Sei sulla pagina 1di 12

Steam Methane Reforming in a Ni/Al

2
O
3
Catalyst: Kinetics and Diffusional Limitations
in Extrudates
Eduardo L. G. Oliveira, Carlos A. Grande and Alrio E. Rodrigues*
LSRE, Laboratory of Separation and Reaction Engineering, Associate Laboratory LSRE/LCM, Department of Chemical
Engineering, Faculty of Engineering, University of Porto, Rua Dr. Roberto Frias, s/n, 4200-465 Porto, Portugal
The true kinetics of methane steam reforming was measured in powder of Ni/Al
2
O
3
catalyst (Octolyst 1001 from Degussa) at different
temperatures (733890K) for several operating conditions. New reaction rate constants were determined for this catalyst. The observed reaction
rate was measured on catalyst extrudates to determine diffusion effects within the porous structure of the particle. A non-isothermal model with
diffusion was used to determine effectiveness factors for each reaction. The objective was to measure all necessary data to model the performance
of the catalyst in a Sorption Enhanced Reaction Process (SERP) for H
2
production with in situ CO
2
capture.
On a mesur e la cin etique r eelle du reformage ` a la vapeur de m ethane sur un catalyseur Ni/Al
2
O
3
sous forme de poudre ( Octolyst 1001 de
Degussa) ` a diff erentes temp eratures (733890K) dans plusieurs conditions dexploitation. De nouvelles constantes de vitesse de r eaction ont et e
d etermin ees pour ce catalyseur. On a mesur e la vitesse de r eaction observ ee sur des extrudats de catalyseur pour d eterminer les effets de diffusion
` a lint erieur de la structure poreuse de la particule. On a utilis e un mod` ele non isothermique avec diffusion pour d eterminer les facteurs defcacit e
de chaque r eaction. Lobjectif consistait ` a mesurer toutes les donn ees n ecessaires pour mod eliser la performance du catalyseur dans un proc ed e
de r eaction am elior ee de sorption pour la production de H
2
avec captage in-situ de CO
2
.
Keywords: steam methane reforming, kinetics, catalysis, hydrogen, effectiveness factor
INTRODUCTION
S
ignicant reductions of emissions of carbon dioxide in
production of energy and fuels are essential to ensure
sustainable development. Unfortunately, nowadays the mix
of all renewable sources of energy cannot ensure secure energy
supply by itself and use of fossil fuels is required. Carbon dioxide
capture and storage (CCS) on permanent and secure geological
formations appears as a plausible short and middle term solution
to this problem (Metz et al., 2005). Production of hydrogen as
energy carrier for vehicular fuel or energy production in post-
combustion schemes is one of the targeted processes that should
benet from CCS.
One of the new technologies of hydrogen production, that is,
receiving large efforts in research and development is the Sorption
Enhanced Reaction Process (SERP). In this process, reforming of
a fossil fuel with water takes place. The process is composed by a
bunch of parallel reactors containing a catalyst and a CO
2
-selective
sorbent mixed. The objective is to capture CO
2
from the reaction
media and displace the reforming equilibrium according to Le
Chateliers principle to produce more and puried hydrogen. The
reactors are operated under a pre-dened scheduling alternating
between sorption and sorbent regeneration (Hufton et al., 1999;
Ding and Alpay, 2000; Mayorga et al., 2001; Xiu et al., 2002, 2003;
Ochoa-Fernandez et al., 2005; Satrio et al., 2005; Hildenbrand
et al., 2006; Cobden et al., 2008; Lee et al., 2008; van Dijk
et al., 2008). Steam methane reforming (SMR) was the most
studied case, although other fuel have been suggested (Comas
et al., 2004; Essaki et al., 2008; Lysikov et al., 2008).

Author to whom correspondence may be addressed.


