Sie sind auf Seite 1von 10

J.-M.

Klein1 Modeling of a Solid Oxide Fuel


e-mail: jean-marie.klein@lepmi.inpg.fr

Y. Bultel
Cell Fueled by Methane: Analysis
e-mail: yann.bultel@lepmi.inpg.fr

Laboratoire d’Electrochimie et de Physico-Chimie


of Carbon Deposition
des Matériaux et des Interfaces (LEPMI), Natural gas appears to be a fuel of great interest for solid oxide fuel cell (SOFC) systems.
UMR 5631 CNRS-INPG-UJF, It mainly consists of methane, which can be converted into hydrogen by direct internal
ENSEEG, reforming (DIR) within the SOFC anode. However, a major limitation to DIR is carbon
BP 75, 38402 Saint Martin d’Hères, formation within the ceramic layers at intermediate temperatures. This paper proposes a
France model solution using the CFD-ACE software package to simulate the behavior of a tubular
SOFC. A detailed thermodynamic analysis is carried out to predict the boundary of
carbon formation for SOFCs fueled by methane. Thermodynamic equilibrium calcula-
M. Pons tions that take into account Boudouard and methane cracking reactions allow us to
Laboratoire de Thermodynamique et de investigate the occurrence of carbon formation. This possibility is discussed from the
Physicochimie Métallurgique (LTPCM),
values of driving forces for carbon deposition defined as ␣ = PCO2 / 共KB PCO 2
兲 and ␤
UMR 5614 CNRS-INPG-UJF,
ENSEEG, = PH / 共KC PCH4兲, from the equilibrium constants KB and KC of the Boudouard and crack-
2
2
BP 75, 38402 Saint Martin d’Hères, ing reactions, and from the partial pressure Pi of species i. Simulations allow the calcu-
France lation of the distributions of partial pressures for all the gas species (CH4, H2, CO, CO2,
and H2O), current densities, and potentials of both electronic and ionic phases within the
anode part (i.e., gas channel and Cermet anode). Finally, a mapping of ␣ and ␤ values
P. Ozil enables us to predict the predominant zones where carbon formation is favorable (␣ or
Laboratoire d’Electrochimie et de Physico-Chimie ␤ ⬍ 1) or unfavorable (␣ or ␤ ⬎ 1) according to the calculation based on thermodynamic
des Matériaux et des Interfaces (LEPMI), equilibrium. With regard to the values of these different coefficients, we can say that a
UMR 5631 CNRS-INPG-UJF, carbon formation can be supposed for temperature less than 800° C and for ratios
ENSEEG, xH2O / xCH4 smaller than 1. 关DOI: 10.1115/1.2759504兴
BP 75, 38402 Saint Martin d’Hères,
France Keywords: SOFC, modeling, simulation, direct internal reforming, carbon formation

1 Introduction cracking and Boudouard reactions have to be avoided because


carbon deposition leads to a decrease in electrode performance.
The increasing and irreversible consumption of fossil matters
Recent literature shows a significant research and development
makes urgent the gradual replacement of fossil energies by effort focusing on the DIR modeling within the SOFC anode.
sources of renewable energy. To this end, fuel cells appear to offer Lehnert et al. 关3兴 proposed one-dimensional simulations based on
an alternative solution of interest as power generators. In fact, a geometry, integrating an anode supported by a planar substrate
they directly convert chemical energy into electric power with a and taking into account mass transport. Multicomponent transport
high energetic efficiency and a low pollution level from the ex- was described owing to the mean transport pore model 共MTPM兲
haust gases. based on three microstructure parameters 共porosity, tortuosity, and
Solid oxide fuel cells 共SOFCs兲 are promising candidates for mean pore radii兲. Special attention was paid by Suwanwarangkul
power generation while preserving the environment. Their high et al. 关4兴 to detail the concentration polarization with different
operating temperature 共973– 1273 K兲 allows a real flexibility for models 共Fick’s law, dusty-gas model, and Stefan-Maxwell model兲
feeding fuel 共fossil fuel, natural gas, methanol, ethanol, and par- for a H2 / H2O / CO / CO2 mixture. The MTPM model has been
ticularly, biogas that contains methane兲, but methane remains to extended by Ackmann et al. 关5兴 into a two-dimensional thermal
be the most common fuel for SOFC systems. Moreover, direct approach, including electrochemical and chemical reactions as
internal reforming 共DIR兲 within the SOFC anode 关1,2兴 allows the heat sources. More recently, some works offered models of a
conversion of methane into hydrogen without using a separate SOFC from a detailed electrochemical analysis and a fluid-
reformer. Such a concept is convenient for high-temperature fuel dynamic calculation of internal heat transfer conditions using the
cells, wherein the steam reforming reaction can be sustained with finite volume method 关6–8兴. Finally, some general models have
catalysts. The reforming reaction at the anode and the shift reac- been developed, taking into account velocity fields of air and fuel
tion are carried out over a supported catalyst such as nickel. These flows, heat generation by Ohmic and thermodynamic effects, and
convective heat transfer. Other works focused on the mass transfer
reactions provide the system with the dihydrogen required by the
of acting chemical species, including electrochemical processes
electrochemical reaction. The SOFC internal reforming is thus
and the distribution of electric potential within electrodes and
designed by closely coupling reforming and electrochemical oxi-
electrolytes 关9,10兴. All these works assumed that steam reforming
dation reactions within the fuel cell. However, the DIR method and water gas–shift reactions take place within the porous anode,
has to deal with two limitations: thermal stress induced by differ- whereas electrochemical oxidations of hydrogen and carbon mon-
ent dilatations of the ceramic layers and carbon formation at in- oxide occur close to the electrolyte/anode interface.
termediate temperatures. During the conversion process, methane The present study proposes a model of SOFC, including a
model of gas diffusion electrode to simulate the behavior of a
1
Corresponding author.
tubular SOFC, using the CFD-ACE software package. A realistic
Manuscript received November 29, 2005; final manuscript received May 30, triple phase boundary distribution is considered through the po-
2006. Review conducted by Roberto Bove. rous electrode thickness. The electrochemical reactions within the

