Anda di halaman 1dari 9

Transport phenomena in a steam-methanol reforming

microreactor with internal heating


Jeong-Se Suh*, Ming-Tsang Lee, Ralph Greif, Costas P. Grigoropoulos
Department of Mechanical Engineering, University of California at Berkeley, Berkeley, CA 94720-1740, USA
a r t i c l e i n f o
Article history:
Received 16 June 2008
Received in revised form
20 September 2008
Accepted 21 September 2008
Available online 28 November 2008
Keywords:
Fuel cell
Methanol conversion
Steam reforming
Packed-bed reformer
Catalyst
Internal heating
a b s t r a c t
An experimental and theoretical study of steam reforming of methanol is carried out in
a packed-bed microreactor with internal heating. Experimental results of the methanol
conversion and carbon monoxide concentration in an internally heated reformer are
compared with those of an externally heated reformer. Higher methanol conversion and
carbon monoxide concentration are obtained for internal heating at the same conditions.
The results show the conversion efciency of methanol and CO concentration increase
with increasing internal heating rate over the range of operating conditions. A correlation
for the conversion efciency of methanol has been obtained as a function of the internal
heating rate and a dimensionless time parameter which represents the ratio of the char-
acteristic time of the methanol ow to the time for chemical reaction.
2008 International Association for Hydrogen Energy. Published by Elsevier Ltd. All rights
reserved.
1. Introduction
In recent years, there has been considerable interest on the use
of hydrogen as an energy carrier for small fuel cell applications
to provide high volumetric and gravimetric energy density for
miniaturizedportable electronic systems [1]. Inorder touse fuel
cells for portable devices, it is essential to develop a small and
lightweight hydrogen supply; this requires efcient microre-
formers to convert hydrocarbon fuel into hydrogen which can
be delivered to a proton exchange membrane fuel cell to
produce electricity. Anumber of studies indicate that methanol
is an attractive fuel because of its low reforming temperature
(200300

C) [2], good miscibility with water and low content of


sulfur compounds. The methanol reformer operating at
moderate temperature also produces low CO concentration
which is important because of CO contamination of the anode.
Steam reforming of methanol is achieved by the chemical
reaction in a catalyst packed bed and is an endothermic reac-
tion which requires heat supply from an external source. The
gas mixture of methanol and steam is generally heated from
a packed catalyst bed wall with lower temperatures resulting
within the bed. Yoon et al. [3] studied an externally heated
packed-bed steam-methanol reformer at a scale larger than
was used in the present study. Their results showed that the
temperature near the center of the reformer is lower than that
near the wall. Davieau and Erickson [4] studied steam-meth-
anol reformers with different aspect ratios (i.e. reactor length
to diameter ratio) but with the same volume. Their experi-
mental results showed that, at the same wall temperature,
space velocity and temperature of reactants at the inlet to the
reformer, methanol conversion with the smaller diameter
reformer is greater than with the larger diameter reformer.
* Corresponding author. Present address: School of Mechanical and Aerospace Engineering, ERI, Gyeongsang National University, 900
Gaza-Dong, Jinju 660-701, Korea. Tel.: 82 55 751 5312; fax: 82 55 751 5622.
E-mail address: jssuh@gnu.ac.kr (J.-S. Suh).
Avai l abl e at www. sci encedi r ect . com
j our nal homepage: www. el sevi er . com/ l ocat e/ he
0360-3199/$ see front matter 2008 International Association for Hydrogen Energy. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijhydene.2008.09.049
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 4 ( 2 0 0 9 ) 3 1 4 3 2 2
They further carried out numerical analyses to obtain the
temperature proles in the reformers and showed that the
temperature near the core of the packed-bed is lower thanthat
near the wall. The lower temperature in the packed-bed leads
to less catalyst activity. Another consideration is that the
externally heated reformer results in signicant energy
transfer to the environment which decreases the overall ef-
ciency of the fuel cell system[5]. The reduction of the heat loss
can be achieved by the conguration of the system. For
example, Holladay et al. [6] fabricated and tested a microscale
cylindrical methanol reformer which integrated a catalytic
combustor of methanol or hydrogen with a methanol
reformer. They achieved a thermal efciency as high as 19%.
Terazaki et al. [7] fabricated and tested a multilayered micro-
reformer which integrated a CO remover and vaporizer on
a methanol reformer heated by a catalytic combustor. The CO
remover and the vaporizer work at lower temperature
compared to the reformer and also utilize the heat that is
transferredfromthe surface of the reformer. Thermal barriers,
insulation components and complex congurations increase
the cost, complexity and volume of the fuel cell system.
In a packed bed heated from an external source, several
efforts have been made to overcome the adverse lower
temperatures in the core. Perry et al. [8] showed that micro-
wave heating could increase the core temperature in
a packed-bed methanol reformer. In the present study, an
internally heated structure in a steam-methanol reformer is
studied. Internal heating is supplied fromthe centerline of the
packed-bed reformer by suspending a heating element
(resistive heating wire) near the center of the catalyst bed.
A number of kinetics models for steam-methanol reform-
ing have been reported in the literature [914]. Amphlett et al.
[9] suggested a semi-empirical kinetics model for the steam-
methanol reforming process over the CuO/ZnO/Al
2
O
3
catalyst,
which includes the mathematical model of the reaction rates
for the plug ow reactor. Park et al. [15] carried out a one-
dimensional analysis for the mass transport in a steam-
methanol reformer that employs Amphletts reaction kinetics
model. Karimet al. [16] carried out a two-dimensional pseudo-
homogeneous analysis for packed-bed reactors. In this model,
they took the diffusion coefcient of the gas mixture in the
reaction process to be constant and focused on the effect of
heat transfer on the measured methanol conversion. Park
et al. [17] carried out a quasi 3-D analysis of the reacting ow
in a steam-methanol reformer, coupled to 3-D heat transfer in
a silicon wafer. Yuan et al. [18] simulated, in a three-dimen-
sional study, the heat and mass transport in a methane steam
reforming duct considering the catalytic chemical reactions.
In the present study, experiments with an internally heated
reformer arecarriedout under different operatingconditionsand
Nomenclature
A, B pre-exponential term in the Arrhenius expression
for k
c
1
concentration of methanol [mol/m
3
]
C
D
, C
R
modication factors of decomposition and
reforming reaction
c
p,i
specic heat of species i [J/kg K]
D
h
characteristic diffusion length [m]
D
1m
ordinary diffusivity [m
2
/s]
d
b
diameter of catalyst bed [m]
d
i
diameter of internal rod [m]
E activation energy in the Arrhenius expression for k
[J/mol]
F
l
feed rate of liquid mixture [ml/min]
DH enthalpy of reaction [J/mol]
k thermal conductivity [W/mK]
k
%
D
volumetric reaction constant for decomposition
[mol/kg s]
k
%
D
modied reaction constant for decomposition,
1 3r
s
k
%
D
[mol/m
3
s]
k
00
R
surface reaction constant [m/s]
k
%
R
volumetric reactionconstant for reforming [m
3
/kg s]
k
%
R;0
modied reaction constant for reforming at
T 180