E-mail address: arodrig@fe.up.pt
Can. J. Chem. Eng. 87:945956, 2009

2009 Canadian Society for Chemical Engineering


DOI 10.1002/cjce.20223
|
VOLUME 87, DECEMBER2009
| |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| 945 |
The design of SMR-SERP units involves the detailed knowledge
of the CO
2
loading and diffusion kinetics of the sorbent and also
the true reaction kinetics and the diffusion limitations that may
exist in particulate catalysts. We have already determined the CO
2
sorption on alkali-modied hydrotalcites (Oliveira et al., 2008).
Using hydrotalcites we are limited to operating temperatures lower
than 823 K.
In this work we report the performance of a commercial
Ni/Al
2
O
3
catalyst (Octolyst 1001 from Degussa, Essen,
Germany) for the production of hydrogen by steam methane
reforming. Steam methane reforming experiments were per-
formed in the temperature range 733890 K, with a steam to
carbon (S/C) ratio of 4, a CH
4
/H
2
ratio of 1.25 and various inlet
owrates. Two different sets of experiments were performed. The
rst one was carried out using ne powder of the catalysts to
determine the true reaction kinetics. We have employed the model
developed by Xu and Froment (1989a) and tted new kinetic
constants for this catalyst. The second set of experiments was
carried out in the catalyst extrudates to determine the inuence
of mass transfer within the particles. The data reported in this
work together with the data of any other CO
2
-selective sorbent
allows the direct modeling of a SMR-SERP unit for hydrogen
production.
STEAM METHANE REFORMING REACTIONS
Steam reforming is normally the term employed to characterise
the oxidation of any carbon source with water vapour. The
steam reforming process dates back to the beginning of the
20th century. Products of this reaction are H
2
and carbon
oxides from complete and incomplete oxidation: CO and CO
2
,
respectively. The concentrations of carbon oxides are balanced
by the watergas shift reaction (Rostrup-Nielsen, 1984). The
reforming reactions take place at strong conditions: high pressure,
high temperature, and presence of large amounts of steam. The
operating conditions employed require large inputs of energy and
resistant materials and catalysts. A catalyst should resist coking
(Froment, 2001) and decomposition by steam, be inactive for
side-reactions, maintain the activity at high temperatures and
must have high mechanical strength as well as good heat transfer
properties (Rostrup-Nielsen, 1984; Twigg, 1989). Steam-reforming
catalysts are usually made of nickel supported in alumina.
Other active metals such as cobalt, platinum, rhodium present
activities higher than nickel but Ni is more economically viable
(Twigg, 1989).
Many studies have focused on different supports such as ZrO
2
and Ce-ZrO
2
(Dong et al., 2002; Roh et al., 2002) to increase
thermal stability, activity, and resistance to steam (Takahashi
et al., 2004). Table 1 shows a brief summary of different catalysts
and supports studied for steam methane reforming. In the work
from Roh et al. (2002) it is possible to compare the performance of
some Ni catalysts with different supports. The combination of Ce
and ZrO
2
gave the highest conversion and CO
2
selectivity while
maintaining high thermal stability. The catalyst with MgAl
2
O
4
support presented higher methane conversion than the one with
Al
2
O
3
support but the conversion dropped signicantly when the
gas hourly space velocity was increased to 0.288 m
3
/g
cat
h.
Several authors have proposed different kinetic expressions
for the reforming reactions. In 1989, Xu and Froment (1989a,b)
proposed a kinetic model accounting for diffusional limitations,
using a nickel catalyst supported on MgAl
2
O
4
. They have also
observed an initial decrease in the activity of the catalyst attributed
to the sintering of the Ni particles. Additionally, they have carried
out an analysis of the reactions in the presence of CH
4
, H
2
O, H
2
,
Table 1. Metals and supports used in steam-reforming catalysts for different hydrocarbons
Catalyst m (g) T (K) P (kPa) S/C X
CH
4
(%) S
CO
2
Refs.
Ni/MgAl
2
O
4
(15% Ni) 0.4 823 500 5
a
17 Xu and Froment (1989a)
Ni/CeZrO
2
(30% Ni) 0.05 1023 101 3 60.9 3.5 Dong et al. (2002)
Ni/CeZrO
2
(15% Ni) 0.05 1023 101 3 97 48.7 Dong et al. (2002)
Ni/ZrO
2
(15% Ni) 0.05 1023 101 3 75 6.3 Dong et al. (2002)
Ni/CeO
2
(15% Ni) 0.05 1023 101 3 54 4.9 Dong et al. (2002)
Ni/MgAl
2
O
4
(15% Ni) 0.05 1023 101 3 79 7.7 Dong et al. (2002)
Ni/Al
2
O
3
(15% Ni) 0.05 1023 101 3 57 4.7 Dong et al. (2002)
Ni/SiO
2
(9% Ni) 0.2 873 101 2 75 Takahashi et al. (2004)
Ni/Al
2
O (1517% Ni) 0.3 823 120 4
b
12.5 0.87 Hou and Hughes (2001)
Ni/ZrO
2
(20% Ni) 0.3 773 101 2 25.5 6.6 Matsumura and Nakamori (2004)
Ni/CeZrO
2
/-Al
2
O
3
(12% Ni) 2 823 101 3 60 3.17 Liu et al. (2002)
Ni/CeZrO
2
/-Al
2
O
3
(12% Ni) 973 101 3 97 0.78 Oh et al. (2003)
Pd/Ce (1% Pd) 0.05 773 101 2 13.1 Wang and Gorte (2002)
a
H
2
/CH
4
=1.25; S/C=steam to carbon ration; S
CO2
= F
CO2
,F
CO
.
b
H
2
/CH
4
=1.
| 946 |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| |
VOLUME 87, DECEMBER 2009
|
CO, CO
2
, and carbon that could take place. It was observed that
the most important ones are:
SMR CH
4
+H
2
O CO +3H
2
(1)
Water gas shift CO +H
2
O CO
2
+H
2
(2)
Global SMR CH
4
+2H
2
O CO
2
+4H
2
(3)
The kinetic expressions obtained using the Langmuir
Hinshelwood methodology are (Xu and Froment, 1989a):
R
1
=
k
1
P
2.5
H
2
(P
CH
4
P
H
2
O
P
3
H
2
P
CO
,K
1
)
DEN
2
1
(4)
R
2
=
k
2
P
H
2
(P
CO
P
H
2
O
P
H
2
P
CO
2
,K
2
)
DEN
2
1
(5)
R
3
=
k
3
P
3.5
H
2
(P
CH
4
P
2
H
2
O
P
4
H
2
P
CO
2
,K
3
)
DEN
2
1
(6)
DEN
1
= 1 +K
CO
P
CO
+K
H
2
P
H
2
+K
CH
4
P
CH
4
+
K
H
2
O
P
H
2
O
P
H
2
(7)
The dependence of rate coefcients (k
1
, k
2
, k
3
) and adsorption
coefcients of the gases (K
CO
K
H
2
, K
CH
4
, K
H
2
O
) with temperature
was described by an Arrhenius type function:
k
i
= k
0i
e
E
i
,RT
i = 1. 2. 3 (8)
K
j
= K
0j
e
LH
i
,RT
j = CO. H
2
. CH
4
. H
2
O (9)
The activation energies, the enthalpies of adsorption and the
pre-exponential factors used in this work are shown in Table 2.
The values of the constants K
0j
and LH
j
used in the determination
of the adsorption equilibrium of carbon monoxide, hydrogen,
methane, and water were taken from the work of Xu and
Froment (1989a); the reagents and products are adsorbed in nickel
crystallites that are present in both catalysts. The equilibrium
constants for reactions 13 were calculated using the standard
Gibbs energy of each reaction at the corresponding temperature.
Table 2. Experimental conditions used for the steam methane
reforming experiments in the catalyst powder
Experiment Powder
m
cat
(mg) 62.12
Pressure (kPa) 200
L
col
(mm) 0.5
D
col
(mm) 27
Oven temperature (K) 733, 763, 793, 890
S/C 4.1
F
0H
2
1.25 F
CH
4
F
0
(mmol/min) 12.9, 17.9, 25.4, 30.4, 35.4, 40.4
Figure 1. SEM of the Ni/Al
2
O
3
catalyst (Degussa) at 20000
magnication.
CATALYST CHARACTERISATION
We have used a commercial catalyst with nickel as the active metal
supported in alumina (Ni/Al
2
O
3
). The commercial name of the
catalyst is Octolyst 1001 (Degussa). The catalyst was shaped as
extrudates with a constant diameter of 1.6 mm and a length of
from 3 to 5 mm. The composition of the catalyst was determined
by inductively coupled plasma in an ICP-AES model 70 Plus (Jobin
Yvon, Unterhaching, Germany). The Ni content of the catalyst was
determined to be 15.4%.
The catalyst extrudates were analysed by Scanning Electron
Microscopy (SEM) in a JEOL JSM-6301F (Jeol, Japan). Figure 1
shows a typical image of the catalyst. Several images were taken
and Ni clusters with sizes ranging from 170 to 3720 nm were
observed.
The density and pore size distribution of the sample were
measured in a Quantachrome PoreMaster 60 (Quantachrome,
Hook, UK). The catalyst presented an average pore diameter of
8.5 nm, particle density of 1274 kg/m
3
and a solid density of
3630 kg/m
3
. A porosity of 64.9% was calculated based on the
values of the particle and solid densities.
It should be pointed out that this catalyst is quite different
from the commercial large holes catalysts used in SMR processes
actually. The commercial catalyst is much larger (diameter of
16 mm), normally ring-shaped with holes of 3.5 mm diameter
(Rostrup-Nielsen, 1984). The extrudates of this catalyst are much
smaller and comparable with the size of sorbent extrudates
resulting in a better distribution in a SMR-SERP unit for hydrogen
production.
EXPERIMENTAL UNIT FOR SMR
A scheme of the experimental unit employed in steam methane
reforming experiments is shown in Figure 2. A high performance
liquid chromatography (HPLC) pump fromMerck-Hitachi (Tokyo,
Japan) was used to control the liquid water owrate while the
|
VOLUME 87, DECEMBER2009
| |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| 947 |
Figure 2. Experimental apparatus used in the steam-reforming
experiments (powder and extrudates) and the measured oven
temperature prole.
gas owrates of H
2
and CH
4
were controlled with two different
mass ow controllers. The liquid water was vaporised and mixed
with the CH
4
/H
2
stream before entering the reactor. The efuent
gas leaving the reaction column was cooled in a condenser
where liquid water was removed; a VeriowBackpressure (Parker,
Warwick, UK) was used to control the pressure inside the
reactor. After the backpressure, the efuent gas owrate was
measured. The exiting stream nally passes through a six-port
valve system where it can be routed to a multi-port valve to be
analysed in a gas chromatograph, or go to the venting system.
The analysis of the dry efuent stream was done using a Gas
Chromatograph (GC) (Dani, Cologno Monzese, Italy) equipped
with a Carboxen-1010 PLOT (Supelco, St. Louis, MO) capillary
column.
To evaluate the kinetic constants of the proposed mechanism,
it is necessary to avoid all resistances to diffusion within the
pores of the catalyst. For this reason, the Ni/Al
2
O
3
catalyst was
grinded to a ne powder (particle size lower than 0.15 mm). The
SMR experiments were done by inserting the catalyst powder
approximately in the middle of the column, between two thick
layers of quartz wool. The catalyst was distributed in all the
diameter of the column (27 mm) such that the length of the
catalyst bed is 0.5 mm. Within the catalyst bed we have inserted
a thermocouple to register the variations of temperature. Prior
to the steam reforming experiments, the sample was heated to
763 K at 1 K/min under helium ow. The catalyst was activated at
763 K for 1 h in a stream containing 5% H
2
balanced by helium.
After the activation, a long time experiment was performed
at 200 kPa and 763 K to detect deactivation of the catalyst. In
this rst experiment, the owrates were 4.03 mmol/min of CH
4
,
16.3 mmol/min of steamand 5.01 mmol/min of H
2
. Hydrogen was
added to the inlet stream to avoid the formation of carbon deposits
in the catalyst. After no catalyst deactivation was detected,
several experiments were performed at different temperatures
and feed ows. The experimental conditions are shown in
Table 2.
A second set of SMR experiments was made in the catalyst
extrudates to determine the inuence of diffusion within the
pore structure of the particle. For these experiments, the same
experimental set-up was employed, but using a much higher
mass (33.233 g) of Ni/Al
2
O
3
extrudates. The reactor column has
a length of 165 mm and a diameter of 26.6 mm. The length of the
catalyst bed was 81.8 mm, with an initial layer of quartz particles
(47 mm) to accommodate ow variations. To ll the remainder
of the column (36.2 mm), a second layer of quartz particles was
used. Prior to the steam reforming experiments, the catalyst was
Table 3. Operating conditions and dry efuent owrates of the steam
methane reforming experiments performed in Ni/Al
2
O
3
catalyst
extrudates
Run
1 2 3 4 5
Inlet operating conditions
F
0H
2
(mmol/min) 4.79 5.64 5.64 8.40 8.40
F
0CH
4
(mmol/min) 3.81 4.51 4.51 8.40 8.40
F
0H
2
O
(mmol/min) 16.16 19.14 19.14 35.82 35.82
S/C 4.24 4.25 4.25 4.26 4.26
H
2
/CH
4
1.26 1.25 1.25 1.00 1.00
Temperature (K) 806 806 747 762 813
Pressure (kPa) 200 200 200 200 200
Outlet owrates and performance variables
F
CH
4
(mmol/min) 2.63 3.06 3.88 6.48 4.93
F
H
2
F
0H
2
(mmol/min) 4.76 6.05 3.57 7.95 12.54
F
CO
2
(mmol/min) 1.04 1.29 0.58 1.78 3.03
F
CO
(mmol/min) 0.14 0.18 0.06 0.16 0.45
X
CH
4
(%) 30.9 32.3 14.0 23.0 41.4
S
CO
2
7.43 7.17 9.67 11.12 6.73
Y
H
2
(g
1
) 1.25 1.34 0.79 0.94 1.49
Mass of adsorbent: 33.233 g; column diameter: 27 mm; column length:
81.8 mm.
reduced in 5% H
2
for 1 h. The experimental conditions used are
also summarised in Table 3.
All the experiments were performed in a vertical oven with a
single-point controller. We have noticed that the oven has a semi-
parabolic temperature prole, with higher temperatures in the top
of the oven. This prole was measured as a function of the axial
position of the catalyst bed to describe T

