Sie sind auf Seite 1von 10

Chemical Engineering Science 63 (2008) 1761 1770

www.elsevier.com/locate/ces

Optimum washcoat thickness of a monolith reactor for syngas production by


partial oxidation of methane
Michael J. Stutz, Dimos Poulikakos
Department of Mechanical and Process Engineering, Laboratory of Thermodynamics in Emerging Technologies, Institute of Energy Technology,
ETH Zurich, 8092 Zurich, Switzerland

Received 15 August 2007; received in revised form 27 November 2007; accepted 27 November 2007
Available online 4 December 2007

Abstract
In this paper, syngas (hydrogen and carbon monoxide) production was investigated by a numerical model of an adiabatic monolithic reformer
(e.g. for a micro fuel cell system). The study includes the thermal and diffusive properties of a washcoat of nite thickness that is modeled
as a porous layer composed of a ceramic support containing catalytic active rhodium sites. It was combined with a two-dimensional radially
symmetric model of a single tubular mini-channel, considering both the thermal and the diffusive transport phenomena in all domains. It
was found that both the methane conversion and the hydrogen yield depended markedly on the washcoat thickness. An interesting result was
obtained by implementing the common experimental conditions into the numerical model: if the inlet volume ow and the amount of catalyst
per washcoat volume are constant, an optimum washcoat thickness of 70 m was found, where the hydrogen yield is maximal. For a thinner
washcoat, the smaller amount of catalyst is limiting, leading to a low methane conversion. For a thicker washcoat, the limiting effect is the
reduced residence time, which stems inherently from the constraint of constant volume ow, rather than the increased diffusive resistance.
The demonstration of the existence of an optimum washcoat thickness is important, because it can have economic effects due to saving of
the precious catalyst rhodium.
2007 Elsevier Ltd. All rights reserved.

Keywords: Microreactor; Washcoat thickness; Diffusion; Reformer; Syngas; Partial oxidation; Hydrogen; Modeling

1. Introduction mixture. However, in the reformer, not only partial oxidation


(POX) is responsible for H2 production, but also steam reform-
In the last decade, the interest in investigating the reforming ing (SR) that simultaneously takes place with POX (Chaniotis
of hydrocarbons has been strong, also due to research in micro- and Poulikakos, 2005; Stutz and Poulikakos, 2005). Other re-
scale fuel cell systems (Hotz et al., 2007). Methane (CH4 ), action paths also inuence the chemistry inside the reformer.
which is the main component of natural gas, is considered to The most important mechanistic reactions with their enthalpy
be one of the ideal fuels as hydrogen (H2 ) source, because it of formation at standard conditions are listed in Table 1.
is well distributed and abundant. The fuel reformer converts To investigate the fundamental transport phenomena inside
it into syngas (H2 and CO), which is used by the fuel cell a catalytic reactor, numerical simulations with detailed reac-
to produce electric energy. There are different routes for H2 tion schemes were applied (Deutschmann and Schmidt, 1998;
production depending on its further usage. For micro fuel cell Schwiedernoch et al., 2003). Chaniotis and Poulikakos (2005)
systems, catalytic partial oxidation of fuel and dry air is the most investigated the inuence of diffusive and convective transport
promising fuel processing method, because it is simple and no inside the ow channel by comparing the results of an advanced
humidication of the inlet stream is necessary. Furthermore, model, which solves the NavierStokes equation for the uid
Hotz et al. (2007) reported no ashbacks of the ammable inlet ow with simplied models. The effect of wall thermal con-
ductivity on the reforming process was shown to be signicant
(Stutz and Poulikakos, 2005). The micro-reformer was also op-
Corresponding author. Tel.: +41 44 632 2435; fax. +41 44 632 1176. timized by investigating the effects of catalyst loading and the
E-mail address: dimos.poulikakos@ethz.ch (D. Poulikakos). channel geometry (Stutz et al., 2006).
0009-2509/$ - see front matter 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2007.11.032
1762 M.J. Stutz, D. Poulikakos / Chemical Engineering Science 63 (2008) 1761 1770

