05.03.2015 Views

Thesis for degree: Licentiate of Engineering

Thesis for degree: Licentiate of Engineering

Thesis for degree: Licentiate of Engineering

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

SOFC modeling from micro to macroscale:<br />

transport processes and chemical reactions<br />

Hedvig Paradis<br />

<strong>Thesis</strong> <strong>for</strong> the <strong>degree</strong> <strong>of</strong> <strong>Licentiate</strong> <strong>of</strong> <strong>Engineering</strong>, 2011<br />

Division <strong>of</strong> Heat Transfer<br />

Department <strong>of</strong> Energy Sciences<br />

Faculty <strong>of</strong> <strong>Engineering</strong> (LTH)<br />

Lund University<br />

www.energy.lth.se


Copyright © Eva Hedvig Charlotta Paradis, 2011<br />

Division <strong>of</strong> Heat Transfer<br />

Department <strong>of</strong> Energy Sciences<br />

Faculty <strong>of</strong> <strong>Engineering</strong> (LTH)<br />

Lund University<br />

Box 118, SE-221 00, Lund, Sweden<br />

ISRN LUTMDN/TMHP-11/7075 – SE<br />

ISSN 0282-1990


Abstract<br />

The purpose <strong>of</strong> this work is to investigate the interaction between transport processes and<br />

chemical reactions, with special emphasis on modeling mass transport by the Lattice<br />

Boltzmann method (LBM) at microscale <strong>of</strong> the anode <strong>of</strong> a solid oxide fuel cell (SOFC). In<br />

order to improve the per<strong>for</strong>mance <strong>of</strong> an SOFC, it is important to determine the microstructural<br />

effect embedded within the physical and chemical processes, which usually are modeled<br />

macroscopically. Without detailed knowledge <strong>of</strong> the transport processes and the chemical<br />

reactions at microscale it can be difficult to capture their effect and to justify assumptions <strong>for</strong><br />

the macroscopic models with regard to the source terms and various properties in the porous<br />

electrodes. The advantage <strong>of</strong> an anode-supported SOFC structure is that the thickness <strong>of</strong> the<br />

electrolyte can be reduced, while still providing an internal re<strong>for</strong>ming environment. For this<br />

configuration with an enlarged anode, more detailed knowledge <strong>of</strong> the porous domain in<br />

terms <strong>of</strong> the physical processes at microscale is called <strong>for</strong>.<br />

In the first part <strong>of</strong> this study, the current literature on the modeling <strong>of</strong> transport processes and<br />

chemical reactions mechanisms at microstructural scales is reviewed with special focus on the<br />

LBM followed by a report on the emphasis to couple conventional CFD to LBM. In the<br />

second part, two models are described. The first model is developed at microscale by LBM<br />

<strong>for</strong> the anode <strong>of</strong> an SOFC in MATLAB. In the LB approach, the main point is to carefully<br />

model the diffusion and convection at microscale in the porous region close to the threephase-boundary<br />

(TPB). The porous structure is reconstructed from digital images, and<br />

processed by Python. The second model is developed at macroscale <strong>for</strong> the whole unit cell.<br />

For the macroscale model the kinetic model is evaluated at smaller scales to investigate if any<br />

severe limiting effects on the heat and mass transfer occur.<br />

LBM has been found to be an alternative method <strong>for</strong> modeling at microscale and can handle<br />

complex geometries easily. However, there is still a need <strong>for</strong> a supercomputer to solve models<br />

with several physical processes and components <strong>for</strong> a larger domain. The result <strong>of</strong> the<br />

macroscale model shows that the three reaction rate models are fast and vary in magnitude.<br />

The pre-exponential values, in relation to the partial pressures, and the activation energy<br />

affect the reaction rate. The variation in amount <strong>of</strong> methane content and steam-to-fuel ratio<br />

reveals that the composition needs a high inlet temperature to enable the re<strong>for</strong>ming process<br />

and to keep a constant current-density distribution. As experiments with the same chemical<br />

compositions can be conducted on a cell or a re<strong>for</strong>mer, the effect <strong>of</strong> the chosen kinetic model<br />

on the heat and mass transfer was checked so that no severe limitation are caused on the<br />

processes at microscale <strong>for</strong> an SOFC.<br />

For future work, macroscale and microscale models will be connected <strong>for</strong> the design <strong>of</strong> a<br />

multiscale model. Multiscale modeling will increase the understanding <strong>of</strong> detailed transport<br />

phenomena and it will optimize the specific design and control <strong>of</strong> operating conditions. This<br />

can <strong>of</strong>fer crucial knowledge <strong>for</strong> SOFCs and the potential <strong>for</strong> a breakthrough in their<br />

commercialization.<br />

Keywords: mass transport, diffusion, microscale, porous media, kinetics, LBM, CFD, anode<br />

multicomponent, MATLAB.<br />

2


Populärvetenskaplig beskrivning på svenska<br />

Bränsleceller kan bidra till ett mer hållbart och miljövänligt samhälle ur ett<br />

energiutvinningsperspektiv genom hög energieffektivitet och väldigt låga utsläpp av<br />

koldioxid, kväveoxider och hälsoskadliga partiklar. Bränsleceller, speciellt vissa<br />

högtemperaturceller, kan arbeta med en rad olika bränslen förutom väte. I den här studien<br />

har både naturgas och väte använts som bränsle men andra kompatibla kandidater är<br />

etanol, metanol, biogas, och ammoniak. För att det skall leda till en mer miljövänlig<br />

energiproduktion måste bränslena tas fram på ett miljövänligt vis.<br />

Det var först omkring 1950 när bränsleceller användes som en kraftkälla i rymdraketer,<br />

som de blev mer allmänt kända och kompletta bränslecellssystem konstruerades.<br />

Bränslecellernas utveckling har tagit fart bland annat för att energipriserna ständigt ökar<br />

och likaså oron kring växthuseffektens påverkan på jordens klimat. Det är nu möjligt att<br />

tillämpa bränsleceller i en rad olika system i olika storlekar från mobiltelefoner till stora<br />

kraftverk med olika bränslen. Eftersom de kan byggas i olika storlekar, har bränsleceller<br />

en stor potential inom flera områden. De största hindren för en kommersialisering i stor<br />

skala är den höga tillverkningskostnaden, korta livslängden och avsaknaden av en<br />

funktionell infrastruktur för vätgas och biogas/naturgas.<br />

Uppbyggnad av en bränslecell<br />

Det finns olika typer av bränsleceller med olika kemiska processer och material. För att<br />

beskriva en typ av bränslecell; fastoxidbränslecellen som används i den här studien, väljs<br />

den med väte som bränsle i det här fallet. I en sådan bränslecell reagerar syre och väte<br />

med varandra och bildar vatten. En bränslecell är uppbyggd av en anod, en katod och en<br />

elektrolyt. En anod är den del i en elektrolytisk cell som är förbunden med strömkällans<br />

positiva pol, och katoden är sammanbunden med dess negativa pol. Det gas<strong>for</strong>miga<br />

bränslet transporteras till anoden där det reagerar i elektrokemiska reaktioner med<br />

syrejoner. Syrejonerna produceras i katoden där syre reagerar med elektroner till jon<strong>for</strong>m.<br />

Syrejonerna transporteras igenom elektrolyten för att nå bränslet i anoden. Elektronerna<br />

släpps inte igenom elektrolyten, vilket gör att en spänning uppstår. Den specifika<br />

bränslecellen i den här studien har en hög arbetstemperatur och elektrolyten, bestående av<br />

en fastoxid, är ut<strong>for</strong>mad för att endast släppa igenom syrejoner från katoden till anoden.<br />

Skillnaden mellan olika typer av bränsleceller är främst vilken typ av elektrolyt som<br />

används och bränslecellens arbetstemperatur.<br />

Övergångsfas<br />

Bränsleceller producerar elektricitet och värme direkt från kemiska och elektrokemiska<br />

reaktioner mellan bränsle och syret i luften, dvs. inget mekaniskt arbete. När ren vätgas<br />

eller biogas används blir det inga nettoutsläpp av koldioxid, hälsoskadliga partiklar eller<br />

kväveoxider, vilket gör processen helt miljövänlig och koldioxidneutral. För<br />

bränsleceller, såsom fastoxidbränslecellen, där arbetstemperaturen är mellan 600 och 800<br />

°C är det möjligt att använda utöver vätgas mer komplexa bränslen, som naturgas, biogas<br />

eller etanol. Då sker en omvandling (re<strong>for</strong>mering) av bränslet, antingen i en separat enhet<br />

som bränslet får passera innan det kommer till bränslecellen, eller inne i bränslecellens<br />

anod. Det material som vanligtvis används i anoden har visat sig vara lämpligt för<br />

3


katalytisk omvandling av naturgas och biogas till vätgas och kolmonoxid. Dessa reagerar<br />

sedan i bränslecellens anod genom de elektrokemiska reaktionerna med syrejoner.<br />

Dagens <strong>for</strong>skning om bränsleceller med hög arbetstemperatur fokuseras till stor del till att<br />

öka förståelsen kring den porösa mikrostrukturens inverkan och både de kemiska och<br />

elektrokemiska reaktionernas inverkan på de fysikaliska processerna. Mer realistiska<br />

numeriska modeller som fångar upp dessa mikroskaliga processers inverkan på de<br />

närliggande fysikaliska processerna kan bidra till en enorm förbättring i bränslecellens<br />

effektivitet. Målet för dagens <strong>for</strong>skning är att kunna simulera de fysikaliska processerna<br />

på alla nivåer. Det är även viktigt att kunna jämföra experimentella och numeriska<br />

analyser speciellt på mikronivå då det finns ett kunskapsglapp där. De parametrar som har<br />

störst inverkan på de minsta nivåerna kan kopplas till de större nivåerna med de<br />

konventionella parametrarna såsom hastighet och temperatur. Med ökad förståelse om vid<br />

vilken nivå en process har störst inverkan, skapas möjligheten att bidra till förbättrad<br />

prestanda vilket i sin tur kan leda till att kostnaden kan sänkas. För att möjliggöra denna<br />

simulering ställs höga krav på den tillgängliga datorkapaciteten och i takt med att<br />

datorkapaciteten ökar kan mer komplexa simuleringar utföras.<br />

Några av problemen och utmaningarna med dagens energisystem, både globalt och lokalt,<br />

är utsläpp av bland annat koldioxid, hälsoskadliga partiklar och kväveoxider. Det<br />

diskuteras hur länge mänskligheten kan <strong>for</strong>tsätta att utvinna fossila bränslen i samma takt<br />

som idag då det verkar finnas en begränsad mängd att tillgå. Möjligheten för en ren,<br />

miljövänlig och energieffektiv bränsleanvändning driver utvecklingen av<br />

bränslecellssystem framåt i allt snabbare takt. Fastoxidbränslecellen kan fungera<br />

framförallt som en övergång från den konventionella teknologin för energiutvinning med<br />

möjligheten att internt hantera kolvätebränslen till en mer hållbar och miljövänlig<br />

energiproduktion. Det som kommer att bestämma tillväxten inom bränslecellsområdet är<br />

hur mycket tillverkningskostnaden kan sänkas, och livslängden ökas på kort tid.<br />

4


Acknowledgements<br />

I am very grateful to my supervisors Pr<strong>of</strong>essor Bengt Sundén and Docent Jinliang Yuan <strong>for</strong> their<br />

support and guidance during the past two years. Many thanks also to Pr<strong>of</strong>essor Sundén who made my<br />

<strong>for</strong>thcoming research visit at University <strong>of</strong> Cali<strong>for</strong>nia, Berkeley possible. I also want to extend my<br />

thanks to Docent Christ<strong>of</strong>fer Norberg <strong>for</strong> encouraging me to go on to doctoral studies and <strong>for</strong> being a<br />

teacher role model.<br />

To Martin Andersson, I owe particular thanks <strong>for</strong> knowledgeable guidance and constant<br />

encouragement. I want to thank Ol<strong>of</strong> Ekedahl <strong>for</strong> help and encouragement in general and <strong>for</strong> all the<br />

support with work on Python in particular.<br />

The study was made possible through financial support from the Swedish Research Council<br />

(Vetenskapsrådet, VR) and the European Research Council (ERC).<br />

i


List <strong>of</strong> publications<br />

Journal publications:<br />

I. H. Paradis, M. Andersson, J. Yuan, B. Sundén, CFD Modeling: Different Kinetic Approaches<br />

<strong>for</strong> Internal Re<strong>for</strong>ming Reactions in an Anode-Supported SOFC, ASME Journal <strong>of</strong> Fuel Cell<br />

Science and Technology, 8, 031014, 2011<br />

II.<br />

III.<br />

IV.<br />

H. Paradis, M. Andersson, J. Yuan, B. Sundén, Simulation Analysis <strong>of</strong> Different Alternative<br />

Fuels <strong>for</strong> Potential Utilization in SOFCs, International Journal <strong>of</strong> Energy Research, 35, DOI:<br />

10.1002/er.1862, 2011<br />

M. Andersson, H. Paradis, J. Yuan, B. Sundén, Modeling Analysis <strong>of</strong> Different Renewable<br />

Fuels in an Anode-Supported SOFC, ASME Journal <strong>of</strong> Fuel Cell Science and Technology, 8,<br />

031013, 2011<br />

M. Andersson, H. Paradis, J. Yuan B. Sundén, Review <strong>of</strong> Catalyst Materials and Catalytic<br />

Steam Re<strong>for</strong>ming Reactions in SOFC anodes, International Journal <strong>of</strong> Energy Research, 35,<br />

DOI: 10.1002/er.1875, 2011<br />

Conference publications:<br />

I. H. Paradis, M. Andersson, J. Yuan, B. Sundén, The Kinetics Effect in SOFCs on Heat and<br />

Mass Transfer: Interparticle, Interphase, and Intraparticle Transport, submitted to: ASME 9 th<br />

Fuel Cell Science Conference, ESFuelCell2011-54015, 2011<br />

II.<br />

III.<br />

IV.<br />

H. Paradis, M. Andersson, J. Yuan, B. Sundén, CFD Modeling Concerning Different Kinetic<br />

Models For Internal Re<strong>for</strong>ming Reactions in an Anode-Supported SOFC, ASME 8 th Fuel Cell<br />

Science Conference, FuelCell2010-33045; pp. 55-64, 2010<br />

H. Paradis, M. Andersson, J. Yuan, B. Sundén, Review <strong>of</strong> Different Renewable Fuels <strong>for</strong><br />

Potential Utilization in SOFCs, 5 th International Green Energy Conference, 2010<br />

M. Andersson, H. Paradis, J. Yuan, B. Sundén, Modeling Analysis <strong>of</strong> Different Renewable<br />

Fuels in an Anode-Supported SOFC, ASME 8 th Fuel Cell Science Conference, FuelCell2010-<br />

33044; pp. 43-54, 2010<br />

V. M. Andersson, H. Paradis, J. Yuan, B. Sundén, Catalysts Materials and Catalytic Steam<br />

Re<strong>for</strong>ming Reactions in SOFC Anodes, 5 th International Green Energy Conference 2010<br />

ii


Table <strong>of</strong> contents<br />

Abstract ................................................................................................................................................... 2<br />

Populärvetenskaplig beskrivning på svenska .......................................................................................... 3<br />

Acknowledgements .................................................................................................................................. i<br />

List <strong>of</strong> publications .................................................................................................................................. ii<br />

Table <strong>of</strong> contents .................................................................................................................................... iii<br />

Nomenclature .......................................................................................................................................... v<br />

1 Introduction .......................................................................................................................................... 1<br />

1.1 Research objectives ....................................................................................................................... 2<br />

1.2 Methodology ................................................................................................................................. 2<br />

1.3 <strong>Thesis</strong> outline ................................................................................................................................ 2<br />

2 SOFC modeling at smaller scales ......................................................................................................... 3<br />

2.1 Solid Oxide Fuel Cells .................................................................................................................. 3<br />

2.2 SOFC modeling development ....................................................................................................... 5<br />

2.2.1 Multiscale and multiphysics modeling .................................................................................. 5<br />

2.2.2 Lattice Boltzmann concept .................................................................................................... 7<br />

2.2.3 Monte Carlo method and Molecular dynamics ...................................................................... 7<br />

2.2.4 Modeling integration issues ................................................................................................... 8<br />

2.3 Lattice Boltzmann method ............................................................................................................ 9<br />

2.3.1 Boundary conditions in LBM .............................................................................................. 12<br />

2.3.2 Choice <strong>of</strong> units in LBM ....................................................................................................... 12<br />

2.3.3 Mass diffusion in LBM ........................................................................................................ 14<br />

2.3.4 Chemical reactions in LBM ................................................................................................. 15<br />

2.3.5 General comments on CFD and LBM coupling .................................................................. 15<br />

2.4 Previous case studies <strong>of</strong> SOFCs in LBM .................................................................................... 17<br />

2.5 Electrode microstructure remarks ............................................................................................... 19<br />

3 Mathematical models ......................................................................................................................... 22<br />

3.1 Model visualization <strong>for</strong> the microscale model ............................................................................ 22<br />

iii


3.2 Governing equations <strong>for</strong> the macroscale model .......................................................................... 24<br />

3.2.1 Mass transport ...................................................................................................................... 25<br />

3.2.2 Heat transport....................................................................................................................... 26<br />

3.2.3 Momentum transport ........................................................................................................... 27<br />

3.2.4 Electrochemical reactions .................................................................................................... 27<br />

3.2.5 Internal re<strong>for</strong>ming reactions ................................................................................................ 28<br />

3.3 Heat and mass transfer limitations <strong>of</strong> the kinetic model ............................................................. 30<br />

3.3.1 Interparticle transport ........................................................................................................... 31<br />

3.3.2 Interphase transport ............................................................................................................. 32<br />

3.3.3 Intraparticle transport ........................................................................................................... 33<br />

4 Results and discussion ........................................................................................................................ 36<br />

4.1 Microscale model by LBM ......................................................................................................... 36<br />

4.2 Evaluation <strong>of</strong> kinetics at smaller scales ...................................................................................... 40<br />

4.3 Macroscale model by CFD ......................................................................................................... 43<br />

4.3.1 Case study: Internal re<strong>for</strong>ming reaction rates ...................................................................... 43<br />

4.3.2 Case study: Methane content and steam-to-fuel ratio .......................................................... 46<br />

5 Conclusions ........................................................................................................................................ 51<br />

6 Future work ........................................................................................................................................ 52<br />