Journal of Fuel Cell Science and Technology NOVEMBER 2007, Vol. 4 / 425
Copyright © 2007 by ASME

Downloaded 05 Sep 2010 to 130.101.140.126. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Table 1 Geometrical parameters of the tubular SOFC methane. The thermodynamic possibility of carbon formation is
then discussed from the values of the driving forces for carbon
Geometry Thickness 共mm兲 deposition. It is worth mentioning that this driving force for car-
Cathode 0.70 bon deposition is only an indicator of carbon presence in the sys-
Electrolyte 0.06 tem but does not provide any information about the carbon
Anode 0.10 amount and its deposition rate. Here, the H2O / CH4 ratio and tem-
Gas channel 共⫻2兲 2.64 perature are varied in order to identify a suitable region for SOFC
working conditions which does not suffer from carbon formation.

porous electrodes are described using the Butler-Volmer equations 2 Tubular Solid Oxide Fuel Cell Model Descriptions
at the triple phase boundary. This gas diffusion electrode model
can be compared with some works on molten carbonate fuel cells 2.1 Description of the Tubular Solid Oxide Fuel Cell. A
共MCFCs兲 porous modeling research 关11,12兴. In particular, this ap- tubular SOFC is a cylindrical assembly of an electrolyte sand-
proach is very close to the three-phase homogeneous model de- wiched between an anode and a cathode. The oxidant gas is in-
veloped by Prins-Jansen et al. 关13兴. However, we have included troduced through a central injection tube, and the fuel gas is sup-
on the anodic side reforming and gas-shift reactions, which occur plied to the exterior tube side. In this arrangement, the fuel gas
in parallel with hydrogen oxidation. Modeling is based on solving flows past the anode on the exterior of the cell parallel to the
conservation equations of mass, momentum, energy, species, and oxidant gas 共coflow兲. The tube length for modeling is 30 mm, and
electric current by using a finite volume approach on 2D grids of the thicknesses of the different parts of the tubular SOFC are
arbitrary topology. Simulations with the CFD-ACE software pack- given in Table 1.
age allow the calculation of the distributions of partial pressures 共 In the present model, mass and charge transport phenomena,
CH4, H2, CO, CO2, and H2O兲, current density, and potentials of coupled with chemical and electrochemical reactions, are investi-
electronic and ionic phases within the anode part 共i.e., gas channel gated within the inlet of a tubular SOFC, as described in Fig. 1. A
and Cermet anode兲. finite volume method using a computational grid 关15兴 allows us to
Based on previous approaches described by Hou and Hughes in solve mass, charge, energy, momentum balances including trans-
literature 关14兴, a detailed thermodynamic analysis that takes into port through porous media, and chemical and electrochemical re-
account Boudouard and methane cracking reactions is carried out actions within porous electrodes in a gas diffusion electrode
to predict the carbon formation boundary for SOFCs fueled by model. The set of resulting conservation equations is solved ow-

Fig. 1 Geometry of the tubular SOFC cell

426 / Vol. 4, NOVEMBER 2007 Transactions of the ASME

Downloaded 05 Sep 2010 to 130.101.140.126. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
ing to the CFD-ACE commercial software package 关16兴.
This model deals with a 2D geometry in steady state conditions
for a SOFC assumed as isothermal. This later hypothesis is con-
sistent with a previous paper 关6兴 showing a thermal gradient less
than 0.2 K / mm.
The O2 / N2 and CH4 / H2O mixtures are supplied at the air cath-
ode and at the anode gas channel, respectively. The anodic fuel
consumption results from both internal reforming and water gas–
shift reactions. Fuel composition can include any combination of
methane and steam.
In the gas phase, the mass conservation equation is described as
follows 关15兴: Fig. 2 DIR of methane by steam


共␧␳兲 + ⵜ共␧␳v兲 = 0 共1兲
⳵t
CO + H2O ↔ CO2 + H2 共4兲
where ␧ is the porosity of the medium and and represents the ratio
while hydrogen is electrochemically converted into electricity
of the volume occupied by the pores to the total volume of the
within the anode 共Eq. 共5兲兲:
porous medium, ␳ is the density, and v is the velocity vector of the
1 1 − 1
gas mixture. 2 H2 + 2 O 2 → 2 H 2O + e − 共5兲
By neglecting compressibility and turbulence effects, the con-
servation equations for the transport of Nth species in a porous The catalytic steam reforming reaction 共Eq. 共3兲兲 occurs at the
media can be presented in the following vector form: surface of a nickel catalyst with a reaction rate 共Eq. 共6兲兲
共kmol m−2 s−1兲 expressed from an Arrhenius law from literature
⳵ ␧ 2␮ v data 关3,14兴:

冉 冊
共␧␳v兲 + ⵜ共␧␳v · v兲 = − ␧ ⵜ p + ⵜ共␧␦兲 + 共2兲
⳵t ␬ 27.063 ⫻ 103
rE1 = 63.6 ⫻ T2 ⫻ exp − 关CH4兴关H2O兴
where p and v are the pressure and the velocity vector of the gas T
mixture, ␦ is the shear stress tensor, ␳ is the density, and ␮ is the
dynamic viscosity of the fluid. ␳ and ␮ properties are dependent
on temperature. ␬ is the permeability and is a quantity represent-
− 3.7 ⫻ 10−14 ⫻ T4 ⫻ exp − 冉 232.78
T