Table 1 Thus, it is necessary that the characteristic size of the porous


Important mechanistic reactions inside the reformer and their enthalpy of medium is much larger than the size of the elementary volume.
formation at standard temperature and pressure
The convective uid ow inside the porous layer was ne-
HSTP , kJ mol1 glected, due to its very low permeability caused by the small
pores (13 nm) inside the washcoat (Hayes and Kolaczkowski,
Partial oxidation (POX) CH4 + 21 O2 CO + 2H2 35.7
Steam reforming (SR) CH4 + H2 O CO + 3H2 +206.1
1997). The permeability was estimated by the theory of gran-
Water gas shift (WGS) CO + H2 O CO2 + H2 41.2 ular media (Du Plessis and Gray, 1997) and calculated to be
Total oxidation (TOX) CH4 + 2O2 CO2 + 2H2 O 802.3 lower than 3 1017 m2 . The conservatively estimated veloc-
ity of the gaseous mixture inside the washcoat was no more
than 3 nm s1 . The pressure gradient needed for this estima-
tion was obtained from the numerical model without washcoat
Poulikakos and Kazmierczak (1987) showed that a porous by calculating the pressure difference between inlet and outlet
layer at the channel wall has a strong impact on the uid (Stutz and Poulikakos, 2005).
ow and the heat transfer capability of a single channel. The The steady-state conservation equation of energy inside the
conversion rate in catalytic reactors is also affected by the porous layer reduces to
diffusion process in the porous washcoat (Kolaczkowski and
Serbetcioglu, 1996). To investigate catalytic combustion, the q = 0. (1)
washcoat domain was integrated into a numerical model
The effects of heat transport accounts for thermal conduction
(Wanker et al., 2000). Both studies used a single step reaction
and energy transport by species diffusion and reads
mechanism and the diffusivity was modeled as viscous ow.
To include the inuence of washcoat diffusion into a simpler 
NG
model, an effectiveness factor can be employed (Hayes and q = keff T + hi j i . (2)
Kolaczkowski, 1997). However, modeling of the washcoat by i=1
an effectiveness factor is an ad hoc simplication that neglects We neglected the thermal radiation due to the high aspect ratio
differences in the diffusivities of various species and thermal of the channel. Thus, radiation can only affect the vicinity of
effects of the washcoat resulting in a limited model. the source of emission. And the assumption of the gas to be
In this paper the micro-reformer is represented by a radi- opaque is also reasonable (Incropera and DeWitt, 2002).
ally symmetric model. The washcoat is modeled as a porous The effective thermal conductivity in porous media was de-
medium of constant thickness where a detailed model of the termined based on the variational approach using composite
ow channel and a detailed elementary surface reaction mech- spheres for macroscopically homogeneous and isotropic multi-
anism consisting of 38 steps (Schwiedernoch et al., 2003) take phase materials and is written as (Hashin and Shtrikman, 1962)
place. The impact of the washcoat thickness on the reactor per-
formance is documented and discussed in detail. It is important 1
keff = 2ks . (3)
to optimize the thickness of the washcoat because it contains ((1 )/3 ks ) + (/(2 ks + kmix ))
the precious noble metal rhodium, whose amount should be The thermal conductivity of the gaseous mixtures was
minimized for economic reasons. Furthermore, in small sys- computed using the Mix Kinetic Theory (Bird et al., 1960).
tems (e.g. micro fuel cell system) space is scarce. Therefore, a The species conservation equation in the porous washcoat is
high throughput per reactor volume is favorable. dened as

2. Mathematical formulation ji + ri = 0, i [1, 2, . . . , NG ]. (4)


Species ux inside the pores consists only of diffusion, which
In all three domains of interest, the uid channel, the porous is written as
washcoat layer and the solid wall, the conservation equations
of mass, momentum, energy and species can be used as needed ji = Di,eff i . (5)
to model the physics of transport phenomena. This leads to a Diffusion inside the pores was modeled with the Knudsen dif-
set of nonlinear partial differential equations thatat were solved fusion coefcient, because the Knudsen number of the pores
numerically. The inlet Reynolds number was varied between 6 inside the washcoat is around 7 and higher. The diffusion
and 17 for all cases, meaning that the ow inside the channel coefcient (Cussler, 1997) is calculated as
is laminar. The reforming process was modeled at steady-state 
conditions.  8RT
The governing equations of the mathematical model of the Di,eff = dpore . (6)
3 Mi
uid channel and the solid wall are identical to those of a
previous study by Stutz and Poulikakos (2005) and therefore assuming no pore roughness and therefore  = 1.
not repeated. The thermal conductivity of the channel wall is k =
For the porous washcoat, the method of local volume aver- 2.76 W m1 K 1 , which is the value of cordierite, a typical
aging that spatially smoothes the conservation equations over a monolith material. It was assumed to be temperature indepen-
representative elementary volume was applied (Kaviany, 1995). dent. For the sake of simplicity, the thermal conductivity of
M.J. Stutz, D. Poulikakos / Chemical Engineering Science 63 (2008) 1761 1770 1763

the solid material of the washcoat was set to the same value. operating temperatures of a CH4 reformer (Deutschmann and
This is reasonable, because the washcoat consists mainly of a Schmidt, 1998). Since homogeneous reactions only become
catalyst support that is in most applications a ceramic. In this important for high pressure applications and channels with a
study, the washcoat was modeled as a porous layer approxi- larger diameter they were neglected.
mated as a granular medium with a catalytically active surface. The surface reactions in this study were modeled with
Measurements of the pore size of a washcoated monolith usu- a multi-step nite-rate reaction mechanism involving both
ally show a bimodal pore size distribution with peak values of gaseous and surface-adsorbed species. Expressions for the
the pore diameters of 10 nm and 3.5 m. Because the larger reaction rate, adsorption process and the reaction rate coef-
pores belong to the cordierite, which is the material of the cients can be found in a previous study (Stutz and Poulikakos,
monolith matrix, they were not taken into account (Hayes and 2005). Inside the washcoat, the mass ux balance of reacting
Kolaczkowski, 1997). The porosity of the catalyst layer was re- species at the catalyst surface for the gas-phase species i is
ported to be around  = 0.5 (Beeckman, 1991). The pore size written in discretized form as
was calculated with the theory of granular media (Du Plessis Nsteps
and Gray, 1997) and reads  i P ,i
Mi (aij aij )Sj = Di . (10)
1/3

dpore = dparticle ((1 ) 1). (7) j =1

For the surface-adsorbed species i the mass ux balance is


The particle size of the ceramic support was assumed to be
expressed as
dparticle = 50 nm, leading to a pore size of dpore = 13 nm, which
corresponds well with reported pore sizes of around 10 nm Nsteps