(z) in the simulations


performed in this work. The oven temperature prole is shown in
Figure 2.
MATHEMATICAL MODEL FOR POWDER
EXPERIMENTS
We have used a very thin layer of catalyst powder in order
to reduce temperature variations and have a small methane
conversion. However, it was veried that conversion was higher
than 10% in many occasions and treat the reactor as a differential
bed may not be correct. Simulations of the steam reforming
of methane using previously proposed parameters (Xu and
Froment, 1989a) showed that, for the experimental owrates and
temperatures used in this work, the mass of the catalyst had to
be at least 500 times smaller to behave as a differential reactor.
Thus, we have simulated the results obtained using a xed bed
reactor model. The mass balance assumed ideal gas behaviour and
| 948 |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| |
VOLUME 87, DECEMBER 2009
|
that the reactor was isothermal. Axial dispersion was considered
negligible as the contact time was very small. The mass balance
for each component is:
F
i
m
cat
= R
i
i = CO. H
2
. CH
4
. H
2
O (10)
The following boundary condition was considered in the
integration of the differential equations:
F
i
(0) = F
0
i
i = CO. H
2
. CH
4
. H
2
O (11)
The kinetic expressions (Equations 49) were combined to give
the reaction rates of each component:
R
CH
4
= R
1
R
3
(12)
R
CO
2
= R
2
+R
3
(13)
R
H
2
O
= R
1
2R
3
R
2
(14)
R
H
2
= R
2
+3R
1
+4R
3
(15)
R
CO
= R
1
R
2
(16)
The centred nite difference method with second order
approximation was used for the discretisation of the differential
equations. The energies of activation and pre-exponential factors
of the rate coefcients of Equation (8) were predicted by a
sequential quadratic programming algorithm using the software
package gPROMS (Process System Enterprise, London, UK).
MATHEMATICAL MODEL FOR EXPERIMENTS
USING EXTRUDATES
A second model was developed to simulate the experiments
performed using the catalyst extrudates. This model is more
complex than the previous one and assumes that the owis axially
dispersed, that the reactor behaviour is non-adiabatic and non-
isothermal and that the reaction occurs inside the pores of the
extrudates, accounting for diffusional effects. Other assumptions
of the model are: ideal gas behaviour and that the extrudates are
non-isothermal.
With all the assumptions listed before, the mass and energy
balances to the xed bed reactor are:

c
C
i
t
=
(uC
i
)
z
+
c

D
ax.i
C
T
y
i
z

(1
c
)o
c
k
f.i

C
i
C
i.p

r=Rp

(17)

c
C
T
C
vg
T
t
=
c

z
ax
T
z

u
z
C
T
C
pg
T
z
+
c
R
gas
T
C
T
t
(1
c
)o
c
h
f

T T
p

r=Rp

2
h
w
R
column
(T T
w
)
(18)
The description of the behaviour within the catalyst extrudates
is also provided by mass and energy balances given by:

p
C
i.p
t
= D
p

2
C
P.i
r
2
+
1
r
C
P.i
r

+,
cat
(u
1.i
R
1
+u
2.i
R
2
+u
3.i
R
3
)
(19)
(
p
C
T.P
C
vg
+(1
p
),
cat

C
p
cat
)
T
p
t
= z
cat

2
T
p
r
2
+
1
r
T
p
r

+,
cat
(R
1
(LH
1
) +R
2
(LH
2
) +R
3
(LH
3
)) (20)
Additionally, since much heat is exchanged with the
surroundings, an energy balance to the reactor wall was
developed:
,
wall
C
pw
T
w
t
=
2R
column
(u
thik
(2R
column
+u
thik
))
h
w
(T T
w
)

(2R
column
+u
thik
)ln

2R
column
+u
thik
2R
column
U(T
w
T

)
(21)
The Ergun equation was used to describe the pressure drop
inside the reactor:

P
T
z
=

150
,
gas
(1
c
)
2
4R
2
pellet

3
c
u +1.75
,
gas
(1
c
)
2R
pellet

3
c
|u| u

10
5
(22)
To solve these partial differential equations, the following
boundary and initial conditions were used:
z = 0
c
D
ax
C
i
z

z=0
= u
feed

y
feed
P
T
RT
0
C
i.0

(23)
z = 0 z
gas
T
z

z=0
= u
feed
C
T
C
pg
(T
feed
T
0
) (24)
z = 0 u
feed
=
Q
A
col
(25)
z = L
C
C
i
z

z=L
C
= 0.0 (26)
z = L
C
T
z

z=L
C
= 0.0 (27)
z = L
C
P
T.Lc
= P
out
(28)
r = 0.0
C
P.i
r

r=0
= 0.0 (29)
r = 0.0
T
P
r

r=0
= 0.0 (30)
r = R
pellet

P
D
p
C
P.i
r

r=R
pellet
= k
f.i
(C C
P.R
pellet
) (31)
|
VOLUME 87, DECEMBER2009
| |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| 949 |
r = R
pellet
z
cat
T
P
r

r=R
pellet
= h
f
(T T
P.R
pellet
) (32)
According to the evolution of the reaction with the column,
some performance variables can be dened:
X
CH
4.z
=
F
0
CH
4
F
CH
4.z
F
0
CH
4
100 (33)
S
CO
2
=
F
CO
2
F
CO
(34)
Y
H
2
=
F
H
2
F
0H
2
F
CH
4
(35)
where X
CH
4.z
is the methane conversion, S
CO
2
is the carbon dioxide
selectivity and Y
H
2
is the hydrogen yield.
Once the mathematical model is solved, it is possible to
calculate the efciency of each reaction. The efciency is dened
as the ratio of the observed velocity and the real velocity and is
dened by:
j
j
=
2
R
pellet