C, 1 3r
s
k
%
R
[s
1
]
L
b
length of bed [m]
Le Lewis number, Le r
g
c
p;g
D
1m
=k
g
M
i
molecular weight of species i [kg/mol]
m
i
mass fraction of species i
m
s
mass of catalyst [kg]
m% volumetric mass source [kg/m
3
s]
q
i
internal heating rate [W]
q
00
i
internal heating ux [W/m
2
]
q% volumetric heat source [W/m
3
]
r, z radial and axial Cartesian coordinates [m]
r
%
i
volumetric chemical reaction rate of species i
[mol/m
3
s]
S surface area [m
2
]
Sc Schmidt number
Sh Sherwood number
SMR molar ratio of steam to methanol
T temperature [

C]
U velocity of mixture [m/s]
V volume [m
3
]
x
i
mole fraction of species i
x
i
cross-sectional area-weighted average of mole
fraction of species i
Greek symbols
3 porosity
h conversion efciency of methanol,
x
CH3OH;in
x
CH3OH;out
=x
CH3OH;in
r density [kg/m
3
]
Subscripts
0 reference
b bed
D decomposition
g gas mixture
in inlet
i species
[ liquid
m average
p pellet
R reforming
s solid
i nt e r na t i o na l j o ur na l o f hy d r og e n e ne r gy 3 4 ( 2 0 0 9 ) 3 1 4 3 2 2 315
are comparedwithresults for anexternally wall heatedreformer
[19]. Themethanol steamreformer withinternal heatinghas also
been analyzed by solving numerically the transport equations of
the gas mixture associated with the composition of each species
throughout the reformer. The steam reforming process of
methanol is characterized by the internal heating rate, the
operating temperature, velocity of the steammethanol mixture
reactant, the inlet temperature of the mixture, the mass of
the catalyst, etc. The effect of the internal heating rate on the
conversion efciency and CO production from methanol in the
reformer has been investigated with the feed rate of the water
methanol mixture and the mass of the catalyst as parameters
andthe results are comparedwiththeexperimental data. For the
commercial BASF F3-01 (CuO/ZnO/Al
2
O
3
) catalyst, a correlation
for the conversionefciency of methanol has beenpresented for
the microreformer as a function of the rate of internal heating
and a dimensionless time parameter that is related to the ow
rate of the steamwater mixture fed to the reactor, the catalyst
mass and the reaction rate. In addition, the radial variations of
temperature and of the molar fraction of methanol in a micro-
reactor with internal heating have been presented.
2. Experiment
The conguration of the internally heated reformer is shown in
Fig. 1. The heating structure is constructed by embedding
a 0.005
00
diameter chromium alloy (Chromega