7 References .......................................................................................................................................... 53<br />

iv


Nomenclature<br />

a lattice direction, –<br />

AV surface area-to-volume ration, m 2 /m 3<br />

Bi Biot number, –<br />

c lattice speed <strong>of</strong> sound, lu/ts<br />

c T total concentration, mol/m 3<br />

c p specific heat at constant pressure, J/(kg·K)<br />

C concentration, mol/m 3<br />

Da Darcy number, –<br />

D ij Maxwell-Stefan binary diffusion coefficient, m 2 /s<br />

D thermal diffusion coefficient, kg/(m·s)<br />

E activation energy, kJ/mol<br />

e lattice speed, lu/ts<br />

f a particle density distribution function, –<br />

f eq equilibrium distribution function, –<br />

F <strong>for</strong>ce vector, N/m 3<br />

F Faraday constant, 96485 C/mol<br />

F gravitational <strong>for</strong>ce, mu·lu/ts 2<br />

h enthalpy, kJ/mol<br />

h heat transfer coefficient, W/(m 2·K)<br />

h v volume heat transfer coefficient, W/(m 3·K)<br />

i current density, A/cm 2<br />

i 0 exchange current density, A/cm 2<br />

J mole flux, mol/(m 2 s 1 )<br />

J* dimensionless mole flux, –<br />

k thermal conductivity, W/(m·K)<br />

k reaction rate constant, mol/(m<br />

3·bar<br />

2·s)<br />

k c mass transfer coefficient, m/s<br />

k´´ pre-exponential factor, 1/(·m 2 )<br />

K e equilibrium constant, Pa 2 or dimensionless<br />

M j molecular weight <strong>of</strong> species j, kg/mol<br />

N number <strong>of</strong> grid points, –<br />

N iter number <strong>of</strong> iterations, –<br />

Nu Nusselt number, –<br />

n e number <strong>of</strong> electrons transferred per reaction, –<br />

P pressure, bar<br />

p partial pressure, Pa<br />

v


Q source term (heat), W/m3<br />

q heat flux, W/m 2<br />

r reaction rate, mol/(m 3·s), mol/(m 2·s)<br />

r mean pore radius, m<br />

R gas constant, 8.314 J/(mol·K)<br />

Re Reynolds number, –<br />

S i source term <strong>for</strong> the reaction rate, kg/(m 3·s)<br />

T temperature, K<br />

T viscous stress tensor, N/m 2<br />

t time step, –<br />

u, v velocity, m/s<br />

w a weight factor, –<br />

w i mass fraction <strong>of</strong> species i, kg/kg<br />

x, y coordinate system, m<br />

x mole fraction, –<br />

z pixel point, –<br />

Greek symbols<br />

β dimensionless maximum temperature rise, –<br />

γ dimensionless activation energy, –<br />

δ reference step (time or length)<br />

ε porosity, –<br />

η overpotential, V<br />

κ permeability, m 2<br />

κ dv deviation from thermodynamic equilibrium, Pa·s<br />

μ dynamic viscosity, Pa·s<br />

ρ density, kg/m 3<br />

ζ ionic/electronic conductivity, Ω -1 m -1<br />

η tortuosity, –<br />

η relaxation time<br />

υ kinematic viscosity, m 2 /s<br />

χ Damköhler number <strong>for</strong> the heat transfer, –<br />

ω Damköhler number <strong>for</strong> the mass transfer, –<br />

Ω D diffusion collision integral, –<br />

Φ phase function, –<br />

Subscripts<br />

0 reference state or initial state<br />

a lattice direction in LBM<br />

act activation polarization<br />

D dimensionless system<br />

e electrode, ea,c<br />

eff<br />

el<br />

effective<br />

electrolyte<br />

vi


g<br />

i<br />

j<br />

LB<br />

P<br />

r<br />

s<br />

w<br />

gas phase<br />

molecule i<br />

molecule j<br />

discrete system, lattice Boltzmann<br />

physical system<br />

steam re<strong>for</strong>ming reaction<br />

solid phase or water-gas shift reaction<br />

wall<br />

Abbreviations<br />

AFL anode functional layer<br />

CFD computational fluid dynamics<br />

CFL cathode functional layer<br />

FDM finite difference method<br />

FEM finite element method<br />

FVM finite volume method<br />

IEA International Energy Agency<br />

IT intermediate temperature<br />

LBM lattice Boltzmann method<br />

LTNE local temperature non-equilibrium<br />

MC Monte Carlo method<br />

MD molecular dynamics<br />

SEM scanning electron microscopy<br />

SF steam-to-fuel ratio<br />

SMM Stefan-Maxwell model<br />

SOFC solid oxide fuel cell<br />

TPB three-phase boundary<br />

YSZ yttria-stabilized zirconia<br />

XCT x-ray computer tomography<br />

Chemical<br />

CH 4 methane<br />

CO carbon monoxide<br />

CO 2 carbon dioxide<br />

H 2 hydrogen<br />

H 2 O water<br />

Ni nickel<br />

O 2 oxygen<br />

vii


1 Introduction<br />

After some time <strong>of</strong> decreasing interest, fuel cell research is now receiving a lot <strong>of</strong> attention.<br />

Fuel cells are considered a promising future resource <strong>for</strong> both stationary and distributed<br />

electric power stations because <strong>of</strong> their high per<strong>for</strong>mance and high reliability [1, 2]. In order<br />

to explore these aspects in greater depth, there is a need <strong>for</strong> both multiphysics and multiscale<br />

modeling. The approach is by solving the equations <strong>for</strong> momentum, charge, heat and mass<br />

transport and chemical reactions at the same time at corresponding scales. At this point,<br />

simulation <strong>of</strong> mass diffusion and convection as well as chemical and electrochemical<br />

reactions at microscale will <strong>of</strong>fer crucial insight to improve the per<strong>for</strong>mance <strong>of</strong> the fuel cell<br />

[3]. Furthermore, future prospects <strong>for</strong> fuel cell research are to connect microscale to<br />

macroscale and obtain stable solutions <strong>for</strong> multiscale modeling.<br />

Solid oxide fuel cells (SOFCs) are particularly interesting because it operates at high<br />

temperature and can there<strong>for</strong>e handle the re<strong>for</strong>ming <strong>of</strong> hydrocarbon fuels directly within the<br />

cell. SOFCs have a number <strong>of</strong> advantages, e.g., high conversion efficiency, high quality<br />

exhaust heat and flexibility <strong>of</strong> fuel input. However, as expected, SOFCs also have<br />

disadvantages. For instance, because they operate at very high temperatures (800–1100C<br />

[1]), material per<strong>for</strong>mance and manufacturing costs are currently <strong>of</strong> concern. Recently,<br />

attempts have been made to lower the operating temperature <strong>of</strong> SOFCs by adopting a porous,<br />

anode-supported structure to reduce the thickness <strong>of</strong> the electrolyte.<br />

It is important to model all physical processes and chemical reactions simultaneously since<br />

the mass and charge transport depends on the multifunctional material structure in the porous<br />

electrodes, the chemical reactions, the temperature distribution and the species concentrations.<br />

The fluid flow depends on the chemical reactions, temperature and fluid characteristics. The<br />

heat transport depends on the polarization losses, the chemical reactions, the fluid flow and<br />

the material structure. The re<strong>for</strong>ming reactions depend on temperature, concentration and<br />

amount <strong>of</strong> catalyst available. The interaction issues are numerous and here the microscopic<br />

contributions are important to include. Also it is necessary to model SOFCs into detail and<br />

explore them at a microscopic level, in order to fully understand how different parameters<br />

affect the per<strong>for</strong>mance, by connecting different physical phenomena at different scales [2, 4].<br />

The purpose <strong>of</strong> this study is to develop a microscale model <strong>of</strong> an anode <strong>of</strong> an SOFC <strong>for</strong> the<br />

transport processes and the chemical reactions to get a deeper understanding <strong>of</strong> the effect on<br />

the different physical processes at multiple scales. The lattice Boltzmann method (LBM) is<br />

used to model the mass transport at microscale <strong>for</strong> a limited part <strong>of</strong> the porous anode. Also an<br />

evaluation <strong>of</strong> a macroscale model <strong>for</strong> the whole unit cell is carried out.<br />

1


1.1 Research objectives<br />

The knowledge <strong>of</strong> the effect <strong>of</strong> different processes at microscale, such as mass diffusion and<br />

electrochemical reactions, on the unit cell in whole is expected to be clarified when the model<br />

includes microscale phenomena within the electrodes and electrolyte. When these processes<br />

are studied, the effect on the overall per<strong>for</strong>mance <strong>of</strong> the cell can provide useful in<strong>for</strong>mation<br />

<strong>for</strong> improvement [5].<br />

The aim is two-fold. Firstly, it provides a description <strong>of</strong> current research on modeling <strong>of</strong><br />

transport processes and kinetics effect on transfer processes in the electrodes <strong>of</strong> SOFCs.<br />

Focus is put on LBM to model the microstructural phenomena. Further, other modeling<br />

methods and equations are briefly reviewed as well as the coupling <strong>of</strong> LBM to conventional<br />

CFD methods. Secondly, it reveals a microscale model <strong>of</strong> an anode <strong>of</strong> an SOFC using the<br />

LBM to carefully investigate the physical and chemical processes at smaller scales which<br />

simulates the mass diffusion and momentum transport <strong>for</strong> a small part <strong>of</strong> the anode close to<br />

the electrolyte. A macroscale model <strong>of</strong> an anode-supported SOFC is also developed where the<br />

equations <strong>for</strong> mass, heat and momentum transport are solved simultaneously. Different<br />

internal re<strong>for</strong>ming reaction rates are tested to examine the effect on the cell. Also the<br />

parameters inlet methane content and steam-to-fuel ratio are tested <strong>for</strong> different ranges. The<br />

kinetics <strong>for</strong> the macroscale model are carefully investigated to check so no severe heat and<br />

mass transfer limitations occur at microscale. Further, the future step is to model all processes<br />

at microscale where microstructural effects are considered to affect the per<strong>for</strong>mance and also<br />

integrate the LB model and the macroscale CFD model, i.e., coupling different physical<br />

models at different scales, to <strong>for</strong>m a multiscale model which can reproduce even more<br />

realistic simulation results. More precisely, the objectives here are:<br />

<br />

<br />

<br />

To identify whether the LBM can function as a method <strong>for</strong> SOFCs at microscale to<br />

investigate the transport processes and chemical reactions, in terms <strong>of</strong> capabilities and<br />

limitations.<br />

To capture and study the microscale effect <strong>of</strong> mass diffusion and momentum<br />

transport within the anode close to the active area.<br />

To identify whether diffusion and chemical reactions will significantly affect the cell<br />

per<strong>for</strong>mance by modeling them at both macro- and microscale.<br />

1.2 Methodology<br />

In order to analyze the transport processes and chemical reactions in SOFCs in detail, an LB<br />

model <strong>for</strong> the porous region close to the three phase boundary (TPB) is developed in<br />

MATLAB. The LB model uses the single relaxation time BGK (Bhatnagar-Gross-Krook)<br />

method and is developed <strong>for</strong> a D2Q9 (two dimensional domain with eight interconnected<br />

directions and nine interconnected speeds) [3]. The model focuses solely on the mass<br />

diffusion and momentum transport at microscopic level in this region. The simulation<br />

procedure is divided into stepwise cases, from a simple channel to a more complex porous<br />

media. The macroscale model <strong>of</strong> an SOFC is developed in COMSOL where the equations <strong>for</strong><br />

mass, heat and momentum transport are solved simultaneously. Finally, a study <strong>of</strong> limiting<br />

effects on the heat and mass transfer by the kinetics is also per<strong>for</strong>med.<br />

1.3 <strong>Thesis</strong> outline<br />

Chapter 1 contains a short presentation <strong>of</strong> the thesis. Chapter 2 gives a general description <strong>of</strong><br />

SOFCs, and an overview <strong>of</strong> the relevant literature <strong>of</strong> the LB methodology. A detailed<br />

description <strong>of</strong> the mathematical model <strong>of</strong> the LBM is presented in Chapter 3 with governing<br />

equations and boundary conditions. The results are presented in Chapter 4 while Chapter 5<br />

provides conclusions drawn from the results. Finally, Chapter 6 gives reflections over future<br />

work.<br />

2


2 SOFC<br />

modeling at smaller scales<br />

This chapter gives a short description <strong>of</strong> SOFCs and fuel cell modeling at different scales is<br />

described with focus on LBM.<br />

2.1 Solid Oxide Fuel Cells<br />

Fuel cells directly convert the free energy <strong>of</strong> a chemical reactant to electricity and heat. This<br />

is different from a conventional thermal power plant, where the fuel is oxidized in a<br />

combustion process and subsequently a conversion process (thermal-mechanical-electrical<br />

energy) occurs. Fuel cells have high energy conversion efficiency due to the direct<br />

conversion. If pure hydrogen is used, there is no destructive environmental pollution, because<br />

the output from the fuel cells is electricity, heat and water.<br />

Among various types <strong>of</strong> fuel cells, the SOFC has attracted significant interest thanks to its<br />

high efficiency and low emissions <strong>of</strong> pollutants like carbon dioxide and hazardous gases. The<br />

SOFC’s high operating temperature <strong>of</strong>fers many advantages, such as high electrochemical<br />

reaction rates, flexibility in choice <strong>of</strong> fuel and high tolerances <strong>for</strong> impurities. However, fuel<br />

cell systems are still immature technologies, as can be noted in the lack <strong>of</strong> a dominant design,<br />

low number <strong>of</strong> commercial systems and a low market demand. The creation <strong>of</strong> strategic niche<br />

markets is <strong>of</strong> a vital importance <strong>for</strong> further development [4].<br />

The International Energy Agency (IEA) has concluded in many reports that the fuel cell will<br />

be a key component in a future sustainable energy system. About 80 percent <strong>of</strong> the energy<br />

resources traded today are fossil fuels (coal, oil and natural gas) [1] and these resources are<br />

considered limited. Here the SOFC can be a key component facilitating the transition towards<br />

more environmentally friendly energy generation. It is possible to use conventional fuels as<br />

natural gas to ease the transition to hydrogen based power generation without emissions <strong>of</strong><br />

pollutants. During recent years, another promoter <strong>of</strong> interest in using fuel cells as auxiliary<br />

power units (APUs) in on-board transport applications, <strong>for</strong> example in luxury passenger<br />

vehicles, military vehicles and leisure boats has increased immensely [1].<br />

SOFCs can function with a variety <strong>of</strong> fuels, e.g., hydrogen, carbon monoxide, methane and<br />

combinations <strong>of</strong> these [5]. Because SOFCs operate at high temperature, they supply a<br />

sufficiently good environment to internally re<strong>for</strong>m the hydrocarbon-based fuel within the cell.<br />

The fuel flexibility gives the SOFCs a major advantage over pure hydrogen, which is highly<br />

flammable and volatile and there<strong>for</strong>e problematic to handle. Also, hydrogen has low density,<br />

which makes storage costly [4]. It should also be mentioned that pure hydrogen is difficult to<br />

obtain because it has to be extracted from another source, most commonly natural gas or<br />

through electrolysis. Within the cell several reactions may take place and vary depending on<br />

which fuel is used. The overall global reactions <strong>for</strong> methane are stated below. More detailed<br />

surface reactions can be found in the literature [5, 6].<br />

3


(2.1)<br />

(2.2)<br />

(2.3)<br />

(2.4)<br />

(2.5)<br />

Equation (2.1) is the reduction <strong>of</strong> oxygen in the cathode. Equations (2.2) and (2.3) are the<br />

electrochemical reactions at the anodic three phase boundary (TPB). TPB is the region where<br />

the ionic, electronic and gas phases meet close to the electrolyte and the electrode interface.<br />

Equation (2.4) is the steam re<strong>for</strong>ming <strong>of</strong> methane, which needs to be carried out prior to the<br />

electrochemical reactions, and is usually called catalytic steam re<strong>for</strong>ming reaction. Carbon<br />

monoxide can be oxidized as in equation (2.3) or react with water as in equation (2.5).<br />

Equation (2.5) is <strong>of</strong>ten called water-gas shift reaction. Note that methane is not participating<br />

in the electrochemical reactions at the anodic TPB. It is catalytically converted, within the<br />

anode, into carbon monoxide and hydrogen, which are used as fuel in the electrochemical<br />

reactions [5-7]. These reactions can be viewed in the schematic illustration <strong>of</strong> an SOFC in<br />

Figure 2.1.<br />

Figure 2.1: Schematic SOFC structure.<br />

The configuration <strong>of</strong> an SOFC can be <strong>for</strong>med in different ways; electrolyte or electrode<br />

supported cells, shaped in a planar or tubular manner, co or counter directional flow. An<br />

electrolyte supported SOFC has thin anodes and cathodes (~50 m), and the thickness <strong>of</strong> the<br />

electrolyte is more than 100 m. An electrolyte supported SOFC works preferably at<br />

temperatures around 1000 °C. In an electrode-supported SOFC, either the anode or the<br />

cathode is thick enough to serve as the supporting substrate <strong>for</strong> cell fabrication, normally<br />

4


etween 0.3 and 1.5 mm. In this configuration the electrolyte is thin (could be as thin as 10<br />

m) [5-7].<br />

SOFC research in the last years has focused on the electrode-supported configuration to lower<br />

the operating temperature. Positive outcomes <strong>of</strong> development in this direction are a decrease<br />

in start-up and shut-down time and, simplified design and cell material requirement. Electrode<br />

supported design makes it possible to have a very thin electrolyte, i.e., the ohmic losses<br />

decrease and the temperature can be lowered to 600-800 °C. Fuel cells working at those<br />

temperatures are classified as intermediate temperature ones if compared to conventional<br />

SOFCs that operate at 800-1000 °C [6]. Corrosion rates are significantly reduced and stack<br />

lifetime is extended. Lowering the operating temperature to an intermediate range will cause<br />

an increase <strong>of</strong> both ohmic- and polarization losses in the electrodes. This requires the<br />

development <strong>of</strong> a highly active electrolyte that has low polarization loss at intermediate<br />

temperatures. Possible electrolyte materials could be doped ceria or doped lanthanum gallate<br />

[7-8]. The SOFC in this case is built up by an electrolyte containing yttria-stabilized zirconia<br />

(YSZ) and a cathode containing strontium doped lanthanium manganite (LSM), and finally,<br />

the anode nickel/YSZ [1, 7].<br />

2.2 SOFC modeling development<br />

Be<strong>for</strong>e designing and constructing a model, it is important to specify what is needed and why.<br />

The selection <strong>of</strong> computational methods must come from a clear understanding <strong>of</strong> both the<br />

in<strong>for</strong>mation being computed and the actual physical processes implemented. In order to solve<br />

some <strong>of</strong> the remaining issues <strong>for</strong> the understanding <strong>of</strong> the detailed physical phenomena <strong>of</strong><br />

SOFC, the computational modeling is the crucial factor. The microstructure is one <strong>of</strong> the least<br />

understood areas <strong>of</strong> research <strong>of</strong> the SOFC. As increased computational capability opens up<br />

<strong>for</strong> more detailed research, today’s fuel cell research challenge is to fully understand<br />

microscale and nanoscale transport phenomena in the active electrochemical material and<br />

further, connect these to macroscale to <strong>for</strong>m a multiscale model.<br />

2.2.1 Multiscale and multiphysics modeling<br />

While multiphysics modeling involves the coupling and interaction between two or more<br />

physical disciplines, the multiscale approach involves the connection <strong>of</strong> specific physical<br />

processes, based on the different levels <strong>of</strong> scale: micro, meso and macro. The division and<br />

ranges <strong>of</strong> scales varies slightly in the literature but an attempt is made here to divide the scales<br />

to suit the modeling perspective in this study. The microscale model corresponds to the<br />

particle level (~10 -6 – 10 -9 m). Also even smaller scales, nanoscales, may be integrated but are<br />

not further studied in this work. Here the mesoscale and macroscale stretch from a scale<br />

larger than a particle to the global flow field (~10 -5 – 10 1 m) [1, 6]. To understand the<br />

multiscale concept, one needs first to understand the different scales involved. Not only<br />

proper length scales need to be considered, but also different time scales. Convective<br />

transport appears in 10 -1 s, cell heating and anode thermal diffusion are in 10 3 s and cathode<br />

thermal diffusion appears in 10 4 s [5]. A relation between time- and length scales with proper<br />

modeling methods can be seen in Figure 2.2. Only the microscale and mesoscale are shown<br />

here but the division can be done differently or in more intermediate steps.<br />

5


Figure 2.2: Characteristic time and length scales <strong>for</strong> various modeling methods. Data is taken from [1].<br />

At macroscale, homogeneity is assumed throughout the model, which subsequently can cause<br />

errors during the loop <strong>of</strong> the modeling algorithm. Several models are based on the assumption<br />

that the porous structure is isotropic and can be described by a few experimentally estimated<br />

parameters [4, 8]. Porosity, tortuosity, and surface area to volume ratio are examples <strong>of</strong><br />

parameters that are affected by assumptions concerning homogeneity at all scales. Note that<br />

these microstructural parameters are known to have a significant influence on the cell<br />

per<strong>for</strong>mance and durability [9]. As the available computational power increases, it opens up<br />

<strong>for</strong> a more sophisticated and deeper understanding <strong>of</strong> the physical processes and effects <strong>of</strong><br />

chemical reactions within the porous microstructure. This makes it possible to address the<br />

microstructural uncertainties to improve cell per<strong>for</strong>mance, as these are limiting the SOFC<br />

progress [10].<br />

Multiphysics modeling takes into account the interaction between several physical processes,<br />

which can be described by partial differential equations. A good computational design<br />

considers the physical processes and the system at both a microscale and a macroscale level.<br />

Some <strong>of</strong> the limitations are the lack <strong>of</strong> material structure and test data in the literature to<br />

validate the models [11, 12]. The results <strong>of</strong> a numerical simulation cannot guarantee <strong>of</strong> how<br />

well the cell actually will operate in reality. Because <strong>of</strong> the numerical approximations and<br />

arbitrary unknowns implemented in the model, there will most likely be a number <strong>of</strong> errors<br />

and inaccurate results [11]. Still, the use <strong>of</strong> numerical modeling as a predictive tool can be<br />

validated through careful consideration <strong>of</strong> results and comparisons <strong>of</strong> numerical and<br />

experimental data. A great deal <strong>of</strong> computational modeling research, where the results are<br />

obtained from numerical codes, has achieved sufficient accuracy both in comparison with<br />

other different numerical modeling approaches and with experimental data [10-11].<br />

