation. Dedman and
Owen (1962) also reported that the reaction was zero order
with respect to CO2 pressures.
3.2. Reforming and shift reaction kinetics
The reaction kinetic model of Xu and Froment
can be summarized as


PH3 2 PCO
k1
R1 = 2:5 PCH4 PH2 O
(DEN)2 ;
K1
PH2


PH4 2 PCO2
k2
2
(DEN)2 ;
R2 = 3:5 PCH4 PH2 O
K2
PH2


k3
PH2 PCO2
PCO PH2 O
(DEN)2 ;
R3 =
PH 2
K3

(1989)

(10a)
(10b)
(10c)

DEN = 1 + KCO PCO + KH2 PH2 + KCH4 PCH4


+ KH2 O PH2 O =PH2 ;

(10d)

where Rj (j = 1; 2; 3) denotes the reaction rate of the MSR


reactions (1) and (2), and the WGS reaction (3). Pi (i =
CH4 , H2 O, H2 , CO, CO2 ) corresponds to the partial pressure of species i. Expressions for the rate constants kj ,
the equilibrium constants Kj (j = 1; 2; 3), and adsorption

constants Ki (i = CO; H2 ; CH4 ; H2 O) as functions of temperature have been given by Xu and Froment (1989), and
were well summarized by Xiu et al. (2002). The rate of formation or consumption of species i is then calculated using
Eqs. (10a)(10c):
rCH4 = (R1 + R2 );

(11a)

rH2 O = (R1 + 2R2 + R3 );

(11b)

rH2 = 3R1 + 4R2 + R3 ;

(11c)

rCO = R1 R3 ;

(11d)

rCO2 = R2 + R3 :

(11e)

3.3. Governing equations


A dynamic model is developed to describe both the MSR
reaction and the CaO carbonation-enhanced MSR reaction
at non-isothermal, non-adiabatic, and non-isobaric operating
conditions. The model assumptions are summarized as follows: axial plug Mow, ideal gas behavior, Mat radial pro8le
of temperature or concentration, constant bed void fraction
and uniform particle size of the reactor-packed materials,
and constant wall temperature. Based on the above assumptions, the pseudo-homogeneous model (Feyo De Azevedo
et al., 1990) of the di<erential mass balance for a packed
bed can be written as
@Ci
@(uCi )

=
+ (1 )cat ri (1 )CaO rcbn ; (12)
@t
@z
where Ci is the molar concentration of species i,  is the bed
void fraction, u is the super8cial velocity,  is the catalyst
e<ectiveness factor, cat and CaO are the apparent density of
the catalyst and CaO pellets, respectively. The last term on
the right-hand side of Eq. (12) represents the molar rate of
CO2 removed by CaO carbonation per unit bed-volume of
the reactor. The molar rate of CO2 removal per kg CaO, rcbn ,
can be represented as a function of the fractional conversion
of CaO, X , using Eqs. (5) and (9):

2
X

1
rcbn =
;
(13)
MCaO
Xu
where  is a parameter introduced to account for that the
local rate of CaO carbonation along the packed bed may be
governed by kinetic limitation or by mass transfer limitation
of the reactant CO2 . For the former, the rate of the CaO
carbonation at a given bed temperature is determined by
kc given in Eq. (7). For the latter case, it is determined
by the availability of CO2 , which depends on the amount
entrained with upcoming convection Mow and the generation
by catalytic reactions within a local bed zone. Therefore,
 = kc under kinetic limitation;
=

MCaO
[
(1 )CaO

CO2

(14a)

+ (1 )cat rCO2 ]

under mass transfer limitation:

(14b)

D.K. Lee et al. / Chemical Engineering Science 59 (2004) 931 942

For a local bed zone where the CaO carbonation rate is


under mass transfer limitation, the quantity of CO2 removable by the carbonation of CaO corresponds to the
amount of CO2 available within the zone minus that of
CO2 escaping from the zone with a convective Mow in
the equilibrium concentration (Ceq; CO2 ) of CO2 in the
CaO-carbonation/CaCO3 -calcination reaction (4). In Eq.
(14b), CO2 is introduced to take into account part of the
convection-delivered CO2 , which is supposed to be removed by carbonation of CaO within a bed zone of a 8nite
bed length Fz:
uCCO2 uCeq; CO2
:
(15)
CO2 =
Fz
Ceq; CO2 at temperature T in Kelvin (Barin, 1989) is provided
as


1 105
12
:
Ceq; CO2 = 4:137 10 exp(20474=T )
0:082T
(16)
For a time increment Ft, the change in the carbonation
conversion of a local bed CaO at time t can be obtained as
 t+Ft
Xt+Ft = Xt + MCaO
rcbn dt:
(17)
t