0
rR
j
(r)dr
R
j

pellet
R
2
pellet
(36)
The centred nite difference method with second order
approximation was used for the discretisation of the differential
equations using 100 elements for the axial coordinate of the
reactor and 20 elements for the radial coordinate of the extrudate.
The software package gPROMS was also used to solve this model.
RESULTS AND DISCUSSION
Steam Methane Reforming Kinetics in Powder
Catalyst
It was previously reported that there is an initial deactivation of
the catalyst probably due to some sintering of the Ni crystallites
(Xu and Froment, 1989a; Hou and Hughes, 2001). To detect this
deactivation and to avoid its interference in further experiments,
we have performed an initial experiment of 31 h. The other
experiments reported in this work were performed after this long
run using the same sample. The experiment was performed at
763 K and 200 kPa of total pressure and with 4.03 mmol/min
of methane, 16.52 mmol/min of water (S/C ratio of 4.1) and
5.04 mmol/min of hydrogen (H
2
/CH
4
ratio of 1.25). We have
detected an initial decrease in conversion (2%) after the rst 10 h,
but no other reductions were observed after 15 h.
An example of the temperature variation inside the catalyst
bed and outside the reactor column during an experiment can be
seen in Figure 3. In all experiments, the H
2
O owrate was started
before methane to allow a steady ow when the reaction starts.
The initial temperature peak that can be observed in Figure 3 at
about 50 s is due to water adsorption on the catalyst. After 100 s of
introducing water and hydrogen, the methane owrate was also
initiated. A decrease of 17 K was observed in the rst 100 s of
reaction and is stable after reaching steady state.
All the operating conditions of the SMR experiments performed
in the catalyst powder are reported in Table 2. Hydrogen was
used in the feed stream to prevent reoxidation of the catalyst
by steam (Xu and Froment, 1989a; Li et al., 2007). Six different
Figure 3. Catalyst and oven temperature versus time at the start of a
SMR reaction in the catalyst powder. This reaction was performed at the
bed temperature of 749K, total owrate of 24.8mmol/min, 200kPa total
pressure and S/C of 4.2. [Colour gure can be viewed in the online issue,
which is available at www.interscience.wiley.com.]
owrates were employed covering temperatures from 733 to
890 K. The conversion of methane and the selectivity towards
carbon dioxide obtained in all the experiments are shown in
Figure 4. The conversion of methane increases with temperature
Figure 4. Conversion of methane (a) and CO
2
selectivity (b) as a
function of methane inlet owrate for the Ni/Al
2
O
3
catalyst in the feed
temperature range (733890) K, 200kPa total pressure, S/C=4 and
H
2
/CH
4
=1.25 along with the simulated results based on the parameters
of Table 4. Solid pointsexperimental, linessimulations. Temperatures
indicated correspond to feed temperatures. [Colour gure can be viewed
in the online issue, which is available at www.interscience.wiley.com.]
| 950 |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| |
VOLUME 87, DECEMBER 2009
|
Figure 5. Dry outlet owrates of: (a) methane; (b) hydrogen; (c) carbon dioxide; (d) carbon monoxide versus methane inlet owrate for the Ni/Al
2
O
3
catalyst in the feed temperature range (733890) K, 200kPa total pressure, S/C=4 and H
2
/CH
4
=1.25 along with the simulated results based on the
parameters of Table 4. Solid pointsexperimental, linessimulations. Temperatures indicated correspond to feed temperatures. [Colour gure can be
viewed in the online issue, which is available at www.interscience.wiley.com.]
due to the endothermic nature of the steam-reforming reaction.
It was also observed that conversion decreases with increasing
inlet owrate due to a reduction of the contact time between
the catalyst and reactants. Another important feature observed
is that the increase of the amount of CO with temperature is
higher than the increase in the amount of CO
2
formed. This effect
translates in a decrease of the selectivity towards CO
2
as can be
observed in Figure 4. This decrease in selectivity is due to the
exothermic nature of the water gas shift reaction (reaction 2)
that converts less CO to CO
2
when temperature increases. A very
positive aspect in terms of SMR-SERP is that the selectivity is
always higher than one, which is important since CO is a poison
for fuel cells and a catalyst selected to produce fuel cell grade
H
2
should be as selective toward CO
2
as possible (Semelsberger
et al., 2004).
Comparing the catalyst tested in this work with the one tested
by Xu and Froment (1989a), our catalyst presents a lower methane
conversion at 890 K while higher values are observed in the lower
temperature range studied. The selectivity towards CO
2
is always
higher in the catalyst tested in this work.
The molar owrates of the exiting gases in the experiments
is provided in Figure 5. For hydrogen, the gure shows the
value of the molar owrate after subtraction of the inlet H
2
owrate in order to show the H
2
that was in fact formed.
The experimental exit molar owrates at each temperature were
used to t the kinetic rate constants using the mathematical
model. The parameters obtained from the tting are shown
in Table 4 and the results are represented as solid lines in
Figures 4 and 5. The parameters give a good t over the range of
experimental conditions, particularly for methane, hydrogen, and
carbon dioxide owrates. The simulations of the carbon monoxide
owrate follow the same trend as the experimental points but
its owrate is always underestimated. This underestimation may
be due to a smaller inuence in the error function used by the
numerical method employed to nd the appropriate tting values;
the absolute value of CO owrate is much smaller than the other
owrates.
We have compared the energies of activation obtained in this
work with previously reported values (Xu and Froment, 1989a)
(240.1 kJ/mol, 67.1 kJ/mol, and 243.9 kJ/mol for reactions 13,
respectively). It can be observed that the energies of activation of
the catalyst used in this work are smaller indicating that higher
conversion can be obtained operating at lower temperatures. This
is very important for SMR-SERP process where hydrotalcites will
be used as sorbent since the operating temperature is limited to
823 K.
Steam Methane Reforming in Extrudates
After the determination of the true SMR kinetics in the catalyst
powder, it is important to measure the inuence that the diffusion
of the species will have in the real behaviour of SMR using catalyst
extrudates. To evaluate the effect of diffusion, ve different
experiments were performed using a reactor column lled with
catalyst extrudates. The experimental conditions of each run are
reported in Table 3.
Since we have introduced a large amount of catalyst in the
column, a strong temperature variation is expected. For this
reason, it will be necessary to measure the temperature within
|
VOLUME 87, DECEMBER2009
| |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| 951 |
Table 4. Kinetic and adsorption parameters for the true kinetic rates of SMR reactions in a Ni/Al
2
O
3
catalyst: comparison of values obtained in this
work and the values given by Xu and Froment (1989a)
This work Xu and Froment (1989a)
Reaction parameters
k
01
(mol kPa
0.5
/(kg
cat
s)) 5.7910
13
1.1610
16
E
I
(kJ/mol) 217.01 240.1
k
02
(mol/(kg
cat
s kPa)) 9.3310
4
5.4110
4
E
II
(kJ/mol) 68.20 67.13
k
0III
(mol kPa
0.5
/(kg
cat
s)) 1.2910
14
2.7910
15
E
III
(kJ/mol) 215.84 243.9
Adsorption parameters
a
K
CO
(Pa
1
) 8.25 H
CO
(kJ/mol) 70.65
K
H
2
(Pa
1
) 6.1510
4
H
H
2
(kJ/mol) 82.90
K
CH
4
(Pa
1
) 66.6 H
CH
4
(kJ/mol) 38.28
K
H
2
O
(Pa
1
) 1.7710
10
H
H
2
O
(kJ/mol) 88.68
a
Adsorption parameters used in this work are the same as the ones reported by Xu and Froment (1989a).
the catalyst bed. The temperature was measured in two different
axial positions: 45 and 87.5 mm from feed inlet.
In these experiments we have also added hydrogen to the
feed stream to prevent formation of coke in the catalyst. A long
experiment was also performed before the experimental runs
reported here in order to avoid the effect of initial deactivation.
This experiment was performed for 7 h at 806 K and
200 kPa total pressure and with 3.81 mmol/min of methane,
16.16 mmol/min of water (S/C ratio of 4.24) and 4.79 mmol/min
of hydrogen (H
2
/CH
4
ratio of 1.26). This experiment was also
used as reference and repeated several times to conrm that there
was no deactivation of the catalyst activity. It can be noted that
steady state is achieved after 20 min of reaction. A total of ve
experiments were performed. A summary of the results obtained
is shown in Table 3. Comparing the performance of runs 1 and 2
(different total owrate), it can be observed that the conversion,
selectivity and H
2
yield are very close. This fact, together with the
very small decrease in the temperature in the last portion of the
column indicates that the equilibrium concentration is reached
within the reactor. The small differences particularly in methane
conversion can be due to small differences in the operating
conditions (S/C and H
2
/CH
4
ratios). A set of experiments were
performed to evaluate the catalyst performance at different
operating conditions like feed temperature and owrates and
different H
2
/CH
4
. The effect of temperature is very marked in all
these runs. However, the reduction of hydrogen in the feed step
is also important in increasing the conversion and the hydrogen
yield, with the consequence of having a smaller selectivity to CO
2
:
more CO is formed from the reforming reactions than what is
consumed by the water gas shift.
The objective of these experiments using extrudates was to
evaluate the effect of diffusion on the observed rate of reaction. A
comparison between the observed rate and the true rate (without
diffusional limitations) is given by the efciency factor. In this
case, since we have considered three different reactions taking
place we should have three different effectiveness factors. In
order to calculate the efciency factors, we need to know the
reaction rates in the different axial positions to introduce them
in Equation (36). The reaction rates as well as the concentration
and molar owrates of the gases within the column were
obtained by simulating the behaviour of the column for each
experiment using the model described by Equations (1732). In
this model, the only parameter to be tted was the pellet tortuosity.
Using a value of t
p
=1.56 we have obtained a good correlation
between all the experimental runs (exit concentration of gases
and temperatures within the column) and the model predictions.
All other parameters employed in the simulations are detailed in
Table 5. Based on these modelling results, we have obtained the
reaction rates, internal temperature prole and molar owrates of
all gases.
As an example of the results obtained, we will discuss the
observed results for run 1 since a similar behaviour was observed
for all other runs. In Figure 6 we can see the simulated molar
owrate of each gas across the column after steady state was
reached. The simulated temperature proles and reaction rates
can also be observed in Figure 6.
It can be observed that according to the simulations, the molar
owrate of hydrogen and carbon oxides is relatively stable after
the initial portion of the reactor also passing through a maximum.
Simulation results in (Figure 6d) also show that the temperature
variation between the surface and the centre of the catalyst
particles is always lower than 1 K indicating that the particles are
isothermal.
Other important observation is the reaction rate of the water
gas shift reaction (R
2
). It can be observed that this reaction
passes through zero two times: the rst one in the initial
portion of the reactor and the second one in the middle. These
variations are related to temperature changes within the reactor.
| 952 |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| |
VOLUME 87, DECEMBER 2009
|
Table 5. Parameters used in the simulations of the non-isothermal
steam methane reforming reactor with Ni/Al
2
O
3
catalyst extrudates