Fuel cell modeling is complicated due to the interaction <strong>of</strong> physical and chemical processes,<br />

such as multicomponent gas flow with heat and mass transfer, electrochemical and re<strong>for</strong>ming<br />

reactions [10-11]. To model SOFCs, it is common to handle the governing equations in<br />

differential <strong>for</strong>ms by deriving them in <strong>for</strong>ms <strong>of</strong> discretized equations. These equations are<br />

solved numerically by the Gaussian-elimination method or the Tri-diagonal matrix algorithm.<br />

There are several approaches to solve these by numerical methods. For macroscale, in a<br />

simplified manner, one may say that the methods differ in the sense <strong>of</strong> how the flow variables<br />

are approximated. The commercial s<strong>of</strong>tware which is currently available is mainly based on<br />

the Finite Difference Method (FDM), the Finite Element Method (FEM) and the Finite<br />

6


Volume Method (FVM) which adopt the macroscopic structure [11-12]. Examples <strong>of</strong><br />

commercial s<strong>of</strong>tware are FLUENT and STAR-CD, based on FVM, and COMSOL<br />

Multiphysics, based on FEM. Through computational modeling, the output can provide<br />

details <strong>of</strong> the processes, such as the fuel cell species distribution, flow patterns, current<br />

density, temperature distribution and pressure drop, etc. [11]. The simulation environment in<br />

commercial s<strong>of</strong>tware facilitates all steps in the modeling process. It is easy to define the<br />

geometry and the mesh as well as specify the physics <strong>for</strong> the domain. As fuel cell testing is<br />

expensive and time consuming, a careful simulation study be<strong>for</strong>e testing can lower the cost<br />

<strong>for</strong> research activities. For this reason, numerical modeling <strong>of</strong> SOFCs is necessary.<br />

2.2.2 Lattice Boltzmann concept<br />

Advances in micro-modeling have been made thanks to better availability <strong>of</strong> computational<br />

power, where lattice Boltzmann mass diffusion models have been found to predict and<br />

visualize the phenomena well in the microstructure and the capability to simulate not only<br />

single phase or multiphase flow but also both <strong>of</strong> these in complex geometries [3]. Based on<br />

the Boltzmann equation, the model is considered to be an alternative to the traditional CFD<br />

based on the Navier-Stokes equations, without any empirical modifications. The idea <strong>of</strong> the<br />

LB model is that it is viewed from a particle perspective and based on a statistical approach to<br />

track a large number <strong>of</strong> particles. The framework is built upon interaction between the<br />

particles, e.g., collision, either particle-to-particle or particle-to-surface interaction, and<br />

streaming [3].<br />

LBM has served as a numerical in<strong>for</strong>mation bank with detailed simulation results <strong>for</strong> a large<br />

number <strong>of</strong> physical processes. In most cases, previous work on LBM in both the continuum<br />

and non-continuum flow regime has focused on single component flow problem. However,<br />

there is still a lack <strong>of</strong> results by applying the LBM with more than one species especially at<br />

high Kn in complex geometries. This is the case <strong>for</strong> the porous electrodes in SOFCs where H 2<br />

and H 2 O diffuse or internal re<strong>for</strong>ming <strong>of</strong> hydrocarbons occurs [11, 14-16].<br />

Lattice gas cellular automation models have acted as the <strong>for</strong>erunners <strong>of</strong> the LBM. Like them<br />

the LBM is based on a concept with an algorithmic entity at a position connected to its<br />

neighbors. The next step in the development <strong>of</strong> LBM was when the basic idea <strong>of</strong> Boltzmann’s<br />

work was included. The idea is based on a gas, composed <strong>of</strong> interacting particles as described<br />

by classical mechanics [3]. The LBM simplifies Boltzmann’s original view by reducing the<br />

number <strong>of</strong> participating and possible spatial particle positions. LBM tracks the statistical<br />

behavior <strong>of</strong> the molecules at each lattice point. Distinct steps have been developed <strong>for</strong> the<br />

microscopic momentum and distribution paths. The spatial positions are confined to the<br />

lattice nodes, and the variations <strong>of</strong> momentum due to velocity changes, are reduced to 8<br />

directions, 3 magnitudes and a single mass in the 2D case [3]. LBM is built up on lattice units<br />

which need to be converted to actual physical properties after the simulation [14]. This is<br />

<strong>of</strong>ten handled through the mole fraction when the mass diffusion is the main focus. The<br />

electrochemical reactions and also the re<strong>for</strong>ming reactions are then coupled with the mass<br />

diffusion to LBM via mole flux boundary conditions at the active surface. To obtain these<br />

dimensionless values, the methodology <strong>for</strong> the LBM needs to be described in detail which is<br />

carried out in a latter section (see section 2.3.2).<br />

2.2.3 Monte Carlo method and Molecular dynamics<br />

A number <strong>of</strong> nano- and microscale models have been developed to simulate microscopic<br />

transport phenomena in SOFCs. The atomistic model refers to a broad group <strong>of</strong> algorithms<br />

and can provide detail in<strong>for</strong>mation to make realistic boundary and interface definitions <strong>for</strong> a<br />

larger scale model [17]. Monte Carlo (MC) methods are based on algorithms which through<br />

repeated random steps can be used in simulating physical and mathematical systems [18]. The<br />

system is propagated through time by stochastically establishing the coming event based on a<br />

relative probability <strong>of</strong> each possible event [18]. The probability is <strong>of</strong>ten determined by the<br />

7


intrinsic rate constant <strong>of</strong> each event. These methods are most suited <strong>for</strong> computational<br />

calculations when it is unfeasible to compute an exact result with a deterministic algorithm<br />

[19]. The MC methods is especially useful <strong>for</strong> simulating systems with many coupled<br />

parameters, such as fluids and disordered materials, but un<strong>for</strong>tunately also tend to have a high<br />

computational cost. MC methods vary, but tend to follow a particular pattern: Define a<br />

domain <strong>of</strong> possible inputs, generate inputs randomly from a probability distribution over the<br />

domain, per<strong>for</strong>m a deterministic computation on the inputs and aggregate the results.<br />

MC is a modeling method <strong>for</strong> the dynamic behavior <strong>of</strong> molecules by comparing the rates <strong>of</strong><br />

individual steps with random numbers. The method is used to investigate non-equilibrium<br />

systems or the time evolution <strong>of</strong> some specific processes occurring in nature, typically<br />

processes that occur with a known rate. The probability needs to be defined prior to the<br />

simulation [19]. MC was proven functional by a study by Lau et al. [19], conducted on a<br />

cathode <strong>of</strong> an SOFC <strong>for</strong> the oxygen reduction reaction. Lau et al. [19] found that the<br />

temperature had a great effect throughout the simulation which in this case was on the ionic<br />

current density.<br />

One difference between Monte Carlo and Molecular dynamics (MD) modeling is the different<br />

time scales. MC can handle longer time scales, typically seconds, whereas MD handles time<br />

scales around microseconds or even smaller [19]. To make statistically valid conclusions from<br />

the simulations, the time span simulated should match the kinetics <strong>of</strong> the natural process. The<br />

modeling should allow <strong>for</strong> different reaction pathways [17]. The relevant parameters are<br />

possible to be captured by both MC and MD. However, one would prefer to use the models<br />

<strong>for</strong> different processes depending on the needed elapsed time.<br />

MD simulates physical movements <strong>of</strong> atoms and molecules by computer modeling based on<br />

statistical mechanics. MD allows insight into molecular motion on an atomic scale and<br />

detailed time and space resolution into representative behavior <strong>for</strong> carefully selected systems<br />

[20]. MD can be used to model local structures at elevated temperatures where it is<br />

challenging to per<strong>for</strong>m experiments with conclusive outcomes. MD has been used to per<strong>for</strong>m<br />

simulations on diffusion and ion transport with successful outcome [20].<br />

2.2.4 Modeling integration issues<br />

Physical problems can <strong>of</strong>ten be described with a set <strong>of</strong> partial differential equations. The<br />

coupled partial differential equations can be solved simultaneously in physical domains <strong>for</strong><br />

corresponding physical phenomena. The integration issues occur because the physical and<br />

chemical processes are linked to each other and so also the equations in the model [11]. The<br />

mass, heat, momentum transport as well as ionic/electronic transport and the reaction rate are<br />

dependent on each other. The fluid properties and the momentum transport (flow field)<br />

depend on the temperature and concentrations and so does the chemical reactions. The<br />

chemical reactions generate and consume heat, i.e., the temperature distribution depends on<br />

the chemical reaction rate, as well as on the solid and the gas properties, <strong>for</strong> example the<br />

specific heat and the thermal conductivity [1, 11].<br />

Until now, multicomponent gas diffusion has been modeled in the continuum flow regime <strong>for</strong><br />

the porous electrodes using the Stefan-Maxwell equations by several research groups. Some<br />

<strong>of</strong> the equations used in our CFD-model, e.g., the Stefan-Maxwell model (SMM), are<br />

described based on a few empirical parameters, which are difficult to measure [14]. To<br />

enhance the knowledge <strong>of</strong> the impact <strong>of</strong> the transport process on the per<strong>for</strong>mance within the<br />

electrodes, the microstructure needs to be modeled in detail. Recent advances have made it<br />

possible to evaluate the microstructure using LBM without any modification <strong>of</strong> empirical<br />

parameters. Models have been developed <strong>for</strong> all the three spatial dimensions, but so far the<br />

main focus has been on particular parts <strong>of</strong> the fuel cell, e.g., anode, and also simple<br />

geometries [14-16].<br />

8


2.3 Lattice Boltzmann method<br />

LBM has shown promising simulation results <strong>of</strong> fluid flows and mass diffusion through<br />

complex geometries and this is an attractive characteristic <strong>for</strong> fuel cell modeling.<br />

Conventional CFD methods use fluid density, velocity and pressure as their primary<br />

variables, while the LBM uses a more fundamental approach with a so-called particle velocity<br />

distribution function (PDF) [16]. The PDF is here denoted f a and originates from the basic<br />

Boltzmann gas concepts where a derived simplified <strong>for</strong>m <strong>of</strong> the Boltzmann equation is<br />

described by classical mechanics with statistical treatment based on the high number <strong>of</strong><br />

particles. The distribution is described by the coordinates <strong>of</strong> the position and momentum<br />

vectors, and the time step [3].<br />

The LBM framework is built on lattice points, which are given locations placed all over the<br />

solution domain. The lattice unit (lu) is a fundamental measure <strong>of</strong> length and the time step (ts)<br />

is the measure <strong>of</strong> the time unit. The neighboring particles are connected to the main focused<br />

particle at the time with the velocity magnitude e a , schematically described in Figure 2.3 from<br />

a 2D point <strong>of</strong> view. This type <strong>of</strong> structural framework is called D2Q9 and stands <strong>for</strong> two<br />

dimensions with nine velocities marked as e a , where a represents the direction (= 0, 1, 2... 8)<br />

[3, 16]. For the 2D case, there exist two appropriate choices <strong>of</strong> system structure, namely<br />

D2Q5 and D2Q9 with 5 and 9 directional velocities, respectively. The D2Q9 framework has<br />

the possibility to capture more in<strong>for</strong>mation but will be computationally heavier than the<br />

smaller one D2Q5.<br />

Figure 2.3: Schematic structural framework <strong>for</strong> the D2Q9 lattice and velocities.<br />

The PDF is defined as the number <strong>of</strong> particles <strong>of</strong> the same species travelling along a particular<br />

direction with a particular velocity. The single-particle distribution function f a can essentially<br />

be seen as a histogram representing a frequency <strong>of</strong> occurrence. The frequencies can be<br />

considered to be direction specific fluid densities. The LBM is described by two different<br />

actions taking part at each lattice point (site); namely streaming and collision. Streaming<br />

describes the movement <strong>of</strong> the particles <strong>of</strong> each species and the collision describes the<br />

interactions between the particles <strong>of</strong> the same or different species. Furthermore, these actions<br />

are combined in the LB equation called the distribution function [3, 14-16].<br />

9


(2.6)<br />

where f a is the PDF, e a the velocity and a the collision term at any spatial location x and time<br />

t along the direction a. The time is increased by the time step Δt. The macroscopic fluid<br />

density is [3]:<br />

(2.7)<br />

The macroscopic velocity u is evaluated by the microscopic velocities e a and the PDF f a and<br />

divided by the macroscopic fluid density ρ as [3]:<br />

(2.8)<br />

This allows the LBM to recover the continuum macroscopic parameters from the discrete<br />

microscopic ones, in this case by the velocities. The distribution function presented in<br />

equation (2.6), called the single relaxation time BGK (Bhatnagar-Gross-Krook) LBM, is one<br />

<strong>of</strong> the simplest models [3, 14]. The BGK is described by using one relaxation time <strong>for</strong> the<br />

collision term. The collision term consists <strong>of</strong> the present PDF and the relaxation toward the<br />

local equilibrium. The collision term Ω a and the D2Q9 equilibrium distribution function f a<br />

eq<br />

are defined as [3]:<br />

(2.9)<br />

(2.10)<br />

where w a is 4/9 <strong>for</strong> the particle a = 0, 1/9 <strong>for</strong> a = 1, 2, 3, 4 and 1/36 <strong>for</strong> a = 5, 6, 7, 8, and τ is<br />

the relaxation number [14-16]. In the simplest implementation the basic speed on the lattice c,<br />

which is also called the lattice speed <strong>of</strong> sound, is 1 lu/ts [16].<br />

When the mass diffusion is modeled in LBM, two approaches are <strong>of</strong>ten used; pure diffusion<br />

or advection-diffusion (also called convection-diffusion). Both pure diffusion and advectiondiffusion<br />

is simulated by another equilibrium distribution f ζ,a eq which is very much alike the<br />

normal fluid distributions function but with a simpler equilibrium equation. For the first case<br />

with pure diffusion only the equilibrium function is defined as [3]:<br />

(2.11)<br />

In the second case, advection-diffusion, which is applied here, the equilibrium function will<br />

include a second term to handle the convective velocity. The equilibrium function is defined<br />

as [3]:<br />

(2.12)<br />

The mixing due to density variations and buoyant effects in porous media can here be handled<br />

as advective and diffusive components rather than an input parameter (such as porosity). For a<br />

porous media, the collision term is considered as a second intermediate step after the<br />

10


streaming [3]. The concentration ρ ζ is defined similarly as the fluid density in equation (2.7)<br />

[3]:<br />

(2.13)<br />

To avoid numerical instability, it is recommended to keep the relaxation times at the order <strong>of</strong><br />

unity. This would mean that the diffusivity values are adjusted by scaling them up to ensure<br />

relaxation times <strong>of</strong> unity [14-16].<br />

The LB equation can be extended to include several components or species. The equations<br />

remain the same <strong>for</strong> each species but the interaction and combination <strong>of</strong> the streaming and<br />

collision action <strong>for</strong> species i is defined as [3, 14-16].<br />

(2.14)<br />

where f i a is the PDF, e i a the velocity and<br />

time t along the direction a <strong>for</strong> species i.<br />

a i the collision term at any spatial location x and<br />

Streaming is described in two steps. The PDFs, <strong>for</strong> example <strong>for</strong> species 1, are streamed from<br />

one lattice point to the adjacent lattice points, while PDFs <strong>for</strong> the remaining species, i.e.,<br />

species 2 and 3, are streamed from one lattice point to <strong>of</strong>f-lattice points in the first step. Offlattice<br />

points are defined as sites which are not related to the current start point. In the second<br />

step, the PDF values <strong>for</strong> species 2 and 3 are determined by interpolation at lattice points. The<br />

collision concept is divided into self-collision, i.e., collision between particles <strong>of</strong> the same<br />

species and cross-collision, i.e., collision between particles <strong>of</strong> different species when the<br />

relative velocity between particles <strong>of</strong> the different species is non-zero [14-16].<br />

For the porous media, the collision term is considered as a second intermediate step after the<br />

streaming [3]. Finally if there are multiple species, the LB model is adjusted by upgrading the<br />

distribution function by also including a composite velocity:<br />

(2.15)<br />

where τ is the relaxation time and i represents the different components involved [3].<br />

To <strong>for</strong>m a realistic model, it is possible to include an external <strong>for</strong>ce, e.g., gravitational <strong>for</strong>ce,<br />

on the particles or interaction <strong>for</strong>ces between the particles. This <strong>for</strong>ce is incorporated in a<br />

velocity term which is added to the velocity calculation. These parts are defined as [3]:<br />

(2.16)<br />

(2.17)<br />

where F is the <strong>for</strong>ce acting on the particle, m the mass and a the acceleration <strong>of</strong> the particle.<br />

Further, u is the velocity, ρ the density and τ the relaxation time.<br />

11


2.3.1 Boundary conditions in LBM<br />

A common choice <strong>of</strong> boundary conditions <strong>for</strong> the mole fractions are specified at x = 0 and the<br />

mole flux is specified at x = L. The boundary conditions <strong>for</strong> velocity and density need to be<br />

described indirectly by the PDFs in the LBM. This differs from the Navier-Stokes equations<br />

in conventional CFD as the boundary conditions are prescribed directly <strong>for</strong> velocity and<br />

density [15]. For the implementation <strong>of</strong> the electrochemical activity in the mass transfer<br />

framework <strong>of</strong> LBM, the specific boundary nodes need to be specified as discrete Neumann or<br />

Dirichlet boundary conditions. The Dirichlet boundary condition is associated with<br />

concentrations specified at the pores along the inlet <strong>of</strong> the domain. The Neumann boundary<br />

condition or flux boundaries are employed similarly to the Dirichlet boundaries. Instead <strong>of</strong><br />

defining the concentration, the particle velocities are explicitly defined at the boundary nodes.<br />

The electrochemical kinetic mechanism can be defined as active and assigned a unique flux at<br />

each node. In turn this affects the particle velocity as a function <strong>of</strong> location within the domain<br />

[14, 21-23].<br />

To treat the electrode structure realistically, it will be built up <strong>of</strong> open space and solid<br />

obstacles where the obstacles are treated like impermeable solid surfaces. The velocities at the<br />

solid obstacles need to be reset to zero at each time step <strong>for</strong> all species The obstacles are<br />

suited <strong>for</strong> three different boundary conditions <strong>for</strong> the velocities; no-slip, free-slip and diffuse<br />

reflection. The no-slip condition is simply a bounce-back definition, as the particles are<br />

reflected back in the same direction as they arrive. For the free-slip, the particles are reflected<br />

back in the angle <strong>of</strong> reflection, which is set to be the same as the angle <strong>of</strong> incidence. For the<br />

diffuse reflection, the particles are reflected with the same probability in every direction [14].<br />

Joshi et al. [15] showed that the implementation <strong>of</strong> the three different boundary conditions<br />

obtained approximately the same results. LBM makes no conceptual distinction between the<br />

transport <strong>of</strong> single species over an obstacle or multiple species over the same object in<br />

opposite directions. The only difference will be the impact <strong>of</strong> the motion <strong>of</strong> species by the<br />

collision [15].<br />

Finally to summarize the whole procedure, the mole fractions are initially assigned specific<br />

values over the whole domain. The velocities <strong>of</strong> the species in the perpendicular direction to<br />

the flow direction (y-direction) are initially set to zero, and are always set to zero <strong>for</strong> the<br />

boundary <strong>of</strong> the obstacle. The following algorithm is repeated until a steady solution has been<br />

reached. The calculation <strong>of</strong> the equilibrium functions is followed by that <strong>of</strong> the collision terms<br />

and subsequently by streaming and interpolation <strong>of</strong> the PDF values. The unknown values at x<br />

= 0 and x = L are calculated and followed by a calculation <strong>of</strong> the values at the boundaries <strong>of</strong><br />

the obstacles. Finally, all the physical parameters are calculated, such as density, velocity etc.<br />

When the total concentration has been checked and a steady solution has been reached, the<br />

modeling is done. Obviously, all the restriction and constraints applied to the model need to<br />

be fulfilled [21, 23-24].<br />

2.3.2 Choice <strong>of</strong> units in LBM<br />

As the LBM parameters are not in physical units when simulated but in so-called lattice units,<br />

this part <strong>of</strong> the algorithm procedure needs extra attention. The approach here is divided into<br />

two steps. First, the physical system is converted to a dimensionless one. This system is<br />

independent <strong>of</strong> both the physical scales and the modeling parameters. Second, this system is<br />

converted into a discrete modeling system.<br />

To exemplify, the incompressible Navier-Stokes equation depends only on a single<br />

dimensionless parameter, namely the Reynolds number Re. Consequently, both the physical<br />

and the dimensionless system will be connected to have the same Re. But there will still be a<br />

need <strong>for</strong> a transition between the two systems, and this is done by a characteristic length scale<br />

l 0 and a characteristic time scale t 0 . The transition from the dimensionless to the discrete<br />

system is done by implementing a discrete length step δ x and a discrete time step δ t [25]. To<br />