The pseudo-homogeneous energy balance for a packed


bed in a plug Mow can be written as
@T
[(1 )s Cps + g Cpg ]
@t

@(ug Cpg T )
(1 )cat
Rj HRj
=
@z
j
4
;
(18)
DR
where s , the average apparent density of the two solids
in the reactor, is given by s = (Wcat + WCaO )=(Wcat =cat +
WCaO =CaO ), with Wcat and WCaO denoting the weights of the
catalyst and CaO pellets packed in the reactor, respectively.
g is the gas phase density, and Cps and Cpg are the solid
and gas heat capacities, respectively. HRj denotes the heat
of reaction j in reactions (1)(3), and Hcbn denotes the heat
of CaO carbonation reaction in reaction (4). The last term in
the right-hand side of Eq. (18) accounts for the heat transfer
through the reactor tube wall at a constant temperature TW .
Wall-bed heat transfer coeScient, hW , is given for a packed
tube reactor of inside diameter DR as Eq. (19a) by Li and
Finlayson (1977), and as Eq. (19b) by De Wash and Froment
(1972) at very low-feed Mow rates.

 

6dp
kg
Rep0:8 exp
hW = 2:03
DR
DR
(1 )CaO rcbn Hcbn + hW (TW T )

(Rep = 20 7600; dp =DR = 0:05 0:3);


 0
kz
(as Rep 0);
hW = 6:15
DR

(19a)
(19b)

where Rep = ug dp =%, and kz0 = kg [ + (1 )={0:139


0:0339 + 2=3(kg =ks )}], in which % is the viscosity of gas,

935

Table 1
Values of parameters used
Parameters

Values

Cpg
Cps
dp
DR
kg
ks
MCaO
WCaO
Wcat

%
CaO
cat
s


8:45 kJ=(kg K)
0:98 kJ=(kg K)
3 103 m
2:4 102 m
2:59 104 kJ=(m s K)
1 103 kJ=(m s K)
56 kg=kmol
83:6 103 kg
16:4 103 kg
0:5
2:8 105 Pa s
1257 kg=m3
246 kg=m3
1503 kg=m3
0.3

and kg and ks denote the thermal conductivity of gas and of


the packed solid, respectively.
Pressure distribution in the packed bed is described by
the Ergun equation (Ergun, 1952)

g u2 1  150(1 )%
dP
+ 1:75 105 : (20)
=
dz
dp

d p g u
3.4. Numerical method
The initial and boundary conditions in Eqs. (12), (18) and
(20) are set as follows:
Ci = Ci; init ;
Ci = Ci; f ;

T = Tinit
T = Tf ;

at t = 0;
P = Pf

at z = 0;

where i denotes the gas species CH4 , H2 O, H2 , CO, CO2 .


The initial bed concentrations of H2 and H2 O were set equal
to those of CH4 and H2 O in the feed, respectively, and the
initial concentrations of other gas species were set to zero.
The feed concentration of H2 was in fact zero, however, it
was assumed as 106 of the feed concentration of CH4 so
that the denominator in Eq. (10d) would not be zero.
The model equations were solved with MATLAB programming. Using the method of lines, the partial di<erential
equations (12) and (18) were converted to a set of ordinary
di<erential equations with initial conditions by discretizing
the spatial derivative in z-direction using backward 8nite
di<erences. For 8nite di<erences, the reactor of bed length
L = 0:29 m was divided into 50 sections with 51 nodes. The
initial value ordinary di<erential equations comprised of 306
equations and other explicit algebraic equations at a time t
were simultaneously solved using ode15s.m, a MATLAB
built-in solver for initial value problems for sti< ordinary
di<erential equations. A summary of the values of the parameters used in the simulation is given in Table 1.

936

D.K. Lee et al. / Chemical Engineering Science 59 (2004) 931 942

4. Results and discussion


4.1. Veri:cation of the model
Simulated pro8les of the product gas composition on a
dry basis are shown in Fig. 2 in comparison with experimental data obtained from the in situ CaO carbonation-enhanced
MSR reaction under the operating conditions: 700 C, 3 bar,
CH4 feed rate = 11:2 NL=h, molar H2 O=CH4 feed ratio = 3.
Total feed Mow rate through the reactor bed corresponds
to 341:5 h1 in gas hourly space velocity (GHSV). Dotted
lines represent the simulated results for the reaction in the absence of the CaO pellets under the same operating conditions
as above, indicating that the reaction reaches at a steady state
shortly after the start of run. In the simulation, feed temperature, initial bed temperature, and reactor wall temperature
were all set to 700 C, and the feed pressure was 3 bar. Except for some scattered experimental data in CH4 composition, excellent agreement exists between the experiment and