关CO兴关H2兴3 共6兲

ing the square of the effective volume to surface area ratio of the where 关i兴 is the concentration of gas species i 共kmol m−3兲 and T is
porous matrix. The last term in Eq. 共2兲 represents Darcy’s drag the temperature 共K兲.
force imposed by the pore walls on the fluid within the pores and Contrary to the steam reforming reaction 共Eq. 共3兲兲, the shift
usually results in a significant pressure drop across the porous reaction 共Eq. 共4兲兲 occurs wherever the gas is present, either in a
medium. gas channel or in pores. Its reaction rate 共kmol m−3 s−1兲 may be
Mass and charge balances are numerically solved with the fol- expressed from literature 关3,14兴 as
lowing boundary conditions. Potentials are set to zero for the an-
odic collector and to 0.6 V for the cathodic one. The ionic current
is supposed to be zero on the surface of both collectors, but vari-
rE2 = 1199 ⫻ T2 ⫻ exp − 冉 12.5 ⫻ 103
T

关CO兴关H2O兴

able elsewhere. Electronic Ohmic drops along the gas channel can
be neglected because of the high electronic conductivity assumed
for the electronic conductor.
− 1119 ⫻ T2 ⫻ exp − 冉 12.5 ⫻ 103
T

关CO2兴关H2兴 共7兲

Finally, gas transport within the porous electrode is described Finally, the kinetics of electrochemical reaction 共Eq. 共5兲兲 within
by using the classical model of Fick diffusion rather the usual the porous electrode may be described using the Butler-Volmer
combination of Stefan-Maxwell and Knudsen diffusions with con- equation at the triple phase boundary 共Eq. 共8兲兲:

冋 冉 冊 冉 冊 册
vective transport. Nevertheless, this simplified approach leads us
to predict a rather similar behavior, as shown by the slight differ- ␣ aF 关H2兴 ␣ cF 关H2O兴
jat = ja0 exp ␩a − exp − ␩a 共8兲
ence between both cases. RT 关H2兴0 RT 关H2O兴0
2.2 Anode. The anode material consists of a Cermet of nickel Here, the exchange current density is expressed in A m−2. The
and yttrium stabilized zirconia 共Ni-YSZ兲. The anode is considered overpotential ␩ 共Eq. 共9兲兲 is defined as the difference of the elec-
from modeling as a porous gas diffusion electrode wherein the tronic and ionic potentials:
electrochemical reaction occurs at the triple phase boundary, i.e., ␩a = ␾aM − ␾as − Ea0 共9兲
at the interface between electronic conductor 共nickel兲, ionic con-
ductor 共YSZ兲, and gas phase. Current density 共related to charge and is locally determined within the porous electrode through a
transports兲 is thus the sum of two indissociable but different con- separate solving of the equations related to electronic and ionic
tributions: one related to ionic species and the other to electron potentials. Ea0 is the potential difference between the electrolyte
transport. Mass transport occurs within the gas pores, and charge and the nickel in equilibrium, i.e., when no current is produced.
transport phenomena depend on four electric parameters: ionic The other parameters are the symmetry factors ␣a and ␣c de-
and electronic conductivities ␴as and ␴aM 共⍀−1 m−1兲 and poten- termined from the experimental Taffel slopes, the reference cur-
tials ␾as and ␾aM 共V兲. rent density ja0 共A m−2兲, the Faraday constant F, and the ideal gas
The DIR consists in adding the function of a reformer to the law constant R. 关H2兴 and 关H2O兴 are interfacial concentrations,
SOFC anode 共Fig. 2兲 关17兴, a direct production of hydrogen from while 关i兴0 refers to concentration values at the reference state at
methane being possible due to the high temperature in the fuel which the reference current density is prescribed.
cell. Within a porous electrode, current can be split into two parts:
Methane can thus be converted into H2 and CO by the steam one flowing through the electrolyte phase and the other one flow-
reforming reaction 共Eq. 共3兲兲 and water gas shift 共Eq. 共4兲兲: ing through the electronic phase of the porous matrix. During
electrochemical reactions, electrons are then either transferred
CH4 + H2O ↔ CO + 3H2 共3兲 from the ionic phase to the electronic one or in the opposite way.

Journal of Fuel Cell Science and Technology NOVEMBER 2007, Vol. 4 / 427

Downloaded 05 Sep 2010 to 130.101.140.126. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Table 2 Material properties at 1173 K

␴s ␬ kM A/V dpore ␴M
Element 共⍀−1 m−1兲 Sc ␧ 共m2兲 共W m−1 K−1兲 共m−1兲 共m兲 共⍀−1 m−1兲

Anode 8 0.7 0.40 1 ⫻ 10−12 6.23 3 ⫻ 106 1 ⫻ 10−6 1.56⫻ 107


Ni-YSZ
Cathode 8 0.7 0.50 1 ⫻ 10−12 9.6 — 1 ⫻ 10−6 1.56⫻ 107
LaMnO3-YSZ
Electrolyte 8 0.7 0.01 1 ⫻ 10−18 2.7 — 1 ⫻ 10−6 —
YSZ


If we consider an electrochemical reaction occurring at the anode, 1/4O2 + e− → 1/2O2 共17兲
charge conservation may thus be expressed from Ohm’s law as:

冉冊
The exchange current jct at the cathode is then obtained from a
ⵜ共␴as ⵜ ⌽as兲 = − ⵜ共␴aM ⵜ ⌽aM 兲 = jt
S
共10兲 Butler-Volmer equation similar to Eq. 共8兲:

冋 冉 冊 冉 冊 册
V eff
On the other hand, mass balances for each gas phase species i is ␣ aF ␣F 关O2兴
jct = jc0 exp ␩c − exp − c ␩c 共18兲
given at a steady state by: RT RT 关O2兴0
3

ⵜ · 共␧␳vwi兲 = ⵜ · Ji + 兺Z
p=1
ip 共11兲
2.4 Electrolyte. The electrolyte material is YSZ, which is a
suitable ionic conductor at high temperature. Conditions of zero
electronic current and zero electronic voltage are imposed in the
Here, the first member corresponds to convection, the term Zip electrolyte to ensure that electrolyte is completely impermeable to
stands for the creation or consumption rate of species i 共per unit electron circulation. The electrolyte potential is thus expressed by
volume of the porous medium兲, and p represents the three reac- a classical Ohm’s law without any charge creation or consumption
tions that takes place at the same time at the anode. Thus, this within the electrolyte:
latter term can be expressed in three forms, from the anodic elec-
trochemical reaction, the steam reforming reaction, or the chemi-
cal gas–shift reaction within the pores: ␴es 冉 ⳵2␾es ⳵2␾es
⳵x2
+
⳵y2
=0 冊 共19兲

⬙ − ␯i3
M i共␯i3 ⬘兲 冉冊 S
V eff
jat
F
= Zi3 共12兲
2.5 Simulation Parameters. Table 2 provides a brief descrip-
tion of the properties 关21兴 for the materials currently used in the
⬙ − ␯i1
共␯i1 ⬘ 兲rE1冉冊 A
V eff
= Zi1 共13兲
various cell components of the tubular SOFC at 1173 K.
Among them, Sc is the Schmidt number that enables the soft-
ware to calculate the diffusion coefficient.
⬙ − ␯i2
共␯i2 ⬘ 兲rE2 = Zi2 共14兲 For species i, we can write

where ␯i⬘ and ␯i⬙ are the normalized stoichiometric coefficients for ␮
reactants and products, respectively, M i and wi are the molecular Di = 共20兲
weight and the mass fraction of species i, respectively, 共S / V兲eff Sc
共m2 m−3兲 is used to take into account the effective surface-to- The kinetic theory 关22兴 of gases is used to calculate the viscosity
volume ratio, which is directly linked to the triple phase boundary ␮ of the gas or mixture of gases. For a gas i, we can write
surface area per unit electrode volume, and 共A / V兲eff is the specific
catalyst surface area of the catalyst per unit volume of the elec- 冑M iT
trode, which is a direct representation of catalyst loading. To re- ␮i = 2.6693 ⫻ 10−5 共21兲
sume, we can say that the mass conservation equation of species d2c ⍀␮
takes into account the three reactions located at the anode side
共steam reforming, water gas shift, and electrochemical reaction兲 at where ␮i is the dynamic viscosity of species i, M i is the molecular
the same time. Note that the second term of Eq. 共11兲 standing for weight of species i, T is the temperature in kelvin, dc is the char-
the diffusion of species i uses the effective mass diffusion coeffi- acteristic diameter of the molecule in angstroms, and ⍀␮ is the
cient Di,eff within the porous medium 共Eq. 共15兲兲, to be deduced collision integral.
from the free stream diffusion coefficient Di 共Eq. 共16兲兲 by the The collision integral ⍀␮ is given by:
so-called Bruggeman model 关18兴:
1.16145 0.52487 2.16178
Ji = ␳Di,eff ⵜ wi 共15兲 ⍀␮ = + + 共22兲
T*0.14874 exp共0.77320T*兲 exp共2.43787T*兲
Di,eff = Di␧␶ 共16兲
where T* is the dimensionless temperature and is given by
depending on porosity ␧ and tortuosity ␶. Simulations were per-
formed here with a tortuosity of 1.5, a classical value from litera-
kT
ture 关19,20兴 in the absence of more accurate data. T* = 共23兲

2.3 Cathode. The cathode consists of a porous lanthanum
manganite 共LaMnO3兲 doped with YSZ solid phases. Its behavior where ␹ is the characteristic energy, k is Boltzmann’s constant,
is described in the same way as is previously done for the anode. and T is the temperature.
Oxygen reduction reaction occurs at the cathode from the elec- To calculate the mixture viscosity using the kinetic theory, the
trochemical reaction following equation is used:

428 / Vol. 4, NOVEMBER 2007 Transactions of the ASME

Downloaded 05 Sep 2010 to 130.101.140.126. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Table 3 Kinetics electrochemical parameters Table 4 Gas flows
a
Anode Cathode Anode 共kg s−1兲 Cathode 共kg s−1兲

共S / V兲 j0 共A m−3兲 1 ⫻ 1011 1 ⫻ 1010 Air — 2.12⫻ 10−3


␣a 0.7 0.7 Fuel 3.415⫻ 10−5 —
␣c 0.7 0.7

a
References 27 and 28.

cal reaction. Moreover, the Boudouard and cracking reactions at


n the anode can also be favored at these temperatures. Carbon for-
w i␮ i
␮mix = 兺 共24兲
mation can thus occur and lead to a deposit capable of polluting
i=1 兺w⌰
j
j i,j
the anode. This carbon layer on the anode surface could obviously
block fuel feeding and oxygen ion transfer, resulting in a decreas-
ing power density.
where wi and w j are the mass fractions of species i and species j,
␮i is the dynamic viscosity of species i, and ⌰i,j is a dimension- 3.1 Standard Case at 1173 K. The fuel 共52.9 wt % steam,
less quantity given by 47.1 wt % methane兲 is fed at the anode, and a gas similar to air