 
(Hayes and Kolaczkowski, 1997). The tortuosity was calcu- (bij bij )Sj = 0. (11)
lated according to the theory of granular media (Du Plessis and j =1
Gray, 1997) as
Thus, the mass ux of reacting species to the surface equals
 = /(1 (1 )2/3 ). (8) the rate at which the species is consumed by the reaction on
the surface.
We chose deliberately relatively simple correlations for tortuos- The reaction mechanism employed in this study consisted
ity, effective diffusivity and pore size because they are derived of 38 elementary reaction steps, 7 gas-phase species and 12
rather by mathematical and physical theories than by tting surface-adsorbed site species (Schwiedernoch et al., 2003). The
of experiments. The values of effective diffusivity reported in inlet uid was CH4 and dry air that reacted on a rhodium coated
the literature differ up to one order of magnitude, which has surface. Deutschmann et al. (2001) successfully validated the
been explained by different washcoat structures mainly due reaction mechanism with experiments.
to different preparation methods (Hayes et al., 2000; Zhang
et al., 2004). However, the diffusion coefcients used in this 3. Numerical solution
study lie in between the reported results. Therefore, the sim-
ple correlations lead to an accurate description of the diffu- 3.1. Models and boundary conditions
sion inside the washcoat, which is the only species transport
mechanism, since convection in the porous layer was neglected The computational domain employed in this study consisted
herein. of a single cylindrical channel, coated with a catalytically active
The source term ri in Eq. (4) accounts for the catalytic re- substance containing rhodium. It is a good representation of a
action inside the washcoat. It was balanced with the diffusive monolith reactor, which is a bundle of identical parallel straight
ux at the surface of the catalyst. The balance equation is given channels (Fig. 1). The monolith reactor is often employed as
in discretized form by a catalytic reactor due to low fabrication cost, high surface to
volume ratio and relatively low pressure drop. Inlet and outlet
i P ,i
ri = Di,eff SV ,eff . (9) conditions of the monolith reactor were assumed to be uniform
 and the outer boundary was assumed to be adiabatic. Therefore,
This includes the difference of mass fraction P ,i of species the conditions in each channel are identical considering that
i in the pore uid and the mass fraction i of species i at the the monolith reactor was approximated as radially symmetric,
catalyst interface. The diffusion length scale  was set equal which is reasonable because the number of channels is large.
to the pore diameter dpore in order to account for the diffusive Although there may be a small heat leak to the surroundings,
path of the reacting species from the pore uid to the pore it is a reasonable approximation to model the outer wall as
wall. The effective surface to volume ratio SV ,eff is the ratio adiabatic.
of catalytically active surface to the volume of the washcoat. Even when the real geometry of a channel inside a mono-
It is a true representation of catalyst loading (Mazumder and lith reactor is not exactly radially symmetric, it is reasonable
Sengupta, 2002). to approximate the channel as radially symmetric due to the
Species conservation of the individual gas-phase components following reasons. First, the washcoat lls up the edges of the
in the ow channel neglects any source terms, because no ho- channels, which makes the approximation of a cylindrical ow
mogeneous reactions occur at atmospheric pressure and usual channel realistic (Fig. 1). Second, it was shown that the axial
1764 M.J. Stutz, D. Poulikakos / Chemical Engineering Science 63 (2008) 1761 1770