100
simulated with CaO
simulated without CaO
experimental

90

H2

80

70

Product gas composition, %

12
9
6
3

CO2

0
6
4
2

CH4

0
12
9
6

CO

3
0
0

40

80

120

160

200

Time on stream, min


Fig. 2. Product gas compositions on a dry basis with the reaction
time on stream: comparison of the simulation with the experiment
(temperature = 700 C, pressure = 3 bar, feed CH4 = 11:2 NL=h, feed
H2 O=CH4 = 3, GHSV = 341:5 h1 ).

the simulation for the CaO carbonation-enhanced MSR. The


concentration of CO2 in the product gas is kept low for the
8rst 28 min of reaction due to its in situ removal by the carbonation of CaO pellets in the reactor bed. During such a prebreakthrough period, reaction enhancements are observed in
the H2 production, the CH4 conversion and the CO reduction. As the degree of carbonation of all the CaO pellets in
the bed approaches the ultimate conversion, which depends
on the local bed temperature, the eKuent concentration of
CO2 rises to a steady value, with the enhancement e<ects
being diminished. For both the prebreakthrough and postbreakthrough periods, the product gas composition is found
close to that of equilibrium at each period. Equilibrium composition for a steady-state period of prebreakthrough nearly
coincides with the simulation results represented by solid
lines while that of postbreakthrough by dotted lines.
Fig. 3 shows the reaction time-dependent axial distributions of CH4 conversion, reactor-bed temperature, and CaO
carbonation conversion. As shown in Fig. 3(A), CH4 conversion towards equilibrium has been accomplished mostly
in the 8rst quarter of the reactor bed. The reactor bed temperature at this MSR reaction-active region rapidly drops to
about 610 C from the initial 700 C, as shown in Fig. 3(B),
due to the large endothermic heat of the MSR reaction. Endothermic MSR in this region is fast enough to reach at
the equilibrium conversion of CH4 , whereas the exothermic
CaO carbonation reaction is quite slow, resulting in a large
temperature decrease. A more reactive CaO (larger porosity) would result in these rates being more equal, would
lessen the large temperature decrease near the reactor entrance. But, the heat e<ect by the exothermic carbonation of
CaO would not last for a long period of time because the
carbonation of CaO at near the reactor entrance would reach
at the state of breakthrough, and after that, no more heat
would be generated. In this cool bed region, the rate of CO2
removal by the carbonation of CaO before its exhaustion is
mostly kinetically limited because of the lowered temperature regardless of an abundance of CO2 available for the
carbonation. This also leads to a gradual increase in the CaO
carbonation conversion up to a certain position along the
bed length, as shown in the pro8les of Fig. 3(C). In the subsequent high-temperature bed after this cool bed region, the
endothermic MSR does not occur to a degree of practical
signi8cance, but only the carbonation reaction of the packed
CaO proceeds as long as it has not been exhausted. The rate
of CaO carbonation in this bed region changes to be limited
by the small amount of CO2 available for the reaction. Local temperature of a bed region where only the carbonation
reaction is occurring increases due to the exothermic heat
of the carbonation reaction. A maximum increase to 25 C
above the initial temperature is observed at the reactor position 0:05 m from the entrance after 1 min of the reaction. It
can be observed that the bed zone with such a temperature
increase where the CaO carbonation is occurring moves toward the reactor exit with the reaction time on stream. The
temperature rise at the reactor exit position is 15 C after

D.K. Lee et al. / Chemical Engineering Science 59 (2004) 931 942

937

100
90

H2

80

Temperature, C
600
650
700
750

70
60

Product gas composition, %

15
10
5

CO2

0
15

CH4

10
5
0
15
10
5

CO
0
0

40

80

120

160

200

Time on stream, min


Fig. 4. Simulated pro8les of the product gas composition on a dry basis
with the change in temperature (pressure = 3 bar, feed CH4 = 11:2 NL=h,
feed H2 O=CH4 = 3, GHSV = 341:5 h1 ).

Fig. 3. Time dependent spatial distribution of CH4 conversion (A), the


reactor-bed temperature (B) and the carbonation conversion of CaO (C)
in the simulated run shown in Fig. 2.

about 28 min of the reaction time on stream, which can be


regarded as a breakthrough time of CaO in the packed bed.
Like the present simulation, Han and Harrison (1994) reported a bed temperature increase of 1015 C in the WGS
reaction using dolomite as a CO2 acceptor in the temperature range of 500600 C.
The carbonation conversion of CaO in the bed after
180 min of run ranges from 9.7% in the vicinity of the reactor entrance to 20.7% at the exit. Such simulated behavior
of the CaO carbonation can be reasonably acceptable. It
can be therefore stated that the kinetics of the MSR reaction and that of the CaO carbonation, and the assumptions
made in the current modeling are adequate to describe the
transient behavior of the in situ CaO carbonation-enhanced
MSR reaction in a packed bed reactor.