12


ease the correspondence between the three systems, the physical system is titled P, the<br />

dimensionless D and the discrete LB. One could also go directly from the physical system (P)<br />

to the discrete system (LB). But the variables δ x and δ t are important <strong>for</strong> the accuracy and<br />

stability <strong>of</strong> the numerical simulation, and in<strong>for</strong>mation regarding them may be lost if the direct<br />

approach is used. The procedure <strong>for</strong> the transition is schematically summarized in Figure 2.4.<br />

Figure 2.4: Transition scheme between the three systems [25].<br />

To illustrate the process <strong>of</strong> choosing units, an example <strong>for</strong> a flow <strong>of</strong> an incompressible fluid<br />

developed by Latt [25] is outlined here. For an incompressible fluid, the density can be<br />

assumed to stay constant; does not vary in time and space, throughout the domain. The<br />

Navier-Stokes equation is <strong>of</strong>ten chosen <strong>for</strong> describing the motion <strong>of</strong> the fluid and is governed<br />

by the conservation <strong>of</strong> mass and momentum. The conservation <strong>of</strong> mass is here written as:<br />

(2.18)<br />

where u is the velocity <strong>of</strong> the fluid and the index P stands <strong>for</strong> the physical system. The<br />

conservation <strong>of</strong> momentum is written as:<br />

(2.19)<br />

where p is the pressure, ρ the density and υ the kinematic viscosity [25]. The next step is to<br />

convert the system into a dimensionless one. Further, two scales are introduced; a length scale<br />

l 0 , e.g., approximate size <strong>of</strong> an obstacle, and a time scale t 0 , e.g., time to pass the obstacle.<br />

These are used to scale the variables between the systems.<br />

(2.20)<br />

(2.21)<br />

where x is a position vector in a numerical modeling environment [25]. Similarly, the other<br />

variables are scaled and then replaced in the equation <strong>for</strong> conservation <strong>of</strong> mass and<br />

momentum.<br />

(2.22)<br />

(2.23)<br />

where the Reynolds number Re is defined as [25]:<br />

13


(2.24)<br />

where v is the viscosity. For two flows with equivalent geometry and same Re, these will<br />

obey Reynolds transport theorem and provide equivalent solutions which can be converted<br />

from one flow to the other. Finally, the dimensionless system can now be trans<strong>for</strong>med into the<br />

discrete system by defining a reference length step δ x and time step δ t . If the reference<br />

variables, i.e., δ xD and δ tD , in the dimensionless system are defined, t 0D and x 0D will turn out to<br />

have the value <strong>of</strong> unity, respectively. Then the reference variables in the discrete system are<br />

defined as:<br />

(2.25)<br />

(2.26)<br />

where N is the number <strong>of</strong> cells and N iter the number <strong>of</strong> iterations [25]. Now, the conversion is<br />

easily handled through dimensionless analysis and the variables are defined as.<br />

(2.27)<br />

(2.28)<br />

Both δ x and δ t are attached with some constraints in LBM. The value <strong>of</strong> u LB may not be larger<br />

than the basic lattice speed <strong>of</strong> sound c even if the fluid is possibly compressible because LBM<br />

does not support supersonic flows [25]. This leads to the relationship <strong>of</strong> the reference<br />

variables as follows:<br />

(2.29)<br />

The LBM is a quasi-compressible solver which means it enters a slightly compressible regime<br />

when solving the pressure equation without any significant impact on the numerical accuracy.<br />

To specify how to choose δ t , it is important to note that the LB model is <strong>of</strong> second-order<br />

accuracy. This means that the lattice error ε(δ x ) ~ δ x 2 . The compressibility error would take<br />

over if the lattice error is reduced and there<strong>for</strong>e the errors should be kept at the same order,<br />

i.e., ε(Ma) ~ ε(δ x ). This leads to a more specific relationship between the reference variables,<br />

namely δ t ~ δ x 2 [25].<br />

2.3.3 Mass diffusion in LBM<br />

The next step here is to show the connection between variables <strong>for</strong> the mass diffusion. It<br />

should be mentioned that other physical processes can also be handled by LBM. For example<br />

thermal flow would be handled by setting the temperature as T ≡ ρ in equation (2.7).<br />

Mass diffusion in LBM can be divided into two cases, namely; pure diffusion (u = 0) or<br />

advection-diffusion (u ≠ 0) [26]. For LBM, this means that the equilibrium distribution<br />

function will differ by containing the velocity or not in the equation (see equations (2.11) and<br />

(2.12)). The LBM parameters in an equivalent system <strong>of</strong> lattice units should be such that<br />

diffusion fluxes are in the same ratio as the actual ones, only larger in magnitude [15-16].<br />

Mass diffusion in a mixture can be described by Fick’s law <strong>of</strong> diffusion. As Fick’s law is only<br />

14


applicable <strong>for</strong> a mixture <strong>of</strong> two species the equation <strong>for</strong> the mass diffusion is better described<br />

by the SMM [23-24]. The SMM is defined as:<br />

(2.30)<br />

where J i is the mole flux <strong>of</strong> species i, c T the total mole concentration, x i the mole fraction <strong>of</strong><br />

species i and D ij is the mass diffusivity between species i and j. Equation (2.31) is <strong>of</strong>ten<br />

difficult to solve as it is, and there<strong>for</strong>e simplified <strong>for</strong>ms are <strong>of</strong>ten applied to obtain an<br />

analytical or numerical solution. In this case, the connection to the physical spectra is made<br />

by the dimensionless diffusivity ratios, which are still in the same range as the actual<br />

diffusivity but <strong>of</strong> a larger magnitude [14-16]. The other parameters in the LBM, such as J i and<br />

X i are adjusted so that the same value <strong>of</strong> J* is maintained. For species i and j the<br />

dimensionless mole flux J* is defined in equation (2.15) which is used as the main parameter<br />

when the system is evaluated in the continuum regime [21, 23].<br />

(2.31)<br />

where the physical meaning <strong>of</strong> J* can be described as the mole fraction drop along the length<br />

L and D ij is the mass diffusivity between species i and j.<br />

2.3.4 Chemical reactions in LBM<br />

Mass transport and chemical reactions play a significant role in modeling the anode <strong>of</strong> a fuel<br />

cell. Here an approach is presented to solve the transport <strong>of</strong> a passive scalar reacting chemical<br />

species connected to mass diffusion. They will be assigned separate particle density<br />

distributions with different values <strong>of</strong> relaxation time and then they are coupled via the flow<br />

velocity. Only a passive scalar transport is used so there will be no feedback <strong>of</strong> the species<br />

distribution on the flow field [28]. A surface catalytic heterogeneous chemical reaction<br />

between two species A and B is considered with a reaction rate proportional to the<br />

concentrations <strong>of</strong> the species, C A and C B as:<br />

(2.32)<br />

(2.33)<br />

where k is the reaction rate constant. This chemical reaction takes place only on the surface <strong>of</strong><br />

the porous media. The reaction coefficient is then only space-dependent in the LBM. The<br />

differential equation <strong>for</strong> the reaction above is implemented to modify the local distribution<br />

functions after the relaxation process [28]. The surface reactions are handled similarly as the<br />

gravity <strong>for</strong>ce term in the particle density distribution by source term. However, by this<br />

procedure it is only possible to simulate the reaction at the solid surface with a specified rate.<br />

Other processes such as adsorption and desorption have been modeled by LBM with good<br />

results by using simple local rules and explicit discretisation <strong>of</strong> the porous geometry.<br />

2.3.5 General comments on CFD and LBM coupling<br />

For fluid flows the CFD approaches, such as FDM, FVM and FEM, are solvers based on a<br />

macroscopic discrete representation <strong>of</strong> the Navier-Stokes equation and at a mesoscale to<br />

microscale level the LBM has evolved. For this example FVM is coupled with LBM by<br />

connecting the boundaries but is can also be per<strong>for</strong>med with FEM or FDM. The importance<br />

15


lies to correctly set up the interface conditions and also to manage the conversion from the<br />

lattice to the macroscopic variables or vice versa [28].<br />

The positioning <strong>of</strong> the variables <strong>for</strong> the LB model and FV model is presented in Figure 2.5.<br />

The FV model uses a staggered grid where the scheme is explicit <strong>for</strong> the velocity and implicit<br />

<strong>for</strong> the pressure [25]. The choice <strong>of</strong> a staggered grid <strong>for</strong> the FV model is to prevent possible<br />

pressure oscillation. The LB variables are evaluated at the corner positions <strong>for</strong> all lattice<br />

variables at the same positions. For the FVM, the velocity variable is evaluated at the<br />

interface <strong>of</strong> the grid cells between two corner positions and the pressure is evaluated at the<br />

center <strong>of</strong> the grid cell [26]. Note that only a 2D domain is discussed here.<br />

Figure 2.5: The indexes and grid positions <strong>for</strong> the LB and FV models variables.<br />

This leads to the question <strong>of</strong> how to couple the two models at their interacting boundary. The<br />

set <strong>of</strong> boundary nodes <strong>for</strong> the two models are connected by an overlap layer <strong>of</strong> nodes so linear<br />

interpolation can be per<strong>for</strong>med to find the effective boundary condition [26]. This overlap is<br />

illustrated in Figure 2.6. Note that the overlap <strong>for</strong> the interface is about one and a half grid<br />

cell to capture the physical processes at the boundary. This can be chosen arbitrarily<br />

depending on the specific need <strong>of</strong> detailed in<strong>for</strong>mation and high resolution, and the access to<br />

computational power.<br />

16


Figure 2.6: Effective boundary arrangement <strong>for</strong> the coupling <strong>of</strong> FVM and LBM.<br />

The FV model is <strong>for</strong>mulated with adimensionless system and so LBM should also be<br />

converted to a dimensionless system to meet this constraint. The approach to connect the FV<br />

model and LB model at the boundary is by linear interpolation because the variables are not<br />

defined at the same positions in the domain. This gives the following relationships <strong>for</strong> the<br />

velocity.<br />

(2.34)<br />

(2.35)<br />

(2.36)<br />

(2.37)<br />

The procedure would start <strong>of</strong>f, <strong>for</strong> the inner nodes, so that the incoming particle distribution<br />

function f a at time t is used to compute the local density ρ and velocity u or v. For the<br />

boundary nodes all three variables are obtained from the variables <strong>of</strong> the FV model at time t<br />

[26, 28]. Next, all nodes are subject to the collision step. Finally, the inner nodes will per<strong>for</strong>m<br />

the streaming step.<br />

2.4 Previous case studies <strong>of</strong> SOFCs in LBM<br />

Although LBM is a relatively new actor among the numerical modeling methods, some work<br />

has already been carried out on SOFCs. There are some limitations connected to the LBM<br />

that need to be highlighted. These have been detected through previous studies. LBM has<br />

only recently been used as a numerical method <strong>for</strong> transport processes in SOFC and compared<br />

with conventional methods such as FDM, FEM and FVM [24]. According to Joshi et al., there<br />

is still a need <strong>for</strong> a supercomputer to per<strong>for</strong>m the LBM simulations. The method is described<br />

in detail by Joshi et al. [14-16, 23] and a comprehensive discussion <strong>of</strong> the method and various<br />

terms in the LB equation are <strong>of</strong>fered. Hence, the reader is referred to the work <strong>of</strong> Joshi et al.<br />

<strong>for</strong> further discussion.<br />

17


A method at microscale needs detailed geometric in<strong>for</strong>mation, and the size <strong>of</strong> computation<br />

domain cannot be too large due to limited computer resources. Because each pore should<br />

contain several lattice nodes in LBM and if the mass diffusion occurs in a large domain, the<br />

method may be unsuitable. When the LBM is applied to solve the mass diffusion in the pores,<br />

the gas velocity cannot be too high, due to the low Mach number limit <strong>for</strong> LBM [29]. In less<br />

porous media with low pore velocity, the density gradient can be very large. Further problems<br />

can occur when the resulting velocity field is applied to the simulation <strong>of</strong> the transport<br />

process, because a non-physical process and a system which is not in chemical equilibrium<br />

<strong>for</strong> the reaction process may be obtained [29]. The LBM has been shown to be less efficient<br />

<strong>for</strong> steady-state problems. However, this is expected because it is an explicit time-marching<br />

method that solves steady-state as the asymptotic solution in time. Accurate and quantitative<br />

comparisons between numerical methods are complicated because different methods have<br />

different underlying approaches so making statements on the relative merits <strong>of</strong> the numerical<br />

methods must be done with care [29-30].<br />

Table 2.1 is a summary <strong>of</strong> works on SOFCs done by different groups. For each work it is<br />

highlighted how the reconstruction <strong>of</strong> the microstructure was carried out, which modeling<br />

approach was used and at which dimension the work was studied. Further, Table 2.1 is<br />

divided into what analysis level the work was conducted on and which equation methodology<br />

was used in the microscale model. The last column contains some notes <strong>of</strong> conclusions made<br />

in each work to summarize the current situation <strong>of</strong> previous work. It can be concluded from<br />

Table 2.1 that there are basically three different approaches <strong>for</strong> the evaluation and<br />

reconstruction <strong>of</strong> the microstructure <strong>of</strong> the porous electrode. Furthermore, these can be<br />

grouped into a stochastic model where the structure is <strong>for</strong>med by a random statistical method<br />

or grouped into a computer aided scanning model such as X-ray computed tomography<br />

(XCT) and scanning electron microscope (SEM) [31-32]. The different studies are carried out<br />

<strong>for</strong> all three dimensions, but are mainly focused on the porous anode or a simpler geometry<br />

such as a channel. Because <strong>of</strong> the microscale modeling complexity, the focus is on mass<br />

diffusion, charge transport and electrochemical reaction because their behavior dominates at<br />

the microscopic level. The conclusions from previous studies relate mainly to the possibility<br />

to actually per<strong>for</strong>m a realistic reconstruction <strong>of</strong> the microstructure, to per<strong>for</strong>m modeling at<br />

microscale as well as to validate the model against other models. Overall, the LBM has shown<br />

good agreement with both SMM and DGM <strong>for</strong> 2D-cases <strong>of</strong> simple geometries. Improvements<br />

<strong>of</strong> the homogeneity <strong>of</strong> the anode microstructure as well as a better understanding <strong>of</strong> ionicelectronic<br />

phenomena at the active surface are viewed as the next interesting area <strong>of</strong> study<br />

[31]. To summarize, the LBM has proven to be a reliable microscale model <strong>for</strong> the complex<br />

porous anode.<br />

18


Table 2.1: Comparison <strong>of</strong> previous works <strong>of</strong> microscale models <strong>for</strong> SOFCs.<br />

Author<br />

Microstructure<br />

reconstruction<br />

Modeling<br />

Dimension<br />

Analysis level<br />

Equation<br />

analysis<br />

Concluding comments<br />

Suzue et al.<br />

(2008)[27]<br />

Stochastic<br />

model<br />

LBM<br />

3D<br />

Anode<br />

(electrode<br />

per<strong>for</strong>mance)<br />

Mass diff.<br />

Charge tr.<br />

Electrochem.<br />

Decrease in temp. → Current<br />

concentration closer to TPB and<br />

reactive anode becomes thinner.<br />

Grew et al.<br />

(2010)[14]<br />

Not in article<br />

LBM<br />

2D<br />

Anode<br />

Mass diff.<br />

Charge tr.<br />

Electrochem.<br />

Constant overpotential acts<br />

limiting → Large resistivity and<br />

unable to account <strong>for</strong> the coupled<br />

transport processes<br />

Joshi et al.<br />

(2007)[15]<br />

Stochastic<br />

model<br />

LBM, DGM<br />

2D<br />

Channel<br />

Mass diff. (Kn)<br />

LBM suitable <strong>for</strong> a wide range <strong>of</strong><br />

Kn in non- and continuum<br />

regime. DGM not suitable <strong>for</strong><br />

porous geometry.<br />

Joshi et al.<br />

(2007)[16]<br />

Stochastic<br />

model<br />

LBM, SMM<br />

1D, 2D<br />

Channel (straight,<br />

tortuous, <strong>for</strong>ked)<br />

Porous geometry<br />

Mass diff.<br />

Good agreement <strong>for</strong> LBM and<br />

SMM. LBM possible <strong>for</strong> porous<br />

geometry without empirical<br />

modification.<br />

Joshi et al.<br />

(2010)[24]<br />

XCT<br />

LBM<br />

3D<br />

Anode<br />

Mass diff.<br />

Good agreement <strong>for</strong> XCT and 2D<br />

stochastic reconstruction.<br />

Sohn et al.<br />

(2010)[22]<br />

Stochastic<br />

model<br />

LBM, RW<br />

3D<br />

Anode (micro)<br />

Cell (macro)<br />

Electrochem.<br />

Energy (macro)<br />

Mass<br />

diff.(macro)<br />

Nu and Sh similar behavior →<br />

heat and mass transport coupling<br />

needed. Micro/macro model<br />

serves as a good modeling<br />

approach.<br />

Chiu et al.<br />

(2009)[21]<br />

XCT<br />

LBM, SMM<br />

2D<br />

Anode<br />

Mass diff.<br />

Electrochem.<br />

Int. re<strong>for</strong>ming<br />

Good agreement <strong>for</strong> LBM and<br />

SMM with methane re<strong>for</strong>ming<br />

and electrochemical activity at<br />

TPB.<br />

Asinari et al.<br />

(2007)[32]<br />

SEM<br />

LBM<br />

3D<br />

Anode<br />

(electrode<br />

per<strong>for</strong>mance)<br />

Mass diff.<br />

Microstructural reconstruction <strong>for</strong><br />

micro-modeling serves well <strong>for</strong><br />

the macro per<strong>for</strong>mance.<br />

2.5 Electrode microstructure remarks<br />

The particle size in SOFCs is in the range <strong>of</strong> microscale, and the TPBs are in nanoscale. The<br />

morphology and the properties <strong>of</strong> these scales are important <strong>for</strong> the per<strong>for</strong>mance <strong>of</strong> the fuel<br />

cell, because they control how much <strong>of</strong> the Gibbs free energy is available <strong>for</strong> use. The science<br />

at nano and microscale is critical to the per<strong>for</strong>mance at a system scale, but it is problematic to<br />

find a suitable and reliable kinetic model within a simplified framework which still fully<br />

describes the interactions between the particles [12]. For example the kinetic model can be<br />

constructed by combining chemical values <strong>for</strong> each species with computed activation energies<br />

and transition-state properties. An important aspect <strong>of</strong> generation <strong>of</strong> steam re<strong>for</strong>ming kinetics<br />

19


data on YSZ-supported and Ni catalyst anodes is that they are prone to carbon deposition.<br />

Small changes during the process <strong>of</strong> manufacturing can have an effect on the catalytic<br />

characteristics. The structural features, <strong>for</strong> example the particle size distribution, have a strong<br />

influence on the anodes catalytic and electrochemical characteristics [33]. Structural<br />

parameters and conditions <strong>of</strong> the experimental cell are not always clearly stated in the<br />

literature, which makes it hard to reproduce results numerically. Another issue is that the<br />

original kinetic data is <strong>of</strong>ten taken from a variety <strong>of</strong> different catalysis studies which makes<br />

the mechanism thermodynamically inconsistent. Due to this issue, some <strong>of</strong> the original kinetic<br />

parameters are <strong>of</strong>ten modified to ensure the overall consistency <strong>of</strong> the enthalpy and entropy.<br />

Averaged structural parameters, such as porosity and tortuosity, may have the same value <strong>for</strong><br />

many different microstructure topologies but the material structure and path ways may differ<br />

significantly [32]. The correlations between the large numbers <strong>of</strong> operating conditions in<br />

combination with the simplified structural parameters make it difficult to verify the diffusion<br />

in practical ranges [14]. In summary, at the time <strong>of</strong> writing, no macroscopic parameters can<br />

properly describe the microscopic physical behavior. One <strong>of</strong> these macroscopic parameters is<br />

the tortuosity. The tortuosity factor is used to handle the discrepancy <strong>of</strong> the pressure loss <strong>for</strong><br />

the complex path ways in porous media and should include both shear <strong>for</strong>ces and elongation<br />