wall, only the energy equation was solved. The coupling of the
two domains is treated by a conjugate heat transfer analysis,
where the energy ux in the solid at the interface is matched
to that on the uid side at the interface. For the FWT model,
the continuity, species, momentum and energy conservation
were solved in the channel. Inside the washcoat, the species
and energy conservation were solved, since convection was ne-
glected, and inside the wall, only heat conduction was taken
into account. At the channel/washcoat interface, a conjugate
Fig. 1. Sketch of a monolith reactor with washcoated channels. heat transfer analysis was applied together with a species ux
balance, and the no slip condition was implemented for the
momentum and continuity equations. At the washcoat/wall in-
terface, the conduction heat ux was balanced and the species
and mass uxes were set to zero.
The channel wall in the numerical model represented the
catalytically inactive part of the monolith reactor with a half-
thickness of twall =75 m , which is also a typical realistic value
(Hayes and Kolaczkowski, 1997).
The ow entered the entrance region of the reactor with a
uniform inlet velocity, which was for the ZWT models uin =
1 m s1 . This entrance region was needed in order to have,
especially for the FWT models, a more stable numerical so-
lution and faster convergence and to account for upstream
diffusion near the beginning of the catalytic active part of the
reactor.
The inlet temperature T =850 K was constant for all models.
The channel diameter is dened as two times the distance from
the centerline to the inner area of the channel wall. The typical
tube diameter of a monolith reactor was equal to d=1 mm for all
models. The pressure at the outlet was constant 1 bar. The inlet
composition of the gaseous mixture was dened by specifying
the equivalence ratio based on the POX reaction (Chaniotis and
Fig. 2. The computational domains and the boundary conditions of (a) the Poulikakos, 2005). It was constant for all investigated models
Zero Washcoat thickness (ZWT) and (b) the Finite Washcoat thickness (FWT)
( = 0.8).
models.
Four different types of constraints were investigated in this
study for the FWT models, each having two different cases,
distribution of the species inside the ow channel alter the com- one with a high catalyst amount/density (denoted as A) and one
plex endo-/exothermic reaction mechanism leading to a signif- with a low catalyst mount/density (denoted as B). The thickness
icant variation of reactor performance (Stutz and Poulikakos, of the washcoat was then varied for all cases, resulting in a total
2005). of eight FWT cases with seven different washcoat thicknesses
Two different models were investigated in this study. The rst (Table 2).
completely neglected the physical thickness of the washcoat FWT 1 simulated the conditions most commonly used in
(Zero Washcoat thickness (ZWT) model) and contained only practical applications: The monolith geometry was given, the
two computational domains, namely the ow channel and the inlet volume ow was constant and the volumetric density of
thermal conductive wall (Fig. 2a). The catalytic reaction took catalyst was constant. The volume ow of the FWT 1 cases was
place at the interface between wall and channel. The second chosen in such a way that it equaled that of the ZWT model.
model accounted for the physical presence and nite thickness FWT 2 is practically also easily applicable. It had the same
of the washcoat (Finite Washcoat thickness (FWT) model), and constraints as the rst, but the volume ow was not constant.
consisted of three computational domains (Fig. 2b). The rst In this case, the inlet velocity and hence the GSV was constant
domain was the ow channel, the second the thermal conductive and set equal to that of the ZWT model.
channel wall and the third the washcoat. The heterogeneous FWT 3 and FWT 4 were more of theoretical importance and
reaction in the FWT model took place in the washcoat volume. were investigated in order to have a fair comparison between
The washcoat boundaries at the inlet and outlet did not permit the FWT and the ZWT models. The inlet volume ow of the
species ux, because the convective ow inside the washcoat FWT 3 cases (that was equal to that of the ZWT model) and the
is negligible, as discussed earlier. geometry of the monolith were constant. However, the amount
For the ZWT model, the continuity, species, momentum and of catalyst was also set constant, which is difcult to handle
energy conservations were solved in the channel. Inside the in practical applications, leading to a varying catalyst volume
M.J. Stutz, D. Poulikakos / Chemical Engineering Science 63 (2008) 1761 1770 1765

Table 2
Numerical models used in this study with their different constraints

uin (m s1 ) V (ml s1 ) ncat 1010 (mol) ncat /Vcat (mol m3 ) GSV (s1 ) CSV (m3 mol1 s1 )

FWT 1A 1.022.78 0.785 0.9028.8 0.57 102278 8775273


FWT 1B 1.022.78 0.785 0.278.55 0.17 102278 29 560919
FWT 2A 1 0.7700.283 0.9028.8 0.57 100 860198
FWT 2B 1 0.7700.283 0.278.55 0.17 100 28 971331
FWT 3A 1.022.78 0.785 8.55 5.470.17 102278 919
FWT 3B 1.022.78 0.785 2.54 1.620.05 102278 3096
FWT 4A 1 0.7700.283 8.55 5.470.17 100 901331
FWT 4B 1 0.7700.283 2.54 1.620.05 100 30341115
ZWT 1A 1 0.785 8.55 100 919
ZWT 1B 1 0.785 2.54 100 3096

density for a different washcoat thickness. We chose these con- mass. More meaningfully, it can be dened in terms of number
straints in order to have a constant CSV. of accessible, catalytically active sites ncat . Thus, in this study
FWT 4 dealt with constant catalyst amount and a given the CSV is dened as
geometry. The inlet velocity was constant (and equaled that of
the ZWT model) leading to a varying volume ow but a con- CSV = Vin /ncat . (17)
stant GSV, what was intended.
CH4 conversion is an important characteristic for CH4
3.2. Solution method, grid independence
reforming and is dened as the ratio between converted CH4
and the inlet ux of CH4 . It reads
In this study the coupled set of partial differential equations
nCH4 ,in nCH4 was solved with the licensed solver CFD-ACE distributed by