4.2. E;ect of temperature


For the in situ CaO carbonation-enhanced MSR reaction,
temperature is a critical variable by which the extent of the
ultimate conversion of CaO as well as the rate of the carbonation is determined. Kinetic rate and equilibrium gas
composition of the MSR reaction are also a<ected by temperature. Fig. 4 shows the simulated pro8les of the product gas composition with the change of temperatures in a
range of 600750 C. The other operating conditions employed here are the same as those in Fig. 2. At the low
temperature of 600 C, the maximum carbonation conversion of the CaO pellets is as low as 8.4% at the end of
180 min run, resulting in the absence of any discrete prebreakthrough period. At 650 C, the prebreakthrough period
is extended to about 10 min by virtue of the increase in
the maximum conversion of CaO as 13.5%. At 750 C, although there exists a long prebreakthrough period due to
the much enhanced conversion of CaO ranging from 12.8 at
the vicinity of the reactor entrance to 30% at the exit, CO2

938

D.K. Lee et al. / Chemical Engineering Science 59 (2004) 931 942

Table 2
Amounts of CO2 uptake and those of additional H2 production per kg of
CaO in the reactor bed by the in situ carbonation removal of CO2

Temperature, ( C)
at 3 bar, 3 H2 O=CH4 ,
341:5 h1
Pressure (bar)
at 700 C, 3 H2 O=CH4 ,
341:5 h1
Feeding ratio of H2 O=CH4
at 700 C, 7 bar, 341:5 h1
GHSV (h1 )
at 700 C, 7 bar,
3 H2 O=CH4

CO2 uptake
(mol/kg-CaO)

H2 productivity
(mol/kg-CaO)

600
650
700
750

1.43
2.29
3.50
5.09

1.56
2.54
3.91
5.63

3
7
15

3.50
3.52
3.55

3.91
3.98
4.08

3
5
7

3.52
3.53
3.54

3.98
4.00
4.01

3.54
3.50
3.37

4.01
3.92
3.74

341.5
683.0
1366.0

Pressure, bar
3
7
15

90
80
70

H2

60

Product gas composition, %

Simulated operating conditions

100

15

CO2

10
5
0
20

CH4

15
10
5
0

concentration in the product gas during that period is observed as high as 3:1 mol% on a dry basis. This is because
the equilibrium pressure of CO2 in the reversible reaction
(4) of CaO-carbonation/CaCO3 -calcination increases with
increasing temperature, as given by Eq. (16).
Higher reaction temperature favors thermodynamically
the formation of CO by reactions (1) and (3) as well as the
conversion of CH4 , as resulted in Fig. 4. Resultantly, purity
of H2 in the product gas during the prebreakthrough period
decreases with higher concentration of CO as the temperature increases. However, the total amount of H2 produced
additionally per kg of the bed-packed CaO pellets is increased with increasing temperature, as summarized in Table
2. It can be found in Table 2 that corresponding amounts of
CO2 removed per kg CaO in the reactor by the carbonation
increase with temperature. Therefore, if the process has to
be operated for the direct production of high purity H2 while
keeping CO as low as possible, low temperature operation
is preferable at the sacri8ce of the decrease in the amount of
CO2 removed. Operation at higher temperature is preferred
for maximal carbonationsequestration of CO2 and for
the maximal production of H2 regardless of the composition of CO.
4.3. E;ect of pressure
Fig. 5 shows the simulated pro8les of the product gas
composition on a dry basis with the change in pressures between 3 and 15 bar for the in situ CaO carbonation-enhanced
MSR reaction (700 C, CH4 feed rate = 11:2 NL=h, feed
H2 O=CH4 =3, GHSV=341:5 h1 ). The pressure drop along
the total reactor bed in the simulation is small enough to be
neglected. Pressure increase in the MSR reaction exerts to
shift the equilibrium toward suppressing CH4 conversion, as

12

CO

9
6
3
0
0

40

80

120

160

200

Time on stream, min


Fig. 5. Simulated pro8les of the product gas composition on a dry basis
with the change in pressure (temperature=700 C, feed CH4 =11:2 NL=h,
feed H2 O=CH4 = 3, GHSV = 341:5 h1 ).