冑 冉 冊 冋 冉 冊 冉 冊册
共21 wt % dioxygen and 79 wt % dinitrogen兲 is injected at the
1 Mi −0.5
␮i 0.5
Mj 2
cathode. Their flow rates were varied and are given in Table 4.
⌰i,j = 1+ 1+ 共25兲
8 Mj ␮j Mi Firstly, a mole ratio xCH4 / xH2 = 1 is considered at the anode, cor-
0

␧ is the porosity 共ratio of the volume occupied by pores to the responding to the classical limit beyond which no carbon forma-
total volume of the porous solid兲, ␬ represents the permeability in tion occurs. Note that the temperature is set to 1173 K in this part.
m2, k M is the solid phase thermal conductivity in W m−1 K−1, and In further simulations discussed below, the influence of the tem-
dpore is the diameter of the pore considered as cylindrical. perature and steam methane ratio will be investigated by varying
The kinetic data used in the Butler-Volmer equations 共reference one parameter at a time, maintaining the other parameters at the
current density j0 and symmetry factors ␣a and ␣c兲 are listed in values defined for the standard case.
Table 3 for both anode and cathode. Figures 3–5 show the mole fraction distribution for methane
Finally, the description is completed by the ideal gas law 共Eq. and dihydrogen at the anode and that for dioxygen at the cathode
共26兲兲: along the fuel cell channel and the electrode thickness.
Concerning the mole fraction distribution for methane through-
PM out the cell in Fig. 3, it clearly appears that the reaction is rather
␳= 共26兲
RT fast since the equilibrium is reached only 15 mm after the cell
inlet. In this part, water and methane are therefore present in large
N amounts compared to hydrogen. In addition, the reaction is almost
M= 兺xM i=1
i i 共27兲 complete. Therefore, the methane mole fraction is very weak at
the cell outlet. Moreover, the shape of the contour plot of the
methane mole fraction confirms that the steam reforming reaction
where M is the molar mass of the mixture gas and xi is the molar
is localized at the anode. The closer the reaction is to the electro-
fraction of species i.
lyte, the faster this reaction is, as confirmed by literature that
provides similar results 关8,23兴.
3 Results and Discussion The concentration distribution of hydrogen 共Fig. 4兲 reveals that
As said above, the high operating temperature of the fuel cell hydrogen is produced at the anode by steam reforming and water
makes the DIR at the anode possible. This process enables the gas–shift reactions, which instantaneously reach equilibrium. Ini-
generation in situ of the dihydrogen required by the electrochemi- tially, the dihydrogen mole fraction obviously increases due to its

Fig. 3 Mole fraction of methane

Journal of Fuel Cell Science and Technology NOVEMBER 2007, Vol. 4 / 429

Downloaded 05 Sep 2010 to 130.101.140.126. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 4 Mole fraction of dihydrogen

production by steam reforming. At the cell inlet, dihydrogen is 2CO ↔ C + CO2 共28兲
supplied more quickly than it is consumed by the oxidation reac-
tion. At the outlet, the steam reforming reaction is negligible 共low
fraction of methane兲, and then the dihydrogen mole fraction con- CH4 ↔ C + 2H2 共29兲
tinuously decreases by consumption. Therefore, a detailed thermodynamic analysis has to be carried
The mole fraction distribution for dioxygen is very different out to predict the carbon formation boundary for SOFCs fueled by
from the previous one 共Fig. 5兲. A large oxygen concentration gra- methane. The preceding reactions are the most probable reactions
dient is exhibited through the cathode thickness, while a low de- that lead to carbon formation in the reaction system. The Boud-
crease appears along the cathode channel. It should be noted that ouard reaction and the cracking reaction are the major pathways
the air flow rate is selected such that the oxygen utilization rate for carbon formation at high operating temperatures 关24兴. The
remains rather small. Such working conditions have been fixed to results will be discussed from the values of driving forces for
avoid a possible cathode effect, i.e., a decrease of the cell perfor- carbon deposition, defined as the ratios ␣ and ␤ depending on the
mances due to oxygen depletion at the cathode. respective equilibrium constants KB and KC of the Boudouard and
As a conclusion for this section, the simulation results seem to cracking reactions and on the partial pressure Pi of component i
confirm the feasibility of using SOFC in DIR with methane. At the 关14,25兴.
anode side, this conversion process includes steam methane re-
forming 共Eq. 共3兲兲 associated with shift 共Eq. 共5兲兲 reactions, but also PCO2
the parasitic Boudouard 共Eq. 共28兲兲 and cracking 共Eq. 共29兲兲 ␣= 2 共30兲
reactions. PCO KB

Fig. 5 Mole fraction of dioxygen

430 / Vol. 4, NOVEMBER 2007 Transactions of the ASME

Downloaded 05 Sep 2010 to 130.101.140.126. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 6 Distribution of the ␣ value Fig. 7 Distribution of the ␤ value