wire from
Omega Inc.) resistive metal wire (1, Fig. 1(b)) into a borosilicate
glass capillary (2, Fig. 1(b)) having 200 mm i.d. and 330 mm o.d.
with 4 cm length. The glass tube contains and supports the
resistive heating wire. Aroundshapedcapillarytube(3, Fig. 1(b))
with a 500 micron i.d. and 700 mm o.d and a length of 1.3 cm is
inserted in a square shape capillary tube (4, Fig. 1(b)) with inner
side widthof 800 mm anda lengthof 1.3 cm. The combinationof
the square and round shape capillaries provides the gap
betweenthosecapillaries whichallows uidtoowthroughthe
region shown in dots in Fig. 1(b). A capillary tube (5, Fig. 1(b))
with a 1.5 mm i.d. and 1.8 mm o.d. and a length of 4 cm is used
as the reformer tube. The middle part of the reformer tube is
lled with 16 mg of 150 mm BASF F3-01 Al
2
O
3
/CuO/ZnOcatalyst
particles; silica wool is used to seal the catalyst bed on both
ends. The length of the catalyst bed is approximately 1.2 cm.
The chromiumresistive wire is centrally suspended and tightly
xed by bending the wire on both ends of the reformer and
inserting the reformer tube and the wire into a PTFE tube (6,
Fig. 1(b)) havinga1/16
00
i.d. and1/8
00
o.d. Electrical connectionsto
the wire are made on the wire ends that are exposed at the
edges of the PTFE tubes on both ends.
Reforming experiments are carried out in the same
manner as was described for the externally heated reformer
experiments [19]. However, the heat is now internally gener-
ated from the heating wire centrally suspended in the
reformer bed. The wall temperature of the reformer is
between 160 and 200

C and results from the electrical power


froma DCpower supply to the internally heated resistive wire.
The power is in the range of 0.35 q
i
0.65 W. The liquid
mixture of water and methanol with a molar ratio (SMR) of 1.1
ows into the reformer system at the rate of 15 ml/min. The
liquid mixture is preheated to stabilize the ow before
entering the catalytic bed. A small K-type thermocouple wire
of 0.003
00
diameter on the outer wall of the reformer tube at the
middle of the catalytic bed section measures the character-
istic wall temperature of the reformer. Another 0.003
00
diam-
eter K-type thermocouple is inserted into the PTFE tubing at
the location approximately 3 mm upstream to the inlet of the
glass reformer tubing to measure a characteristic inlet
temperature of the steammethanol mixture and yields the
range of 85

CT
in
130

C. At the outlet of the reformer, the


nonreacted water and methanol mixture condenses to liquid.
The mixture of gas phase products, H
2
, CO
2
and CO, ows
through the owmeter and hydrogen sensor for the conver-
sion measurement and is collected by a gas sampling bag for
CO measurement. The CO measurement is carried out with
a Fourier Transform Infrared Spectrometer (FTIR).
3. Analysis of the steam-methanol
reforming process
3.1. Chemical kinetics and reaction rate
Methanol can be reformed by two overall reactions in
a reformer lled with the catalyst CuO/ZnO/Al
2
O
3
as described
PTFE tubing for
fluid connection
Heating structure holder
Suspended
heating wire
Glass tubing space for catalyst
particle packing bed
6
5
4
3
2
1
a
b
Fig. 1 Conguration of the internally heated
microreformer. (a) A three dimensional cross-sectional
sketch, (b) A cross-sectional sketch in which the dots
denotes the inlet ow region (tube thickness is omitted).
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 4 ( 2 0 0 9 ) 3 1 4 3 2 2 316
by Amphlett et al. [9]:
CH
3
OH H
2
O4
kR
CO
2
3H
2
(1)
CH
3
OH4
k
D
CO2H
2
(2)
where the subscript R refers to steam reforming, and D
refers to decomposition. Methanol is primarily reformed via
reaction (1) called reforming. Some portion of the methanol
also decomposes directly to CO via reaction (2). In addition,
the watergas shift reaction adjusts the composition of the
product gas:
CO H
2
O4CO
2
H
2
(3)
The watergas shift reaction [reaction (3)] can be neglected
without a substantial loss in accuracy of methanol concen-
trations as described by Amphlett et al. [9]. The generation
rates of the species can be expressed as
r
%
1
r
%
R
r
%
D
; r
%
2
r
%
R
; r
%
3
r
%
R
; r
%
4
r
%
D
; r
%
5
3r
%
R
2r
%
D
(4)
where the subscripts are 1 CH
3
OH, 2 H
2
O, 3 CO
2
, 4 CO,
and 5 H
2
. The methanol reforming reaction occurs on the
catalyst particles in a packed bed during the reforming of
a steammethanol mixture. It also includes the heat and mass
transfer in the catalyst porous mediumof the packed bed. Our
analysis for reaction rate is undertaken by considering the
mass diffusionof methanol fromthe free streamto the surface
of the catalyst pellet, and into the pores of the pellet. The
methanol steammixture diffuses into the pores of the pellet
and encounters two resistances; one is a convective resistance
between the free stream and pellet surface, and the other is
a diffusive resistance within the pellet pore. The heteroge-
neous reaction rate within the packed bed of catalyst r
%
R
can be
expressed as [20,21]
r
%
R
1 3
S
p
V
p
c
1
1
_
V
p
r
s
=S
p
_
h
p
k
%
R