suggested by reaction (1). The MSR at higher pressure after


the breakthrough of CaO generally gives a product gas containing less CO and more CO2 , as shown in Fig. 5, because
reaction (1) forming CO is suppressed by pressure increase.
Pro8les of CO2 composition in the product gases indicate
that the in situ CaO carbonation-enhanced MSR reaction is
a<ected by the change of pressure. However, no direct e<ect
of the total pressure on the CaO carbonation is considered
because the carbonation kinetics has been modeled assuming temperature dependence only. Instantaneous axial pro8les at 10 min of reaction of CH4 conversion, temperature,
and CaO conversion are displayed in Fig. 6. At that instant,
the bed fronts where the CaO carbonation is taking place are
located in the middle section of the reactor with only a small
e<ect of pressure. The most important thing noteworthy is
the degree of bed temperature rise, as shown in Fig. 6(B).
The maximum temperature rises at that instant are 16, 12
and 8 C at the pressures 3, 7 and 15 bar, respectively. Operation at higher pressure leads to a lower rate of CO2 production by suppressing the conversion of CH4 , and resultantly,

CH4 conversion, %

D.K. Lee et al. / Chemical Engineering Science 59 (2004) 931 942


100

100

80

90

60

20

H2

70

(A)

60

CO2

15

Product gas composition, %

Temperature, oC

H 2O/CH 4 ratio
3
5
7

80
Pressure, bar
3
7
15

40

0
720
700
680
660
640
620

(B)
600

CaO conversion, %

939

15

10
5
0
15

CH4
10
5

10

0
5

12

(C)
0

CO

0
0.05

0.1

0.15

0.2

0.25

0.3

Axial distance,m
Fig. 6. Axial instantaneous pro8les of CH4 conversions (A), temperatures
(B), and CaO conversions (C) at 10 min of the reaction time on stream
in the simulated run shown in Fig. 5.

6
3
0
0

40

80

120

160

200

Time on stream, min

the rate of CaO carbonation is slow. Accordingly, the bed


temperature rise is small. As mentioned earlier, higher bed
temperature rise gives a product stream with higher equilibrium concentration of CO2 while CaO is still available for
further carbonation in the bed, as shown in Figs. 5 and 6.
Operation at higher pressure also contributes to maintaining the reactor temperature close to the set value, as shown
in Fig. 6(B). The degree of reactor bed cooling appears
smaller during operation at higher pressure by suppressing
the endothermic CH4 conversion. Corresponding to this, the
observed degree of CaO carbonation conversion is higher,
as shown in Fig. 6(C). The total amount of H2 produced
additionally per kg of the bed-packed CaO pellets is somewhat increased with pressure, as listed in Table 2. The corresponding amounts of CO2 removed in the reactor by the
CaO carbonation are increased as well.
4.4. E;ects of the feed H2 O=CH4 and the feed space
velocity
Fig. 7 shows the simulated composition pro8les of the
product gas on a dry basis as function of the molar ratio of
feed H2 O=CH4 for the in situ CaO carbonation-enhanced
MSR reaction (700 C, 7 bar). In this simulation, the feed

Fig. 7. Simulated pro8les of the product gas composition on a dry basis


with the change in the molar ratio of feed H2 O=CH4 (temperature=700 C,
pressure = 7 bar, feed CH4 = 11:2, 7.47, 5:6 NL=h for the feed
H2 O=CH4 = 3, 5, 7, respectively, GHSV = 341:5 h1 ).

Mow rates of CH4 are varied to 11.2, 7.47, and 5:6 NL=h corresponding to a change of the feed H2 O=CH4 ratio to 3, 5,
and 7, respectively, to get the total feed rate be the same as
44:8 NL=h, equivalent to 341:5 h1 in GHSV. After the CaO
breakthrough, the MSR at higher feed ratio of H2 O=CH4
gives a product stream containing less CO and more CO2 ,
with higher conversion of CH4 , as shown in Fig. 7. Due to
the decreased feed rate of CH4 , the prebreakthrough period
is extended with increasing feed ratio of H2 O=CH4 . At the
feed ratios of 5 H2 O=CH4 and above, it can be noticed that
the production of H2 with a purity of about 98% is possible
during the prebreakthrough period. The total amounts of H2
produced additionally per kg of the bed-packed CaO pellets
are not largely di<erent by the feed ratios of H2 O=CH4 ,
being given by about 4:0 mol-H2 /kg-CaO, as shown in
Table 2.
Fig. 8 shows the simulated composition pro8les of the
product gas on a dry basis at di<erent space velocities

940

D.K. Lee et al. / Chemical Engineering Science 59 (2004) 931 942

25

100
GHSV, hr-1
341.5
683
1366

90

20

CaO conversion, %

80

H2
70

CO2

Product gas composition, %

15
10
5

15

10

GHSV, hr -1
341.5
683
1366

CH4

0
0
2

0.05

0.1

0.15

0.2

0.25

0.3

Axial distance, m
Fig. 9. Axial bed pro8les of the 8nal carbonation-conversions of the CaO
at the end of the simulated run shown in Fig. 8.