These results are similar to those found in literature 关14兴, where


it is admitted that carbon formation can only be due to cracking
2 reaction.
PH
␤= 2
共31兲 Among the two reactions where carbon acts, cracking reaction
PCH4KC proceeds to the right and could yield carbon, as the ␤ value is
much smaller than unity, whereas the trend of the Boudouard
For each reaction, if ␣ or ␤ ⬍ 1, the system is not at equilib- reaction is to proceed to the left. This latter observation indicates
rium; the reaction will proceed to the right, and carbon formation that it may play a role in decoking to some extent. After these
can be observed. Equilibrium is reached when ␣ or ␤ = 1. On the preliminary considerations, it appears of great interest to investi-
contrary, if ␣ or ␤ ⬎ 1, carbon formation is thermodynamically gate the simultaneous influences of both reactions through coeffi-
impossible and the reaction will proceed to the left. Moreover, an cient ␥ 共Fig. 8兲.
additional coefficient ␥ can be defined to evaluate the possibility Again, white plane indicates the boundary value of unity. Car-
of carbon deposition due to the two combined reactions. If ␣ and bon deposition is favored on the first millimiter of the cell,
␤ are less than 1 or higher than 1, there is no utility in introducing whereas the risk of carbon deposition by the cracking reaction is
the ␥ coefficient because the Boudouard and the cracking reac- inhibited by the Boudouard reaction after 1 mm. It is important to
tions will have the same effect. On the other hand, if one reaction remember that the carbon deposition rate is already unknown.
proceeds to the right and the other proceeds to the left, we must
assess the influence of both reactions at the same time. This is 3.2 Sensitivity Analysis to Temperature and Methane
why the ␥ coefficient is introduced, Steam Mixture. In this part, the modeling is applied to four tem-
␥ = ␣␤ 共32兲 peratures ranging from 973 K to 1273 K 共973 K, 1073 K,
1173 K, and 1273 K兲. The material properties were modified to be
It is worth mentioning that the driving force for carbon deposi- in conformity with the various temperatures, while all the other
tion is only an indicator of the potential carbon presence in the
system. It does not provide any information about the amount of
deposited carbon and its deposition rate or even about whether
there will really be a carbon deposition. It just locates the area
where carbon deposition is thermodynamically favored. The local
values of parameters ␣, ␤, and ␥ are then calculated from the
distributions of partial pressures within the 2D porous anode. Fi-
nally, their mapping enables us to predict the predominance zones
where carbon formation is favored or where carbon formation is
prevented according to thermodynamics.
Large positive values of ␣ always greater than unity are ob-
tained for the Boudouard reaction 共Fig. 6兲. This observation con-
firms that carbon formation is unlikely from this reaction at
1173 K because it would be shifted to the left according to ther-
modynamics under such conditions. The predicted peak of ␣ at
the cell inlet points out the location where CO2 is not yet formed.
Figure 7 shows a monotonical increase of ␤ with cracking re-
action extent. However, ␤ remains lower than unity on the first
17 mm of the cell, where a carbon deposit is favored. In this
figure, a white plane indicates the value of unity beyond which a
carbon deposit is not thermodynamically possible. It is possible to
conclude that the ␤ value close to the gas channel is smaller than
that near the electrolyte. This trend seems to confirm that dihy-
drogen oxidation is more important close to the electrolyte. Fig. 8 Distribution of the ␥ value

Journal of Fuel Cell Science and Technology NOVEMBER 2007, Vol. 4 / 431

Downloaded 05 Sep 2010 to 130.101.140.126. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 9 Distribution of the ␥ coefficient along the fuel cell for four temperatures. „䊏…
1273 K, „쎲… 1173 K, „ⴛ… 1073 K, „䉱… 973 K, and „⽧… carbon deposition limit.

model parameters were unchanged. account this parameter 共Fig. 10兲.


Figure 9 shows the ␥ coefficient profiles along a certain length Whatever the xH2O / xCH4 ratio value is, a carbon deposit should
of the fuel cell since the inlet 共millimeter兲, in the anode, for these always occur in the first millimeter of the anode at temperatures
four simulations. close to 1173 K. These results disagree with the major part of
Whatever the temperature is, Fig. 9 demonstrates that carbon papers published in literature that commonly assumed that carbon
deposition is always thermodynamically possible. In addition, the
higher the temperature is, the more carbon deposition will be re- deposition does not occur beyond the ratio xH2O / xCH4 = 1 关6兴.
stricted to a small length from the anode inlet. It also appears that However, it is important to note that these previous works only
carbon deposition is favored all over the cell length at 973 K. considered the Boudouard reaction. The present study also shows
Here, parameters ␣, ␤, and ␥ are evaluated at 1173 K for dif- that carbon deposition from the Boudouard reaction is thermody-
ferent methane and steam mixtures. Eight mixtures characterized namically impossible for operating temperatures of about 1173 K.
by several xH2O / xCH4 ratios 共0.1, 0.2, 0.3, 0.5, 0.7, 1, 2, and 3兲 are More thorough studies take into account the Boudouard reaction
tested. In all cases, the methane mole number is kept constant. as well as the methane cracking reaction 关14,26兴. They consider
The flow rates are thus modified in all simulations to take into that the possible occurrence of carbon deposition is only due to

Fig. 10 Distribution of the ␥ coefficient along the cell for three ratios. „⽧… R = 3, „쎲…
R = 1, and „䉱… R = 0.4.

432 / Vol. 4, NOVEMBER 2007 Transactions of the ASME

Downloaded 05 Sep 2010 to 130.101.140.126. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 11 Carbon deposition limit for ten ratios, xH2O / xCH4