D
h
ShD
1m
(5)
The effectiveness, h
p
, of the catalyst pellet is the ratio of the
actual reaction rate divided by that for a pellet with an innite
diffusion coefcient for a spherical pellet [22,23]. The diffu-
sivity of the mixture within the packed bed is given in Ref. [22].
The reaction rate constants depend on the properties of the
catalyst surface and on the temperature. These quantities can
be related to the rate constants of Amphlett et al. [9] which are
derived from an Arrhenius relation as follows:
k
%
R
C
R
A
R
B
R
ln SMRe
E
R
=RT
(6)
k
%
D
C
D
A
D
e
E
D
=RT
: (7)
where SMR is the molar ratio of steamto methanol; A
R
, B
R
, and
A
D
are constants specied by Amphlett (ref. [9]) for reforming
and decomposition reactions, respectively. C
R
and C
D
are
correction factors for reforming and decomposition reactions,
respectively and can be determined empirically from the
activity of the catalyst. Fromthe experimental data [19] for the
BASF F3-01 (CuO/ZnO/Al
2
O
3
) catalyst, the factors C
R
and C
D
were determined to have the values of 5.5 in the reforming
and decomposition reactions.
The correlation for a dry packed bed for the Sherwood
number (Sh), dened as a dimensionless mass transfer
conductance in equation (5), is given in Ref. [24].
3.2. Conservation equations and numerical procedures
The analytical conguration of the methanol steam reformer
with internally heated rod is axisymmetric as shown in Fig. 2,
and includes the isotropic and continuously distributed cata-
lyst. The rate of heat generation from the internal rod is
uniform. The ow of the steammethanol mixture is assumed
to be steady with no angular variation. The mixture consists of
ve species, takento be ideal gases, and reacts heterogeneously
withthe catalyst. It is also assumedthat the thermal state of the
catalyst and the gas mixture is locally in equilibrium and the
Lewis numbers of all species are equal to unity,
Le r
g
c
p;g
D
1m
=k
g
1. The coordinate systemis showninFig. 2.
The mass conservation of the gas mixture canbe expressed
as
d
dz
_
3r
g
U
_
0 (8)
where U is the mean velocity of the gas mixture, and 3 is the
catalyst porosity given by 3 1 (V
s
/V). The equation for the
conservation of energy reduces to [25]
3r
g
c
p;g
U
vT
vz

1
r
v
vr
_
rk
m
vT
vr
_

v
vz
_
k
m
vT
vz
_
q
%
m
(9)
where k
m
is a volume-weighted average conductivity of the
catalyst material and gas mixture, and q
%
m
is the volumetric
heat source of the porous medium generated from the
heterogeneous reaction between the gas mixture and catalyst:
k
m
1 3k
s
3k
g
(10)
q
%
m
DH
R
r
%
R
DH
D
r
%
D
: (11)
The specic heat of the gas mixture c
p,g
is obtained from the
mass-weighted average for each species, c
p;g

5
i1
m
i
c
p;i
[22].
The thermal conductivity of the gas mixture k
g
is calculated
from Wilkes mixture rule [26]. The conservation of mass
equation for species i reduces to [25]
3r
g
U
vm
i
vz

1
r
v
vr
_
rr
g
D
m
vm
i
vr
_

v
vz
_
r
g
D
m
vm
i
vz
_
m
%
g;i
(12)
where m
i
is the mass fraction of gas component i dened by
m
i
r
i
/r
g
, D
m
is the mean diffusivity of the gas mixture in the
porous media, and m
%
g;i
is the mass generation of gas compo-
nent i:
z
r
L
b
Catalyst Bed
T
w
d
i
d
b
Internally Heated Rod
i
q
T
in
Fig. 2 Schematic of the modeling domain in the reactor
with internally heated rod.
i nt e r na t i o na l j o ur na l o f hy d r og e n e ne r gy 3 4 ( 2 0 0 9 ) 3 1 4 3 2 2 317
D
m
3D
1m
(13)
m
%
g;i
M
i
r
%
i
(14)
There is uniformgeneration of heat fromthe internal rod so
that the heat ux, q
00
i
constant, is the specied boundary
condition. A constant tube wall temperature, Tr d
b
=2 T
w
and the temperature of the gas mixture at the inlet,
Tz 0 T
in
, are also specied. The mass transfer for each of
the species is impermeable through the tube and internal rod
walls, vm
i
=vrj
rd
b
=2; d
i
=2
0. The liquid mixture of water and
methanol is considered to have been preheated and then
ows as a gas mixture into the packed catalyst at z 0. The
molar ratio of water to methanol is prescribed and the velocity
of the gas mixture is uniform. The physical parameters are
listed in Table 1 and the other physical gas properties of each
species are given in the existing literature [22,23,2729].
The governing equations were solved numerically using
the method of Karki and Patankar [30]. The governing equa-
tions are transformed into the curvilinear cylindrical coordi-
nate system r r(x, h) and z z(x, h) and discretized using the
central difference scheme [31]. The resulting algebraic equa-
tions were solved using a nonuniform grid system with 71
nodes in the x direction and 45 nodes in the h direction. Iter-
ations were continued until changes in the mass fractions
were less than 0.1%. It was found that the results for the
temperatures and mass fractions are differed by less than
0.01% from a grid system of 90 142 nodes.
4. Results and discussion
4.1. Experimental results
Fig. 3 shows the variation of the conversion efciency of
methanol in the internally heated microreformer with respect
to the wall temperature. Experiments were repeated several
times at each wall temperature. In this gure, the range of
experimental results is shown at each wall temperature with
the conversion efciency of methanol increasing with
increasing wall temperature of the microreformer.
Fig. 4 shows the methanol conversion efciencies for the
internally heated and the externally heated microreformers
[19,23]. It is pointedout that the results for the externally heated
reformer are obtained by interpolating the experimental data
0
0.2
0.4
0.6
0.8
1
150 160 170 180 190 200 210
T
w
(C)
Fig. 3 Methanol conversion with respect to the wall
temperature of reformer.
Table 1 Geometric parameters and physical properties.
Diameter of catalyst bed (d
b
) 1.5 10
3
m
Axial length of catalyst bed (L
b
) 12 10
3
m
Diameter of internal rod (d
i
) 0.33 10
3
m
Porosity of catalyst bed (3) 0.35
Density of catalyst (r
s
) 1300 kg/m
3
)
Thermal conductivity of catalyst (k
s
) 0.3 W/mK(Karim et al. [16])
0
0.2
0.4
0.6
0.8
1
140 160 180 200 220 240
Internal Heating
External Heating
T
w
(C)
Fig. 4 Methanol conversion efciencies of internally and
externally heated reformers.
0
500
1000
1500
2000
2500
0.1 0.2 0.3 0.4 0.5 0.6
Internal heating
External heating
C
O