= . (12)
nCH4 ,in ESI US R&D Inc. with a convergence criterion of 1 105 .
The solver was the same as that of a previous study where its
Oxygen (O2 ) conversion is dened in an analogous manner. The details were presented (Stutz and Poulikakos, 2005).
performance of the reactor with respect to the H2 production Grid independence of the FWT models was checked with the
is characterized by the H2 yield. It is the ratio of the produced FWT 2B model with a washcoat thickness twc 50 m. It was
H2 to the theoretical maximum amount of H2 reached with 11 307 cells. The contour plots of the temperature
nH2 and of the H2 mole fraction were compared to those of the
H2 = . (13) FWT model with 17 906 cells and showed no visible differences
2nCH4 ,in
for the washcoat, the channel wall and the channel domain.
CO yield is dened in an analogous manner and reads Comparing the ne with the coarse model, the H2 yield differed
only by 0.71% at the channel outlet. CH4 conversion at the
nCO
CO = . (14) outlet differed only by 0.41% and the temperature at the outlet
nCH4,in was 0.02% higher for the ne model than for the coarse model.
The selectivity of CO toward unwanted CO2 is dened as For the ZWT models grid independence was checked with the
ZWT 1 model. It is reached with 6198 cells. H2 mole fraction
nCO and temperature were compared with those of the model with
SCO = . (15)
nCO + nCO2 19 864 cells. No visible difference could be found in the contour
It was shown that in catalytic reactors, both the gas space ve- plots. H2 yield at the channel outlet differed only by 0.03%,
locity (GSV) and the catalyst space velocity (CSV) are impor- CH4 conversion only by 0.02% and the temperature at the outlet
tant evaluation criteria (Stutz et al., 2006). The GSV is dened only by 0.01%.
as the ratio of volumetric ow through the reactor inlet to the
channel volume (without washcoat volume) and reads 4. Results and discussion
GSV = Vin /Vchannel . (16)
The detection of the optimum washcoat thickness can help
It equals the often used gas hour space velocity (GHSV) if save a large amount of the noble metal rhodium, which is even
converted to SI units. more expensive than platinum.
The CSV is often used in experimental research of catalytic H2 yield dependence on the washcoat thickness twc shows
reactions and is dened as the ratio of volumetric ow through an interesting result: It features a maximum at twc = 70 m for
the reactor inlet to the amount of catalyst inside the washcoat both FWT 1 cases (Fig. 3a). Its value is =70.4% for FWT 1A.
(although it is sometimes misleadingly named GHSV). The That for FWT 1B is signicantly lower, which is explained by
amount of catalyst in experiments is often dened in terms of the larger amount of catalytically active sites of the FWT 1A.
1766 M.J. Stutz, D. Poulikakos / Chemical Engineering Science 63 (2008) 1761 1770

Fig. 3. H2 yield of (a) the FWT 1, FWT 2, (b) FWT 3 and FWT 4 models dependent on washcoat thickness. represents the ZWT 1A and the ZWT 2B case.

The thinnest washcoat leads to the lowest H2 yield. It can The results of the case FWT 3A show a completely different
be associated with the decrease of catalytically active sites behavior (Fig. 3b). The H2 yield decreases monotonically with
with decreasing washcoat thickness. The decrease of H2 yield increasing washcoat thickness. The outcome of FWT 3B is
with increasing washcoat thickness for twc > 70 m can have similar, however, with lower values. Compared to the FWT 1
different reasons. The diffusive resistance is high for a thick and FWT 2 models, the increase of washcoat thickness has a
washcoat; hence the catalytically active sites deep within the negative effect on the performance in any case. It is shown
washcoat are difcult to access from the uid channel. Another below that the diffusive resistance is the main reason for this
inuence stems from the varying interface area between wash- result if the washcoat is thin. For a thick washcoat, it is the
coat and uid channel with varying washcoat thickness. Thus, effect of GSV that prevails.
a thick washcoat has a smaller entrance area for the species The FWT 3 and FWT 4 models are similar for a small wash-
to enter the washcoat. These two effects are combined in the coat and therefore the results follow similar trends. For the
characteristic length of the washcoat and reads FWT 3 cases, there is no effect that enhances the performance
Vwc for an increasing washcoat thickness leading to a monotonic
Lwc = . (18) decrease of H2 yield (Fig. 3b). On the other hand, for the FWT
Auid.wc
4 cases, the effects that increase H2 yield must prevail for a
GSV varies with the washcoat thickness, due to the constant thick washcoat, because the FWT 4 cases reach a minimum at
volume ow constraint. Thus, the diffusion inside the gas phase twc = 100 m. Thus, there is a different mechanism that com-
can also become limiting. The results of FWT 2 show that the petes with the diffusive resistance. It is the CSV, describing
latter effect is responsible for the low H2 yield of the FWT the ratio of inlet volume ow to the available catalytic sites,
1 models with a thick washcoat. This conclusion is discussed which decreases for increasing washcoat thickness. This is un-
below. derpinned by the results of the FWT 3 cases where the CSV is
The H2 yield of the FWT 2 cases show a similar result to constant.
those of FWT 1 for a small washcoat twc 20 m). This is The conversion trend of CH4 in Fig. 4 is similar to that
reasonable, because the constraints of FWT 1 and FWT 2 are of H2 yield. The FWT 1 cases exhibit a maximum of
=
similar for a small twc . On the other hand, for increasing wash- 83.1% at twc = 50 m (FWT 1A). The FWT 2 cases increase
coat thicknesses, the curves deviate. The H2 yield of the FWT monotonically. The results of the FWT 3 and FWT 4 cases also
2 cases increases monotonically to a maximum of = 86.5% show a similar shape to that of H2 yield.
at twc = 200 m for the FWT 2A case. Hence, it is clear that Figs. 3 and 4 indicate that the FWT 3 and FWT 4 cases ap-
the main reason for the decrease of H2 yield of FWT 1 is the proach the ZWT model if the washcoat thickness is reduced
GSV (or residence time) and thus diffusion in the gas phase. continuously. However, it is visible that they do not converge
By further increasing the washcoat thickness, we also expected completely to the ZWT model. This behavior was conrmed
a decrease in H2 yield for the FWT 2 models, because the by additional computations not shown here, reducing the wash-
diffusive resistance, represented by the characteristic length coat thickness down to 1.25 m, although results for such thin
Lwc , rises drastically. It is already large (LC = 0.53 mm) for washcoats are not correct due to the numerical constraints of
a twc = 200 m. For an even thicker washcoat (twc = 300 m) the volume averaging method (Whitaker, 1999).
it becomes very large LC = 1.1 mm) and the numerical solver The reason for that is the different formulation of the diffu-
was not able to produce a converging solution. sion process. In the FWT model, the reacting species rst have
M.J. Stutz, D. Poulikakos / Chemical Engineering Science 63 (2008) 1761 1770 1767