<strong>for</strong>ces <strong>of</strong> the fluid elements in the flow. But even the meaning <strong>of</strong> tortuosity goes apart, where<br />

in some models it represents the effect <strong>of</strong> additional pathways and in some models just a<br />

numerical parameter to fit the experimental data [28]. Microscale models may enhance the<br />

knowledge and bridge over this gap.<br />

Another important remark is that the functional electrode structures are known to work in<br />

favor <strong>of</strong> electrochemical reaction, mass and charge transport. Several earlier microscale<br />

models have adopted a structural schematic spherical approach to reconstruct the porous<br />

media [22]. The approaches have until recently been based upon statistical parameters from<br />

the random spherical particles to create the model-connected physical parameters. The<br />

structure can be created by a random statistical simulation, by size and location, or by<br />

computer tomography X-ray absorption contrast (XCT), or similarly, by scanning electron<br />

microscope (SEM) <strong>of</strong> a real existing microstructure <strong>of</strong> a SOFC electrode. The SEM requires a<br />

great deal <strong>of</strong> ef<strong>for</strong>t in terms <strong>of</strong> measurement and data processing, and still the effect <strong>of</strong><br />

microstructures on the electrochemical activity remains unclear [27]. According to Asinari et<br />

al. [32], the XCT does not provide good enough reconstruction <strong>for</strong> the microscale and the<br />

only viable resolution <strong>for</strong> SOFCs is provided by SEM, but still XCT is cheaper, faster and<br />

demands less ef<strong>for</strong>t. XCT can be improved by statistical regression, which is <strong>of</strong>ten used when<br />

a 3D structure needs to be obtained by a 2D structure [32, 34].<br />

To visualize in a basic schematic way on the microscale structure, it is schematically<br />

described in Figure 2.6 by randomly situated spherical particles and transport <strong>of</strong> the different<br />

species when hydrogen is fed. The electron flow is represented and the ionic and electronic<br />

feature is represented by the binary particles. Part <strong>of</strong> the cell is divided into three microscopic<br />

regions which are named as cathode functional layer (CFL), electrolyte and anode functional<br />

layer (AFL). The functional layer defines the part <strong>of</strong> the anode or the cathode closest to the<br />

electrolyte where most <strong>of</strong> the active reactions take place. A TPB location is enlarged to show<br />

the different material components <strong>of</strong> an SOFC.<br />

20


Figure 2.7: Schematic illustration <strong>of</strong> the microstructure components and the main processes.<br />

The electrode per<strong>for</strong>mance can be increased if there is sufficient porosity so that gas transfer<br />

is not limiting and so that the TPB needs to be maximized. While a fine microstructure and a<br />

high surface area are clearly desirable, this can lead to low mechanical strength [35]. The<br />

anode is <strong>of</strong>ten used as the mechanical support <strong>for</strong> the cell and a change <strong>of</strong> the anode structure<br />

can be problematic. Because the active region in the anode where the electrochemical<br />

reactions takes place, extends less than approximately 10 μm from the anode–electrolyte<br />

interface, a graded porosity (like functional layers) is sometimes used to maximize the<br />

amount <strong>of</strong> TPB in the active region while still maintaining a high mechanical strength <strong>for</strong> the<br />

rest <strong>of</strong> the anode which is used primarily as the cell support [35].<br />

Another effect captured by microstructure considerations is that the cell per<strong>for</strong>mance can be<br />

permanently affected by the electric field. The pore <strong>for</strong>mation in the material can be affected<br />

by oxygen potential gradient at the cathode/electrolyte interface and nickel agglomeration at<br />

the TPB. This will cause a degradation <strong>of</strong> the per<strong>for</strong>mance by the electrochemical reactions at<br />

the TPB due to the sintering <strong>of</strong> metal particles, which causes a decrease in specific contact<br />

area <strong>of</strong> Ni particles [34, 36].<br />

21


3 Mathematical<br />

models<br />

This chapter presents the LB model visualization and equation methodology <strong>for</strong> the transport<br />

processes. Finally, a validity check <strong>of</strong> the kinetic effects on the transport processes is<br />

presented <strong>for</strong> interparticle, interphase and intraparticle transport within the microscopic range.<br />

3.1 Model visualization <strong>for</strong> the microscale model<br />

To visualize the complex geometry flow, the model discretisation is created from a digital<br />

image. The digital image is created by a 3D computer tomography from a real object and is<br />

shown in Figure 3.1. The complex structure <strong>of</strong> the Ni/YSZ anode was printed in grayscale<br />

(256 colors) and through data conversion functional voxel in<strong>for</strong>mation was obtained. The<br />

conversion process was conducted in Python where the voxel data, numerical in<strong>for</strong>mation<br />

about the color at specific positions, was transferred to a matrix functional <strong>for</strong> MATLAB. To<br />

distinguish the two phases (pore/solid) a phase function is defined at each pixel (z) as:<br />

(3.1)<br />

Figure 3.1: Digital image <strong>of</strong> a Ni/YSZ anode.<br />

This border can be adjusted to obtain different values <strong>for</strong> the porosity which <strong>of</strong>fered the<br />

possibility to vary the porosity in the LBM simulation. To visualize the physical quantities<br />

(velocity, concentration and pressure) these are obtained by the Lattice Boltzmann particle<br />

distribution function and updated every iteration loop. In Figure 3.2, the image <strong>for</strong> the LB<br />

22


model is presented after the data conversion from the original digital image (Figure 3.1) and<br />

the choice <strong>of</strong> border between pore and solid, i.e., white and black, was chosen to obtain a<br />

porosity <strong>of</strong> 40%.<br />

Figure 3.2: Image <strong>for</strong> LBM by two colors, black (solid) and white (pore), with a porosity <strong>of</strong> 40%.<br />

The conversion process <strong>for</strong> three colors, namely white, grey and black, was also per<strong>for</strong>med in<br />

Python to create a matrix functional <strong>for</strong> MATLAB. To distinguish the three phases <strong>for</strong> the<br />

pore and the two solid types, a phase function is defined at each pixel (z) as:<br />

(3.2)<br />

In Figure 3.3 the real image <strong>of</strong> the XCT scan is shown. In Figure 3.4 and 3.5 the three-colorconversion<br />

is implemented where the choice <strong>of</strong> border between pore and solid (both grey and<br />

black) was chosen to obtain a porosity <strong>of</strong> 40% and 60%, respectively. The black colored<br />

patches represents YSZ, the grey Ni and the white represents the pores.<br />

Figure 3.3: Digital image <strong>of</strong> a Ni/YSZ anode.<br />

23


Figure 3.4: Image <strong>for</strong> LBM by three colors (black, grey and white) with a porosity <strong>of</strong> 40%.<br />

Figure 3.5: Image <strong>for</strong> LBM by three colors (black, grey and white) with a porosity <strong>of</strong> 60%.<br />

3.2 Governing equations <strong>for</strong> the macroscale model<br />

It is essential to connect the different transport processes when modeling SOFCs. The transfer<br />

<strong>of</strong> fuel gases to the active surface <strong>for</strong> the electrochemical reactions is governed by different<br />

parameters, such as the porous microstructure, the gas consumption/generation, the pressure<br />

gradient between the fuel flow duct and the porous anode, and finally the inlet conditions [9].<br />

The gas molecules diffuse to the TPB, where the electrochemical reactions take place. The<br />

hydrogen concentration depends on the transport within the porous anode and the<br />

heterogeneous re<strong>for</strong>ming reaction chemistry [1]. In the following sections the main transport<br />

processes are briefly described.<br />

24


3.2.1 Mass transport<br />

In the electrodes, mass transfer is dominated by gas diffusion and the transport takes place in<br />

the gas phase, which is influenced by the electrochemical reactions at the solid surface at the<br />

active TPB [33]. The transport phenomena can be classified into some general categories<br />

based on the Knudsen number. For the porous layer, continuum phenomena are predominant<br />

<strong>for</strong> the case with large pores, whose size is much bigger than the free path <strong>of</strong> the diffusion gas<br />

molecules [11]. The Knudsen diffusion is used when the pores are small in comparison to the<br />

mean free path <strong>of</strong> the gas. For Knudsen diffusion, molecules collide more <strong>of</strong>ten with the pore<br />

walls than with other molecules [1].<br />

Mass transport can be calculated using Fick’s law, which is the simplest diffusion model. But<br />

<strong>for</strong> a multicomponent system, the Stefan-Maxwell model is <strong>of</strong>ten implemented to calculate<br />

the diffusion [37-38]. Furthermore, when extended with the Knudsen diffusion term to predict<br />

the collision effect by the molecules it is usually called the Dusty Gas model [5, 37, 39]. The<br />

Stefan-Maxwell equation is a simplified equation <strong>of</strong> the Dusty Gas Model. The Knudsen term<br />

is neglected because the collision between the gas molecules and the porous medium is not<br />

considered. The Stefan-Maxwell equation is defined <strong>for</strong> the electrodes, the fuel and air<br />

channels, as below [5, 22]:<br />

(3.3)<br />

(3.4)<br />

(3.5)<br />

where w is the mass fraction, D ij the Stefan-Maxwell binary diffusion coefficient, x the mole<br />

fraction, D i T the thermal diffusion coefficient and S i the source term. S i is, in this case, zero<br />

because the electrochemical reactions are assumed to take place at the interfaces between the<br />

electrolyte and electrodes. There<strong>for</strong>e, they are defined as an interface condition and not as a<br />

source term. The diffusion coefficient in the electrodes is calculated as [40]:<br />

(3.6)<br />

where is the tortuosity. Moreover, D ij is calculated as:<br />

(3.7)<br />

(3.8)<br />

(3.9)<br />

(3.10)<br />

25


where T is the temperature, P the pressure, σ AB the characteristic length and D the<br />

dimensionless collision integral. The averaged molecular weight M AB between substance A<br />

and B is defined as [40]:<br />

(3.11)<br />

Note that the pressure P is in bar in equation (3.7) and in our case P is set to 1 bar. Also note<br />

that the parameters in this equation are not all in SI units. The ones which differ from the SI<br />

unit standard here is M AB [g/mol] and σ AB [Å].<br />

However, if the non-continuum regime is present, additional dimensionless parameters need<br />

to be introduced in terms <strong>of</strong> the Knudsen diffusivity defined as [21]:<br />

(3.12)<br />

where is the pore mean radius, T the temperature, M i the molecular mass. Furthermore, the<br />

dimensionless Knudsen diffusivity <strong>for</strong> species i in relation to species j is obtained by [19]:<br />

(3.13)<br />

3.2.2 Heat transport<br />

The heat transfer within the whole unit cell consists <strong>of</strong> convection between the solid surface<br />

and the gas flow, conduction in the solid and the porous parts, and heat<br />

generation/consumption occurs due to the electrochemical reactions at the TPB as well as the<br />

internal re<strong>for</strong>ming reactions. The temperature distribution in this study uses a local thermal<br />

non-equilibrium (LTNE) approach due to the low Reynolds number and large differences in<br />

thermal conductivities between the gas and solid phases. The temperature distribution is<br />

calculated separately <strong>for</strong> the gas and the solid phases. The general heat conduction equation is<br />

used to calculate the temperature distribution <strong>for</strong> the solid medium in the porous electrodes [8,<br />

22]:<br />

(3.14)<br />

where is thermal conductivity <strong>for</strong> the solid media, the temperature in the solid phase and<br />

the heat source (heat transfer between the gas and solid phases, the heat generation due to<br />

the ohmic polarization and due to the internal re<strong>for</strong>ming reactions). The temperature <strong>for</strong> the<br />

gas phase in the fuel gas and air channels and the porous electrodes are calculated according<br />

to [22]:<br />

(3.15)<br />

where T g is the gas temperature, c p,g the specific heat, k g the gas thermal conductivity and Q g<br />

the heat transfer between the gas and solid phases. The heat transfer between the gas and solid<br />

phases is defined as:<br />

(3.16)<br />

26


where h v is the volume heat transfer coefficient and AV the active surface area to volume<br />

ratio.<br />

3.2.3 Momentum transport<br />

The approach <strong>for</strong> analysis <strong>of</strong> the momentum transport is to solve the Darcy equation <strong>for</strong> the<br />

porous electrodes and the Navier-Stokes equations <strong>for</strong> the channels. The Darcy-Brinkman<br />

equation is then used to solve the gas flow in the gas phase [5, 33]:<br />

(3.17)<br />

where μ is the dynamic viscosity, κ the permeability <strong>of</strong> the porous medium, ε p the porosity, T<br />

the viscous stress tensor and F the volume <strong>for</strong>ce vector. λ is the second viscosity and <strong>for</strong> gases<br />

it is normally set to λ = -2/3∙μ. κ dv is the deviation from the thermodynamic equilibrium and is<br />

by default set to zero. The Darcy-Brinkman equation is converted into the Darcy equation<br />

when the Darcy number Da 0 in the porous layers and into the Navier-Stokes equation<br />

when κ and ε p = 1 in the fuel and air channels [5].<br />

The velocity pr<strong>of</strong>ile is defined at the air and fuel channel inlets as laminar flow and the outlets<br />

are treated as pressure surfaces. The boundaries at the top and the bottom <strong>of</strong> the cell model<br />

are defined by symmetry because the cell is considered to be surrounded by other similar cells<br />

with the identical temperature distribution. The temperatures at the air and fuel channels inlets<br />

are defined as constant and the outlet boundaries are defined as convective flux surfaces.<br />

3.2.4 Electrochemical reactions<br />

The electrochemical reactions occur at the TPB. The function <strong>of</strong> the electrolyte is on the one<br />

hand to transport the oxygen ions to the anode and on the other hand to prevent the electrons<br />

to cross from the anode to the cathode. The flow <strong>of</strong> electronic charge through the external<br />

circuit balances the flow <strong>of</strong> ionic charge through the electrolyte. This transport is described in<br />

terms <strong>of</strong> the ion transport from the conservation <strong>of</strong> charge [11, 14, 22]:<br />

(3.18)<br />

(3.19)<br />

where i io and i el are charge fluxes <strong>for</strong> ions and electrons, respectively, and ϕ io is the ionic<br />

potential in the electrolyte. The Nernst potential is calculated as the sum <strong>of</strong> the potential<br />

differences across the anode and the cathode as [41]:<br />

(3.20)<br />

where E is the reversible electrochemical cell voltage and ϕ the charge potential. At the<br />

interface between the electrode and electrolyte the Butler-Volmer equation is used to<br />

calculate the volumetric current density [22]:<br />

(3.21)<br />

where i 0 is the exchange current density, F the Faraday constant, β the transfer coefficient, n e<br />

the number <strong>of</strong> electrons transferred per reaction, η act,e the electrode activation polarization<br />

27


over-potential, and finally R the ideal gas constant. If the transfer coefficient β is assumed to<br />

be 0.5, the Butler-Volmer equation is reduced to [2, 5]:<br />

(3.22)<br />

(3.23)<br />

(3.24)<br />

where k´´ is the pre-exponential factor and E the activation energy. The gas species<br />

distributions are implemented by source terms due to the electrochemical reaction as [5]:<br />

(3.25)<br />

(3.26)<br />

(3.27)<br />

where i is the current density and F the Faraday constant.<br />

3.2.5 Internal re<strong>for</strong>ming reactions<br />

The internal re<strong>for</strong>ming reaction rates are taken into account by the source terms in the Stefan-<br />

Maxwell equation. The mass source terms due to the re<strong>for</strong>ming reactions are expressed as:<br />

(3.28)<br />

(3.29)<br />

(3.30)<br />

(3.31)<br />

The equation <strong>for</strong> CO 2 can be solved separately because the sum <strong>of</strong> the mass fractions is equal<br />

to unity. The reaction rate r r is <strong>for</strong> the catalytic steam re<strong>for</strong>ming reaction and r s is <strong>for</strong> the<br />

water-gas shift reaction.<br />

The reaction rates <strong>for</strong> the methane steam re<strong>for</strong>ming reaction are evaluated by kinetic models<br />

and <strong>for</strong> the water-gas shift reaction an equilibrium approach is applied. The three reaction<br />

kinetic approaches applied are from [41-45]. It is worth noting that both Achenbach &<br />

Riensche’s [42-43] (equation (3.32)) together with Leinfelder’s [44] (equation (3.33)) kinetics<br />

are an Arrhenius type kinetics reaction rate, while Dreschers kinetics [45] (equation (3.34)) is<br />

a Langmuir-Hinshelwood type. They are selected on the basis <strong>of</strong> the different order <strong>of</strong> the<br />

partial pressure and the broad range <strong>of</strong> the activation energy. The differences in kinetics<br />

depend on how the experimental configuration is set up and, how the material decomposition<br />

and operating conditions are selected. From the studies <strong>of</strong> Achenbach & Riensche [42] and <strong>of</strong><br />

28


Achenbach [43], it was found that the reaction order <strong>of</strong> the partial pressure <strong>of</strong> methane is<br />

unity and the partial pressure <strong>of</strong> water has no catalytic effect on the reaction [42-43]. Note<br />

that Leinfelder [44] found a positive reaction order <strong>of</strong> water and Achenbach & Riensche [42-<br />

43] found a reaction order <strong>of</strong> zero. The reaction rates from these three different experimental<br />

studies are shown below:<br />

(3.32)<br />

(3.33)<br />

(3.34)<br />

where p is the partial pressure and T s the solid phase temperature. AV is the active surface area<br />

to volume ratio. The units <strong>for</strong> all the steam re<strong>for</strong>ming reaction rates are mol/(s∙m 3 ).<br />

The reaction rate kinetic models, equations (3.32) and. (3.33), consist <strong>of</strong> three parts: partial<br />

pressures, pre-exponential factor and activation energy. These parameters differ substantially<br />

in the literature among different research works. The pre-exponential factor describes the<br />

number <strong>of</strong> collisions between the molecules within the reaction. The exponential expression<br />

including the activation energy describes the probability <strong>for</strong> the reaction to occur. As the<br />

activation energy increases, the catalytic reaction becomes less probable. The activation<br />

energy is based on the catalytic characteristics, such as chemical composition. Even though<br />

the activation energy may be high, leading to a decrease in the reaction rate, the overall<br />

reaction rate can still be high due to the pre-exponential value. The pre-exponential factors<br />

depend strongly on both the temperature and properties <strong>of</strong> the anode material. It is possible to<br />

change the reaction rate, either by changing the particle size <strong>of</strong> the active catalysts or the<br />

porous structure, i.e., the active catalytic area. The large difference between the activation<br />

energies found in the literature, [1, 41-46] suggests that additional parameters have significant<br />

influence on the reaction rate.<br />

According to Nagel et al. [41] a small steam-to-carbon (SC) ratio yields positive reaction<br />

orders and a high SC ratio yields negative reaction orders. For this study, the steam to carbon<br />

ratio is around 2, which agrees with the three kinetic models. Achenbach & Riensche [42-43]<br />

applied a 14 mm thick nickel cermet semi-disk consisting <strong>of</strong> 20 wt.% Ni and 80 wt.% ZrO 2<br />

(stabilized). The active surface area was 3.86 ∙ 10 -4 m 2 . The temperature was varied from 700<br />

to 940 ºC and the system pressure from 1.1 to 2.8 bar. Leinfelder [44] applied a 50 µm thick<br />

anode built up by two layers with 64 wt.% Ni and 36 wt.% YSZ and 89 wt.% Ni and 11 wt.%<br />

YSZ, respectively. The active surface area <strong>for</strong> the anode was 2.5 ∙ 10 -3 m 2 . The test was<br />

conducted <strong>for</strong> temperatures from 840 to 920 ºC and at a pressure <strong>of</strong> 1 bar. Drescher [45]<br />

studied an anode consisting <strong>of</strong> 50 wt% Ni and 8 mol% YSZ. Achenbach & Riensche’s model<br />

is based on work on a re<strong>for</strong>mer while the other two, Leinfelder’s and Drescher’s, are based on<br />

a unit cell.<br />

In this study, the temperature is varied from 727 to 827 ºC (1000 to 1100 K) because this is in<br />

the range which the experiments were carried out. The active surface area to volume ratio is<br />

varied 10 ∙10 4 - 5 ∙10 5 m 2 /m 3 . The active surface area to volume ratio has been adopted<br />

according to a commonly used value in the literature [5-6, 39, 45]. Several authors have<br />

29


applied an active surface area to volume ratio <strong>of</strong> 5 ∙ 10 5 <strong>for</strong> modeling work. Janardhanan and<br />