(
p
p
m
)
Fig. 5 Comparison of the carbon monoxide concentration
to methanol conversion for internally and externally
heated reformers. The externally heated reformer data
were interpolated from the previous study [19].
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 4 ( 2 0 0 9 ) 3 1 4 3 2 2 318
[19] at the same experimental conditions of the internally
heated reformer except for the heating method. For both the
internally heated and the externally heated reformers, the
conversion efciency of methanol increases with increasing
temperature with internal heating attaining larger efciencies.
As is discussed in the next section, at the same wall tempera-
ture the core temperatures for the internally heated reformer
are higher than for the externally heated reformer which result
in higher conversion efciencies.
The CO concentrations of the internally and externally
heated reformers are shown in Fig. 5 and correspond to the
same steammethanol ow rate and catalyst mass. The
concentration of CO is seen to increase with increasing
conversion efciency of methanol. At the same methanol
conversion, the CO concentration of the internally heated
reformer is higher than that of the externally heated reformer.
It is important to point out that both methanol conversion
efciency and CO concentration increase with increasing
reformer wall temperature (cf. Figs. 3 and 5). However,
increasing reformer wall temperature increases the tempera-
ture difference between the reformer and the environment or
the fuel cell main stack which results in increased energy los-
ses. High reformer wall temperature also yields melting of the
PTFE tubes. Therefore, the experimental tests in the present
study were performedat temperature below200

C. However, it
is important to point out and to emphasize that with internal
heating, the core temperature of the catalyst bed is much
greater thanthewall temperaturewhichyieldsreasonablylarge
methanol conversion efciencies (up to w60%, cf. Fig. 3).
4.2. Numerical results
Numerical results are obtained for a methanol steam
reformer with an internally heated rod over the range of
conditions: 0 q
00
i
pd
i
L
b
q
i
0:75 W, 160

CT
w
200

C,
80

CT
in
180

C, 10 mg m
s
18 mg, 6 ml/minF
[,0
22 ml/
min. Note that the condition for no internal heating(i.e. q
i
0)
corresponds to the external heating condition with a small
rod present.
The numerical results for the conversion efciency of
methanol, along with the experimental results, are shown in
Fig. 6(a) as a function of the wall temperature of the reformer.
There is generally good agreement betweenthe numerical and
the experimental results with the numerical results being
slightly lower than the experimental results. The numerical
results for the mole fraction of CO produced from reaction (2)
in the catalyst and the experimental results are shown in
Fig. 6(b) as a function of the wall temperature of the reformer.
The numerical results are signicantly lower than the exper-
imental data. The signicant difference between the numer-
ical results and experimental data for the carbon monoxide
0.0
0.2
0.4
0.6
0.8
1.0
160 170 180 190 200
Experiment
Analysis
T
in
(C)
q
i
(W)
T
w
(C)
103 0.41 160
108 0.44 170
116 0.48 180
89 0.56 190
114 0.60 200
0
500
1000
1500
2000
2500
3000
160 170 180 190
C
O
(
p
p
m
)
T
in
(C)
q
i
(W)
T
w
(C)
98 0.37 161
95 0.41 171
106 0.49 181
116 0.55 191
Experiment
Analysis
T
w
(C)
T
w
(C)
a
b
Fig. 6 Comparison of the numerical results with
experimental data for internal heating for the feed water
methanol ow rate of 15 ml/min (space velocity [41 ml/
(ml hr), Ref. [33]) and m
s
[16 mg: (a) conversion efciency of
methanol and (b) Molar fraction of CO with respect to the
wall temperature of microreformer.
0.0
0.2
0.4
0.6
0.8
1.0
160 170 180 190 200
0
0.25
0.5
0.75
q
i
[W]
T
in
=120C
T
w
(C)
Fig. 7 Calculated conversion efciency of methanol for
four cases of internal heating rate for m
s
[16 mg.
i nt e r na t i o na l j o ur na l o f hy d r og e n e ne r gy 3 4 ( 2 0 0 9 ) 3 1 4 3 2 2 319
concentration can be attributed to the choice of the chemical
reaction kinetics for the carbon monoxide formation. Studies
[12,32] have suggested that in steam-methanol reforming with
the CuO/ZnO/Al
2
O
3
based catalyst, carbon monoxide is
produced as a secondary product from the reverse watergas
shift (RWGS) reaction:
CO
2
H
2
4CO H
2
O (15)
rather than frommethanol decomposition that was utilized in
the present study.
The calculated conversion efciency of methanol as
a function of the wall temperature is shown in Fig. 7 for
several values of the heat generation rate from an internally
heated rod for a feed water-methanol ow rate of 15 ml/min. A
space velocity can also be dened as [33]
Space velocity
Reactant flow rate
Reactor volume