Fig. 4. CH4 conversion of (a) the FWT 1, FWT 2, (b) FWT 3 and FWT 4 models dependent on washcoat thickness. represents the ZWT 1A and the
ZWT 2B case.

to diffuse into the porous layer (Eq. (5)) and then from the
pore uid to the reacting wall described by Eq. (9). The latter
process employs an additional parameter, the diffusion length
scale , which was not needed in the ZWT model. Although the
washcoat thickness of the FWT model was continuously low-
ered, the diffusive resistance from the pore to the catalytically
active surface was always present. This result was supported
by the comparison of the A to the B models. For the A models,
the reaction kinetics is dominant, leading to a good approxima-
tion of the FWT to the ZWT model. If the diffusion becomes
more important, (i.e. for the B models) the approximation is
weaker, as seen in Figs. 3b and 4b.
A surprising result is the dependence of the CO selectivity
on the catalyst density: the low density cases show a higher
CO selectivity than the respective high density cases for the
FWT 1 and FWT 2 cases (Fig. 5). CO can also be used as a
fuel in the SOFC, either by a direct electrochemical reaction
of CO or by the WGS reaction that produces H2 , which is then
electrochemically used. It is known that a high CO content in
the fuel can be advantageous at relatively low temperatures of
the fuel cell (Inui et al., 2006).
The CO selectivity features a minimum for both FWT 1 cases
(Fig. 5). For the FWT 2 cases, it decreases monotonically with
Fig. 5. CO selectivity of the FWT 1 and FWT 2 models dependent on
increasing washcoat thickness, which was also unexpected. An washcoat thickness.
investigation of the CO yield (not shown here for brevity) shows
that the difference in CH4 conversion outweighs the differences
in CO selectivity leading to a similar picture to that of H2 yield weak (Fig. 6a). By increasing the thickness of the porous layer,
or CH4 conversion. It is worth note that the dependence of the the CH4 mole fraction becomes gradually more dependent on
CO yield on the washcoat thickness is weaker compared to that the radial position (e.g. twc = 50 m ), hence CH4 decomposi-
of H2 yield. tion. For a thick washcoat (e.g. twc = 200 m) the characteristic
In order to detect the regions of the washcoat that are active length LC becomes high, leading to isolines following the
or unused, the individual components were investigated further x-axis near the washcoatchannel interface and to isolines
in detail. Furthermore, it is of importance to locate the dominant following the r-coordinate near the washcoatwall interface.
reaction steps inside the washcoat. Hence, a large amount of the catalyst cannot be accessed
For a small washcoat thickness (e.g. twc = 5 m), the CH4 sufciently by the CH4 . The thin washcoats of FWT 1A and
mole fraction does not vary signicantly in the radial direction FWT 2A (Fig. 6b) have a nearly identical distribution of CH4
for all types of models, hence the diffusion inuence is rather because their constraints are similar and feature no regions of
1768 M.J. Stutz, D. Poulikakos / Chemical Engineering Science 63 (2008) 1761 1770

Fig. 6. CH4 mole fraction inside the washcoat for (a) the FWT 1A and (b) for the FWT 3A cases with a washcoat thickness of twc = 5, 50 and 200 m
(from top to bottom). The interface between channel wall and washcoat thickness lies at the upper end and that between washcoat and uid channel lies at
the lower end of each plot.