c
(m
3
gas
/m
3
column
) 0.43

cat
(kg
cat
/m
3
cat
) 1274
m
cat
(g) 33.233
R
pellet
(mm) 0.8

p
(m
3
gas
/m
3
pore
) 0.64
R
pore
(nm) 4.25
a
c
2500

p
1.56

gas
(W/mK) 0.56
h
f
(W/(m
2
cat
K)) 109
U (W/(m
2
column
K))
a
200
C
pw
(J/(kg
wall
K)) 500
h
w
(W/(m
2
column
K))
a
400
w
thik
(mm) 9.1

wall
(kg
wall
/m
3
wall
) 7750
Component CO
2
CH
4
H
2
O CO H
2
k
f
(mm/s)
b
86 114 156 107 296
D
k
(mm
2
/s)
b
1.76 2.93 2.76 2.21 8.28
D
m
(mm
2
/s)
b
59.6 82.6 116 76.6 228
D
pore
(mm
2
/s)
b
1.10 1.81 1.72 1.38 5.11
D
ax
(mm
2
/s)
b
44.8 61.3 85.3 56.9 167
a
Fitted parameter using the data from run 1.
b
Calculated based on the feed temperature composition of run 1.
These temperature changes are very important in controlling the
hydrogen formation. The temperature decreases strongly in the
initial portion of the reactor where a large portion of methane
is converted: net balance of the SMR is a strongly endothermic
process. After that point, the gas is being heated while passing
through the reactor and also converting some methane. After the
middle of the reactor, thermodynamic equilibrium was achieved.
At that time, a reduction of the temperature of the oven (according
to the semi-parabolic prole shown in Figure 2) results in a
reduction of the temperature of the column that is reected
in the gas distribution. When the temperature decreases, the
thermodynamic equilibrium dictates that the amount of formed
hydrogen should be smaller and the reactions proceed towards
the consumption of hydrogen, instead to its formation. In this
experiment, we can say that the WGS reaction regulates the ratio
of carbon oxides (COand CO
2
). In the initial portion of the reactor,
the WGS reaction proceeds to form H
2
and CO
2
, consuming some
CO. As the catalyst forms more CO
2
than CO (the velocity of
Figure 6. Simulated results of SMR in catalyst extrudates in steady state:
(a) molar owrates of CO
2
, CH
4
, CO, and H
2
; (b) temperature prole;
(c) reaction rates; and (d) temperature prole inside the catalyst
extrudate. Results for run 1 (P =200kPa, S/C=4.24,
F
0
=24.76mmol/min). Symbols are experimental results and lines are
simulation results. [Colour gure can be viewed in the online issue,
which is available at www.interscience.wiley.com.]
|
VOLUME 87, DECEMBER2009
| |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| 953 |
Figure 7. Simulated effectiveness factors of reactions 13 as a function
of the catalyst bed length. Results for run 1 (P =200kPa, S/C=4.24,
F
0
=24.76mmol/min) at t =10000s after start of reaction. [Colour
gure can be viewed in the online issue, which is available at
www.interscience.wiley.com.]
reaction 3 is faster than the velocity of reaction 1), there comes a
position where the concentration of CO
2
is high enough to reverse
the direction of the WGS to form CO.
The calculated reaction rates can be employed to calculate
the effectiveness factors. The determination of the effectiveness
factors is important to consider simplications to the reactor
model, that is, eliminating the mass and energy balances to
catalyst extrudate and therefore reducing the computation time.
According to Equation (36), when the reaction rate changes
direction, it passes through zero creating a discontinuity in the
effectiveness factor. The results of the integration of the reaction
rates can be observed in Figure 7. Two different asymptotes
are observed. In the case of the second asymptote after the
maximum in the temperature, the change in the direction of
the reaction rates happens as the global composition of the
system is above the equilibrium values due to the decrease in the
temperature. Asymptotes in the effectiveness factor were already
reported by other authors (Xu and Froment, 1989b; Elnashaie
et al., 1990; Pedernera et al., 2003) for a steam methane reforming
reactor.
The experiments performed using the extrudates have shown
that it is possible to describe the results with the mathematical
model proposed. Average values of the effectiveness factor of
each reaction were calculated for each experimental run and are
reported in Table 6. The data provided in this work provided
the necessary information to model the behaviour of the catalyst
Table 6. Average values of the effectiveness factor of reactions 13 for
the Ni/Al
2
O
3
catalyst extrudates in the temperature range of
747813K
Reaction Effectiveness factor
Run 1 Run 2 Run 3 Run 4 Run 5
SMR (1) 0.32 0.32 0.40 0.38 0.32
Watergas shift (2) 0.08 0.08 0.03 0.05 0.08
Global SMR (3) 0.33 0.33 0.41 0.39 0.32
in the temperature range of a SMR-SERP process operated using
hydrotalcites as CO
2
sorbent.
CONCLUSIONS
An experimental study for the steam methane reforming (SMR)
was performed using a commercial Ni/Al
2
O
3
catalyst. Different
experiments using catalyst powder were performed to determine
the true kinetics and experiments using catalyst extrudates were
also done to evaluate the effects of diffusion and evaluate the
effectiveness factors.
The experiments using catalyst powder were performed in the
temperature range of 773890 K. It was observed that the CO
2
selectivity decreases as temperature increases and was always
above 6 at temperatures below 793 K. The kinetic model proposed
by Xu and Froment (1989a) was selected to t the experimental
data. Energies of activation of 217, 68, and 216 kJ/mol were tted
for each of the three proposed kinetic expressions.
A non-isothermal, non-adiabatic xed bed model was used
to describe the experimental results obtained with the catalyst
extrudates. The experimental results were well approximated by
the kinetic parameters determined in catalyst powder using a
tortuosity factor of t
p
=1.56. Based on the mathematical model,
the effectiveness factors of each of the three proposed equations
were calculated. The effectiveness factors have shown asymptotic
behaviour within two different positions along the catalyst bed.
The SMR and global SMR reaction show average values of the
effectiveness factors of j
1
, j
3
=0.4 below 762 K and j
1
, j
3
=0.33
above 806 K. For the watergas shift reaction, the average values
of the effectiveness factors were close to j
2
=0.04 at temperatures
lower than 762 K but increased to double at temperatures higher
than 806 K.
The results presented in this work allowthe complete modelling
of a SMR reactor and they can also be used in hybrid reaction-
separation systems like sorption enhanced reaction process
(SERP) for H
2
production.
ACKNOWLEDGEMENTS
Financial support for this work was in part provided by national
research grant POCTI/EQU/44513/2002 and by LSRE nancing by
FEDER/POCI/2010, for which the authors are thankful. Eduardo
L. G. Oliveira acknowledges his Ph.D. scholarship by FCT
(SFRH/BD/17132/2004).
NOMENCLATURE
a
c
area/volume ratio of the extrudates (m
1
)
A
col
reactor column area (m
2
)
C
i
gas concentration of component i (mol/m
3
)
C
T
total gas concentration (mol/m
3
)
C
vg
gas caloric capacity at constant volume
(J/mol K)
C
pg
gas caloric capacity at constant pressure
(J/mol K)