900 ml=hr
22 ml
41
_
ml=hr
ml
_
(16)
The effect of increasing wall temperature and increasing heat
generation rate is to increase the conversion efciency of
methanol inthereformer. Theconversionefciencyof methanol
is higher for the internally heated reformer; note that the wall
heated reformer with no internal heating (q
i
0) corresponds to
the externally heated reformer and is the solid curve in Fig. 7.
The axial and radial distributions of temperature and
the methanol mole fraction are shown in Fig. 8 for the
wall temperature, T
w
180

C, and the inlet temperature,


T
in
120

C, for four heat generation rates, q


i
, including q
i
0.
Fig. 8(a) shows the axial variation for temperature and the
mole fraction of methanol on the surface of the internal rod.
The surface temperature of the internal rod increases sharply
near the inlet of the catalyst bed, while the mole fraction of
120
160
200
240
280
0.3
0.4
0.5
0.6
0 2 4 6
T
[

C
]
T
[

C
]
x
C
H
3
O
H
x
C
H
3
O
H
z / d
b
q
i
(W)
0
0.25
0.5
0.75
T
w
=180C
T
in
=120C
180
200
220
240
260
0.30
0.40
0.50
0.60
0.2 0.4 0.6 0.8 1
r / r
b
T
w
=180C
T
in
=120C
q
i
[W]
0
0.25
0.5
0.75
a
b
Fig. 8 Proles for temperature and molar fraction of
methanol at several cases of internal heating rate: (a) along
the axial direction of the reactor on the surface of the
internal rod, r [d
i
/2, and (b) in the radial direction of the
reactor at z [L
b
/2 and T
in
[120 8C.
0.0
0.2
0.4
0.6
0 2 4 6 8
CH
3
OH
H
2
O
CO
2
CO
H
2
T
w
=180C
q
i
=0.5W
z / d
b
x
x
0.0
0.2
0.4
0 0.2 0.4 0.6
Heat generation rate from inner rod, q
i
(W)
CH
3
OH
H
2
O
CO
2
CO
H
2
T
w
=180C
/ min(Space Velocity 41 /( hr), Ref .[33]) 15 F
l,0
a
b
Fig. 9 Bulk molar fraction variations of the components
for the feed watermethanol ow rate of 15 ml/min (space
velocity [41 ml/(ml hr), Ref. [33]): (a) Molar fraction
variations of species along the axial direction of the reactor
and (b) Molar fraction variations of species at the exit of the
reformer (z [L
b
) with respect to heating rate q
i
from the
internal rod.
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 4 ( 2 0 0 9 ) 3 1 4 3 2 2 320
methanol continuously decreases downstream. As shown in
Fig. 8(b) at the central axial location of the bed, z L
b
/2, the
temperature in the packed bed increases with increasing heat
generation, q
i
. Also, note that for the same wall temperature,
the bed temperatures for internal heating (q
i
s0) are greater
than for external (wall) heating (q
i
0). This results in greater
conversion for the internally heated reformer (higher bed
temperature) than for the externally heated reformer (cf. Fig. 4).
Note that for q
i
0 there is little variation of the temperature in
the radial direction (which is not discernible in Fig. 8(b)), but
withinternal heating the temperature decreases signicantlyin
the radial directionfromthe core to the wall. Note that the mole
fraction of methanol varies very little in the radial direction.
The results suggest that the transport of methanol in the radial
direction is very effective and thereby yields almost uniform
prole of the methanol concentration in the radial direction.
Fig. 9 shows the bulk molar fraction variations of the
components. As shown in Fig. 9(a), when the steammethanol
mixture decomposes at the constant wall temperature,
T
w
180