CO

4.5. Operating options of the in situ CaO


carbonation-enhanced MSR

4
2
0
0

40

80

120

160

200

Time on stream, min


Fig. 8. Simulated pro8les of the product gas composition on a dry basis
with the change in the GHSV (temperature=700 C, pressure=7 bar, feed
H2 O=CH4 = 7, feed CH4 = 5:6, 11.2, 22:4 NL=h for the GHSV = 341:5,
683, 1366 h1 , respectively).

(341.5, 683, and 1366 h1 in GHSV), 700 C, 7 bar, feed


H2 O=CH4 = 7. As can be expected, the increase of the
GHSV results in a decrease of the prebreakthrough period.
The temperature distribution of the reactor bed along the axial distance is a<ected by the GHSV because higher GHSV
makes the cool region of the bed wider in the axial direction, where endothermic MSR takes place. As a result, the
carbonation-conversion pro8les are di<erent due to the different temperature distribution of the bed. Fig. 9 shows the
axial pro8les of the 8nal carbonation-conversions of the
CaO at the end of the simulated run shown in Fig. 8. It can
be seen that lower conversion of CaO is attained locally
along the reactor bed with higher GHSV. In this simulated
run, the amounts of CO2 removed and those of H2 produced
additionally by the carbonation of CaO throughout the bed
are decreased with increasing GHSV, as shown in Table
2. From the standpoint of carbonation-utilization of CaO
packed in the bed, it is better to operate the process at lower
GHSV.

As a summary for some 8ndings above about the transient in situ CaO carbonation-enhanced MSR reaction, it
should be 8rst mentioned that the reaction at lower temperatures than 650 C fails to give a signi8cant conversion of
the CaO pellets prepared in this study. As shown in Table
2, the amount of CO2 uptake and the productivity of H2
per kg-CaO by CaO carbonation are increased with increasing temperature up to 750 C. However, operation at higher
temperatures su<ers not only from higher outlet CO2 concentration due to the increased partial pressure in the equilibrium of CaO-carbonation/CaCO3 -calcination, but also from
the production of CO in higher concentration. If the process is operated in pursuit of as large CO2 sequestration as
possible and large production of H2 regardless of CO concentration, higher temperature up to a maximum 750 C will
be better. The amount of CO2 uptake by CaO carbonation,
as shown in Table 2, appears much larger than those by the
hydrotalcite-based chemisorbents with about 0:45 mol=kg
(Hufton et al., 1999). For a process option of the direct production of H2 with as low carbon oxides concentrations as
possible, operation at 650 C is the best choice with the disadvantage of very limited carbonation conversion at lower
temperature than that.
As shown in Table 2 and discussed above, operation at
higher pressure is bene8cial for increasing the conversion
of CaO in the reactor-bed at a given condition, and also for
lowering the concentration of CO in the product. Raising
the feed ratio of H2 O=CH4 is helpful for increasing the
conversion of CH4 as well as lowering the product CO

D.K. Lee et al. / Chemical Engineering Science 59 (2004) 931 942

941

Table 3
Analysis for the reaction performance and the eKuent product gases during the prebreakthrough period in the simulated run with the changes in pressure
Pressurea (bar)

15

30

50

80

CH4 conversion (%)


H2 production rate (mol/h)
H2 purityb (mol%)
CO compositionb (ppm)
CO2 compositionb (ppm)
CH4 compositionb (mol%)

98.7
1.10
99.3
1230
3020
0.32

97.4
1.08
99.1
571
1430
0.66

95.3
1.05
98.7
282
727
1.22

93.0
1.01
98.1
170
456
1.87

90.0
0.94
97.1
112
315
2.84

a The

other operating conditions: temperature = 650 C, feed CH4 = 5:6 NL=h, and feed H2 O=CH4 = 7 (GHSV = 341:5 h1 ).
and compositions on a dry basis.