C
ps
extrudate caloric capacity (J/kg
cat
K)
C
Pw
wall caloric capacity (J/kg
wall
K)
D
col
diameter of reactor column (m)
D
ox.i
axial dispersion (m
2
/s)
D
p
pore diffusivity (m
2
/s)
E
j
energies of activation of reactions j (J/mol)
F
i
molar owrate of component i (mol/min)
h
f
lm heat transfer coefcient (V,(m
2
cat
K))
| 954 |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| |
VOLUME 87, DECEMBER 2009
|
h
w
wall heat transfer coefcient (V,(m
2
cat
.K))
k

mass transfer coefcient in the lm surround-


ing the particle (m/s)
k
01
, k
03
pre-exponential factors of reaction 1 and 3
(mol Pa
0.5
/(kg
cat
s))
k
02
pre-exponential factor of reaction 2
(mol/(kg
cat
s Pa))
K
0CO
. K
0H
2
. K
0CH
4
pre-exponential factor of adsorption of CO, H
2
,
and CH
4
(Pa
1
)
K
0H
2
O
pre-exponential factor of adsorption of H
2
O
K
1
, K
3
equilibrium constants of reactions 1 and 3
(Pa
2
)
K
2
equilibrium constant of reaction 2
L
col
length of reactor (m)
m
cat
mass of catalyst (kg)
P
T
total pressure (Pa)
Q volumetric owrate entering the reactor (m
3
/s)
R
j
reaction rates of reactions j (j =1, 2, 3)
(mol/(kg
cat
s))
R
i
reaction rate of component i (mol/(kg
cat
s))
R
p
pellet radius (m)
R
gas
ideal gas constant (J/mol K)
R
col
column radius (m)
S/C steam to carbon ratio
S
CO
2
selectivity towards carbon dioxide
T temperature (K)
U overall heat transfer coefcient (V,(m
2
col
K))
u gas velocity (m/s)
X
CH
4
methane conversion (%)
w
thik
reactor column thickness (m)
y
i
is the molar fraction of component i
Greek Letters