the cracking reaction. These results thus confirm our simulation Nomenclature
data. However, no study has been carried out before to localize the 共A / V兲eff ⫽ specific catalyst surface area of catalyst
carbon deposit, as is done here. However, other works are obvi- 共m2 m−3兲
ously necessary because the rate of carbon formation is not yet
Di ⫽ diffusion coefficient of species i 共m2 s−1兲
taken into account, and therefore, there is still no information
available concerning the rate of anode pollution. Dieff ⫽ effective diffusion coefficient of species i
Figure 11 presents the evolution of the ratio xH2O / xCH4 versus 共m2 s−1兲
the carbon deposition limit corresponding to the length of the cell F ⫽ Faraday constant 共96,500兲 共C mol−1兲
共millimeter兲 where ␥ is less than 1. It shows that the carbon depo- Ji ⫽ diffusion flux of the ith species 共kg m−2 s−1兲
sition is negligible when xH2O / xCH4 is at least equal to unity. For KB ⫽ equilibrium constant of the Boudouard reaction
xH2O / xCH4 ⬍ 1, a sharp increase is observed in the carbon deposi- 共Pa−1兲
tion limit. KC ⫽ equilibrium constant of the cracking reaction
共Pa兲
M ⫽ molecular weight of the mixture of gases
4 Conclusion 共kg kmol−1兲
A model using the CFD-ACE software package has been devel- M i ⫽ molecular weight of species i 共kg kmol−1兲
oped for DIR at the anode of a tubular SOFC. Moreover, a de- Pi ⫽ partial pressure of species i 共Pa兲
tailed thermodynamic analysis based on three characteristic pa- R ⫽ universal gas constant 共8314兲 共J mol−1 K−1兲
rameters ␣, ␤, and ␥ has been carried out to predict the carbon Sc ⫽ Schmidt number
formation limit for a SOFC due to methane cracking and Boud- 共S / V兲eff ⫽ effective surface to volume ratio 共m2 m−3兲
ouard reactions.
Simulation allows the prediction of the distributions of partial T ⫽ temperature 共K兲
pressures 共CH4, H2, CO, CO2, and H2O兲, current density, and T* ⫽ dimensionless temperature
potentials of electronic and ionic phases within the anode part Zip ⫽ production rate of species i due to reactions
共i.e., gas channel and Cermet anode兲. As expected, the steam re- 共kg m−3 s−1兲
forming reaction of CH4 should largely occur within the electrode dpore ⫽ pore diameter 共m兲
volume close to the gas inlet, thus leading to a sharp increase in dc ⫽ characteristic diameter of the molecule 共Å兲
the hydrogen partial pressure. In this part, water and methane are 关i兴 ⫽ molar concentration of species i 共kmol m−3兲
therefore present in a large amount compared to hydrogen. ja0 ⫽ reference current density of the anode 共A m−2兲
Finally, mapping ␣, ␤, and ␥ values enabled us to predict that jat ⫽ exchange current due to the anodic reaction
in the classical working conditions of a SOFC, there is a signifi- 共A m−2兲
cant tendency for methane cracking, which could result in carbon jc0 ⫽ reference current density of the cathode
formation. However, the reverse Boudouard reaction seems to be 共A m−2兲
favorable, which could form CO from the carbon deposited; there- jct ⫽ exchange current due to the cathodic reaction
fore, a decoking is possible. The ␥ values are indicators to assess 共A m−2兲
the contributions of both reactions in the system. With regard to k ⫽ Boltzmann’s constant
the ␥ values, we can say that a carbon formation can be supposed k M ⫽ solid phase thermal conductivity 共W m−1 K−1兲
for temperatures less than 800° C and for ratios xH2O / xCH4 fewer
v ⫽ fluid velocity 共m s−1兲
than 1.
wi ⫽ mass fraction of species i
xi ⫽ molar fraction of species i
Acknowledgment
The authors would like to warmly thank Prof. J. Fouletier, Dr. Greek
S. Georges, and M. Ucar for many fruitful discussions. ⍀␮ ⫽ collision integral