Deutschmann [5] applied a slightly smaller surface area to volume ratio <strong>of</strong> 102500 m 2 /m 3 ,<br />

whereas Klein et al. [46] applied a much larger value <strong>of</strong> 2.2 ∙ 10 6 m 2 /m 3 . Note that only a<br />

small part <strong>of</strong> the whole active surface acts as a locus <strong>for</strong> the chemical reactions. The trend <strong>for</strong><br />

the development during the last years is in the direction <strong>of</strong> employing smaller particles to get<br />

larger AV.<br />

The water-gas shift reaction is considered to be very quick and to remain in equilibrium by<br />

active several authors in the literature [2, 46-47]. The equilibrium approach in the fuel<br />

channel and the anode can be defined as:<br />

(3.35)<br />

(3.36)<br />

where k s is the reaction rate constant and K e,s the equilibrium constant <strong>for</strong> the water-gas shift<br />

reaction. The value <strong>for</strong> k s is calculated according to Haberman and Young [48]. The unit <strong>for</strong><br />

the water-gas shift reaction rate is mol/(s∙m 3 ). The heat generation and heat consumption are<br />

defined as source terms in the governing equations. The heat generation in the fuel channel<br />

enters in the gas phase. The heat generation and the heat consumption are assumed to occur<br />

on the solid surface. The heat generation and heat consumption due to the re<strong>for</strong>ming reactions<br />

are implemented in equation (3.14) and are defined as:<br />

(3.37)<br />

where Δh reac is the enthalpy change due to the reactions and Q int,ref the heat generation.<br />

3.3 Heat and mass transfer limitations <strong>of</strong> the kinetic model<br />

For fuel cell research it is interesting to identify whether any transport process at any level <strong>of</strong><br />

scale limits the whole process. An analysis on the basis <strong>of</strong> interparticle, interphase and<br />

intraparticle heat and mass transport is per<strong>for</strong>med to provide knowledge <strong>of</strong> the limiting steps<br />

at each level. The different domains are explained by the division below [49]:<br />

<br />

<br />

<br />

Interparticle, also called intrareactor, is defined between the local fluid regions or<br />

catalyst particles.<br />

Interphase is defined between the external surfaces <strong>of</strong> the particles and fluid adjacent<br />

to them.<br />

Intraparticle is defined within individual catalyst particles.<br />

The structure and equation methodology <strong>of</strong> evaluating the limiting steps <strong>for</strong> heat and mass<br />

transfer at different scales in SOFCs consists <strong>of</strong> catalytic kinetic equations in terms <strong>of</strong> criteria<br />

obtained from experimental work by Mears [49]. The analysis explained by Mears [49] was<br />

per<strong>for</strong>med on a reactor bed. Here, it is transferred to the anode <strong>for</strong> the steam re<strong>for</strong>ming<br />

reaction and <strong>for</strong> the electrochemical reactions at the anode and the cathode <strong>of</strong> the SOFC. The<br />

main difference between the two reactor environments is that a reactor bed has walls along<br />

the flow direction, compared to an anode and a cathode that are supplied with fuel and air,<br />

respectively. This means that special consideration needs to be applied to calculate the<br />

interparticle heat and mass transport limitations along the flow direction. For the<br />

interphase and the intraparticle heat and mass transport limitations the SOFC anode and<br />

30


cathode are assumed to be similar to the case <strong>of</strong> the reactor bed. The aim <strong>of</strong> this case study is<br />

to examine whether the kinetics used <strong>for</strong> previous models [1] fulfill these criteria so no<br />

limiting effects occur <strong>for</strong> the heat and mass transport.<br />

The macroscale (2D) computational fluid dynamics (CFD) model <strong>of</strong> an intermediate<br />

temperature anode-supported SOFC operating on 30% pre-re<strong>for</strong>med natural gas is the base<br />

<strong>for</strong> the calculations per<strong>for</strong>med here. First, the criteria <strong>for</strong> the different domains are described<br />

and defined below. Then, the results <strong>of</strong> the analysis <strong>for</strong> the SOFC are presented in the next<br />

chapter.<br />

3.3.1 Interparticle transport<br />

The largest scale in this analysis is <strong>for</strong> the interparticle transport which is also sometimes<br />

called intrareactor scale because it applies to gradients within the reactor as a whole.<br />

Transport phenomena can occur both radially and axially within the reactor and these are hard<br />

to control and evaluate. For the SOFC the axial direction refers to the main flow direction (xdirection)<br />

and the radial direction refers to the direction normal to the main flow direction (ydirection).<br />

But if neglected, radial temperature gradients can <strong>for</strong>ce the reaction rates to be<br />

thousandfold greater <strong>for</strong> parts <strong>of</strong> the reactor <strong>of</strong>ten close to the axis [49]. For the SOFC this<br />

would occur in the anode and near the electrolyte close to the inlet <strong>of</strong> the cell and would mean<br />

a risk <strong>for</strong> disturbing “hot spots”. This can be checked by radial dispersion [50]:<br />

(3.38)<br />

where Bi R is the Biot number based on the reactor diameter, ΔH the enthalpy change <strong>of</strong><br />

reaction, r r the reaction rate, R o the reactor diameter, k e the thermal conductivity <strong>for</strong> the solid<br />

porous media, T w the temperature at the solid surface and γ the dimensionless activation<br />

energy. The axial dispersion is a less frequent limitation in a severe manner. Axial<br />

temperature gradients and axial conduction are possible to neglect if the length-to-particle<br />

diameter ratio is large enough (L/d p > 30) which is the case <strong>for</strong> SOFCs [50]. The criterion <strong>for</strong><br />

the limitation <strong>for</strong> the temperature gradient across the reactor diameter is defined as [49]:<br />

(3.39)<br />

where the parameters are the same as <strong>for</strong> equation (3.38) except the Biot number, R the gas<br />

constant and E a the activation energy.<br />

Mears [49] described an approach to adopt a differential reactor and this seems to be a very<br />

useful approach <strong>for</strong> the SOFC reactor beds. A differential reactor consists <strong>of</strong> different<br />

amounts <strong>of</strong> catalyst throughout the reactor bed to compensate <strong>for</strong> unfavorable effects such as<br />

extremely high reaction rates in parts <strong>of</strong> the reactor. For the SOFC one would wish to level<br />

out the reactions and the electricity generating action over the whole bed. This can be<br />

achieved by either increasing the amount <strong>of</strong> catalyst or to use finer particles to increase the<br />

reaction rate. However, the SOFC has contradicting needs <strong>for</strong> the reaction rates depending on<br />

whether the focus is on the methane steam re<strong>for</strong>ming reaction or electrochemical reactions.<br />

For the steam re<strong>for</strong>ming reaction, the reaction rate is very high at the inlet and then gradually<br />

decreases in the flow direction <strong>for</strong> the cell. But the reaction rate <strong>for</strong> the electrochemical<br />

reactions requires a higher reaction activity right at the inlet <strong>for</strong> a limited area which would<br />

increase if more catalyst material was provided or finer particles were present. By adjusting<br />

the reaction rate activity <strong>for</strong> its needs, severe effects <strong>of</strong> temperature or concentration gradients<br />

could be minimized.<br />

31


3.3.2 Interphase transport<br />

The limitation <strong>of</strong> heat transport at the interphase transport level is normally less severe than<br />

the interparticle transport level and the greater part <strong>of</strong> the resistance is <strong>of</strong>ten in the boundary<br />

layer around the particle rather than within it [48]. This can be expected as the thermal<br />

conductivity <strong>of</strong> the solid is <strong>of</strong>ten much larger than that <strong>for</strong> the gas. For a low Reynolds<br />

number Re, Bizzi et al. [51] described the equation <strong>for</strong> the mass transfer coefficient k c as:<br />

(3.40)<br />

(3.41)<br />

(3.42)<br />

where ρ is the fluid density, v the fluid velocity, κ the permeability, μ the dynamic viscosity<br />

and ψ the shape factor. For spherical particles it is assumed that the shape factor is ψ = 1 [51].<br />

Then equation (3.41) reduces to:<br />

(3.43)<br />

The definition <strong>of</strong> the criterion <strong>for</strong> the concentration gradient across the gas boundary film<br />

along the particle is [49]:<br />

(3.44)<br />

where r r is the reaction rate, r p the particle diameter, C the concentration, D eff the effective<br />

diffusivity and n reaction order. The Damköhler number <strong>for</strong> interphase mass transport is<br />

defined as [49]:<br />

(3.45)<br />

where, besides those mentioned in equation (3.44), k c is the mass transfer coefficient. The<br />

Damköhler number <strong>for</strong> interphase heat transport is defined as [49]:<br />

(3.46)<br />

where ΔH is the enthalpy change <strong>of</strong> reaction, h heat transfer coefficient and T the bulk<br />

temperature. Further, the dimensionless activation energy is defined as [49]:<br />

(3.47)<br />

where E a is the activation energy, R the gas constant and T the bulk temperature. The<br />

dimensionless maximum temperature rise is used in the criterion <strong>for</strong> the concentration<br />

gradient across the gas film and the criterion is defined as [49]:<br />

32


(3.48)<br />

where, besides the parameters mentioned above which are the same, k s is the thermal<br />

conductivity <strong>of</strong> the particle. Mass transport cannot be a significant limitation unless the<br />

effectiveness factor is low <strong>for</strong> the intraparticle range. A criterion <strong>for</strong> the heat transport can be<br />

<strong>for</strong>mulated as [49]:<br />

(3.49)<br />

where all the parameters are the same as mentioned above.<br />

To detect the limitation <strong>for</strong> the boundary layer it is necessary to find out if transport<br />

limitations exist in the particle or not. Heat transport limitations can be expected when the<br />

reaction rates are high and the flow rates are low (small h) [49]. The criterion <strong>for</strong> the<br />

interphase transport <strong>for</strong> the heat transfer is defined as [49]:<br />

(3.50)<br />

where h is the heat transfer coefficient, d p the particle diameter and k s the thermal<br />

conductivity <strong>of</strong> the particle. It can be safely assumed a uni<strong>for</strong>m temperature distribution in the<br />

solid part if the Bi is less than 0.1 <strong>for</strong> an SOFC. But if Bi is larger than 10, then the<br />

conduction resistance dominates which will generate temperature gradients in the solid<br />

particle. The convective gas particle heat transfer coefficient h is defined as [52]:<br />

(3.51)<br />

where Nu is the Nusselt number, k f the thermal conductivity <strong>of</strong> the gas and D o the hydraulic<br />

diameter <strong>of</strong> the whole reactor bed, e.g., anode or cathode. The Nu in this case is calculated as<br />

[52]:<br />

(3.52)<br />

where Re is the Reynolds number, and Pr the Prandtl number which are defined as in [52]:<br />

(3.53)<br />

where μ f is the dynamic viscosity <strong>of</strong> the gas, c pf the specific heat <strong>of</strong> the gas and k f the thermal<br />

conductivity <strong>of</strong> the gas. Mears [49] pointed out that the heat transfer over the boundary layer<br />

causes larger deviation from the criterion long be<strong>for</strong>e the mass transfer limitations.<br />

3.3.3 Intraparticle transport<br />

The smallest scale <strong>for</strong> the analysis in this work is <strong>for</strong> the intraparticle transport and it has been<br />

more widely studied <strong>for</strong> reactor beds than the two previous scales. When diffusion might have<br />

a strong influence, the objective is <strong>of</strong>ten to calculate the effectiveness factor which is shown<br />

to be inversely proportional to the characteristic dimensions <strong>of</strong> the particle. Though, this<br />

33


approach <strong>of</strong>ten requires detailed kinetic behavior which is close in representation to the<br />

realistic kinetic catalysis.<br />

The heat transport limitation is evaluated first by the larger scale interparticle transport. If the<br />

criterion is fulfilled <strong>for</strong> the heat transport in the range <strong>for</strong> the interparticle scale, then there is<br />

no risk <strong>for</strong> too high temperature gradient across the reactor y-axis <strong>for</strong> the intraparticle<br />

transport, since R o >> r p and k e approaches the value <strong>of</strong> λ at low Reynolds numbers. The<br />

interparticle transport criterion <strong>for</strong> heat transport is considered much stricter than that <strong>for</strong><br />

intraparticle transport [49].<br />

The mass transport at the intraparticle scale range is analyzed <strong>for</strong> the internal diffusion within<br />

the SOFC anode and the Knudsen diffusion is taken under consideration in the calculations.<br />

The effective diffusivity by Knudsen diffusion is defined as [51]:<br />

(3.54)<br />

where d p is the particle diameter and the molecular weight M AB <strong>of</strong> substance A and B is<br />

defined as in equation (3.11). The effective diffusivity which is based on the ordinary<br />

diffusion is defined here as in equation (3.7). Both <strong>of</strong> the effective diffusivities are then<br />

averaged as below [51]:<br />

(3.55)<br />

The averaged effective diffusivity is needed in the Thiele Modulus and is here defined as in<br />

[53]:<br />

(3.56)<br />

where r p is the particle radius, k c the mass transfer coefficient, C the concentration, n the<br />

reaction order and D eff the effective diffusivity.<br />

The effectiveness factor is defined in words as [49, 51]:<br />

It is calculated as:<br />

(3.57)<br />

where Φ is the Thiele Modulus. Also, to ensure η ≥ 0.95 it is required that [49]:<br />

(3.58)<br />

34


<strong>for</strong> an isothermal spherical particle where r r is the reaction rate, r p particle radius, C the<br />

concentration and D eff the effective diffusivity. The results <strong>for</strong> the control <strong>of</strong> the criteria <strong>for</strong><br />

the heat and mass transfer limitations are provided in the next chapter.<br />

35


4 Results<br />

and discussion<br />

This section presents the results from the LB model and the criteria <strong>for</strong> the kinetic parameters<br />

are checked so that no critical effects occur on the transport processes. Both the LB model<br />

and the validation <strong>of</strong> the kinetic effects are viewed from a microscale perspective. Also the<br />

results <strong>of</strong> the macroscale model are provided and divided into two parts; change <strong>of</strong> internal<br />

re<strong>for</strong>ming reaction rate model and change <strong>of</strong> amount <strong>of</strong> methane content and steam-to-fuel<br />

ratio.<br />

4.1 Microscale model by LBM<br />

The LB model stops when the maximum deviations <strong>of</strong> the mean velocity differ with less than<br />

10 -10 over the last iteration. Reynolds number is calculated based on the velocity and is<br />

relatively low Re typically in the order <strong>of</strong> 0.1 to 1. The physical geometry <strong>of</strong> the LB model<br />

and material data <strong>for</strong> the anode is presented in Table 4.1. In the LB model discrete units are<br />

used <strong>for</strong> the length and time. The lattice unit lu represents the fundamental measure <strong>of</strong> length<br />

and time step ts the measure <strong>of</strong> time.<br />

Table 4.1: Anode geometry and relevant parameters [1].<br />

Anode<br />

Size<br />

Length 40 lu, 200 lu 1<br />

Height 10 lu, 50 lu 1<br />

Porosity (ε) 0.4<br />

Inlet mole fraction<br />

H 2 0.9<br />

H 2 O 0.1<br />

The study is conducted in an order <strong>of</strong> increasing complex geometries to validate the method<br />

<strong>for</strong> future modeling <strong>of</strong> all the physical processes in an SOFC. First <strong>of</strong> all, a small test is<br />

1 Two different values <strong>for</strong> this parameter are tested.<br />

36


Velocity<br />

carried out to check whether the LBM shows comparable results with an analytical solution.<br />

For this case, the velocity pr<strong>of</strong>ile <strong>for</strong> a channel is modeled by LBM and compared to the<br />

analytical solution <strong>of</strong> a Poiseuille flow. The test can be seen as comparable to a fuel cell<br />

channel. The results can be seen in Figure 4.1 and the agreement between the two velocity<br />

pr<strong>of</strong>iles is good.<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

u x<br />

(Analytical Poiseuille)<br />

u x<br />

(LBM)<br />

0.02<br />

0<br />

0 5 10 15 20 25 30 35 40<br />

Distance [nodes]<br />

Figure 4.1: Velocity pr<strong>of</strong>ile [lu/ts] <strong>for</strong> a cross section in the middle <strong>of</strong> a channel domain compared to<br />

the analytical solution <strong>of</strong> a Poiseuille flow.<br />

Secondly, a circular obstacle is placed in the channel to increase the complexity <strong>of</strong> geometry.<br />

In Figure 4.2 shows the flow field past a circular obstacle in a channel. The case was able to<br />

be simulated by LBM with good results in MATLAB. Both the wake and the no-slip at the<br />

walls are obtained efficiently.<br />

37


0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

Figure 4.2: Velocity contours [lu/ts] in a channel with a circular obstacle.<br />

0.525<br />

0.52<br />

0.515<br />

0.51<br />

0.505<br />

0.5<br />

Figure 4.3: Mole fraction in a channel with a circular obstacle.<br />

In Figure 4.3 a simple mass diffusion case is provided <strong>for</strong> the channel with a circular obstacle<br />

where the convection-diffusion approach is applied. The component enters with almost its<br />

maximum <strong>of</strong> 0.525 and diffuses continuously through the channel. Around the obstacle, there<br />

is a slight bounce-back be<strong>for</strong>e and a wake behind. This case is just a test case with fictitious<br />

numbers and provides useful in<strong>for</strong>mation <strong>of</strong> LBMs functionality.<br />

38


Figure 4.4: Velocity field [lu/ts] in part <strong>of</strong> the porous media domain with a porosity <strong>of</strong> 0.40.<br />

Thirdly, a porous geometry is tested <strong>for</strong> simulation <strong>of</strong> an SOFC anode. In Figure 4.4, a porous<br />

domain is provided with a porosity <strong>of</strong> 0.4. Here the velocity field is given to illustrate that<br />

LBM can easily handle a porous domain. Note that only part <strong>of</strong> the porous domain is shown<br />

from Figure 3.2 to see the velocity arrows better and the following Figure 4.5 and 4.6 show<br />

the whole modeled domain as can be seen in Figure 3.2. The velocity arrows provide an<br />

understanding <strong>of</strong> the bounce-back theory and provide intuitive feeling <strong>for</strong> the flow process in<br />

the porous media.<br />

0.8<br />

0.75<br />

0.7<br />

0.65<br />

0.6<br />

0.55<br />

0.5<br />

0.45<br />

0.4<br />

Figure 4.5: Mole fraction distribution <strong>of</strong> H 2 in a porous media with a porosity <strong>of</strong> 0.40.<br />

39


In Figure 4.5 shows the whole modeled porous domain <strong>for</strong> the mole fraction distribution <strong>of</strong><br />

H 2 . The inlet mole fraction is specified as x H2 = 0.9 and the mole flux is specified at the<br />

outlet. Note that no reaction effects are included in the model. The mass diffusion <strong>of</strong><br />

hydrogen predicts, in a similar manner as the simpler channel case, the continuous reduction<br />

<strong>of</strong> hydrogen along the flow direction and the contours <strong>of</strong> the mole fraction around the<br />

obstacles normal to the surface indicates that mass diffusion occurs parallel to the surface.<br />

0.45<br />

0.4<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

Figure 4.6: Mole fraction distribution <strong>of</strong> H 2 O in a porous media with a porosity <strong>of</strong> 0.40.<br />

In Figure 4.6 the mole fraction distribution <strong>of</strong> H 2 O is shown <strong>for</strong> the whole modeled domain.<br />

The inlet mole fraction is specified as x H2O = 0.1 and the mole flux is specified at the outlet in<br />

a negative direction, i.e., normal to the boundary interface in opposite flow direction. Note<br />

that no reaction effects are included. This is a heavier species where the diffusion is not as<br />

fast as <strong>for</strong> the hydrogen case. Only the interaction between the two species is illustrated in this<br />

case. The local spots with higher mole fraction are because these are a closed pores and the<br />

velocity effect is very low there. For future studies, the production and consumption <strong>of</strong><br />

species will be included and the effect <strong>of</strong> the chemical reactions on the mass density<br />

distribution.<br />

4.2 Evaluation <strong>of</strong> kinetics at smaller scales<br />

As mentioned be<strong>for</strong>e, the analysis is carried out <strong>for</strong> the steam re<strong>for</strong>ming reaction in the anode<br />

and the electrochemical reactions in the anode and cathode. The most significant parameters<br />

concerning the cell structure and catalytic activity, which are used <strong>for</strong> the calculations, are<br />

presented in Table 4.2, while the inlet and operating conditions are presented in Table 4.3.<br />