c
catalyst bed porosity

p
porosity of the catalyst extrudate
LH
i
enthalpy of adsorption of CO, H
2
, CH
4
, and H
2
O (J/mol)
LH
j
heat of reactions 13
z thermal conductivity (W/mK)
, density (kg/m
3
)
u
j.i
stoichiometric coefcients of component i in reactions
j
Subscripts
ax axial properties
cat conditions in the catalyst extrudate
feed feed conditions
gas conditions in the gas phase
i CH
4
, H
2
O, H
2
, CO, CO
2
j reactions 13
P conditions inside the pores of the catalyst extrudates
wall conditions at the column wall
conditions in the heating oven
REFERENCES
Cobden, P. D., G. D. Elzinga, S. Booneveld, J. W. Dijkstra, D.
Jansen and R. W. van den Brink, Sorption-Enhanced
Steam-Methane Reforming: CaO-CaCO
3
Capture Technology,
in GHGT-9-9th International Conference on Greenhouse Gas
Control Technologies, Washington DC, 1620 November
(2008).
Comas, J., M. Laborde and N. Amadeo, Thermodynamic
Analysis of Hydrogen Production from Ethanol Using CaO as
a CO
2
sorbent, J. Power Sources 138, 6167 (2004).
Ding, Y. and E. Alpay, Adsorption-Enhanced Steam-Methane
Reforming, Chem. Eng. Sci. 55, 39293940 (2000).
Dong, W.-S., H.-S. Roh, K.-W. Jun, S.-E. Park and Y.-S. Oh,
Methane Reforming Over Ni/Ce-ZrO
2
Catalysts: Effect of
Nickel Content, Appl. Catal. A 226, 6372 (2002).
Elnashaie, S. S. E. H., A. M. Adris, A. S. Al-Ubaid and M. A.
Soliman, On the Non-Monotonic Behaviour of
Methane-Steam Reforming Kinetics, Chem. Eng. Sci. 45,
491501 (1990).
Essaki, K., T. Muramatsu and M. Kato, Hydrogen Production
From Ethanol by Equilibrium Shifting Using Lithium Silicate
Pellet as CO
2
Absorbent, J. Jpn. Inst. Energy 87, 7275
(2008).
Froment, G. F., Modeling of Catalyst Deactivation, Appl. Catal.
A 212, 117128 (2001).
Hildenbrand, N., J. Readman, I. M. Dahl and R. Blom, Sorbent
Enhanced Steam Reforming (SESR) of Methane Using
Dolomite as Internal Carbon Dioxide Absorbent: Limitations
Due to Ca(OH)
2
Formation, Appl. Catal. A 303, 131137
(2006).
Hou, K. and R. Hughes, The Kinetics of Methane Steam
Reforming Over a Ni/-Al
2
O Catalyst, Chem. Eng. J. 82,
311328 (2001).
Hufton, J. R., S. Mayorga and S. Sircar, Sorption-Enhanced
Reaction Process for Hydrogen Production, AIChE J. 45,
248256 (1999).
Lee, K. B., M. G. Beaver, H. S. Caram and S. Sircar, Production
of Fuel-Cell Grade Hydrogen by Thermal Swing Sorption
Enhanced Reaction Concept, Int. J. Hydrogen Energy 33,
781790 (2008).
Li, D., I. Atake, T. Shishido, Y. Oumi, T. Sano and K. Takehira,
Self-Regenerative Activity of Ni/Mg(Al)O Catalysts With
Trace Ru During Daily Start-Up and Shut-Down Operation of
CH
4
Steam Reforming, J. Catal. 250, 299312 (2007).
Liu, Z.-W., K.-W. Jun, H.-S. Roh and S.-E. Park, Hydrogen
Production for Fuel Cells Through Methane Reforming at Low
Temperatures, J. Power Sources 111, 283287 (2002).
Lysikov, A. I., S. N. Trukhan and A. G. Okunev, Sorption
Enhanced Hydrocarbons Reforming for Fuel Cell Powered
Generators, Int. J. Hydrogen Energy 33, 30613066
(2008).
Matsumura, Y. and T. Nakamori, Steam Reforming of Methane
Over Nickel Catalysts at Low Reaction Temperature, Appl.
Catal. A 258, 107114 (2004).
Mayorga, S. G., T. R. Gaffney, J. R. Brzozowski and S. J. Weigel,
Carbon Dioxide Adsorbents Containing Magnesium Oxide
Suitable for Use at High Temperatures, European Patent
1074297 (2001).
Metz, B., D. Ogunlade, H. Connink, M. Loos and L. Meyer,
Special Report on Carbon Dioxide Capture and Storage, in
Special Report of the Intergovernmental Panel on Climate
Change, (2005).
Ochoa-Fernandez, E., H. K. Rusten, H. A. Jakobsen, M. Ronning,
A. Holmen and D. Chen, Sorption Enhanced Hydrogen
Production by Steam Methane Reforming Using Li
2
ZrO
3
as
Sorbent: Sorption Kinetics and Reactor Simulation, in
International Conference on Gas-Fuel 05, Brugge,
November (2005), pp. 4146.
Oh, Y.-S., H.-S. Roh, K.-W. Jun and Y.-S. Baek, A Highly Active
Catalyst, Ni/Ce-ZrO
2
/-Al
2
O
3
, for On-Site H
2
Generation by
|
VOLUME 87, DECEMBER2009
| |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| 955 |
Steam Methane Reforming: Pretreatment Effect, Int. J.
Hydrogen Energy 28, 13871392 (2003).
Oliveira, E. L. G., C. A. Grande and A. E. Rodrigues, CO
2
Sorption on Hydrotalcite and Alkali-Modied (K and Cs)
Hydrotalcites at High Temperatures, Sep. Purif. Technol. 62,
137147 (2008).
Pedernera, M. N., J. Pina, D. O. Borio and V. Bucala, Use of a
Heterogeneous Two-Dimensional Model to Improve the
Primary Steam Reformer Performance, Chem. Eng. J. 94,
2940 (2003).
Roh, H.-S., K.-W. Jun, W.-S. Dong, J.-S. Chang, S.-E. Park and
Y.-I. Joe, Highly Active and Stable Ni/Ce-ZrO
2
Catalyst for H
2
Production From Methane, J. Mol. Catal. A Chem. 181,
137142 (2002).
Rostrup-Nielsen, J. R., Catalytic Steam Reforming,
Springer-Verlag, New York (1984).
Satrio, J. A., B. H. Shanks and T. D. Wheelock, Development of
a Novel Combined Catalyst and Sorbent for Hydrocarbon
Reforming, Ind. Eng. Chem. Res. 44, 39013911 (2005).
Semelsberger, T. A., L. F. Brown, R. L. Borup and M. A. Inbody,
Equilibrium Products From Autothermal Processes for
Generating Hydrogen-Rich Fuel-Cell Feeds, Int. J. Hydrogen
Energy 29, 10471064 (2004).
Takahashi, R., S. Sato, T. Sodesawa, M. Yoshida and S.
Tomiyama, Addition of Zirconia in Ni/SiO
2
Catalyst for
Improvement of Steam Resistance, Appl. Catal. A 273,
211215 (2004).
Twigg, M. V., Catalyst Handbook, Wolfe Publishing Ltd,
England (1989).
van Dijk, H. A. J., S. Walspurger, P. D. Cobden, D. Jansen, R. W.
Van den Brink and F. G. de Vos, Performance of Water-Gas
Shift Catalysts under Sorption-Enhanced Water-Gas Shift, in
GHGT-9 9th International Conference on Greenhouse Gas
Control Technologies, Washington DC, 1620 November
(2008).
Wang, X. and R. J. Gorte, A Study of Steam Reforming of
Hydrocarbon Fuels on Pd/Ceria, Appl. Catal. A 224, 209218
(2002).
Xiu, G.-H., P. Li and A. E. Rodrigues, Sorption-Enhanced
Reaction Process With Reactive Regeneration, Chem. Eng.
Sci. 57, 38933908 (2002).
Xiu, G.-H., P. Li and A. E. Rodrigues, New Generalized Strategy
for Improving Sorption-Enhanced Reaction Process, Chem.
Eng. Sci. 58, 34253437 (2003).
Xu, J. and G. F. Froment, Methane Steam Reforming,
Methanation and Water-Gas shift: I. Intrinsic Kinetics,
AIChE J. 35, 8896 (1989a).
Xu, J. and G. F. Froment, Methane Steam Reforming: II.
Diffusional Limitations and Reactor Simulation, AIChE J. 35,
97103 (1989b).
Manuscript received January 15, 2009; revised manuscript
received May 19, 2009; accepted for publication May 26, 2009.
| 956 |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| |
VOLUME 87, DECEMBER 2009
|

Potrebbero piacerti anche