b Purity

concentration. The degree of overall utilization of CaO pellets in the reactor-bed could be increased by lowering the
GHSV. Increased pressure, high ratio of H2 O=CH4 , and decreased GHSV are favorable for both of the two process
options at the temperatures investigated here.
For the case of a process option to produce a pure H2
stream, there is a need to investigate the e<ect of pressure on
the composition of CO in the product stream at the operating conditions as follows: temperature=650 C, feeding rate
of CH4 = 5:6 NL=h, and feed H2 O=CH4 = 7 (GHSV of total
feed gas = 341:5 h1 ). Table 3 provides the analysis of the
gas products obtained during the prebreakthrough period in
the simulated run at such conditions with changing pressure
up to 80 bar. The prebreakthrough periods are maintained
for 30, 33, 35, 37, and 40 min at the pressures of 7, 15, 30,
50, and 80 bar, respectively. At these simulation conditions,
the data show that a very high pressure is required for reducing the product concentration of CO down to the level of a
few hundreds ppm. Hufton et al. (1999) demonstrated that
it was possible to produce 95 mol% H2 stream containing
5 mol% CH4 and less than 50 ppm carbon oxide impurities
in their SERP using a CO2 chemisorbent at the operation
conditions: temperature=450 C, pressure=4:8 bar, and the
feed H2 O=CH4 = 6. A key to the production of a H2 stream
with such a low carbon oxide concentration in their experiment is that the chemisorbent was working at reaction temperatures as low as 450 C. At 650 C of the simulated reaction temperature, the pressure required for the product with
about 50 ppm of carbon oxides is expected above 100 bar.
Consequently, it is thought that targeting the direct production of pure H2 with carbon oxides less than 50 ppm is not
appropriate in the in situ CaO carbonation-enhanced MSR
reaction as long as any CaO-based chemisorbent with a considerable working capacity at temperatures much lower than
650 C has not been developed.
5. Conclusions
A dynamic model derived for the CaO carbonation-enhanced MSR reaction at non-isothermal, non-adiabatic,
and non-isobaric operating condition was shown to accurately depict the transient behavior of the in situ CaO

carbonation-enhanced MSR reaction. The reaction at lower


temperatures than 650 C failed to give a practical conversion of the CaO pellets used in this study. The amount of
CO2 uptake and the productivity of H2 per kg-CaO by the
CaO carbonation were increased with increasing temperature up to 750 C. High CO2 uptake by the CaO carbonation
would be very advantageous in the process application for
the maximal sequestration of CO2 and production of H2
compared to the hydrotalcite-based chemisorbent. Operation at higher temperatures, however, resulted not only
in the release of CO2 in higher concentration but also the
production of CO in higher concentration during the prebreakthrough period. Operations at high pressure, high ratio
of feed H2 O=CH4 , and decreased feed rate at a given temperature were shown favorable for increasing the degree of
overall carbonation-utilization of CaO pellets in the reactor
bed, and for lowering the product concentration of CO.
In particular, the operation at higher pressure was helpful
for maintaining the reactor temperature close to the initial
set value by adequately suppressing the endothermic MSR
reactions and the exothermic CaO carbonation reaction,
by which a front cool zone and a rear hot zone along the
reactor bed were brought about, respectively. For the application of the in situ CaO carbonation-enhanced MSR to the
direct production of pure H2 with carbon oxides less than
50 ppm, a CaO-based chemisorbent with a considerable
working capacity at temperatures much lower than 650 C
should be developed.
Notation
b
Ceq; CO2
Ci
Ci; f
Ci; init
Cpg
Cps

time taken for half the ultimate conversion


of CaO, s
equilibrium concentration of CO2 in reaction (4), kmol=m3
concentration of component i, kmol=m3
concentration of component i in the feed,
kmol=m3
initial concentration of component i,
kmol=m3
heat capacity of gas, kJ/kmol K
heat capacity of solid, kJ/kmol K

942

dp
DR
hW
Hcbn
HRj
kc
kg
kj
ks
kz0
Ki
Kj
MCaO
P
Pf
Pi
rcbn
ri
Rj
Rep
t
T
Tf
Tinit
TW
u
WCaO
Wcat
X
Xu
z

D.K. Lee et al. / Chemical Engineering Science 59 (2004) 931 942

particle diameter, m
inside diameter of reactor, m
heat transfer coeScient through reactor
wall, kJ=m2 s K
reaction heat of CaO carbonation, kJ/kmol
reaction heat of reaction j (j = 13),
kJ/kmol
apparent rate constant of CaO carbonation, s1
thermal conductivity of gas, kJ/m s K
rate constant of reactions j, j = 1; 2:
kmol bar 0:5 =kg-cat s; j = 3:
kmol/kg-cat s bar
thermal conductivity of solid, kJ/m s K
static e<ective thermal conductivity,
kJ/m s K
adsorption constant for component i,
i = CO, H2 , CH4 : bar 1 ; i = H2 O: dimensionless
equilibrium constant of reactions j,
j = 1, 2: bar 2 ; j = 3: dimensionless
molecular weight of CaO, kg/kmol
pressure, bar
pressure at reactor entrance, bar
initial reactor bed pressure, bar
rate of CO2 removal by CaO carbonation,
kmol/kg-CaO s
rate of formation of component i,
kmol/kg-cat s
rate of reaction j (j = 13), kmol/kg-cat s
particle Reynolds number, dimensionless
time, s
temperature, K
feed gas temperature, K
initial temperature of the reactor bed, K
reactor wall temperature, K
super8cial velocity of gas, m/s
weight of the CaO pellets packed in the
reactor, kg
weight of catalyst pellets packed in the
reactor, kg
fractional carbonation conversion of CaO,
dimensionless
ultimate fractional carbonation conversion
of CaO, dimensionless
axial distance of reactor from the entrance,
dimensionless