C, the main products are H


2
and CO
2
, and only
a small amount of CO is produced, which is not discernible in
the gure. As the steammethanol mixture is gradually
reformed through the reactor, the molar fractions of hydrogen
and CO
2
increase along the axial length of the reactor. Fig. 9(b)
shows the mole fraction variations of species at the exit of the
reformer (z L
b
) with the heat generation rate. The increase of
H
2
produced in the reformer is attributed to the activation of
the chemical reaction of methanol with increasing tempera-
ture resulting from the increasing heat generation.
The conversion efciency of methanol can be expressed as
a function of a dimensionless parameter which is related to the
ratio of the characteristic time of the methanol ow, m
s
/F
l,0
r
s
to
the time for the methanol chemical reaction in the reformer,
1=k
%
R;0
[23]. The characteristic time was also presented in Ref.
[33]. Results for the conversion efciency of methanol are
shown in Fig. 10 for four internal heat generation rates in terms
of the dimensionless time parameter, m
s
k
%
R;0
=F
l;0
r
s
. The
conversion efciency of methanol increases with increasing
values of m
s
k
%
R;0
=F
l;0
r
s
. From the numerical results, the conver-
sion efciency of methanol is correlated as follows:
h 0:70 0:67q
i
W
_
0:33 0:38q
i
W
_
ln
_
m
s
k
%
R;0
F
l;0
r
s
_
(17)
The conversion efciency of methanol is shown in Fig. 11 for
several values of the inlet temperature in terms of the dimen-
sionless timeparameter. The conversionefciency of methanol
decreases slightly with increasing values of the inlet tempera-
ture. The reduction in the conversion efciency of methanol at
the higher inlet temperature results from the increasing volu-
metric ow rate of the steammethanol mixture at the inlet to
the reformer and the resulting smaller residence time.
5. Conclusions
The effects of internal heating on the conversion character-
istics of steam reforming of methanol in a microreactor tube
packed with a commercial BASF F3-01 (CuO/ZnO/Al
2
O
3
) cata-
lyst have been experimentally and numerically investigated.
Experimental results show that at the same condition of wall
temperature, inlet steammethanol ow rate and catalyst
mass, the conversion efciency of methanol is higher for the
internally heated reformer than for the externally heated
reformer. The concentration of CO increases with tempera-
ture and methanol conversion. The predicted results for the
conversion efciency of methanol agree with the experi-
mental data over the range of operating conditions. However,
the predicted results for the CO concentration do not agree
with the experimental data. From the numerical results,
a correlation for the conversion efciency of methanol over
the range of operating conditions is presented as a function of
the internal heating rate and a dimensionless time parameter
0.0
0.2
0.4
0.6
0.8
1.0
200 400 600 800
0
0.25
0.5
0.75
'''
/
s R,0 s
m k F
l,0
q
i
[W]
T
w
=180C
T
in
=120C
Fig. 10 Results for the conversion efciency of methanol
as a function of the dimensionless time parameter.
0.0
0.2
0.4
0.6
0.8
1.0
200 400 600 800
80
120
180
T
in
(C)
T
w
=180C
q
i
=0.5W
'''
/
s R,0 s
m k F
l,0
Fig. 11 Results for the conversion efciency of methanol
as a function of the dimensionless parameter for several
values of inlet temperature of packed bed with T
w
[180 8C.
i nt e r na t i o na l j o ur na l o f hy d r og e n e ne r gy 3 4 ( 2 0 0 9 ) 3 1 4 3 2 2 321
which represents the ratio of the characteristic time of the
methanol ow to the time for the chemical reaction of
methanol. The numerical results show that the conversion
efciency of methanol increases with internal heating rate
over the range of operating conditions and decreases slightly
with increasing inlet temperature of packed bed. The inter-
nally heated reformer improved the conversion of methanol
and also increased the production of CO in the reformer.
In addition, the numerical results show that the molar
fraction of each species varies very little in the radial direc-
tion. There is signicant variation of the gas temperature in
the radial directions for internal heating; but for no internal
heating (q
i
0), there is little temperature variation in the
radial direction. For the internally heated reformer, the higher
temperatures in the inner region of the packed bed yield
higher CO production.
Acknowledgements
We thank Stephen Sharratt and Allan Chen of the University
of California at Berkeley who assisted in carrying out the
measurements. This work was partially supported by the
Industrial Technology and Research Institute (ITRI) in Taiwan,
R.O.C., and the Berkeley ITRI Research Center (BIRC). Professor
J.-S. Suh wishes to acknowledge support from the 2nd BK and
the NURI Projects of Gyeongsang National University in Korea.
r e f e r e n c e s
[1] Ghenciu AF. Review of fuel processing catalysts for hydrogen
production in PEM fuel cell systems. Current Opinion in Solid