Journal of Fuel Cell Science and Technology NOVEMBER 2007, Vol. 4 / 433

Downloaded 05 Sep 2010 to 130.101.140.126. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
␣ ⫽
Boudouard coefficient Finite Volume SOFC Model for a Tubular Cell Geometry,” J. Power Sources,
132, pp. 113–126.
␣a ⫽
anodic Taffel constant 关9兴 Aguiar, P., Adjiman, C. S., and Brandon, N. P., 2004, “Anode-Supported In-
␣c ⫽
cathodic Taffel constant termediate Temperature Direct Internal Reforming Solid Oxide Fuel Cell. I:
␤ ⫽
cracking coefficient Model-Based Steady-State Performance,” J. Power Sources, 138, pp. 120–
␹ ⫽
characteristic energy 136.
关10兴 Li, P. W., and Suzuki, K., 2004, “Numerical Modeling and Performance Study
␧ ⫽
porosity of a Tubular SOFC,” J. Electrochem. Soc., 151共4兲, pp. A548–A557.
␾as ⫽
ionic phase potential at the anode 共V兲 关11兴 Yuh, C. Y., and Selman, J. R., 1992, “Porous-Electrode Modeling of the
␾aM ⫽
electronic phase potential at the anode 共V兲 Molten-Carbonate Fuel-Cell Electrodes,” J. Electrochem. Soc., 139共5兲, pp.
␾es ⫽
ionic phase potential of the electrolyte 共V兲
关12兴
1373–1379.
Fontes, E., Lagergren, C., and Simonsson, D., 1997, “Mathematical Modelling
␥ ⫽
Boudouard+ cracking coefficient of the MCFC Cathode on the Linear Polarisation of the NiO Cathode,” J.
␩a ⫽
anode overpotential 共V兲 Electroanal. Chem., 432, pp. 121–128.
␩c ⫽
cathode overpotential 共V兲 关13兴 Prins-Jansen, J. A., Fehribach, J. D., Hemmes, K., and De Wit, J. H. W., 1996,
“A Three-Phase Homogeneous Model for Porous Electrodes in Molten-
␬ ⫽
permeability 共m2兲 Carbonate Fuel Cells,” J. Electrochem. Soc., 143共5兲, pp. 1617–1628.
␮ ⫽
dynamic viscosity 共Pa s兲 关14兴 Hou, K., and Hughes, R., 2001, “The Kinetics of Methane Steam Reforming
⬙ , ␯ip
␯ip ⬘ ⫽
normalized stoichiometric coefficients Over a Ni/ ␣-Al2O Catalyst,” Chem. Eng. J., 82, pp. 311–328.
关15兴 Mazumder, S., and Cole, J. V., 2003, “Rigorous 3-D Mathematical Modeling
␳ ⫽
mass density 共kg m−3兲 of PEM Fuel Cells. I. Model Predictions With Liquid Water Transport,” J.
␴as ⫽
ionic phase conductivity at the anode Electrochem. Soc., 150共1兲, pp. A1503–A1509.
共⍀−1 m−1兲 关16兴 ESI group, CFD-ACE⫹, Version 2006, Manuals.
关17兴 Vernoux, P., Guindet, J., and Kleitz, M., 1998, “Gradual Internal Methane
␴aM ⫽ electronic phase conductivity at the anode Reforming in Intermediate-Temperature Solid-Oxide Fuel Cells,” J. Electro-
共⍀−1 m−1兲 chem. Soc., 145共10兲, pp. 3487–3492.
␴es ⫽ ionic phase conductivity of the electrolyte 关18兴 Bruggeman, D. A. G., 1935, Ann. Phys., 24, pp. 636–664.
共⍀−1 m−1兲 关19兴 Springer, T. E., Zawodinski, T. A., and Gottesfeld, S., 1991, “Polymer Elec-
trolyte Fuel Cell Model,” J. Electrochem. Soc., 138共8兲, pp. 2334–2342.
␶ ⫽ tortuosity of pores 共m m−1兲 关20兴 Natarajan, D., and Nguyen, T. V., 2001, “A Two-Dimensional, Two-Phase,
Multicomponent, Transient Model for the Cathode of a Proton Exchange
Membrane Fuel Cell Using Conventional Gas Distributors,” J. Electrochem.
References Soc., 148共12兲, pp. A1324–A1335.
关1兴 Ahmed, K., and Foger, K., 2000, “Kinetics of Internal Steam Reforming of 关21兴 EG&G Services, 2000, Fuel Cell Handbook, 5th ed., DOI 10.2172/769283, pp.
Methane on Ni/YSZ Based Anodes for Solid Oxide Fuel Cells,” Catal. Today, 8.5–8.6.
63, pp. 479–487. 关22兴 Bird, R. B., Stewart, W. E., and Lightfoot, E. N., 2002, Transport Phenomena,
关2兴 Vernoux, P., Guillodo, M., Fouletier, J., and Hammou, A., 2000, “Alternative 2nd ed., Wiley, New York.
Anode Material for Gradual Methane Reforming in Solid Oxide Fuel Cell,” 关23兴 Matelli, J. A., and Bazzo, E., 2005, “A Methodology for Thermodynamic
Solid State Ionics, 135, pp. 425–431. Simulation of High Temperature, Internal Reforming Fuel Cell Systems,” J.
关3兴 Lehnert, W., Meusinger, J., and Thom, F., 2000, “Modelling of Gas Transport Power Sources, 142, pp. 160–168.
Phenomena in SOFC Anodes,” J. Power Sources, 87, pp. 57–63. 关24兴 Amor, J. N., 1999, “The Multiple Roles for Catalysis in the Production of H2,”
关4兴 Suwanwarangkul, R., Croiset, E., Fowler, M. W., Douglas, P. L., Entchev, E., Appl. Catal., A, 176, pp. 159–176.
and Douglas, M. A., 2003, “Performance Comparison of Fick’s, Dusty-Gas 关25兴 Sangtongkitcharoen, W., Assabumrungrat, S., Pavarajarn, V., Laosiripojana,
and Stefan-Maxwell Models to Predict the Concentration Overpotential of a N., and Praserthdam, P., 2005, “Comparison of Carbon Formation Boundary in
SOFC Anode,” J. Power Sources, 122, pp. 9–18. Different Modes of Solid Oxide Fuel Cells Fuelled by Methane,” J. Power
关5兴 Ackmann, T., de Haart, L. G. J., Lehnert, W., and Stolten, D., 2003, “Modeling Sources, 142, pp. 75–80.
of Mass and Heat Transport in Planar Substrate Type SOFCs,” J. Electrochem. 关26兴 Zang, X., Ohara, S., Chen, H., and Fukui, T., 2002, “Conversion of Methane to
Soc., 150共6兲, pp. A783–A789. Syngas in a Solid Oxide Fuel Cell With Ni-SDC Anode and LSGM Electro-
关6兴 Morel, B., Laurencin, J., Bultel, Y., and Lefebvre-Joud, F., 2005, “Anode- lyte,” Fuel, 81, pp. 989–996.
Supported SOFC Model Centered on the Direct Internal Reforming,” J. Elec- 关27兴 Costamagna, P., Panizza, M., Cerisola, G., and Barbucci, A., 2002, “Effect of
trochem. Soc., 152共7兲, pp. A1382–A1389. Composition on the Performance of Cermet Electrodes. Experimental and The-
关7兴 Larrain, D., Van Herle, J., Maréchal, F., and Favrat, D., 2003, “Thermal Mod- oretical Approach,” Electrochim. Acta, 47, pp. 1079–1089.
eling of a Small Anode Supported Solid Oxide Fuel Cell,” J. Power Sources, 关28兴 Cosatmagna, P., Costa, P., and Arato, E., 1998, “Some More Considerations on
118, pp. 367–374. the Optimization of Cermet Solid Oxide Fuel Cell Electrodes,” Electrochim.
关8兴 Campanari, S., and Iora, P., 2004, “Definition and Sensitivity Analysis of a Acta, 43, pp. 967–972.

434 / Vol. 4, NOVEMBER 2007 Transactions of the ASME

Downloaded 05 Sep 2010 to 130.101.140.126. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Das könnte Ihnen auch gefallen