Greek letters



%
CaO
cat

parameter de8ned in Eq. (14), s1


bed void fraction, dimensionless
catalyst e<ectiveness factor, dimensionless
viscosity of gas, Pa s
apparent density of CaO pellets, kg=m3
apparent density of catalyst pellets, kg=m3

s
g
CO2

average apparent density of the two mixed


solids in reactor, kg=m3
density of gas, kg=m3
parameter de8ned in Eq. (15), kmol=m3 s

Acknowledgements
This work was carried out through a Strategic National
R&D Program with 8nancial support from the Ministry of
Science and Technology (MOST), Republic of Korea.
References
Abanades, J.C., 2002. The maximum capture eSciency of CO2 using a
carbonation/calcinations cycle of CaO=CaCO3 . Chemical Engineering
Journal 90, 303306.
Balasubramanian, B., Ortiz, A.L., Kaytakouglu, S., Harrison, D.P., 1999.
Hydrogen from methane in a single-step process. Chemical Engineering
Science 54, 35433552.
Barin, I., 1989. Thermochemical data of pure substance. Weinheim, VCH.
Bhatia, S.K., Perlmutter, D.D., 1983. E<ect of the product layer on the
kinetics of the CO2 -lime reaction. A.I.Ch.E. Journal 29 (1), 7986.
Brun-Tsekhovoi, A.R., Zadorin, A.N., Katsobashvili, Y.R., Kourdyumov,
S.S., 1986. The process of catalytic steam-reforming of hydrocarbons
in the presence of carbon dioxide acceptor. In: Proceedings of the
World Hydrogen Energy Conference, Vol. 2, Pergamon Press, New
York, pp. 885900.
Carvill, B.T., Hufton, J.R., Anand, M., Sircar, S., 1996.
Sorption-enhanced reaction process. A.I.Ch.E. Journal 42 (10),
27652772.
Dedman, A.J., Owen, A.J., 1962. Calcium cyanamide synthesis Part 4.
The reaction CaO+CO2 =CaCO3 . Transactions of the Faraday Society
58, 20272035.
De Wash, A.P., Froment, G.F., 1972. Heat transfer in packed bed.
Chemical Engineering Science 27, 567576.
Ding, Y., Alpay, E., 2000. Adsorption-enhanced steam-methane
reforming. Chemical Engineering Science 55, 39293940.
Ergun, S., 1952. Fluid Mow through packed columns. Chemical
Engineering Progress 48, 8994.
Feyo De Azevedo, S., Romero-Ogawa, M.A., Wardle, A.P., 1990.
Modeling of tubular 8xed bed catalytic reactors: a brief review.
Transactions of the Institution of Chemical Engineers, Part A 68,
483502.
Han, C., Harrison, D.P., 1994. Simultaneous shift reaction and carbon
dioxide separation for the direct production of hydrogen. Chemical
Engineering Science 49, 58755883.
Hufton, J.R., Mayorga, S., Sircar, S., 1999. Sorption-enhanced reaction
process for hydrogen production. A.I.Ch.E. Journal 45, 248256.
Lee, D.K., 2004. An apparent kinetic model for the carbonation of calcium
oxide by carbon dioxide. Chemical Engineering Journal, in press.
Li, C., Finlayson, B.A., 1977. Heat transfer in packed bedsA
reevaluation. Chemical Engineering Science 32, 10551066.
Ortiz, A.L., Harrison, D.P., 2001. Hydrogen production using
sorption-enhanced reaction. Industrial and Engineering Chemistry
Research 40, 51025109.
Waldron, W.E., Hufton, J.R., Sircar, S., 2001. Production of hydrogen by
cyclic sorption enhanced reaction process. A.I.Ch.E. Journal 47 (6),
14771479.
Xiu, G., Li, P., Rodrigues, A.E., 2002. Sorption-enhanced reaction
process with reactive regeneration. Chemical Engineering Science 57,
38933908.
Xu, J., Froment, G.F., 1989. Methane steam reforming, methanation and
watergas shift: I. Intrinsic kinetics. A.I.Ch.E. Journal 35, 8896.

Вам также может понравиться