State and Materials Science 2002;6(5):38999.
[2] Brown LF. A comparative study of fuels for on-board
hydrogen production for fuel-cell-powered automobiles.
International Journal of Hydrogen Energy 2001;26(4):38197.
[3] Yoon CY, Otero J, Erickson PA. Reactor design limitations for
the steam reforming of methanol. Applied Catalysis B:
Environmental 2007;75:26471.
[4] Davieau DD, Erickson PA. The effect of geometry on reactor
performance in the steam-reformation process.
International Journal of Hydrogen Energy 2007;32:1192200.
[5] Hotz N, Lee MT, Grigoropoulos CP, Senn SM, Poulikakos D.
Exergetic analysis of fuel cell micropowerplants fed by
methanol. International Journal of Heat and Mass Transfer
2006;49(1516):2397411.
[6] Holladay JD, Jones EO, Dagle RA, Xia GG, Cao C, Wang Y. High
efciency and low carbon monoxide micro-scale methanol
processors. Journal of Power Sources 2004;131:6972.
[7] Terazaki T, Nomura M, Takeyama K, Nakamura O,
Yamamoto T. Development of multi-layered microreactor
with methanol reformer for small PEMFC. Journal of Power
Sources 2005;145:6916.
[8] Perry WL, Datye AK, Prinja AK, Brown LF, Katz JD. Microwave
heating of endothermic catalytic reactions: reforming of
methanol. AIChE Journal 2002;48(4):82031.
[9] Amphlett JC, Creber KAM, Davis JM, Mann RF, Peppley BA,
Stokes DM. Hydrogen-production by steam reforming of
methanol for polymer electrolyte fuel-cells. International
Journal of Hydrogen Energy 1994;19(2):1317.
[10] Peppley BA, Amphlett JC, Kearns LM, Mann RF. Methanol-
steam reforming on Cu/ZnO/Al
2
O
3
part 1: the reaction
network. Applied Catalysis 1999;179:219.
[11] Peppley BA, Amphlett JC, Kearns LM, Mann RF. Methanol-
steam reforming on Cu/ZnO/Al
2
O
3
catalysts part 2:
a comprehensive kinetic model. Applied Catalysis A: General
1999;179(12):3149.
[12] Agrell J, Birgersson H, Boutonnet M. Steam reforming of
methanol over a Cu/ZnO/Al
2
O
3
catalyst: a kinetic analysis
and strategies for suppression of CO formation. Journal of
Power Sources 2002;106(12):24957.
[13] Lee JK, Ko JB, Kim DH. Methanol steam reforming over Cu/
ZnO/Al
2
O
3
catalyst: kinetics and effectiveness factor. Applied
Catalysis A: General 2004;278(1):2535.
[14] Choi YT, Stenger HG. Kinetics, simulation and optimization
of methanol steam reformer for fuel cell applications.
Journal of Power Sources 2005;142(12):8191.
[15] Park HG, Lee MT, Hsu FK, Grigoropoulos CP, Greif R. Transport
in a methanol steam reformer as the fuel processor for fuel
cell systems. In: Proc. IMECE 2004; 2004. p. 17.
[16] Karim A, Bravo J, Datye A. Nonisothermality in packed bed
reactors for steam reforming of methanol. Applied Catalysis
A: General 2005;282(12):1019.
[17] Park HG, Malen JA, Piggott WT, Morse JD, Greif R,
Grigoropoulos CP, et al. Methanol steam reformer on
a silicon wafer. Journal of Microelectromechanical Systems
2006;15(4):97685.
[18] Yuan J, Ren F, Sunden B. Analysis of chemical-reaction-
coupled mass and heat transport phenomena in a methane
reformer duct for PEMFC. International Journal of Heat and
Mass Transfer 2007;50:687701.
[19] Lee MT, Greif R, Grigoropoulos CP, Park HG, Hsu FK.
Transport in packed-bed and wall-coated steam-methanol
reformers. Journal of Power Sources 2007;166:194201.
[20] Schmidt LD. The engineering of chemical reactions. New
York: Oxford University Press; 2005.
[21] Hill CG. An introduction to chemical engineering kinetics
and reactor design. John Wiley and Sons; 1977.
[22] Mills AF. Mass transfer. Upper Saddle River, N.J.: Prentice
Hall; 2001.
[23] Suh JS, Mt Lee, Greif R, Grigoropoulos CP. A study of steam
methanol reforming in a microreactor. Journal of Power
Sources 2007;173:45866.
[24] Mills AF. Heat transfer. Upper Saddle River, N.J.: Prentice
Hall; 1999.
[25] Nield DA, Bejan A. Convection in porous media. New York:
Springer Science-Business Media; 2006.
[26] Wilke CR. A viscosity equation for gas mixtures. Journal of
Chemical Physics 1950;18:5179.
[27] Vargaftik NB. Tables of thermophysical properties of liquids
and gases. New York: McGraw-Hill; 1975.
[28] Incropera FP, Dewitt DP. Fundamentals of heat and mass
transfer. John Wiley and Sons; 2002.
[29] Touloukian YS. Thermal conductivity: nonmetallic solids.
New York: IFI/Plenum; 1970.
[30] Karki KC, Patankar SV. Calculation procedure for viscous
incompressible ows in complex geometries. Numerical
Heat Transfer 1988;14:295307.
[31] Patankar SV. Numerical heat transfer and uid ow.
Washington, DC: Hemisphere; 1980.
[32] Breen JP, Ross JRH. Methanol reforming for fuel-cell
applications: development of zirconia-containing CuZnAl
catalysis. Catalysis Today 1999;51:52133.
[33] Liao C, Erickson PA. Characteristic time as a descriptive
parameter in steam reformation hydrogen production
processes. International Journal of Hydrogen Energy 2008;33:
165260.
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 4 ( 2 0 0 9 ) 3 1 4 3 2 2 322

Anda mungkin juga menyukai