Note that the x-direction is the main flow direction and the direction set at the inlet. The y-<br />

direction is normal to the main flow direction.<br />

40


Table 4.2: Parameters in the SOFC analysis.<br />

Parameter<br />

Value<br />

Cell length<br />

Anode thickness<br />

Cathode thickness<br />

Particle diameter<br />

0.1 m<br />

500 μm<br />

50 μm<br />

0.34 μm<br />

Anode TPB thickness (y-dir) 10 μm, 1 μm 2<br />

Cathode TPB thickness (y-dir) 10 μm, 1 μm 2<br />

Tortuosity 3<br />

Porosity 0.3<br />

Velocity v x,A<br />

Velocity v y,A<br />

Velocity v x,C<br />

Velocity v y,C<br />

Enthalpy change (MSR 3 )<br />

Activation energy (MSR 3 )<br />

Activation energy (CE 4 )<br />

Activation energy (AE 5 )<br />

Reaction rate (MSR 3 ), close to inlet<br />

Reaction rate (CE 4 )<br />

Reaction rate (AE 5 )<br />

6.9 ∙10 -5 m/s<br />

1.5∙10 -3 m/s<br />

1∙10 -3 m/s<br />

7.5 ∙10 -4 m/s<br />

226 kJ/mol<br />

82 kJ/mol<br />

71 kJ/mol<br />

185 kJ/mol<br />

15 mol/m³/s<br />

0.53 kmol/m³/s,<br />

5.3 kmol/m³/s 2<br />

1.06 kmol/m³/s,<br />

10.1 kmol/m³/s 2<br />

It should be noted that <strong>for</strong> the thickness <strong>of</strong> the anode and cathode TPB as well as the reaction<br />

rate <strong>for</strong> the anode and the cathode have been tested <strong>for</strong> two different values <strong>of</strong> each<br />

parameter.<br />

2 Two different values <strong>for</strong> this parameter are tested.<br />

3 MSR stands <strong>for</strong> methane steam re<strong>for</strong>ming reaction<br />

4 CE <strong>for</strong> the electrochemical reactions in the cathode<br />

5 AE <strong>for</strong> the electrochemical reactions in the anode<br />

41


Table 4.3: Inlet and operating conditions.<br />

Parameter<br />

Value<br />

Fuel utilization 80 %<br />

Oxygen utilization 20 %<br />

Inlet mole fraction <strong>of</strong> methane 0.171<br />

Inlet mole fraction <strong>of</strong> hydrogen 0.2626<br />

Inlet mole fraction <strong>of</strong> water 0.4934<br />

Inlet mole fraction <strong>of</strong> carbon monoxide 0.0294<br />

Inlet mole fraction <strong>of</strong> carbon dioxide 0.0436<br />

Inlet temperature<br />

Pressure<br />

Average current density<br />

1000 K<br />

1 atm<br />

3000 A/m²<br />

The results are presented in Table 4.4 and commented below <strong>for</strong> all levels <strong>of</strong> the analysis <strong>of</strong><br />

the kinetic criteria.<br />

Table 4.4: Results from the criteria analysis.<br />

Transport domain Heat transport Mass transport<br />

Interparticle<br />

No limitation<br />

–<br />

Equation (3.40): ~10 -5 –10 -9 < ~10 -2<br />

Interphase<br />

No limitation<br />

Bi = 2 – 4∙10 -2 < 10<br />

No limitation<br />

Equation (3.45):<br />

~10 -7 –10 -8 < ~1–10 -2<br />

Intraparticle No limitation No limitation<br />

Equation (3.61):<br />

η ≈ 0.999 > 0.95<br />

The results are homogeneous with no limitations <strong>for</strong> any <strong>of</strong> the involved scales and reactions.<br />

As the criterion is fulfilled <strong>for</strong> the heat transport at the interparticle transport level then no<br />

limitation occurs <strong>for</strong> the heat transport at interphase or intraparticle level. The main difference<br />

between the steam re<strong>for</strong>ming reaction and the electrochemical reactions is that the latter ones<br />

have a somewhat higher reaction rate. These reactions occur only in a small part <strong>of</strong> the<br />

anode/cathode domain and the reaction rate is highest at the boundary between the anode and<br />

electrolyte or between cathode and electrolyte. But the criteria <strong>for</strong> the electrochemical<br />

42


eactions are safely fulfilled. This is a good verification <strong>of</strong> the chosen parameters <strong>for</strong> the<br />

computational model examined in previous studies.<br />

However, it is interesting to reflect over which parameters can <strong>for</strong>m a potential risk <strong>for</strong><br />

limiting the transport processes. For the heat transport at all the domains, a larger enthalpy<br />

change or an increased reaction rate can increase the risk. Also, a lowering <strong>of</strong> the temperature<br />

can cause an increased risk but this is less significant. For the mass transport, an increased<br />

reaction rate or a decreased concentration <strong>of</strong> reactant can cause a higher risk. As mentioned<br />

be<strong>for</strong>e the effectiveness factor is affected by the characteristic dimensions <strong>of</strong> the particle<br />

which can also be confirmed here. If a change in particle diameter causes a change in the<br />

reaction rate, this may have a strong influence on the transport processes [6].<br />

To summarize, the elimination <strong>of</strong> transport gradients which limit the reaction and catalytic<br />

kinetics is complex to study. This case study, by tools to assess the transport limiting issues,<br />

seeks to locate the limiting sources and improve these <strong>for</strong> the desired outcome. The reaction<br />

rate is the most direct risk <strong>for</strong> limitation on the transport processes. If the reaction rate is<br />

increased it will affect every criterion in the analysis and can cause severe gradients which<br />

will create transport limitations. The anode and cathode structure and catalytic characteristics<br />

have an impact on the reaction rates, especially on the steam re<strong>for</strong>ming reaction, which will in<br />

turn affect the cell per<strong>for</strong>mance.<br />

4.3 Macroscale model by CFD<br />

A two-dimensional model <strong>for</strong> an anode-supported SOFC has been developed and<br />

implemented in the commercial s<strong>of</strong>tware COMSOL Multiphysics (version 3.5a). Equations<br />

<strong>for</strong> momentum, mass and heat transport are solved simultaneously. The cell geometry and<br />

SOFC operating parameters are defined in Table 4.3. It should be mentioned that this<br />

macroscale model is 2D only, and the connection between the electrodes and interconnect<br />

cannot be explicitly observed in this case.<br />

4.3.1 Case study: Internal re<strong>for</strong>ming reaction rates<br />

The flow direction is set to be from left to right <strong>for</strong> air and fuel channels as well as the anode<br />

and the cathode. It is also possible with counter flow but this is not included in this study. It<br />

should be explicitly mentioned that the length <strong>of</strong> the cell is 100 times longer than the height <strong>of</strong><br />

the air or the fuel channel.<br />

43


Figure 4.7: Temperature distribution <strong>for</strong> Leinfelder’s kinetics.<br />

The predicted gas phase temperature in the cell is plotted in Figure 4.7 <strong>for</strong> Leinfelder’s<br />

kinetics. Achenbach & Riensche’s and Drescher’s kinetics are not shown here due to<br />

similarity in the plots. In the fuel and air channels, there is a decrease in temperature after a<br />

short distance from the inlet. In the fuel channels it is due to the steam re<strong>for</strong>ming reaction,<br />

which consumes the heat when the methane is re<strong>for</strong>med to hydrogen and carbon monoxide.<br />

The temperature on the air side is lower due to a higher air flow rate which affects the<br />

convective heat transfer. The decrease in temperature close to the inlet is 50 K <strong>for</strong> both<br />

Achenbach & Riensche’s and Leinfelder’s kinetics. The temperature distribution <strong>for</strong><br />

Drescher’s kinetics does not drop initially as much as the other two. The area <strong>of</strong> the<br />

temperature drop is larger <strong>for</strong> Achenbach & Riensche’s kinetics than those from Leinfelder’s<br />

and Drescher’s kinetics. But the recovery to a higher temperature occurs faster <strong>for</strong> both<br />

Leinfelder’s and Drescher’s kinetics than <strong>for</strong> Achenbach & Riensche’s kinetics. This might be<br />

due to the fact that the latter is affected by the fast conversion <strong>of</strong> methane to hydrogen and<br />

carbon monoxide.<br />

44


Figure 4.8: Mole fraction distribution in the middle <strong>of</strong> the anode along the flow direction <strong>for</strong><br />

Achenbach & Riensche’s (left) and Drescher’s kinetics (right).<br />

The effect on mole fraction distribution <strong>for</strong> the different gas species is similar <strong>for</strong> both<br />

Achenbach & Riensche’s and Leinfelder’s kinetics and there<strong>for</strong>e only the mole fraction<br />

distribution <strong>for</strong> Achenbach & Riensche’s (left) along with Drescher’s kinetics are presented in<br />

Figure 4.8. Drescher’s kinetics obtained the maximum mole fraction <strong>of</strong> hydrogen faster and a<br />

higher maximum mole fraction than Achenbach & Riensche’s and Leinfelder’s kinetics. The<br />

initial consumption <strong>of</strong> water and the initial generation <strong>of</strong> hydrogen <strong>for</strong> Drescher’s kinetics<br />

result in larger gradients <strong>of</strong> the mole fractions. All three kinetics are fast although Drescher’s<br />

kinetics, expressed by a Langmuir-Hinshelwood type, differs slightly more from the others. It<br />

deserves to be pointed out that Drescher’s kinetics includes both positive and negative orders<br />

<strong>of</strong> the partial pressure <strong>of</strong> methane and water, as well as two different activation energies <strong>for</strong><br />

the denominator and the numerator, which can have some effect on the results.<br />

Figure 4.9: Reaction rate distribution in the middle <strong>of</strong> the anode along the flow direction <strong>for</strong> Achenbach<br />

& Riensche’s (left) and Drescher’s model (right).<br />

45


The reaction rates <strong>for</strong> both the steam re<strong>for</strong>ming reaction and the water-gas shift reaction are<br />

plotted in Figure 4.9 <strong>for</strong> Achenbach & Riensche’s (left) and Drescher’s kinetics (right). It<br />

should be clearly noted that the reaction rates are only plotted <strong>for</strong> the entrance region, through<br />

0.01 m. Close to the inlet where the concentration <strong>of</strong> methane is high the reaction rate <strong>for</strong> the<br />

steam re<strong>for</strong>ming reaction is high. The reaction rate <strong>for</strong> the steam re<strong>for</strong>ming and the water-gas<br />

shift is much higher <strong>for</strong> Drescher’s kinetics than <strong>for</strong> both Achenbach & Riensche’s and<br />

Leinfelder’s kinetics. Leinfelder’s kinetics is higher than Achenbach & Riensche’s. Close to<br />

the inlet in the anode where the carbon monoxide generation is high, the reaction rate <strong>for</strong> the<br />

water-gas shift reaction is at the highest. The high generation <strong>of</strong> carbon monoxide is due to<br />

the steam re<strong>for</strong>ming reaction. Furthermore, more hydrogen is produced when steam is<br />

generated due to the fact that the water-gas shift equation is in equilibrium through the<br />

process. As hydrogen is consumed, steam is generated thanks to the electrochemical reaction<br />

at the TPB. The reaction rate <strong>for</strong> the water-gas shift reaction reaches a higher value due to the<br />

faster reaction rate <strong>for</strong> the steam re<strong>for</strong>ming reaction <strong>of</strong> the Drescher’s and Leinfelder’s<br />

kinetics compared to Achenbach & Riensche’s. The comparison between the different kinetic<br />

models needs to be evaluated on a more detailed level as it cannot be correctly explained by<br />

just a few empirical parameters, such as the activation energy and the pre-exponential value.<br />

To fully understand the effect and dependence <strong>of</strong> the parameters, microscale modeling is<br />

needed. What can be concluded from this study is that the configuration and geometrical<br />

properties <strong>of</strong> the anode and the chemical composition and catalytic characteristics are<br />

important. To draw firm conclusions from the modeling work, it is important to reveal the<br />

difference <strong>of</strong> the kinetic models from experimental work carried out on SOFC and a re<strong>for</strong>mer<br />

based on the same properties.<br />

A parameter study was carried out <strong>for</strong> Leinfelder’s kinetics by increasing the inlet<br />

temperature by 50 K. The other parameters were kept the same as in the base case. The<br />

temperature distributions <strong>for</strong> both cases result in similar effects but obviously resulted in a<br />

higher temperature range. The mole fraction distribution <strong>of</strong> the fuel gas species was<br />

maintained in the same range and trend as the base case at 1000 K. The reaction rates are<br />

slightly higher <strong>for</strong> a higher inlet temperature. Due to the increased inlet temperature the<br />

maximum reaction rates are almost doubled compared to the base case at the inlet.<br />

Another parameter study was conducted <strong>for</strong> the active surface area-to-volume ratio (AV) from<br />

10 ∙10 4 to 5 ∙10 5 m 2 /m 3 , which is a frequently used interval in the literature. All the other<br />

parameters were kept the same as the base case. The temperature pr<strong>of</strong>ile and each mole<br />

fraction pr<strong>of</strong>ile were distributed in a similar overall trend as <strong>for</strong> the base case. The mole<br />

fractions reach approximately the same maximum value <strong>for</strong> different AV but occur at different<br />

distances from the inlet. A higher ratio moves in the maximum <strong>of</strong> all the mole fraction species<br />

closer to the inlet. The characteristics <strong>of</strong> reaction rates <strong>for</strong> Leinfelder’s kinetics with an<br />

increased AV are distributed similarly to the base case, but the maximum value is more or less<br />

doubled <strong>for</strong> an increased active surface area to volume ratio.<br />

4.3.2 Case study: Methane content and steam-to-fuel ratio<br />

A case study was per<strong>for</strong>med to simulate biogas as a fuel by varying the amount <strong>of</strong> methane<br />

and SF. Three cases were per<strong>for</strong>med, all with 60% CH 4 content but the SF was varied<br />

between 1, 3, and 5. Additionally, three cases with SF equal to 3 but CH 4 varied as 45%, 60%<br />

and 75%. For all the cases, a small fraction <strong>of</strong> hydrogen is added to enable electrochemical<br />

reactions close to the inlet as well. CO is also added, similarly as H 2 , to achieve numerical<br />

stability. Steam was added to avoid carbon deposition and also to be used in the re<strong>for</strong>ming<br />

reactions. The amount <strong>of</strong> H 2 O was calculated from the relationship SF = [H 2 O]/[CH 4 ] as SF<br />

is specified <strong>for</strong> the different cases. In Table 4.5 the inlet mole fractions are presented <strong>for</strong> the<br />

different cases with varying amount <strong>of</strong> methane content and SF.<br />

46


Table 4.5: Inlet mole fractions <strong>for</strong> the different case studies.<br />

Case models CH 4 CO 2 H 2 O<br />

CH 4 0.45, SF=3 0.191 0.235 0.574<br />

CH 4 0.60, SF=3 0.214 0.143 0.643<br />

CH 4 0.75, SF=3 0.231 0.076 0.692<br />

CH 4 0.60, SF=1 0.375 0.25 0.375<br />

CH 4 0.60, SF=5 0.15 0.1 0.75<br />

Similar to the previous section, the fuel flow rates <strong>for</strong> the different cases were calculated to<br />

keep the fuel utilization at 80 percent. Note that each molecule <strong>of</strong> methane corresponds to a<br />

generation <strong>of</strong> four molecules <strong>of</strong> hydrogen by both the steam re<strong>for</strong>ming and water-gas shift<br />

reaction and each molecule <strong>of</strong> carbon monoxide corresponds to one molecule generation <strong>of</strong><br />

hydrogen. The flow rate <strong>of</strong> air was kept constant <strong>for</strong> all cases and the oxygen utilization is set<br />

to 20 percent. The inlet temperature was set to 1100 K to ensure functional conversion <strong>of</strong> the<br />

fuel and the current density was kept constant at 3000 A/m 2 . The Knudsen diffusion was<br />

neglected in this model to reduce computational cost. For both the steam re<strong>for</strong>ming reaction<br />

and the water-gas shift reaction, the equilibrium model is chosen <strong>for</strong> the kinetic model.<br />

Figure 4.10: Mole fraction distribution in the middle <strong>of</strong> the anode <strong>for</strong> 45% CH 4 (left) and 75% CH 4<br />

(right).<br />

In Figure 4.10, the mole fractions at the centerline <strong>of</strong> the anode along the whole cell length<br />

are presented. The change <strong>of</strong> H 2 O between the different cases is directly connected to the<br />

methane content, in this case, to ensure significant steam at the inlet <strong>of</strong> the cell. Depending on<br />

the inlet fraction a decrease in the mole fraction <strong>of</strong> water can be observed close to the inlet.<br />

The biogas contains no hydrogen in the collected data but <strong>for</strong> numerical calculations the<br />

hydrogen is initially set to have a small inlet value to enable the electrochemical reactions at<br />

the inlet <strong>of</strong> the cell. The variations <strong>of</strong> the mole fractions are however quite small and it is<br />

mostly CO 2 that changes. The reason <strong>for</strong> this is that CO 2 is linked to the choice <strong>of</strong> CH 4 and<br />

47


H 2 O. Both CO and H 2 are initially set to small values to enable the numerical calculations, but<br />

CO 2 and H 2 O are decided in accordance with the chosen methane content and SF.<br />

Figure 4.11: Reaction rate distribution in the middle <strong>of</strong> the anode <strong>for</strong> 45% CH 4 (left) and 75% CH 4<br />

(right).<br />

The reaction rates <strong>for</strong> 45% and 60% methane contents are rather similarly distributed, but in<br />

the case <strong>of</strong> 75 % methane the steam re<strong>for</strong>ming reaction rate has initially higher values and<br />

much lower values at the outlet <strong>of</strong> the cell. Also, the water-gas-shift reaction rate <strong>for</strong> high CH 4<br />

content increases to the maximum value closer to the inlet <strong>of</strong> the cell than it does <strong>for</strong> the other<br />

two cases. For the situations 45% and 75% content <strong>of</strong> methane shown in Figure 4.11, the<br />

steam re<strong>for</strong>ming reaction rate (within the anode) is high as long as a high concentration <strong>of</strong><br />

methane is available. The reaction rate increases as the temperature and concentration <strong>of</strong><br />

steam increase, and decreases as the concentration <strong>of</strong> methane decreases. Note that there is a<br />

difference in scale between the x- and y-axes. It is possible to change the reaction rate, either<br />

by changing the particle size <strong>of</strong> the active catalyst, catalytic material composition or the<br />

porous structure, i.e., the active catalytic area. The limitation to be considered is that the<br />

probability <strong>of</strong> carbon deposition increases where there is almost no hydrogen present. A<br />

higher risk <strong>for</strong> carbon deposition occurs when there is a high temperature gradient close to the<br />

cell inlet. This is not the case here as the gradients are not so high.<br />

48


Figure 4.12: Mole fraction distribution in the middle <strong>of</strong> the anode along the flow direction <strong>for</strong> SF = 1<br />

(left) and SF = 5 (right).<br />

Only the two extreme cases are shown here to visualize the effect <strong>of</strong> a change in SF. In Figure<br />

4.12 the mole fraction <strong>for</strong> SF=1 can be viewed to the left and SF=5 to the right. It is<br />

important to verify that there is a sufficient amount <strong>of</strong> H 2 O throughout the cell or else no<br />

efficient reaction will occur and there will be a risk <strong>for</strong> carbon deposition. In the figure <strong>for</strong> the<br />

mole fractions, it can be seen that a drop <strong>of</strong> H 2 O exists slightly downstream the inlet. Instead<br />

<strong>of</strong> putting all focus on the inlet mole fraction <strong>of</strong> H 2 O, one should also consider whether there<br />

is a sufficient amount to handle this drop and adjust the inlet mole fraction subsequently. This<br />

mole fraction drop increases when a faster reaction rate is applied which was shown in the<br />

previous section.<br />

Figure 4.13: Reaction rate distribution in the middle <strong>of</strong> the anode <strong>for</strong> SF = 1 (left) and SF = 5 (right).<br />

49


The reaction rates <strong>for</strong> SF=1 (left) and SF=5 (right) are presented in Figure 4.13. It should be<br />

mentioned that the temperature distribution was overall quite similar <strong>for</strong> the cases. When<br />

SF=1 the temperature was overall lower <strong>for</strong> the larger part <strong>of</strong> the cell. For the case with<br />

SF=5, the lowest temperature was much closer to the inlet than <strong>for</strong> SF=1 and the three cases<br />

with varying methane content. The temperature distribution has an effect on the reaction rates.<br />