Fig. 7. (a) O2 and (b) H2 mole fraction inside the washcoat for the FWT 1A cases with a washcoat thickness of twc = 5, twc = 50 and 200 m (from top to
bottom).

low CH4 mole fraction. However, the resulting CH4 conversion The O2 mole fraction is of interest because it acts as a re-
is low, because there are not enough catalytic sites present to actant for the POX and the TOX reaction. It varies notably
process the CH4 . in radial direction even for the thinnest computed washcoat
Based on the above, two conditions can be formulated that (Fig. 7a). The thicker the washcoat, the stronger is its radial
result in a high CH4 conversion. First, there must be enough variation. The diffusion process of O2 is slow due to its rela-
catalytically active sites existent inside the washcoat and, sec- tively high molecular weight (MO2 = 32.0 kg kmol1 ). This is
ond, these sites have to be accessible within the residence time the reason for the stronger radial dependence of O2 compared
of each species. to that of CH4 . For even thicker washcoats (e.g. twc = 200 m),
These two conditions also explain the difference in perfor- the isolines of O2 mole fraction are nearly following the x-axis
mance of thick washcoats of the FWT 1 and FWT 2 cases. underlining the slow transport mechanism of O2 .
They feature a large region that is nearly unpenetrated by CH4 Because the decomposition process of O2 on rhodium is fast,
and therefore unused. The FWT 2A, on the other hand, has a large part of the washcoat contains no O2 molecules. The O2
a smaller GSV and additionally a smaller inlet volume ow. decient region was dened as the location inside the reformer
Therefore, the accessible volume of the washcoat is larger for at the catalytic active surface, where hardly any O2 is left to feed
this case and it has to cope with less CH4 molecules. the exothermic O2 consuming reaction (Stutz and Poulikakos,
M.J. Stutz, D. Poulikakos / Chemical Engineering Science 63 (2008) 1761 1770 1769

2005). It is dened (ad hoc) in this study as the location where computational cost is signicantly lower and less parameters are
the O2 mole fraction is below a mole fraction of 0.001, which needed.
corresponds to an O2 conversion of around 99.4%.
For a thin washcoat this region is far downstream or non- Notation
existent. For a thicker washcoat (e.g. twc = 50 m) this volume
a gas-phase species stoichiometric coefcient
is larger. Thus, a large part of the washcoat is not used for the
Auid.wc interfacial area between channel and washcoat
oxidation of CH4 . Hence, in the entire volume where CH4 is
b surface-adsorbed species stoichiometric coef-
still existent but O2 is decient, H2 is produced by SR.
cient
The H2 mole fraction does not vary signicantly in the radial
CSV catalyst space velocity, m3 mol1 s1
direction and is well distributed for a small washcoat thickness
d diameter, m
(twc = 5 m , Fig. 7b). This is connected to the fact that the
Di multicomponent diffusivity of species i, m2 s1
H2 transport inside the porous layer is fast, due to its low dif-
fusive resistance (MH2 = 2.0 kg kmol1 ). The H2 yield is low GSV gas space velocity, s1
due to the low conversion of methane. For a thicker washcoat h specic enthalpy, J kg1
twc =50 m) H2 is also distributed everywhere inside the wash- j mass ux, kg m2 s1
coat. Near the inlet, close to the uidwashcoat interface, H2 is k thermal conductivity, W m1 K 1
produced by POX. Farther downstream and deeper inside the Lwc characteristic length of washcoat, m
washcoat in the O2 decient region, H2 is produced by SR. M molecular weight, kg kmol1
Thus, the entire washcoat is active leading to a high H2 yield. n mole ow, kmol s1
For a thick washcoat (twc = 200 m), H2 is again distributed NG total number of gas-phase species
everywhere inside the washcoat, although it is only produced Nsteps total number of reaction steps
near the uidwashcoat interface because CH4 is non-existent ncat total number of catalytic active sites, kmol
deep within the washcoat. q heat ux, W m2
r species production rate, kg m3 s1
5. Conclusions R universal gas constant, 8.314510 J K 1 mol1
S selectivity
The effect of washcoat thickness on reforming performance S reaction/adsorption rate, mol m2 s1
of a monolith reactor was investigated. Two different axisym- SV surface to volume ratio,m1
metric models were employed, a limited model consisting only twc washcoat thickness, m
of the ow channel and the channel wall and a more complete Vchannel channel volume, m3
model consisting of an additional domain, the porous washcoat. V volume ow, m3 s1
In the limited model, the catalytic reaction takes place at the xi mole fraction of species i
interface between channel and channel wall. In the extended
Greek letters
model, the surface reaction occurs also inside the washcoat. A
detailed surface reaction mechanism described the reaction of  diffusion length scale, m
CH4 and dry air over rhodium catalyst.  porosity
Modeling the thermal and diffusive transport inside the wash-
CH4 conversion
coat proved to be important in predicting correctly the reactor  pore roughness
performance. Dependent on washcoat thickness, the CH4 con-  tortuosity
version was shown to vary up to 48% and H2 yield up to 62%. equivalence ratio
An interesting result was obtained by simulating the com- H2 yield
mon experimental conditions by the numerical model: if the  mass fraction
inlet volume ow and the amount of catalyst per washcoat vol-
Indices
ume are constant, an optimum washcoat thickness of 70 m
was found. For that washcoat, H2 yield, both CH4 conversion  reactant
and CO yield are maximal. If the washcoat is thinner, all the  product
catalytically active sites are accessible but the amount is not eff effective
sufcient to process the inowing reactants. This leads to a low in inlet
CH4 conversion, hence low H2 yield. If the washcoat is thicker, mix mixture
the amount of catalytically active sites is sufcient but not all P pore uid
of them are accessible due to the reduced residence time, which s solid
comes from the constraint of constant volume ow. Thus, both wc washcoat
CH4 conversion and H2 yield are lower in this case.
The FWT model proved to be powerful in determining Acknowledgments
trends and performing optimization with respect to washcoat
parameters. However, the ZWT model can still be useful We gratefully acknowledge the nancial support by the fol-
for rough estimations of reactor performance, because its lowing Swiss Institutions: Commission for Technology and
1770 M.J. Stutz, D. Poulikakos / Chemical Engineering Science 63 (2008) 1761 1770