The pr<strong>of</strong>iles <strong>of</strong> the reaction rates are quite different. For SF=1, the maximum rate value <strong>of</strong> the<br />

water-gas-shift reaction is not reached until just upstream from the outlet but, on the other<br />

hand, <strong>for</strong> SF=5, it is reached close to the inlet. The steam re<strong>for</strong>ming reaction shows the same<br />

tendency but it is slightly higher <strong>for</strong> SF=1 to begin with. The water-gas-shift reaction rate was<br />

low because the mole fraction <strong>of</strong> CO is initially so small. Furthermore, the water-gas shift<br />

reaction is connected to the steam re<strong>for</strong>ming reaction, which affects the pr<strong>of</strong>ile throughout the<br />

cell.<br />

50


5 Conclusions<br />

The physics and the transport processes in SOFCs can be described at different length and<br />

time scales. This constitutes a challenge <strong>for</strong> the development <strong>of</strong> multiscale models <strong>for</strong> fuel<br />

cell simulations. In this study, a LBM microscale model was developed <strong>for</strong> the D2Q9 case<br />

(two-dimensional nine speed case). The kinetic model was examined so that no severe<br />

limiting effects on heat and mass transport occurred. Also, a FEM based model <strong>for</strong> an anodesupported<br />

SOFC was developed to better understand the internal re<strong>for</strong>ming reactions <strong>of</strong><br />

methane and the effects on the transport processes. The model was implemented in COMSOL<br />

Multiphysics <strong>for</strong> the analysis <strong>of</strong> three different kinetic models found in the literature. An<br />

equilibrium equation was employed <strong>for</strong> the water-gas shift re<strong>for</strong>ming reaction rate. Parameter<br />

studies were also conducted <strong>for</strong> the methane content and SF.<br />

Five conclusions can be made in this study. First, LBM was found to be a functional method<br />

to microscale modeling predicting the velocity pr<strong>of</strong>ile and mass diffusion well. LBM could<br />

handle both a simple geometry as a channel to a more complex geometry such as a porous<br />

media. For the velocity field, the LBM was able to illustrate the flow correctly around the<br />

obstacles. The mass diffusion <strong>for</strong> hydrogen was reduced from the inlet to the outlet as<br />

expected and contours were seen around the obstacles where mass diffusion <strong>of</strong> hydrogen<br />

occurred parallel to the surface. The detailed in<strong>for</strong>mation from LBM at microscale regarding<br />

the transport processes and chemical reactions can improve the macroscale model by<br />

including this in<strong>for</strong>mation <strong>for</strong> the TPB areas.<br />

Second, it was shown that the reaction rates were very fast and differed slightly across the<br />

three models due to the great differences <strong>of</strong> the pre-exponential value and the activation<br />

energy. The model was found to be sensitive to variation <strong>of</strong> the steam re<strong>for</strong>ming reaction rate.<br />

Both the inlet temperature and active surface area to volume ratio showed an effect on the<br />

reaction rates in terms <strong>of</strong> the maximum value.<br />

Third, it was found that a fuel containing a high percentage <strong>of</strong> methane in combination with a<br />

high inlet temperature produced a steep temperature gradient close to the cell inlet. Fourth, a<br />

higher steam-to-fuel ratio showed a decreased risk <strong>of</strong> carbon deposition at the anode catalytic<br />

active area.<br />

Finally, there was no direct significant risk <strong>for</strong> heat and mass transport limitations <strong>for</strong> the<br />

SOFC model with the kinetic parameters in this study. Care should be taken if the reaction<br />

rate is increased since this will affect almost every criterion in the analysis. It transpired not to<br />

be sufficient only to describe the reaction rates with a few empirical parameters. It was<br />

necessary to develop a suitable microscale model <strong>for</strong> the SOFC. However, the global kinetic<br />

models have still predicted valuable behaviors. The reason why the kinetics models differed<br />

to a large extent is that they were sensitive to how the experiment was designed.<br />

51


6 Future<br />

work<br />

Future work will involve a study <strong>of</strong> an SOFC at multiscale which will <strong>of</strong>fer promising<br />

knowledge to understand in detail the effect <strong>of</strong> design. To approach a successful electricity<br />

producing device with improved durability and life time, the understanding <strong>of</strong> multiscale<br />

transport and reaction phenomena within the cell is crucial. The next step is to model the<br />

involvement <strong>of</strong> microscale thermal diffusion through LBM connected to a macroscale CFD<br />

model. In the extended model the Knudsen diffusion, which describes collisions between the<br />

gas molecules and the porous structure (inside the porous electrodes), is also taken into<br />

account. Also the electrochemical reactions are prospected to contribute to capture valuable<br />

microstructural effects. These reactions occur at a limited part <strong>of</strong> the cell, the TPB, which can<br />

only be captured if modeled at microscale or smaller. Here the Monte Carlo method could<br />

<strong>of</strong>fer advantages to improve the multiscale development.<br />

Another extension <strong>of</strong> the model is to include catalytic chemical surface reactions (instead <strong>of</strong><br />

global kinetics expressions). These surface reactions can provide knowledge <strong>of</strong> the interaction<br />

between the transport processes and the reactions which will be valuable <strong>for</strong> fuel cell model<br />

development. The re<strong>for</strong>ming reaction rate is dependent on temperature, concentrations, type<br />

and catalyst available. If the chemical reactions can be simulated on a microscale level, it<br />

would open up to involve all the detailed multistep chemical reactions. More knowledge and<br />

understanding <strong>of</strong> the effect behind the activation energy is important to enable reduction <strong>of</strong><br />

the operating temperature. Physical and material properties are calculated from data found in<br />

literature and there<strong>for</strong>e experimental work is desired <strong>for</strong> validation <strong>of</strong> the model.<br />

52


7 References<br />

[1] Andersson M., Yuan J., Sundén B., Review on Modeling Development <strong>for</strong><br />

Multiscale Chemical Reactions coupled Transport Phenomena in Solid Oxide Fuel<br />

Cells, Applied Energy, 87, pp. 1461-1476, 2010<br />

[2] Lehnert W., Meusinger J., Thom F., Modelling <strong>of</strong> Gas Transport Phenomena in<br />

SOFC Anodes, J. Power Sources, 87, pp. 57-63, 2000<br />

[3] Sukop M.C., Thorne D.J. Jr., Lattice Boltzmann Modeling – An Introduction <strong>for</strong><br />

Geoscientists and Engineers, Springer, NY, U.S., 2007<br />

[4] Sanchez D., Chacartegui R., Munoz A., Sanchez T., On the Effect <strong>of</strong> Methane<br />

Internal Re<strong>for</strong>ming in Solid Oxide Fuel Cells, Int. J. Hydrogen Energy, 33, pp. 1834-<br />

1844, 2008<br />

[5] Janardhanan V., Deutschmann O., CFD Analysis <strong>of</strong> a Solid Oxide Fuel Cell with<br />

Internal Re<strong>for</strong>ming, J. Power Sources, 162, pp. 1192-1202, 2006<br />

[6] Hecht E., Gupta G., Zhu H., Dean A., Kee R., Maier L., Deutschmann O., Methane<br />

Re<strong>for</strong>ming Kinetics within a Ni-YSZ SOFC Anode Support, Applied Catalysis A:<br />

General, 295, pp. 40-51, 2005<br />

[7] Yang Y., Du X., Yang L., Huang Y., Xian H., Investigation <strong>of</strong> Methane Steam<br />

Re<strong>for</strong>ming in Planar Porous Support <strong>of</strong> Solid Oxide Fuel Cell, Applied Thermal<br />

<strong>Engineering</strong>, 29, pp. 1106-1113, 2009<br />

[8] Yuan J., Huang Y., Sundén B., Wang W.G., Analysis <strong>of</strong> Parameter Effects on<br />

Chemical Coupled Transport Phenomena in SOFC Anodes, Heat Mass Transfer, 45,<br />

pp. 471-484, 2009<br />

[9] Yuan J., Sundén B., Analysis <strong>of</strong> Chemically Reacting Transport Phenomena in an<br />

Anode Duct <strong>of</strong> Intermediate Temperature SOFCs, ASME J. Fuel Cell Sci. Tech., 3,<br />

pp. 687-701, 2006<br />

[10] Nam J. H., Jeon D. H., A Comprehensive Micro-scale Model <strong>for</strong> Transport and<br />

Reaction in Intermediate Temperature Solid Oxide Fuel Cells, Electrochim. Acta, 51<br />

pp. 3446-3460, 2006<br />

53


[11] Kakac S., Pramuanjaroenkij A., Zhou X., A Review <strong>of</strong> Numerical Modeling <strong>of</strong> Solid<br />

Oxide Fuel Cells, Int. J. Hydrogen Energy, 32, pp. 761-786, 2007<br />

[12] Andersson M., Paradis H., Yuan J., Sundén B., Review <strong>of</strong> Catalyst Materials and<br />

Catalytic Steam Re<strong>for</strong>ming Reactions in SOFC Anodes, Int. J. Energy Research,<br />

2011 (in press)<br />

[13] Li. P. W., Schaefer L., Chyu M. K., Multiple Transport Processes in Solid Oxide<br />

Fuel Cells, Chapter 1 in Sundén B., Faghri M. (eds.), Transport Phenomena in Fuel<br />

Cells, WIT Press, UK, 2005<br />

[14] Grew K.N., Joshi A.S., Peracchio A.A., Chiu W.K.S., Pore-Scale Investigation <strong>of</strong><br />

Mass Transport and Electrochemistry in a Solid Oxide Fuel Cell Anode, J. Power<br />

Sources, 195, pp. 2331-2345, 2010<br />

[15] Joshi A.S., Peracchio A.A., Grew K.N.,Chiu W.K.S., Lattice Boltzmann Method <strong>for</strong><br />

Multi-Component, Non-Continuum Mass Diffusion, J. Phys. D: Appl. Phys., 40, pp.<br />

7593-7600, 2007<br />

[16] Joshi A.S., Grew K.N., Peracchio A.A., Chiu W.K.S., Lattice Boltzmann Modeling<br />

<strong>of</strong> a 2D Gas Transport in a Solid Oxide Fuel Cell Anode, J. Power Sources., 164, pp.<br />

631-638, 2007<br />

[17] Wang X., Lau K.C., Turner C.H., Dunlap B.I., Kinetic Monte Carlo Simulation <strong>of</strong><br />

the Elementary Electrochemistry in a Hydrogen-powered Solid Oxide Fuel Cell, J.<br />

Power Sources, 195, pp. 4177-4184, 2010<br />

[18] Voter A.F., Introduction to the Kinetic Monte Carlo Method, Theoretical Division,<br />

Los Alamos National Laboratory, Report <strong>for</strong> U.S. DOE, 2005<br />

[19] Lau K.C., Turner C.H., Dunlap B.I., Kinetic Monte Carlo Simulation <strong>of</strong> the Yttria<br />

Stabilized Zirconia (YSZ) Fuel Cell Cathode, Solid State Ionics, 179, pp. 1912-1920,<br />

2008<br />

[20] Tarancón A., Morata A., Peiró F., Dezanneau G., A Molecular Dynamics Study on<br />

the Oxygen Diffusion in Doped Fluorites: The Effect <strong>of</strong> the Dopant Distribution,<br />

Fuel Cells, 11, pp. 26-37, 2011<br />

[21] Chiu W.K.S., Joshi A.S., Grew K.N., Lattice Boltzmann Model <strong>for</strong> Multi-<br />

Component Mass Transfer in a Solid Oxide Fuel Cell Anode with Heterogeneous<br />

Internal Re<strong>for</strong>mation and Chemistry, Eur. Phys. J. Special Topics, 171, pp. 159-165,<br />

2009<br />

[22] Sohn S., Nam J.H., Jeon D.H., Kim C., A Micro/Macroscale Model <strong>for</strong> Intermediate<br />

Temperature Solid Oxide Fuel Cells with Prescribed Fully-Developed Axial Velocity<br />

Pr<strong>of</strong>iles in Gas Channels, Int. J. Hydrogen Energy, 35, pp. 11890-11907, 2010<br />

[23] Joshi A.S., Peracchio A.A., Grew K.N., Chiu W.K.S., Lattice Boltzmann Method <strong>for</strong><br />

Multi-Component Mass Diffusion in Complex 2D Geometries, J. Phys. D: Appl.<br />

Phys., 40, pp. 2961-2971, 2007<br />

[24] Joshi A.S., Grew K.N., Izzo J.R. Jr., Peracchio A.A., Chiu W.K.S., Lattice<br />

Boltzmann Modeling <strong>of</strong> Three-Dimensional, Multicomponent Mass Diffusion in a<br />

Solid Oxide Fuel Cell Anode, ASME J. Fuel Cell Sci. Tech., 7, pp. 1-8, 2010<br />

54


[25] Latt J., Hydrodynamic Limit <strong>of</strong> the Lattice Boltzmann Equations, PhD <strong>Thesis</strong>,<br />

University <strong>of</strong> Geneva, 2007<br />

[26] Haslam I.W., Crouch R.S., Seaïd M., Coupled Finite Element-Lattice Boltzmann<br />

Analysis, Comput. Methods Appl. Mech. Eng.., 197, pp. 4505-4511, 2008<br />

[27] Suzue Y., Shikazono N., Kasagi N., Micro Modeling <strong>of</strong> Solid Oxide Fuel Cell Anode<br />

Based on Stochastic Reconstruction, J. Power Sources, 184, pp. 52-59, 2008<br />

[28] Bernsdorf J., Simulation <strong>of</strong> Complex Flows and Multi-Physics with the Lattice<br />

Boltzmann Method, PhD <strong>Thesis</strong>, University <strong>of</strong> Amsterdam, 2008<br />

[29] Kang Q. J., Lichtner P. C., Janecky D.R., Lattice Boltzmann Method <strong>for</strong> Reacting<br />

Flows in Porous Media, Adv. Appl. Math. Mech., 5, pp. 545-563, 2010<br />

[30] Servan Camas B., Lattice Boltzmann Modeling <strong>for</strong> Mass Transport Equations in<br />

Porous Media, PhD <strong>Thesis</strong>, Dept. Civil & Environmental <strong>Engineering</strong>, Louisiana<br />

State University, USA, URN etd-11072008-084923, 2008<br />

[31] Kim S.D., Lee J.J., Moon H., Hyun S.H., Moon J., Kim J., Lee H.W., Effects <strong>of</strong><br />

Anode and Electrolyte Microstructures on Per<strong>for</strong>mance <strong>of</strong> Solid Oxide Fuel Cells, J.<br />

Power Sources, 169, pp. 265-270, 2007<br />

[32] Asinari P., Calí Quaglia M., von Spakovsky M.R., Kasula B.V., Direct Numerical<br />

Calculation <strong>of</strong> the Kinematic Tortuosity <strong>of</strong> Reactive Mixture Flow in the Anode<br />

Layer <strong>of</strong> Solid Oxide Fuel Cells by the Lattice Boltzmann Method, J. Power Sources,<br />

170, pp. 359-375, 2007<br />

[33] Nam J. H., Jeon D. H., A Comprehensive Micro-scale Model <strong>for</strong> Transport and<br />

Reaction in Intermediate Temperature Solid Oxide Fuel Cells, Electrochim. Acta, 51<br />

pp. 3446-3460, 2006<br />

[34] Janardhanan V. M., Deutschmann O., Numerical Study <strong>of</strong> Mass and Heat Transport<br />

in Solid-Oxide Fuel Cells Running on Humidified Methane, Chem. Eng. Sci., 62, pp.<br />

5473-5486, 2007<br />

[35] Gorte R.J., Vohs J.M., Nanostructured Anodes <strong>for</strong> Solid Oxide Fuel Cells, Current<br />

Opinion in Colloid & Interface Science, 14, pp. 236–244, 2009<br />

[36] Sadykov V.A., Mezentseva N.V., Bunina R.V., Alikina G.M., Lukashevich A.I.,<br />

Kharlamova T.S, Rogov V.A, Zaikovskii V.I.,. Ishchenko A.V, Krieger T.A.,<br />

Bobrenok O.F., Smirnova A., Irvine J., Vasylyev O.D., Effect <strong>of</strong> Complex Oxide<br />

Promoters and Pd on Activity and Stability <strong>of</strong> Ni/YSZ (ScSZ) Cermets as Anode<br />

Materials <strong>for</strong> IT SOFC, Catalysis Today, 131, pp. 226-237, 2008<br />

[37] Yuan J., Faghri M., Sundén B., On Heat and Mass Transfer Phenomena in PEMFC<br />

and SOFC and Modeling Approaches, Chapter 4 in Sundén B., Faghri M.(eds.),<br />

Transport Phenomena in Fuel Cells, WIT Press, UK, 2005<br />

[38] Yakabe H., Hishinuma M., Uratani M., Matsuzaki Y., Yasuda I., Evaluation and<br />

Modeling <strong>of</strong> Per<strong>for</strong>mance <strong>of</strong> Anode-Supported Solid Oxide Fuel Cell, J. Power<br />

Sources, 86, pp. 423-431, 2000<br />

55


[39] Hussain M. M., Li. X., Dincer I., Mathematical Modeling <strong>of</strong> Transport Phenomena in<br />

Porous SOFC Anodes, Int. J. Thermal Sciences, 46, pp. 48-86, 2007<br />

[40] Reid R.C., Prausnitz J.M., Poling B.E., The Properties <strong>of</strong> Gases and Liquids, 4th ed.,<br />

McGraw Hill, USA ISBN 0-07-051799-1, 1987<br />

[41] Nagel F.P., Schildhauer T.J., Biollaz S.M.A., Stucki S., Charge, Mass and Heat<br />

Transfer Interactions in Solid Oxide Fuel Cells Operated with Different Fuel Gases-<br />

A Sensitivity Analysis, J. Power Sources, 184, pp. 129-142, 2008<br />

[42] Achenbach E., Riensche E., Methane/Steam Re<strong>for</strong>ming Kinetics <strong>for</strong> Solid Oxide<br />

Fuel Cells, J. Power Sources, 52, pp. 283-288, 1994<br />

[43] Achenbach E., Three-Dimensional and Time-Dependent Simulation <strong>of</strong> a Planar Solid<br />

Oxide Fuel Cell Stack, J. Power Sources, 49, pp. 333-348, 1994<br />

[44] Leinfelder R., Reaktionskinetische Untersuchungen zur Methan-Dampf-<br />

Re<strong>for</strong>mierung und Shift-Reaktion an Anoden Oxidkeramischer Brennst<strong>of</strong>fzellen”,<br />

Doctoral <strong>Thesis</strong>, Der Technischen Fakultät, Universität Erlangen-Nürnberg,<br />

Erlangen-Nürnberg, Deutschland, 2004 (in German)<br />

[45] Drescher I., Kinetik der Methan-Dampf-Refomierung, Institut für Werkst<strong>of</strong>fe und<br />

Verfahren der Energietechnik, Forschungszentrums Jülich, Jülich, Deutschland,<br />

ISSN 0944-2952, 1999 (in German)<br />

[46] Klein J.M., Bultel Y., Georges S., Pons M., Modeling <strong>of</strong> a SOFC Fuelled by<br />

Methane: From Direct Re<strong>for</strong>ming to Gradual Internal Re<strong>for</strong>ming, Chem. Eng. Sci.,<br />

62, pp. 1636-1649, 2007<br />

[47] Aguiar P., Adjiman C.S., Brandon N.P., Anode-Supported Intermediate Temperature<br />

Direct Internal Re<strong>for</strong>ming Solid Oxide Fuel Cell I: Model-Based Steady-State<br />

Per<strong>for</strong>mance, J. Power Sources, 132, pp. 113-126, 2004<br />

[48] Haberman B.A., Young J.B., Three-Dimensional Simulation <strong>of</strong> Chemically Reacting<br />

Gas Integrated-Planar Solid Oxide Fuel Cell”, Int. J. Heat and Mass Transfer, 47, pp.<br />

3617-3629, 2004<br />

[49] Mears D.E., Tests <strong>for</strong> Transport Limitations in Experimental Catalytic Reactors, Int.<br />

Eng. Chem. Process Des. Develop., 10, pp. 541-547, 1971<br />

[50] Mears D.E., On Criteria <strong>for</strong> Axial Dispersion in Nonisothermal Packed-Bed Catalytic<br />

Reactors, Ind. Eng. Chem., Fundam., 15, pp. 20-23, 1976<br />

[51] Nso<strong>for</strong> E.C., Adebiyi G.A., Measurement <strong>of</strong> the Gas-Particle Heat Transfer<br />

Coefficient in a Packed Bed <strong>for</strong> High-Temperature Energy Storage, Exp. Thermal<br />

and Fluid Sci., 24, pp. 1-9, 2001<br />

[52] Bizzi M., Basini L., Saracco G., Specchia V. Short Contact Time Catalytic Partial<br />

Oxidation <strong>of</strong> Methane: Analysis <strong>of</strong> Transport Phenomena Effects, Chem. Eng. J., 90,<br />

pp. 97-106, 2002<br />

56

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!