Innovation (CTI) (Grants: KTI 7085.2 DCPP-NM and KTI Incropera, F.P., De Witt, D.P., 2002. Fundamentals of Heat and Mass Transfer.
8446.1 DCPP-NM), Center of Competence Energy and Mobil- Wiley, New York.
ity (CCEM) and Swiss Electric Research. Inui, Y., Urata, A., Ito, N., Nakajima, T., Tanaka, T., 2006. Performance
simulation of planar SOFC using mixed hydrogen and carbon monoxide
gases as fuel. Energy Conversion and Management 47, 17381747.
References
Kaviany, M., 1995. Principles of Heat Transfer in Porous Media. Springer,
New York.
Beeckman, J.W., 1991. Measurement of the effective diffusion-coefcient Kolaczkowski, S.T., Serbetcioglu, S., 1996. Development of combustion
of nitrogen monoxide through porous monolith-type ceramic catalysts. catalysts for monolith reactors: a consideration of transport limitations.
Industrial & Engineering Chemistry Research 30, 428430. Applied Catalysis A-General 138, 199214.
Bird, R.B., Stewart, W.E., Lightfoot, E.N., 1960. Transport Phenomena. Wiley, Mazumder, S., Sengupta, D., 2002. Sub-grid scale modeling of heterogeneous
New York. chemical reactions and transport in full-scale catalytic converters.
Chaniotis, A.K., Poulikakos, D., 2005. Modeling and optimization of catalytic Combustion and Flame 131, 8597.
partial oxidation methane reforming for fuel cells. Journal of Power Sources Poulikakos, D., Kazmierczak, M., 1987. Forced-convection in a duct partially
142, 184193. lled with a porous material. Journal of Heat Transfer-Transactions of the
Cussler, E.L., 1997. Mass Transfer in Fluid Systems. Cambridge University ASME 109, 653662.
Press, Cambridge. Schwiedernoch, R., Tischer, S., Correa, C., Deutschmann, O., 2003.
Deutschmann, O., Schmidt, L.D., 1998. Modeling the partial oxidation of Experimental and numerical study on the transient behavior of partial
methane in a short-contact-time reactor. A.I.Ch.E. Journal 44, 24652477. oxidation of methane in a catalytic, monolith. Chemical Engineering
Deutschmann, O., Schwiedernoch, R., Maier, L.I., Chatterjee, D., 2001. Science 58, 633642.
Natural gas conversion in monolithic catalysts: interaction of chemical Stutz, M.J., Poulikakos, D., 2005. Effects of microreactor wall heat conduction
reactions and transport phenomena. In: Natural Gas Conversion VI, Studies on the reforming process of methane. Chemical Engineering Science 60,
in Surface Science and Catalysis.Elsevier, Amsterdam, pp. 251258. 69836997.
Du Plessis, J.P., Gray, W.G., 1997. Pore-scale modelling of interstitial transport Stutz, M.J., Hotz, N., Poulikakos, D., 2006. Optimization of methane
phenomena. In: Fluid Transport in Porous media.Computational Mechanics reforming in a microreactoreffects of catalyst loading and geometry.
Publication, Southampton, pp. 61104. Chemical Engineering Science 61, 40274040.
Hashin, Z., Shtrikman, S., 1962. A variational approach to theory of effective Wanker, R., Raupenstrauch, H., Staudinger, G., 2000. A fully distributed
magnetic permeability of multiphase materials. Journal of Applied Physics model for the simulation of a catalytic combustor. Chemical Engineering
33, 31253131. Science 55, 47094718.
Hayes, R.E., Kolaczkowski, S.T., 1997. Introduction to Catalytic Combustion. Whitaker, S., 1999. The Method of Volume Averaging. Kluwer Academic
Gordon and Breach Science Publishers, Amsterdam. Publishers, Dordrecht.
Hayes, R.E., Kolaczkowski, S.T., Li, P.K.C., Awdry, S., 2000. Evaluating the Zhang, F., Hayes, R.E., Kolaczkowski, S.T., 2004. A new technique to
effective diffusivity of methane in the washcoat of a honeycomb monolith. measure the effective diffusivity in a catalytic monolith washcoat. Chemical
Applied Catalysis B-Environmental 25, 93104. Engineering Research & Design 82, 481489.
Hotz, N., Stutz, M.J., Loher, S., Stark, W.J., Poulikakos, D., 2007. Syngas
production from butane using a ame-made Rh/Ce0.5 Zr0.5 O2 catalyst.
Applied Catalysis B: Environmental 73, 336344.

Das könnte Ihnen auch gefallen